You are on page 1of 505

David Evans Walter

Heather C. Proctor

Mites: Ecology,
Evolution &
Behaviour
Life at a Microscale
Second Edition
Mites: Ecology, Evolution & Behaviour
David Evans Walter • Heather C. Proctor

Mites: Ecology, Evolution


& Behaviour
Life at a Microscale

Second Edition
David Evans Walter Heather C. Proctor
Invertebrate Zoology Biological Sciences
University of the Sunshine Coast University of Alberta
Royal Alberta Museum Edmonton, AB, Canada
Edmonton, AB, Canada

ISBN 978-94-007-7163-5 ISBN 978-94-007-7164-2 (eBook)


DOI 10.1007/978-94-007-7164-2
Springer Dordrecht Heidelberg New York London

Library of Congress Control Number: 2013948357

© Springer Science+Business Media Dordrecht 2013


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of
the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation,
broadcasting, reproduction on microfilms or in any other physical way, and transmission or information
storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology
now known or hereafter developed. Exempted from this legal reservation are brief excerpts in connection
with reviews or scholarly analysis or material supplied specifically for the purpose of being entered and
executed on a computer system, for exclusive use by the purchaser of the work. Duplication of this
publication or parts thereof is permitted only under the provisions of the Copyright Law of the Publisher’s
location, in its current version, and permission for use must always be obtained from Springer.
Permissions for use may be obtained through RightsLink at the Copyright Clearance Center. Violations
are liable to prosecution under the respective Copyright Law.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
While the advice and information in this book are believed to be true and accurate at the date of
publication, neither the authors nor the editors nor the publisher can accept any legal responsibility for
any errors or omissions that may be made. The publisher makes no warranty, express or implied, with
respect to the material contained herein.

Printed on acid-free paper

Springer is part of Springer Science+Business Media (www.springer.com)


Preface to the Second Edition

Most mites are minuscule, barely perceptible to even the sharpest eyes, but size is a
poor guide to importance. When a crop disappears under a blanket of silk or stored
food turns into a seething mass of hairy motes, then the numbers of mites will be
legion. If the mite is an itch mite burrowing in one’s skin or a tick or chigger inject-
ing virulent rickettsiae into one’s blood, then a single mite may be enough to affect
one’s life dramatically.
Some akari were known to the more perceptive of ancient Greeks (Kevan 1985)
and some of their effects were understood long ago (e.g. the cause of scabies has
been known since 1689), but few mites caused comment until the late 1700s (Prasad
1982). At that time, fewer than 100 species of ‘ticks’, stored-product pests and
brightly coloured plant and water mites had been discussed in the literature before
Linnaeus (Baker and Wharton 1952), and Linnaeus himself included only 29 spe-
cies of Acarus in the tenth edition of his Systema Naturae. Not until the late 1800s
was the term ‘acarologist’ coined to describe those who study mites, and it was
another 50 years before the word appeared in dictionaries (Krantz 1996).
Advances in microscopy contributed to the birth of this new science, but even
more important was the industry and imagination of acarologists such as G. Canestrini,
A.D. Michael, A. Nalepa, E. Trouessart and, in particular, Antonio Berlese in the
late 1800s. It is largely through Berlese’s work that scientists first began to appreciate
the ubiquity and diversity of the Acari – and how intricately they are linked to every
aspect of terrestrial life.
Acarology prospered during its first century, nurtured by the excitement of dis-
covery and by medical, veterinary and agricultural needs. In the last few decades,
however, like other important disciplines that require an understanding of whole
organisms (such as entomology and nematology), acarology has been eclipsed by a
fascination with technology, subcellular processes and genomes. As a result, no
matter how well versed they may be in cellular biology and genetics, many of
today’s students are not trained to deal with living organisms or the problems those
organisms can cause.
The repercussions of this ignorance are already being felt, especially in the pro-
liferation of pests, the spread of disease vectors and the alarming incidence of

v
vi Preface to the Second Edition

imaginary mite infestations (Delusional Parasitosis) in the general public. Still, not
all is gloomy. The transformation of molecular biology from a separate field of
study to a tool that is becoming increasingly integrated into ecology, systematics
and other fields of evolutionary biology is a promising change. For example, there
is increasingly strong evidence from molecular studies that the Acari are diphyletic
and that ‘miteness’ has evolved independently in two relatively distantly related
lineages of arachnids. As well, molecular techniques are increasingly useful in
determining species boundaries, population structuring, feeding biology and impor-
tant life history characteristics from reproductive biology to virulence. New com-
puter technology is supporting a renaissance in interactive identification tools. The
Internet has transformed our ability to retrieve and exchange information with col-
leagues around the world and to make taxonomic tools available to an international
audience. As retirement and death continue to thin the ranks of trained acarologists,
we take hope in these innovations and in the burgeoning number of young acarolo-
gists outside of North America and Europe and the continuing role of the OSU
Acarology Summer Program in training biologists in mite taxonomy and contribut-
ing to the global fellowship of acarologists.
Our goal in revising this book is to attract new students to acarology by showing
them that mites do interesting and exciting things. To that end, we have doubled the
number of illustrations and poured over more than 1,600 scientific publications look-
ing for examples of why mites are interesting to study. We highlight the roles that
mites have played in the development of important theoretical concepts in ecology
and evolution (e.g. local mate competition, prey refugia, multi-level selection and
tritrophic level interactions) and also emphasize that, as Dana Wrensch has said,
‘mites do things they shouldn’t’. In many cases, the lives of mites clearly demon-
strate that currently accepted theory is flawed. We also reveal the depth and breadth
to which mites have permeated the world – from sequoias to seamounts to sea lions –
but most of all, we strive to transmit the wonder and affection that we feel for these
most miniature of arthropods, with the hope that our acarophilia will infect others.
Quite a few biologists with at least mild cases of this affliction have contributed
valuable suggestions after reading drafts of chapters, when reviewing earlier works
or during discussions (semi-sober and otherwise). We think it safe to say that none
of these friends and colleagues would agree with everything we say in this book, but
all have contributed to its more lucid and informative sections. We would like to
thank everyone who has provided help and support, and though it is easy to forget
one’s debts, these names spring immediately to mind: Jenny Beard, Fred Beaulieu,
Tyler Cobb, Valerie Behan-Pelletier, Andre Bochkov, Kaylee Byers, John Clark,
Matt Colloff, Rob Colwell, John Cooke, Dac Crossley, Ashley Dowling,
Doug Craig, Jacek Dabert, Jason Dunlop, Qinghai Fan, Farid Faraji, Norm Fashing,
Ray Fisher, Gordon Gordh, Bruce Halliday, Mark Harvey, Marjorie Hoy, Mark
Judson, Jim Kierans, Hans Klompen, Wayne Knee, Jerry Krantz, Evert Lindquist,
Zoë Lindo, Rogelio Macías-Ordóñez, Serge Mironov, Masha Minor, Jim McMurtry,
Helen Nahrung, Jeffrey Newton, Roy Norton, Barry OConnor, Norm Platnick, the
late Frank Radovsky, Susanne Randolf, Heinz Schatz, Owen Seeman, Matthew
Shaw, Bruce Smith, Ian Smith, Diana Wall, Gerd Weigmann, Jay Yoder and
Preface to the Second Edition vii

Zhi-Qiang Zhang. We would also like to thank Michael Caldwell for allowing us to
use his striking drawing of a rampant eurypterid and our students Catherine (née
Bryant) Harvey and Juanita Choo for their help with many of the line drawings.
Anthony O’Toole, Caroline Meacham and George Braybrook enhanced our skills at
the scanning electron microscope. We would also like to thank Zuzana Bernhart at
Springer for her editorial support. Finally, the senior author would like to express
his appreciation to the entire Nahrung family of Miva, Queensland, for their kind-
ness in allowing him to spend weeks immersed in solitude and an extraordinary
natural diversity while a large part of this second edition was taking form.
Additionally, our knowledge and appreciation of mites has grown in direct pro-
portion to our ability to study these fascinating animals. We have funded some of
our studies out-of-pocket, but that largesse has never been sufficient to tackle more
than interesting sidelines. Fortunately, mites are major components of all terrestrial
ecosystems and there are few research projects that cannot be used to learn some-
thing exciting about mites. The senior author is especially appreciative of the sup-
port he received from the National Science Foundation (Graduate Research
Scholarship and Postdoctoral Fellowship), the USDA Agricultural Research Service
(Postdoctoral Fellowship), the Australian Research Council (Postdoctoral
Fellowship, Large and Small Grant Schemes), the Australian Biological Resources
Survey, the Department of Entomology at the University of Queensland, the Alberta
Biodiversity Monitoring Institute and the Royal Alberta Museum. The junior author
wishes to thank the Natural Sciences and Engineering Research Council of Canada,
the Department of Biology at Queen’s University, the Australian Research Council’s
Small Grant Scheme, the Australian School of Environmental Studies at Griffith
University and the Department of Biological Sciences at the University of Alberta
for their support.

David Evans Walter


Heather C. Proctor
Contents

1 What Good Are Mites? ........................................................................... 1


What Is a Mite? ......................................................................................... 1
Why Study Mites?..................................................................................... 5
What Follows? .......................................................................................... 7
References ................................................................................................. 9
2 The Origin of Mites: Fossil History and Relationships ....................... 11
The Cambrian Explosion and the Rise of the Arthropoda ........................ 11
The First Major Dichotomy: Mandibulata Versus Chelicerata ................. 15
A Review of Arthropod Limb Structure, Metamerism and Tagmosis ...... 18
Marine Euchelicerates ............................................................................... 20
Scorpionida: The First Arachnids?............................................................ 21
The Origin of the Arachnids: A Palaeofantasy.......................................... 22
Arachnids and the Colonisation of Land .................................................. 25
Fossil Mites ............................................................................................... 26
Fossil Acariformes ............................................................................... 27
Fossil Parasitiformes ............................................................................ 27
Potential Arachnid Relatives of Mites ...................................................... 28
Palpigradi ............................................................................................. 29
Opiliones .............................................................................................. 31
Ricinulei ............................................................................................... 32
Pseudoscorpionida................................................................................ 33
Solifugida ............................................................................................. 34
Summary and Preview .............................................................................. 34
References ................................................................................................. 35
3 Systematic and Morphological Survey.................................................. 39
What Is ‘Acari’? The Question of Mite Monophyly ................................. 39
Parasitiformes: Ticks and Their Relatives................................................. 40
Acariformes: The Mite-Like Mites ........................................................... 50

ix
x Contents

How Do Mites Do the Things They Do? .................................................. 55


Sensing, Feeding, Silk and Sex: The Gnathosoma............................... 55
Moving, Sensing and Interacting: The Legs ........................................ 59
Reproduction ........................................................................................ 60
Digestion and Excretion ....................................................................... 61
Keeping It All In: The Cuticle .............................................................. 62
Identifying Mite Superorders and Orders ................................................. 62
Key to the Superorders and Orders of the Acari .................................. 64
Summary ................................................................................................... 66
References ................................................................................................. 66
4 Life Cycles, Development and Size ........................................................ 69
Oviposition................................................................................................ 70
Parental Care ............................................................................................. 70
Egg Number and Egg Size ........................................................................ 72
Postembryonic Development .................................................................... 74
Prelarva and Larva .................................................................................... 76
Suppression and Skipping of Stages ......................................................... 80
Life Cycle of the Parasitengona ................................................................ 82
Paedomorphosis, Progenesis and Neoteny................................................ 84
Size, Developmental Rate and Generation Times ..................................... 86
Overview of Mite Size Patterns ................................................................ 86
Developmental Rates and Generation Times ............................................ 88
Dissociation Between Body Size and Developmental
Rate in Mesostigmata................................................................................ 91
Dispersal, Migration and Phoresy ............................................................. 92
Migratory Stages .................................................................................. 93
Phoresy ................................................................................................. 94
Summary ................................................................................................... 97
References ................................................................................................. 98
5 Sex and Celibacy ..................................................................................... 105
Modes of Sperm Transfer.......................................................................... 106
Distribution of Sperm-Transfer Modes Among Non-Acarine Animals .... 107
Diversity of Sperm-Transfer Behaviours in Mites .................................... 109
Reproductive Anatomy ......................................................................... 109
The Parasitiformes: Elaborations on a Theme ..................................... 111
The Adventurous Acariformes ............................................................. 116
Spermatophore Structure and Function................................................ 124
Exploding Sperm Packets..................................................................... 125
Fields of Fragrant Spermatophores ...................................................... 126
Sexual Selection ........................................................................................ 128
Intrasexual Competition: Male Modifications for Mate
Monopolisation .................................................................................... 128
Intersexual Selection as an Agent of Morphological
and Behavioural Change ...................................................................... 138
Contents xi

Parthenogenesis......................................................................................... 142
Why Have Sex? .................................................................................... 142
Distribution of Parthenogenesis in Mites ............................................. 144
Sex-Ratio Manipulation ............................................................................ 146
Immaculate Conception: Did Sexual Astigmatans Arise
from Asexual Oribatids? ........................................................................... 149
Summary ................................................................................................... 150
References ................................................................................................. 151
6 Mites in Soil and Litter Systems ............................................................ 161
The Enigma of Soil Biodiversity .............................................................. 161
What Is Soil? ............................................................................................. 164
Forest Floor Habitats ............................................................................ 164
Ephemeral Versus Stable Soil-Litter Habitats ...................................... 166
Mites, the Rhizosphere and Mycorrhizae............................................. 166
How Deep Is Soil?................................................................................ 168
Mites and Decomposition .................................................................... 169
Soil Mites in a Simple System: Antarctica ............................................... 170
Antarctic Mites ..................................................................................... 171
An Antarctic Food Web ........................................................................ 172
Feeding Guilds and Functional Groups .................................................... 177
Comminuting Microbivore–Detritivores: Grazers and Browsers ........ 178
Piercing-Sucking Microbivores............................................................ 181
Filter-Feeding Microbivores................................................................. 182
Direct Plant Parasites ........................................................................... 182
Mites and Moss .................................................................................... 184
Indirect Plant Parasites ......................................................................... 185
The Worm-Eaters: Nematophages........................................................ 185
Predators of Arthropods ....................................................................... 187
Predation in the Soil .................................................................................. 190
Cruise and Pursuit Predators ................................................................ 190
Ambush or Sit-and-Wait Predators ...................................................... 191
Saltatory Search ................................................................................... 192
Constraints and Variations .................................................................... 193
Intraguild Predation.............................................................................. 194
Cannibalism.......................................................................................... 196
Avoiding Predation: Defences of Mites and Mite Prey ............................ 197
Jumping ................................................................................................ 198
Chemical Defence ................................................................................ 198
Autotomy, Armour, Hairs, Dirt and Thanatosis ................................... 199
Acarophagy: Mites as Food for Larger Animals ....................................... 202
Eating Armoured Mites ........................................................................ 202
Vertebrates That Eat Mites ................................................................... 204
Poison Frogs and Cleptotoxins............................................................. 206
Body Size Patterns .................................................................................... 206
xii Contents

Sensitivity and Diversity: Soil Mites as Environmental Indicators .......... 208


Mites and Earthworms ......................................................................... 210
Palaeoacacrology.................................................................................. 211
Summary ................................................................................................... 211
References ................................................................................................. 212
7 Acari Underwater, or, Why Did Mites Take the Plunge? .................... 229
Taxonomic Distribution of Secondarily Aquatic Arthropods.................... 229
Repeated Invasions of Water ..................................................................... 232
Parasitiformes ....................................................................................... 232
Oribatida............................................................................................... 232
Astigmata ............................................................................................. 234
Prostigmata........................................................................................... 234
Number of Invasions into Different Aquatic Habitats .............................. 235
Phytotelmata......................................................................................... 235
Temporary Freshwater Bodies ............................................................. 236
Standing Fresh Water ........................................................................... 236
Running Fresh Water ............................................................................ 237
Interstitial Fresh Water ......................................................................... 238
Brackish Water ..................................................................................... 239
Marine Intertidal Zone ......................................................................... 239
Marine Subtidal Zone (Including Abyssal) .......................................... 239
(Pre)Adaptations to Subaquatic Life ......................................................... 240
Gas Exchange ....................................................................................... 240
Feeding ................................................................................................. 243
Osmoregulation .................................................................................... 254
Sperm Transfer ..................................................................................... 256
Predation: The Correlation Between Foul Taste and Bright Colour ......... 257
Locomotion ............................................................................................... 262
Swimming ............................................................................................ 262
Levitation .................................................................................................. 265
Sensitivity and Diversity: Water Mites as Environmental Indicators........ 266
Temperature.......................................................................................... 267
Depth .................................................................................................... 267
Substrate ............................................................................................... 268
Standing Versus Running Water ........................................................... 269
pH ......................................................................................................... 270
Organic Pollution ................................................................................. 270
Summary ................................................................................................... 271
References ................................................................................................. 272
8 Mites on Plants ........................................................................................ 281
Mites on Plants: Where Do They Come From? ........................................ 282
Plant Parasites ........................................................................................... 284
Rust, Gall and Erinose Mites: Eriophyoidea ........................................ 284
Earth Mites: Penthaleidae and Its Kin.................................................. 287
Contents xiii

Spider Mites and Their Kin .................................................................. 289


Duckweeed and Water Hyacinth Mites ................................................ 293
Fruit and Fig Mites ............................................................................... 294
Venereal Diseases of Plants.................................................................. 295
Hunting on Leaves .................................................................................... 297
Predatory Prostigmata .......................................................................... 298
Foliar Mesostigmata ............................................................................. 300
Development and Reproduction of Phytoseiid Mites .......................... 300
Feeding Biology of Phytoseiid Mites................................................... 301
Mites and Leaf Domatia............................................................................ 304
Structure and Distribution of Leaf Domatia......................................... 305
What Lives in Leaf Domatia? .............................................................. 307
Domatia as a Constitutive Plant Defence ............................................. 310
What’s in It for the Mites? ................................................................... 312
Arboreal Scavengers and Fungivores........................................................ 313
Scavenging on Leaves .......................................................................... 313
Moss and Lichen Mites ........................................................................ 317
Fungal Sporocarps................................................................................ 318
Under Bark ........................................................................................... 320
Mites and Biological Control .................................................................... 320
Infochemicals ....................................................................................... 321
Induced Resistance ............................................................................... 321
Transgenic Mites .................................................................................. 322
Biocontrol of Weeds ............................................................................. 323
Summary ................................................................................................... 324
References ................................................................................................. 325
9 Animals as Habitats ................................................................................ 341
Types of Ecological Interactions ............................................................... 342
Evolutionary Pathways Between Interactions ...................................... 345
Life with Invertebrates .............................................................................. 352
Taxonomic Survey of Associates ......................................................... 352
Phoresy and Dispersal .......................................................................... 355
Commensalism ..................................................................................... 360
Parasitism and Parasitoidism................................................................ 361
Mutualism ............................................................................................ 368
Acarinaria ............................................................................................. 370
Life with Vertebrates ................................................................................. 373
Mammals and Their Homes ................................................................. 373
Mites on, in and Around Birds ............................................................. 379
Fish, Amphibians, Reptiles and the Mystery of Mite Pockets ............. 387
Effects of Parasitic Mites on Their Hosts ................................................. 389
Differential Host Susceptibility to Parasitism ...................................... 389
The Evil That Mites Do: Adverse Effects of Acarine Symbionts ........ 393
Parasitic Mites and Mate Choice by Hosts .......................................... 398
xiv Contents

Mite–Host Coevolution: Any Evidence? .................................................. 401


Coevolution by Mutual Adaptation ...................................................... 402
Cospeciation ......................................................................................... 403
Summary ................................................................................................... 408
References ................................................................................................. 409
10 Mites That Cause and Transmit Disease ............................................... 423
Critical Concepts and Terminology .......................................................... 424
Mite-Caused Diseases ............................................................................... 427
The Human Itch Mite: A Life in the Skin ............................................ 428
Demented Dermanyssoidea: Biting Mites of Birds, Rodents,
and Whatever Else Is Nearby ............................................................... 429
Perverse Prostigmata: Whirligigs, Straw Itch,
and Walking Dandruff .......................................................................... 431
Mite- and Tick-Borne Diseases................................................................. 434
Trombiculoidea (Chiggers): Scrub Typhus .......................................... 434
Ixodoidea (Ticks) ................................................................................. 435
Diseases That Mites Do Not Cause........................................................... 437
Mystery Bites ....................................................................................... 439
Delusions of Mite-Bites ....................................................................... 440
Summary ................................................................................................... 442
References ................................................................................................. 443
11 Mites and Biological Diversity ............................................................... 447
Mites and Microhabitats ........................................................................... 448
Mites and Complementarity...................................................................... 450
Size and Biodiversity ................................................................................ 452
Host Specificity, Size and Diversity .......................................................... 454
Summary ................................................................................................... 457
References ................................................................................................. 458
12 Mites as Models ....................................................................................... 461
Theoretical and Applied Population Ecology ........................................... 462
Microcosms ............................................................................................... 462
Moss Islands.............................................................................................. 463
Biomonitoring ........................................................................................... 464
Transgenic Releases .................................................................................. 464
The Evolution of Host Specificity and Virulence...................................... 465
Sexual Selection and Diversification ........................................................ 465
Sex-Ratio Control and the Devolution of Sex........................................... 466
Pushing the Limits of Physiology and Morphology ................................. 467
Selection at More Than One Level ........................................................... 467
Summary ................................................................................................... 468
References ................................................................................................. 468

Index ................................................................................................................. 471


Chapter 1
What Good Are Mites?

Man is certainly crazy. He could not make a mite, yet he makes


gods by the dozen.
(Michel de Montaigne, 1580)
Nature uses only the longest threads to weave her patterns, so
that each small piece of her fabric reveals the organization of
the entire tapestry.
(Richard P. Feynman, The Character of Physical Law, 1965)
A mite makes the seas roar.
(Richard P. Feynman, The Value of Science, address to the
National Academy of Sciences, Autumn 1955)

Imagine animals so small, that even if you could see them, you would assume they
must be insignificant to your life. Then imagine you could see them up close, watch
what they do, and learn that, rather than being insignificant, these tiny mites were
part of the very fabric of Nature. Moreover, imagine when magnified into visibility,
mites were bizarrely beautiful, did incredibly interesting things, and were critical
components of ecosystem function, agricultural production and human health.
Sound farfetched? Well, once we too thought so, but now we know better. The pur-
pose of this book is to share with you what we have learned about the unexpectedly
fascinating world of mites.

What Is a Mite?

The word ‘mite’ (Figs. 1.1 and 1.2) comes to us from Old English and means a very
small creature. Compared to the spider, or even the scorpion, the mite’s contribution to
idiomatic English is minute. Bartlett’s Familiar Quotations (1992) has 17 bons mots
about spiders and three about scorpions but (except for the quotation above from

D.E. Walter and H.C. Proctor, Mites: Ecology, Evolution & Behaviour: Life at a Microscale, 1
DOI 10.1007/978-94-007-7164-2_1, © Springer Science+Business Media Dordrecht 2013
2 1 What Good Are Mites?

Fig. 1.1 An acariform mite


(Acariformes,
Trombidiformes,
Dasythyreidae,
Xanthodasythyreus toohey)
(SEM by DE Walter)

Fig. 1.2 A parasitform mite


in lateral view
(Parasitiformes,
Mesostigmata, Zerconidae).
This mite is about a half
millimetre long from one end
to another, about the median
body length for a mite (SEM
by DE Walter)

Michel de Montaigne in French) manages merely one mention of a mite in English.


A search of the World Wide Web today for quotable mites is hardly more productive
(even the list of quotes from the DC Comics’ character Bat-Mite is miniscule).
This spot on the language is mirrored in the more familiar meaning of mite – a
very small quantity. The ‘widow’s mite’ from the King James’ version of The
Gospel According to Saint Mark 12:42 was actually borrowed in translation from a
Flemish coin worth a third of a penny – a usage that captures the essence of being a
mite. Above all, mites are very small. Most are less, usually much less, than a mil-
limetre in length, and none grow larger than a few centimetres long. Among the
larger mites are those that are best known – the blood-sucking fiends we call ticks,
again a word derived from Old English.
What Is a Mite? 3

Fig. 1.3 Mites are members


of the Chelicerata along with
scorpions, pseudoscorpions,
spiders etc., but are
distinguished from their
relatives by their body
organisation, diversity of
ecologies, and generally
very small size. SEMs
by DE Walter

Although tiny themselves, mites belong to the largest and most impressive
lineage of animals, the arthropods. Within the Arthropoda are two ancient lineages,
the Mandibulata, which includes crustaceans, myriapods and insects, and the
Chelicerata (Regier et al. 2010). Mites are the most diverse and successful of the
chelicerates (Fig. 1.3).
All arthropods are composed of serially repeated segments that each originally
bore a pair of jointed limbs (arthropod means ‘jointed foot’), but the chelicerate
body plan is profoundly different from that of mandibulate arthropods. Perhaps the
most striking difference is that chelicerates do not have a head. Instead, an anterior
body region, the prosoma (‘front body’), combines the functions of sensing, think-
ing, feeding, and locomotion. Antennae, the distinctive sensory limbs of mandibu-
late arthropods, are absent, as are the mandibles that give the Mandibulata its name.
Instead, the genes that would regulate the production of the first pair of antennae in
a mandibulate produce a pair of often pincer-like mouthparts, the chelicerae
(Thomas and Telford 1999). The second pair of limbs are sensing/feeding append-
ages called pedipalps, which can be pincer-like, leg-like, or even antenniform. The
next four pairs of limbs form the walking legs. Digestion, gamete production and
gas exchange largely occur in the legless posterior body region or opisthosoma
(‘hind body’). Although all other arachnids share this basic two-region chelicerate
body plan, mites have their own unique idiosoma (‘peculiar body’) composed of all
of the segments behind the pedipalps and the special feeding region called the gna-
thosoma (‘jaw body’) or capitulum (‘little head’) (Figs. 1.4 and 1.5).
Many chelicerates are familiar inhabitants of our homes, gardens and parks.
Spiders are ubiquitous – weaving, scurrying, and jumping throughout our lives to the
consternation of some, but the enjoyment of others. Daddy longlegs, scorpions, vine-
garoons, sun-spiders, whip-scorpions and horseshoe crabs, are often large and look
menacing enough to cause comment and attract common names. Ticks and their
disease-causing microbes have their own special place in our list of unpleasant things.
Spider mites attack our crops and houseplants and the odd red velvet mite scurrying
4 1 What Good Are Mites?

Fig. 1.4 Mites have only


two body regions an anterior
gnathosoma (= capitulum)
and a posterior idiosoma.
SEM of venter of male
Promegistus armstrongi
by DE Walter

Fig. 1.5 Mites and spiders are related, but have very different ground plans (Illustration by
DE Walter and C. Harvey)

along the sidewalk may momentarily attract attention. The remainder – most mites,
pseudoscorpions, palpigrades and ricinuleids – mostly scurry about in minute
obscurity or have left traces only as fossils (e.g. the extinct taxa Eurypterida,
Haptopoda, Phalangiotarbida, Trigonotarbida, Urananeida).
Why Study Mites? 5

Lurking at the margins of the Chelicerata is a bizarre marine group, the


Pycnogonida or sea spiders. Although they seem to possess the basic chelicerate
characters – chelicerae rather than antennae and a body divided into a prosoma and
opisthosoma – their many unique features, such as a proboscis, a strongly reduced
opisthosoma and extraneous limbs, leave some workers uncomfortable with their
placement in the Chelicerata (Arnaud and Bamber 1987).
The application of modern molecular phylogenetic techniques to the problem of
chelicerate phylogeny is a recent phenomenon and the relationships among the vari-
ous groups of chelicerates are still less clear than among the major mandibulate
groups (Regier et al. 2010). Traditionally, chelicerate arthropods that live on land
have been treated as members of the Class Arachnida. Living arachnids have been
variously classified in 10–20 major groups (orders), many of which are easily rec-
ognised. For example, the order Araneae includes the spiders and the order
Scorpionida the scorpions. However, mites are so diverse in form that in order to
accommodate them they require two superorders: Parasitiformes (including the
enigmatic Opilioacarida) and Acariformes (Krantz and Walter 2009). In our first
edition of this book we adopted the convention for referring to these groups that has
been long used for other zoological groups (e.g. crustacean, hymenopteran, dip-
teran). Thus, for the three groups of mites just mentioned, replacing their endings
with ‘-an’ (which means ‘of or belonging to’) results in acariformans, parasitifor-
mans, and opilioacarans. We also are happy to convert adjectives like ‘oribatid’ to
nouns where it will cause no confusion. Although it may make some traditionalists
cringe, uses such as ‘arachnids’ have a long tradition and in a book about mites, that
word will be repeated enough times as it is.
Speaking of arachnids, spiders share characters of general distribution within
the Arachnida, e.g. four pairs of legs, a prosoma and opisthosoma, and a carapace
(a shield covering the prosoma). But the best way to distinguished spiders is by the
characters unique to spiders: their fangs with poison glands, their opisthosomal
spinnerets that produce silken threads, a narrow waist (pedicle) and the bulbous
male palps used in transferring sperm. Characterising mites, however, is much more
difficult and requires an understanding of the basic morphology for each superorder
(Chap. 3). In this book, we continue the tradition of treating the two superorders of
mites as a subclass of the Arachnida called the Acari, or less concisely, the Acarina.
But as we will see later, being a mite may be more a grade of evolution than mem-
bership in a natural group (Fig. 1.6).

Why Study Mites?

In a world simultaneously fascinated by the wonders of biological diversity and


aghast at its accelerating destruction, it is typically mammals and birds that are the
icons of our loss. A case is sometimes made for saving a butterfly or fish or frog.
Rainforests and their plants are argued to be worth preserving for their potential as
pharmacological cornucopias. But what of the worms, slugs, bugs, maggots and
6 1 What Good Are Mites?

Fig. 1.6 Lateral view of a mygalomorph spider with a typical arachnid ground plan of prosoma
bearing six pairs of limbs and covered by a carapace and an opisthosoma. Special spider characters
include poison fangs, pedicle, and spinnerets (Illustration by DE Walter and C. Harvey)

mites that feed on and are the food for the rainforests’ larger inhabitants? Can these
scarcely seen symbionts be worth consideration?
A vertebrocentric view of animal diversity is rarely questioned, but it cannot
tolerate close scrutiny. Approximately 40,000 species of vertebrates, half of them
fish, are alive today (Wheeler 1990). Compare this paltry sum to the number of
named invertebrate species – more than a million! Thus when measured simply as
species richness, invertebrates make up 95 % of animal biodiversity. Some esti-
mates suggest that there may be as many as 30 million species of arthropods alone
(Erwin 1982). Insects are the most obvious of the invertebrate groups but other
minute animals, some of which can spend their entire lives on an individual insect,
are also rich in undescribed taxa. This contrasts with vertebrates, where the discov-
ery of a new species is a rare thing indeed. For example, in the decade from 1978 to
1987 only five new species of birds were described compared to more than 2,300
beetles (Groombridge 1992).
One might expect that with so little known about invertebrates most biologists
would be rushing to work on them. This is not the case. Despite their overwhelming
numerical superiority, disproportionately few systematists study invertebrates.
Consider the ratios of the number of named species of animals to the number of
taxonomists in the United States that study them. For fish and birds, this ratio is
about 30:1, while for mammals it is a lower 7:1 (calculated from Wheeler 1990).
For the group that is both most diverse and least described, the invertebrates, the
ratio is a staggering 307:1. The most diverse group within the invertebrates, insects,
has an even higher ratio of 425:1.
Vertebrate bias is not limited to systematists: it is also pervasive among ecolo-
gists, ethologists and evolutionary biologists. For the first edition of this book we
examined 15 issues of three major journals – Ecology, Animal Behaviour and
Evolution – and classified articles according to the taxa they discussed. An accurate
reflection of species diversity would be a ratio of about one article mentioning
What Follows? 7

vertebrates to every 24 that mentioned invertebrates. Yet, of the 1,221 articles


examined, 615 involved birds, fish, mammals, reptiles or amphibians, while only
410 included invertebrates of any kind – a ratio of 1.5 vertebrate:1 invertebrate. We
repeated this exercise for the current edition, again examining the abstracts of 15
recent (2011–2012) issues of each of the three same journals (1,163 articles in total).
The ratio is remarkable similar: 1.47 vertebrate:1 invertebrate paper. The distribu-
tion was uneven among journals, though, with Ecology and Evolution having a very
slight invertebrate bias (helped by Drosophila papers for Evolution), and Animal
Behaviour being strongly vertebrate oriented (2.13 vertebrate: 1 invertebrate).
Scientific bias is not spread evenly among the invertebrates. Gaudy or medically
important groups like butterflies and mosquitoes are relatively well studied, while
many other equally worthy taxa have been given short shrift. In this book we attempt
to redress the balance for one particularly species-rich and biologically fascinating
group of invertebrates – the mites.
Although not the equal of insects, mites do rank high in species richness. There
are over 50,000 named species of Acari and, when estimates of unnamed species are
included, this rises towards one million species (Barnes 1989; Groombridge 1992;
Walter and Proctor 1999). There are only about 50 mite systematists in the USA
(data from Welbourn 1992), that is one for every 1,000 described species! As the
insect:entomologist ratio is 425:1, this suggests that on a per species basis, mites are
less well known than insects. This is certainly true in Australia, where about 3,000
mite species have been recorded, which Halliday (1998) estimates as perhaps 3 %
of the total acarofauna. Even in snowy Canada, it is thought that there are records
for only 20 % of the estimated 10,000 species of mites (Lindquist et al. 1979).
Biodiversity does not simply refer to taxonomic diversity: it also includes the
range of behaviours and lifestyles exhibited by a taxon. Here again, mites are very
poorly represented in ecological, ethological and evolutionary studies. Our original
survey found that only nine of the 1,221 articles mentioned acarines, and only five
of these had mite biology as their primary focus. In our new survey, mites had an
even poorer showing: only a single paper of the 1,163 was focused on mites. This
scientific neglect may be partly the result of ignorance, for acarology is seldom
taught in universities and few modern texts discuss mite biology, but biologists
would do well to pay attention to these little arachnids, as much of their behaviour
and ecology puts vertebrate-based theory to the test. In this volume we present an
overview of mite behaviour, evolution and ecology. We hope that readers will be
both enlightened by past research and stimulated to pursue their own theoretical
interests using mites as models.

What Follows?

Mites are ubiquitous in every sort of aquatic, terrestrial, arboreal and parasitic habitat
but, being among the smallest of arthropods, even those inhabiting well-studied
systems are often overlooked. Ignoring mites, however, is a mistake. They are not
8 1 What Good Are Mites?

passive inhabitants of ecosystems; rather they are strong interactors, important


indicators of disturbance in both aquatic and terrestrial systems, and major compo-
nents of biological diversity (Chaps. 6, 7, 8, 9, 10, 11, and 12). Most taxonomic
effort has been spent on the groups that directly affect humans, our crops and live-
stock, or on those mites that are associated with vertebrates (Chaps 8, 10, and 11).
To mites, larger animals represent habitats and a given species of bird, beetle or
butterfly may host a multispecies community of a dozen or more mite species
(Chap. 9). Even a single feather from a bird may have three or four mite species
living in strictly circumscribed microhabitats and exhibiting as many different inter-
actions with their host (Chap. 10).
Plants, too, are mite habitats (Chap. 8). Plant-parasitic mites use stylet-like
mouthparts to stab individual plant cells and to suck out their contents, and some
species are able to modify the growth of host tissues to provide themselves with a
home. In agricultural systems, plant-parasitic mites are extremely damaging pests
with rapid generation times, high fecundity and a tendency to over-exploit their
hosts. Fortunately, acarine predators have followed their prey onto plants and many
are important biological control agents. Collections from amber and fossil leaf
packs indicate that mites have lived on trees at least since the Cretaceous. Today, an
incredible diversity and abundance of mites, including microherbivore-detritivores,
plant-parasites and predators, are suspended on vegetation as they scurry through
the canopies of rainforests like an arboreal plankton, many even drifting on air cur-
rents from tree to tree.
The original home of mites was probably decaying vegetation and soil, and that
is where one still finds mites in their most dazzling diversity (Chap. 6). A handful
of forest humus can literally contain mites from a hundred species, each earning
a living in a complex world that is usually beneath our notice but essential for our
survival. Mites are engineers of soil structure, indicators of the health of soil sys-
tems and major interactors with nematodes and microbes in decomposition. Any
bit of decomposing matter, from a penguin corpse in the Antarctic to a maggot-
riddled mushroom in a tropical rainforest, is likely to contain an unexpected diver-
sity of mites.
Although we think of mites as being quintessentially terrestrial, many acarine
lineages have successfully invaded aquatic habitats (Chap. 7). We are not speaking
merely of the surface-film dwellers or of the many species that dwell in the ooze of
rotting vegetation. Rather, we speak of those mites that have become true sub-
aquatic denizens of streams, lakes and seas. These tiny arthropods may be found in
the chilly runoff of melting glaciers, or the steaming margins of hot springs, in the
shallow water trapped by plants or the still and frigid depths of the oceanic abyss.
In order to understand the complexity of their interactions with other organisms,
it is first necessary to know something about mites. Chapter 3 gives a brief overview
of the major kinds of mites, shows what they look like, and provides an introduction
to their morphology, that is the structures mites use to interact with their world.
Chapter 4 explores the limits to which mites have pushed their physiology and size,
and shows clearly that theory based on large vertebrates often tells us little about life
at a smaller scale. Chapter 5, Sex and Celibacy, continues this theme in lurid detail
References 9

Fig. 1.7 The hairs on mites


and other arthropods are
called setae (singular, seta).
Most function as
mechanoreceptors – touch
sensors – but others are used
to smell and taste. Setae can
help protect a mite from
attack, e.g. the pincushion-
like array of long setae on the
mites at the beginning and
end of this chapter make it
difficult for a predator to
get at their body (SEM by
DE Walter)

and demonstrates both that mites make excellent models for investigating evolutionary
patterns in sexual selection and that sex is not as essential as we like to think. The
final chapter reviews what we think is our most important point – that mites are
excellent organisms for comparative and manipulative experimentation (Fig. 1.7).
But before all this, we need to explore the origins of mites (Chap. 2) and that means
starting with a bang.

References

Arnaud, F., & Bamber, R. N. (1987). The biology of Pycnogonida. Advances in Marine Biology,
24, 1–96.
Bartlett, J. (1992). In J. Kaplan (Ed.), Familiar quotations. A collection of passages, phrases, and
proverbs traced to their sources in ancient and modern literature (16th ed.). London: Little/
Brown & Company.
Barnes, R. D. (1989). Diversity of organisms: How much do we know? American Zoologist, 29,
1075–1084.
Erwin, T. (1982). Tropical forests: Their richness in Coleoptera and other arthropod species. The
Coleopterists Bulletin, 36, 74–75.
Groombridge, B. (Ed.). (1992). Global biodiversity, status of the earth’s living resources. London:
Chapman & Hall.
Halliday, R. B. (1998). Mites of Australia, a checklist and bibliography. Monographs on inverte-
brate taxonomy (Vol. 5). Collingwood: CSIRO Publishing.
Krantz, G. W., & Walter, D. E. (Eds.). (2009). A manual of acarology. (3rd ed.). Texas Tech
University Press, 807 p. 338 b/w illustrations, 60 figures, ISBN 978-0-89672-620-8. http://
www.ttup.ttu.edu/BookPages/9780896726208.html.
Lindquist, E. E., with contributions by Ainscough, B. D., Clulow, F. V., Funk, R. C., Marshall, V. G.,
Nesbitt, H. H. J., O Connor, B. M., Smith, I. M., & Wilkinson, P. R. (1979). Acari. In H. V. Danks (Ed.),
Canada and its insect fauna (Memoirs of the Entomological Society of Canada, Vol. 108,
pp. 252–263, 267–284).
10 1 What Good Are Mites?

Regier, J. C., Shultz, J. W., Zwick, A., Hussey, A., Ball, B., Wetzer, R., Martin, J. W., & Cunningham,
C. W. (2010). Arthropod relationships revealed by phylogenomic analysis of nuclear protein-
coding sequences. Nature, 463, 1079–1084.
Thomas, R. H., & Telford, M. J. (1999). An SEM study of appendage development in embryos of
the oribatid mite, Archegozetes longisetosus (Acari, Oribatei, Trhypochthoniidae). Acta
Zoologica, 80, 193–200.
Walter, D. E., & Proctor, H. C. (1999). Mites: Ecology, evolution and behaviour. Sydney/
Wallingford: University of NSW Press/CABI. 322 p. ISBN 0 86840 529 9.
Welbourn, C. (1992). Acarologists of the world: A directory. Columbus, OH: Ohio State University.
Wheeler, Q. (1990). Insect diversity and cladistic constraints. Annals of the Entomological Society
of America, 86, 1031–1047.
Chapter 2
The Origin of Mites: Fossil History
and Relationships

Mites are members of the arthropod subphylum Chelicerata, a group with a long
fossil history of about 500 million years (Dunlop 2010) (Fig. 2.1). The earliest
chelicerate fossils are known from the Cambrian and appear to be related to a
group of anomalous arthropods with large, raptorial anterior appendages in place of
antennae (Haug et al. 2012). Within the Chelicerata, the mites are placed among
the terrestrial lineages known as the Arachnida. Thus, to understand the origin of
mites, we must peer back through the mists of time and seek answers to these basic
questions: What is an arthropod? Why are chelicerate arthropods different from
other arthropods? What does it mean to be an arachnid? Which arachnids are the
closest relatives of mites?

The Cambrian Explosion and the Rise of the Arthropoda

Multicellular life may have arisen as early as 2.1 billion years ago (Donoghue and
Antcliffe 2010), The earliest evidence of complex multicellular life appears in rocks
formed about 600 million years ago (mya) but the first fossil records of something
that could be a close ancestor to mites begin more recently (Fig. 2.2), during the
so-called ‘Cambrian Explosion’ about 530–550 mya (Gould 1989; Raff 1996).
During the Cambrian, most major animal phyla underwent explosive diversification.
Entombed at more than 30 sites around the world are entire, unusually well preserved
communities of often bizarre organisms. These rocks have been quarried and
studied extensively. Recently developed methods for reconstructing these fossils
in three-dimensions and increasingly sophisticated molecular/morphological analyses
of extant arthropods have combined to provide a coherent hypothesis on the origins
of arthropods (Giribet and Edgecomb 2012).
Recognising an arthropod today is usually easy (Fig. 2.3). As well as being
equipped with a chitinous external skeleton (exoskeleton) that is periodically shed
during development (a process called moulting or ecdysis, which gives rise to cast

D.E. Walter and H.C. Proctor, Mites: Ecology, Evolution & Behaviour: Life at a Microscale, 11
DOI 10.1007/978-94-007-7164-2_2, © Springer Science+Business Media Dordrecht 2013
12 2 The Origin of Mites: Fossil History and Relationships

Fig. 2.1 In the mists of time,


a giant marine scorpion
(Eurypterida) attacks a group
of agnathan fish (Illustration
courtesy of Michael
Caldwell)

skins called exuviae [both singular and plural, like ‘clothes’]), arthropods exhibit
metamerism. That is, arthropods are composed of serially repeated body segments
(or somites) each of which is composed of the tissues and organs needed for life: an
outer skin and sensory structures and inner muscles, gut, nerve mass, excretory
organs, and often a means of circulating blood. Many kinds of animals exhibit
metamerism, including earthworms, velvet worms (Fig. 2.4) and people, but
arthropod segments are unique in that each originally came with a pair of jointed
limbs. These ‘jointed feet’ not only coined the name ‘arthropod’, but they and their
diverse modifications are the best way of understanding how arthropods interact
with their world and why we classify them in various taxonomic groups.
We have only a vague idea about what the first animal that deserved to be called
an arthropod may have looked like. But, if you can imagine an animal composed of
a series of nearly identical segments, each with the genetic capability to produce a
pair of limbs, and a mouth at one end and an anus at the other, then in a cartoonish
way, you have your ‘dawn arthropod’. Of course, life is never that simple, but this
gives us a model with which to start. To feed, this dawn arthropod could just plod
ahead with its mouth open, but it might be helpful to be able to (a) see what is in
front of it, (b) taste or smell what is there, and (c) grasp what seems tasty. Since
arthropods use their limbs to deal with their external environment, it makes sense
that a primary function of the limbs on the anterior most segments would be recog-
nising and capturing food.
The Cambrian Explosion and the Rise of the Arthropoda 13

Fig. 2.2 Mites in relation to


the geological time scale
(Image by HC Proctor and
DE Walter)

Each segment, including its limbs, develops under the control of a series of
genes. At the front end of our series of segments is a structure called the labrum
that may be the specialized expression of a pair of limbs of a long lost segment
(Scholtz and Edgecombe 2006). In our dawn arthropod, it appears that the next seg-
ment went through a series of changes that eventually resulted in the expression of
a pair of simple eyes lenses (ocellus, ocelli) and then, through multiplication of lenses,
14 2 The Origin of Mites: Fossil History and Relationships

Fig. 2.3 A hypothetical early


land chelicerate considers an
early myriapod for dinner
(Image by HC Proctor)

Fig. 2.4 A peripatus or velvet worm (Onychophora) moults its skin and is composed of a long
series of segments, but does not have jointed legs. Onychophorans may be the closest living rela-
tives to arthropods (Image by DE Walter and C. Harvey)
The First Major Dichotomy: Mandibulata Versus Chelicerata 15

more complex light-sensitive organs called compound eyes (Paterson et al. 2011).
Since many modern crustaceans have their compound eyes on movable stalks, and
the earliest known compound eyes are stalked, it is possible that these organs of
sight developed from the limbs of this hypothesized segment. In any case, current
evidence seems to support a single origin of arthropod eyes on an anterior region
called the ocular segment that also contains the most anterior neural ganglion,
the protocerebrum (Maxmen et al. 2005; Scholtz and Edgecombe 2006; Giribet
and Edgecomb 2012). Segments are usually numbered from the front to the back
using Roman Numerals and starting with the first clearly limb-bearing segment
(ignoring any labrial or ocular segments that have been lost in the mists of time). We
follow that convention here.

The First Major Dichotomy: Mandibulata Versus Chelicerata

Once the job of seeing had been taken over by the ocular segment, the next segment
with its limbs and nerves (deuterocerebrum) was available for improving the dawn
arthropod’s ability to acquire food. Since the ocular segment was long hypothetical
(and is still open to contention) and lacks apparent limbs, this first clearly limb-
bearing somite is usually called ‘Segment I’. It is here that the first major dichotomy
(‘in two’ + ‘cut’) of current arthropods seems to have originated. In one group,
including the trilobites and early mandibulates, the appendages on the first limb-
bearing segment became elongate, sensory structures called antennae. The descendants
(Figs. 2.5 and 2.6) of these early mandibulates eventually gave rise to modern crus-
taceans, myriapods, and hexapods. A number of enigmatic Burgess Shale arthropods
lack obvious antennae, and instead have a pair of ‘great appendages’: large spiny,
elbowed limbs at the front of the body. It has long been thought that these strange
Megacheira (‘large hands’) might be closely related to what we today consider
the Chelicerata (Briggs and Fortey 1989; Dunlop and Seldon 1998; Dunlop 2010).
The recent analysis of Haug et al. (2012) of one of the Burgess Shale megacheiran,
Yohoia tenuis, provides a detailed hypothesis of how the 6-segmented megacheiran
great appendage may have been transformed into both the 4-segmented chelifore of
the sea spiders (Pycnogonida) and the 3-segmented chelicera of the Euchelicerata
(Weygoldt and Paulus 1979).
These Euchelicerata or ‘true chelicerates’ represent our first step on the road to
recovering the mites from the mists of time. Most analyses agree that the Euchelicerata
is monophyletic (see Box 2.1). Technically, this means the Euchelicerata is com-
posed of a common ancestor (unknown) and all of its descendants (including mites).
The Euchelicerata does not include the sea spiders, although they may be the closest
relatives (sister group) of the true chelicerates and together form the Chelicerata
sensu lato (i.e. in its broad sense). Our definition of the Euchelicerata depends on
evolutionarily new characters (apomorphies) shared in one form or another
(i.e. synapomorphies), including secondary loss, among all members of the group.
Since we are discussing arthropods, the most useful of these shared derived
16 2 The Origin of Mites: Fossil History and Relationships

Fig. 2.5 Hypothesis of segmental homology: Mandibulate (Crustacea) versus Chelicerate; and a
modern ant head for comparison (insects have lost Antenna 2) (Image by DE Walter)

Fig. 2.6 Simple cladogram


showing some dichotomies
(splits between two lineages)
and derived character
supporting lineages (Image
by DE Walter)

characters will be the form of the limbs and the organisation of the body segments
into functional units (Fig. 2.7). Sea spiders, for example, have a number of unique
characters not found in other chelicerates: a proboscis, a pair of egg-carrying limbs
called ovigers, and a greatly reduced posterior body whose organs have been dis-
placed into the legs (Dunlop 2010).
The First Major Dichotomy: Mandibulata Versus Chelicerata 17

Box 2.1: Phylogenetic Terminology


Evolutionary relationships among taxa can be represented by branching
diagrams called phylograms (if the lengths of the branches represent degree
of differentiation) or cladograms (if they do not). The illustration here is a
cladogram. In a phylogenetic tree, taxa are shown at the tips of the trees
branches (here A–F and X). The taxa can be species or higher taxa, such as
genera, families, etc. More closely related taxa are shown with more recently
shared branch points, or nodes. For example, here A and C share a more
recent common ancestor (node, see arrow) than do A and F. A clade is a subset
of a phylogenetic tree that includes all descendent taxa and their common
ancestor as represented by a node. ACF is a clade, as are EBD and ACFEBD.
A clade is a monophyletic group. If a group were composed of distantly
related taxa (e.g., AD), it is polyphyletic. If a group includes most but not
all members of a clade (e.g., ACEBD), it is paraphyletic. One goal of
modern phylogenetic taxonomy is to avoid creating poly- and paraphyletic
taxa. Phylogenetic trees are constructed using features shown by terminal
taxa, including morphology, behaviour, DNA, etc. The type of feature is
called a character (e.g., ‘eyes’) and the expression of that feature in a taxon
it its state (e.g., ‘eyes present’). A character state that is shared among two
or more taxa, but is different from that in the ancestor (here represented by X),
is a shared, derived character state or synapomorphy, and is evidence that
members of that group shared a common ancestor. Here, the presence of
wings is a synapomorphy of EBD. A character state that is expressed in only
one of the terminal taxa is called an autapomorphy and is not helpful in
reconstructing phylogenetic relationships. In this tree, presence of hair is an
autapomorphy of D. Derived character states can undergo reversal to the
original state. On this tree, eyes are gained in the ancestor of ACFEBD
and lost in F. If eyes had also been lost in B, that would be an example of
convergence. Taxa that share a most recent common ancestor not shared by
any other taxon are called sister taxa. In this tree, the sister taxon of AC is
F, and of BD is E.
18 2 The Origin of Mites: Fossil History and Relationships

Fig. 2.7 A modern scorpion with a prosoma with six pairs of limbs covered by a carapace bearing
a pair of median eyes and two clusters of lateral eyes. The opisthosoma is composed of 12 visible
segments, the first 7 of which are protected dorsally by tergites and the last 5 (the metasoma) in the
form of ring segments. A tail-spine in the form of a bulbous sting terminates the body after the anal
opening (Picture by HC Proctor)

A Review of Arthropod Limb Structure,


Metamerism and Tagmosis

The first arthropod limbs were probably uniramous (i.e. with only one branch or
ramus), but in the early arthropod ground plan only Segment I retained this feature:
all subsequent segments gave rise to biramous limbs, i.e. each had two branches
(Briggs et al. 2012). The inner branch of the biramous limb was usually leg-like and
is called the walking or leg branch. The outer branch was modified for gas exchange
(the gill branch) and often for swimming as well. Each ancestral appendage had a
basal piece that articulated with the body, the coxa (sometimes called the peduncle).
The inner faces of the coxae (plural) were opposed along the ventral mid-line of the
body and typically had spines and processes that allowed them to act as grinding or
filtering surfaces. In the dawn arthropod, all post-ocular body segments and pairs of
limbs were probably very similar in design, but over evolutionary time they have
been variously fused, suppressed or modified for special functions.
When segments join together to form discrete functional units of an arthropod’s
skeleton they are called tagmata (singular, tagma) and the process is called tag-
mosis. Megacheirans had two body tagmata: a 4-segmented head-like region fol-
lowed by a long trunk with all but the most distal segments bearing pairs of biramous
limbs (Dunlop 2010). Euchelicerates also are recognised by having two body
tagmata (Fig. 2.8), but these are differently formulated: an anterior prosoma with
six limb-bearing segments and a posterior opisthosoma with a few pairs of highly
modified limbs. The prosoma is not just a sensing-decision making-feeding head,
as in mandibulate arthropods, but rather it is a body region that combines the
A Review of Arthropod Limb Structure, Metamerism and Tagmosis 19

Fig. 2.8 Venter of mygalomorph spider Antrodiaetus pacificus showing (a) prosoma with six pairs
of limbs around a sternal plate and (b) spinnerets – opisthosomal limbs modified for spinning silk
(SEMs by HC Proctor and DE Walter)

food-acquisition functions of a head with the locomotory functions of a thorax and


is sometimes, especially when covered by a shield (carapace), referred to as a
cephalothorax. The prosoma is composed of the ocular segment (Giribet and
Edgecomb 2012) and six limb-bearing segments (I–VI). The first pair of limbs, the
chelicerae, give their name to the Chelicerata and are the primary organs of feeding.
Generally they are short (2–3 segments) and pincer-like but they can be modified
into fangs, stylets or slicing organs. The second pair of limbs, the pedipalps or
palps, are usually sensory and/or feeding appendages, and come in various forms that
range from simple leg-like limbs to the chelate pinchers of scorpions. The third
through to the sixth pair of limbs are usually walking legs. Antennae are absent in
the Chelicerata because their ancestor apparently found grabbing a neighbour with
raptorial front limbs more useful than exploring its environment with antennae.
Body Segment VII occurs at the juncture of the prosoma and the opisthosoma
and is usually strongly reduced or apparently absent. In spiders, segment VII forms
the pedicel or waist, while in scorpions this segment and its appendages appear
fleetingly in the embryo but are absent in other stages (Hjelle 1990). The composi-
tion of the chelicerate opisthosoma is more contentious, but 12 segments makes a
good working model. Like the abdomen of insects, the opisthosoma contains most
of the digestive, excretory, respiratory and reproductive organs. The limbs of the
opisthosomal segments are strongly reduced, lost, or variously modified (book gills,
book lungs, spinnerets, genital papillae, pectines) for special functions. An anus
opens at the end of the opisthosoma and may be followed by a tail-spine (e.g. on the
horseshoe crab) that can be modified for special functions as in the scorpion’s sting
or the whipscorpion’s flagellum.
20 2 The Origin of Mites: Fossil History and Relationships

Marine Euchelicerates

Traditionally, the Euchelicerata has been subdivided into three major groupings: the
Xiphosurida (horseshoe crabs), the Eurypterida (marine scorpions) and the
Arachnida (terrestrial scorpions and arachnids). The first two groups are almost
entirely extinct; only four species of horseshoe crabs are alive today. Some tantalis-
ing fossils from the Cambrian (550 mya) may represent the earliest xiphosurans, but
the earliest confirmed horseshoe crabs are from the Ordovician (Briggs et al. 2012).
Perhaps the most intriguing of the fossil horseshoe crabs is Euproops danae, small
animals, 2 cm long, from the Mazon Creek beds (300 mya) that are commonly
found in association with the fossils of terrestrial and freshwater animals and even
on the trunk of a fossilised tree (Fisher 1979). Although modern horseshoe crabs
live in the oceans and crawl ashore only to lay eggs, there is no reason to dismiss
the possibility that, during the time of their greatest diversity and abundance, some
xiphosurans colonised the land. But if so, it was a failed experiment: horseshoe
crabs left no extant terrestrial descendants and do not appear to be the sister group
of the Arachnida. Another mysterious marine group of true chelicerates, the
Chasmataspidida, has been variously interpreted as sharing characters with both
the Xiphosurida and the Eurypterida (Dunlop et al. 2004), but again has no obvious
relationship to modern arachnids.
The marine ‘scorpions’ in the Eurypterida (Fig. 2.1) were mostly small to
medium-sized (10–20 cm long) predators that thrived in the oceans during the
Ordovician and Silurian periods (Bergström 1979) when land was first being colo-
nised by plants and animals (Kenrick et al. 2012). This diversity declined during the
Devonian with most survivors being restricted to brackish or freshwater habitats
(Lamsdell and Braddy 2010). Many species inhabited shallow aquatic environments
and appear to have been adapted for walking on land and possibly for both terres-
trial and aquatic gas exchange (Manning and Dunlop 1995). Within the surviving
lineages, however, many groups became gigantic – often with bodies over a metre
in length with some Hibbertopteroidea approaching 2 m and some Pterygotidae
estimated to reach 2.5 m in length with meter long raptorial chelicerae, the largest
arthropods known to have lived (Lamsdell and Braddy 2010). At least 16 lineages
of eurypterids are known from the Silurian and within this diversity may lie the
origin of the arachnids, including the group that has traditionally been considered
the earliest derivative Arachnida, the scorpions.
The opisthosoma of terrestrial scorpions is unique among arachnids: the last
five segments are ring-like and are followed by a tail-spine modified into a sting
with venom glands. Although this metasoma (‘after body’) is not present in
other arachnids, it is characteristic of some eurypterids (Dunlop 1997, 2010).
One such is Mixopterus, which has been reconstructed in a very scorpion-like
pose that suggests that the tail-spine in these animals may have been used as a
sting (Hanken and Størmer 1975). The ecology of ancient Scorpionida is prob-
lematic: some may have been marine and breathed through gills, but whether the
similarities between scorpions and eurypterids reflect ancestry or convergence
Scorpionida: The First Arachnids? 21

remains unresolved (Dunlop 2010). Interestingly, recent research has suggested


that the ability to produce spermatophores, a characteristic of arachnids but not
of horseshoe crabs or sea spiders, may have first arisen within the Eurypterida
(Kamenz et al. 2011).

Scorpionida: The First Arachnids?

A single origin of the Arachnida from a marine chelicerate that colonized land
would be neat and tidy (but see ‘Paleofantasy’, below), and scorpions have long
been the hopeful monster of such a scenario. Modern terrestrial scorpions are a
monophyletic group (Dunlop 2010), and ancient scorpions are known from the
Silurian (428 mya), but the earlier Palaeozoic fossil scorpions lack many of the
derived characters that define the modern scorpions (Kühl et al. 2012). The major
differences between early and modern scorpions are largely due to adaptations to
the terrestrial environment. For example, book lungs replaced book gills, trichoboth-
ria (‘hair + pit’) evolved to sense the movement of air currents, and the lateral com-
pound eye of ancient scorpions gave way to five or fewer simple eyes. Marine
chelicerates have compound eyes, as did early apparently terrestrial forms, but over
time n terrestrial forms these eyes became disorganised and decayed into small
arrays of simple eyes, possibly because the chelicerate compound eye requires a
film of water to function (Levi-Setti 1975).
Relating scorpions to other terrestrial chelicerates is problematic. None of the
defining characters of modern scorpions, such as pectines (Fig. 2.10c), metasoma
(Fig. 2.7), advanced embryonic development (scorpions give birth to live young)
and the structure of the buccal region, are present in other arachnids and it is diffi-
cult to imagine a hypothetical ancestor of both (Snodgrass 1952; Sissom 1990;
Weygoldt 1998). Book lungs may have been one of the earliest of the terrestrial
adaptations in scorpions (Kühl et al. 2012); but if so, it is unclear how the arachnids
that lack lungs (and apparently never had them) would be derived from pulmonate
(i.e. having book lungs) scorpions. Also, other pulmonate arachnids have their book
lungs on Segments VIII and IX, but scorpions have pectines on a subdivided
Segment VIII and book lungs (or gills) on Segments IX–XII. Other non-acarine
arachnids share the division of the body into a prosoma with six pairs of limbs
(Figs. 2.8, 2.9, and 2.10) and an opisthosoma without limbs or with only limb rem-
nants (e.g. the spinnerets of spiders), but lack a well defined metasoma and pectines.
Of greater relevance to our quest for the ancestor of the mites is that neither modern
nor ancient scorpions have characters suggesting that they have a close relationship
to mites. Recent analyses (e.g. Regier et al. 2010; Dabert et al. 2010) support a
rather middle-level derivation of scorpions and do not suggest a close relationship
to either lineage of Acari. So, we must search for an acarine origin among some
other lineage of enigmatic chelicerates that lived in the Ordovician or Silurian. A
single, intrepid founder of the Arachnida would be an inspiring way to start the
story of the successful rise of the mites, but is this a reasonable hypothesis?
22 2 The Origin of Mites: Fossil History and Relationships

Fig. 2.9 Anterior of


mygalomorph spider
Antrodiaetus pacificus
showing how the chelicerae,
pedipalps and anterior portion
of the sternum
(labium = arrow) form the
region around the opening to
the digestive system (SEM
by HC Proctor and DE
Walter). Scale bar = 0.1 mm

The Origin of the Arachnids: A Palaeofantasy

‘Arachnid’ is the name we give to animals with chelicerae that live on land and to
those that have secondarily colonised fresh or salt waters from terrestrial ancestors.
Typically, the Arachnida is treated as a class of equal rank to the Xiphosurida and
Eurypterida and is assumed to have arisen from a single origin. But must this be
true? Could modern arachnids instead be the descendants of several lineages of
small, probably poorly sclerotized, and mysterious chelicerates that swam or
crawled through the Ordovician or Silurian seas?
By way of analogy, it is useful to consider the crustaceans that live in a habitat
we know well – the rainforests of Queensland, Australia. Terrestrial isopods (Class
Malacostraca, Order Isopoda) are well-known litter inhabitants throughout the
world, and finding them in Queensland rainforests is not surprising. In these forests,
however, isopods are usually outnumbered by other crustaceans, talitrid amphipods
(Class Malacostraca, Order Amphipoda, Family Talitridae) – better known as beach
hoppers – are important processors of leaf litter (Friend and Richardson 1986),
including suspended humus many metres up in the rainforest canopy. Other unex-
pected crustaceans also inhabit the rainforest floor, often in large numbers. For
someone used to the litter fauna of the Northern Hemisphere, the sight of hundreds
of small, white, bivalved shells crawling across the bottom of a live-extraction vial
is disconcerting. These animals are ostracods (Class Ostracoda), well known for the
The Origin of the Arachnids: A Palaeofantasy 23

Fig. 2.10 (a) Marine scorpion (Eurypterida); (b) Horseshoe crab (Xyphsurida) leg II (arrow
points to gnathobase); (c) Scorpion (Scorpionida) pectine; (d) trigonotarbid (Trigonotarbida);
(e) Ricinuleid (Ricinulei, Cryptocellus); (f) Camel spider (Solfugida); (g) spider (Aranea);
(h) Whipscorpion (Amblypygida) (Illustrations by HC Proctor and DE Walter)

several thousand species that live in fresh and marine waters (McLaughlin 1980)
but, in this case, adapted to life on land (Fig. 2.9). Less well adapted to terrestrial
existence and by far the smallest of the rainforest crustaceans are the copepods
(Class Copepoda). Although common in leaf litter, these animals are active only in
24 2 The Origin of Mites: Fossil History and Relationships

Fig. 2.11 Austromesocypris – a


terrestrial ostracod from
rainforest litter. The blobs
on the side are phoretic
rotifers. Scale bar = 0.1 mm
(SEM by DE Walter)

films of water in leaf litter, decaying fungi and other wet organic remains. In contrast
are the largest (15 cm long) and most vividly coloured of rainforest crustaceans, the
blue and yellow crayfish or yabbies (Class Malacostraca, Order Decapoda, Family
Parastacidae). Although these animals are primarily stream inhabitants, they often
forage on land and can be found well away from water (Jones and Morgan 1994).
The terrestrial crustacean fauna in Queensland has been derived from at least
five independent colonisations of land and probably by three independent routes.
Isopods and amphipods are believed to have colonised land directly from the
ocean via beach wrack (Labandeira and Beall 1990). Terrestrial ostracods
(Fig. 2.11) and copepods probably started out in fresh water and then colonised
land via soil interstices (Gordon and Olson 1995). Yabbies may be in the process
of moving directly to land from fresh water, perhaps much like scorpions or
eurypterids did long ago. Because all of these animals have living aquatic rela-
tives and because the Crustacea has a long and rich fossil history, no one is likely
to suggest that terrestrial Crustacea came from a single origin. Terrestrialization
is clearly a grade of evolution that has been achieved a number of times (with
varying degrees of success) and it would be meaningless to erect a ‘Class
Geocrustacea’. Why then do we recognise a class called the Arachnida for what
seems an even more diverse array of terrestrial chelicerates?
Unlike crustaceans, chelicerates have neither a rich fossil history nor many living
marine relatives. The sparse array of known fossils may relate to the basic mode of
earning a living in the Chelicerata: eating other animals. Behaviours often corre-
lated with high population densities, such as scavenging, filter-feeding and detritus
feeding, seem to be the dominant ways of earning a living in the Crustacea, but
marine chelicerates were always primarily predators and existed in relatively low
numbers. Although comparatively few fossils of these marine predators have been
found, they include a fair diversity of lineages that were alive when chelicerates first
colonised land. For example, at least 54 genera of Eurypterida (Tollerton 1989),
what appear to be aquatic Scorpionida (Sissom 1990) and several enigmatic chelic-
erates of unknown affinities (Bergström 1979) are known from the Ordovician
Arachnids and the Colonisation of Land 25

through to the Lower Devonian periods (ca 500–400 mya), a minimum date for
when the colonisation of land by chelicerates most likely occurred (Gordon and
Olson 1995; Labandeira 2005; Schaefer et al. 2010). In the late Silurian alone, 70
species of eurypterids representing 28 genera are known (Plotnick 1996). Did one
and only one of this great diversity of forms make the successful transition to land?
Well, perhaps, but we shouldn’t be too upset if it was more complicated than that.

Arachnids and the Colonisation of Land

Although fossil records from the apparent dawn of terrestrial life continue to accu-
mulate, exactly when animals began leaving the oceans for life on land is still
unclear. Track fossils attributed to millipedes are known from the mid-Ordovician
(~450 mya) and millipede fossils with distinct spiracles (indicative of air-breathing)
have been found from early in the second half of the Silurian (~423 mya) (Selden
and Read 2008; Kenrick et al. 2012). The oldest known terrestrial ecosystem, from
the late Silurian (~419 mya), provides our first clear records of terrestrial arachnids
in the form of a tiny (1.3 mm long) member of the order Trigonotarbida (Jeram
et al. 1990), Eotarbus jerami (Dunlop 1996). The ordinal name comes from the
Greek words for triangle (trigono – probably in reference to the triangular carapace
of some) and terror (tarbo). The binomial could be translated as ‘Jeram’s Dawn
Terror’. Currently, trigonotarbids (Fig. 2.10d) are the earliest known pulmonate
terrestrial arachnids: they had two pairs of well-developed book lungs for breathing
air (Gordon and Olson 1995), as do modern whipscorpions (Fig. 2.10h) and myga-
lomorph spiders (Fig. 2.10g).
Trigonotarbids appeared to have a body composed of (excluding the ocular
segment) six prosomal and 12 opisthosomal segments (Dunlop 2006). A carapace
with a projecting rostrum protecting the mouthparts covered the trigonotarbid
prosoma and usually bore a median simple eye tubercle and often two lateral eye
complexes (probably remnants of compound eyes). The walking legs and pedipalps
were rather simple in structure, but the mouth opening was surrounded by a filter-
like array of dense setae and each chelicera had a fang-like digit opposed to a base
with large teeth. This suggests that these ‘triangular terrors’, like modern whipscor-
pions with similar chelicerae, impaled a prey on their fangs, mashed it against the
toothed bases of the chelicerae, and digested it externally, sucking in the prey fluids.
This is similar to how whipscorpions and some spiders feed: Garwood and Dunlop
(2011) have suggested that some Carboniferous trigonotarbids may have been
ambush predators rather like today’s crab spiders.
Opisthosomal segments were protected by a complicated arrangement of sclero-
tized plates called tergites, if dorsal; sternites, if ventral; and ring segments if
reduced to small, sclerotized rings (Dunlop 2006). The first tergite (Segment VII)
was narrow and bore a ridge that formed a locking mechanism with the carapace
and the last (Segment IX) was expanded and covered the remaining segments. In
between, Tergites 2 + 3 (Segments VIII + IX) were fused into a single unit and
26 2 The Origin of Mites: Fossil History and Relationships

Tergites 2 + 3 through 8 (and sometimes 9) were subdivided into a median plate and
one (early forms) or two (later forms) lateral plates. This characteristic subdivision
of the tergites into median and lateral plates is considered a defining character of
the Trigonotarbida. Ventrally, Segment VII is not armoured, but the remainder are
protected by undivided plates or ring segments. Segments VIII and IX bore a pair
of opercula (plural form of the Latin operculum, meaning a cover or lid) over the
entrances to the book lungs. The genital pore is thought to open under the oper-
cula on Segment VIII. The last three segments form a pygidium (‘rump’) with
Segment X a plate; XI–XII small and ring-like; and the anus opening at the tip of
Segment XII.
Trigonotarbid fossils show up again in the Devonian Rhynie Chert of Scotland
(407–411 mya), the black shale of Alken der Mosel of Germany (390 mya) and the
black shales of Gilboa in New York (392 mya). They appear to have reached their
highest diversity in the Carboniferous Period and persisted into the early Permian
before disappearing from the fossil record (Dunlop 2010), after fossils of the extant
order Ricinulei began to appear in the late Carboniferous (Selden 1993), perhaps,
when parasitiform mites first made their appearance (Mans et al. 2011). Ricinulei is
one of the arachnid groups that have been proposed as close relatives of the mites
(see below).
The early trigonotarbids appear to have been quite small (1–14 mm long) and
lightly armoured. Later specimens were much larger, up to 5 cm in length, and were
heavily armoured (Shear and Kukalová-Peck 1990). This may represent an example
of Cope’s Rule, i.e. when new taxa arise they tend to be small relative to their even-
tual size range (Stanley 1973). If so, then mites seem to flout this rule, as they do so
many others. Although the earliest known mites were, indeed, very small (half a
millimetre or less in length), for the most part they have stayed small or become
even smaller. In fact, miniaturisation may have been the key to their invasion of land
through beach sands and soil pores.

Fossil Mites

Acariform mites are among the earliest of the fossil arachnids (at least 411 mya and
possibly older – see Bernini et al. 2002; Schaefer et al. 2010) and are known from
both the Devonian Gilboa Shales and the Rhynie Chert (Dunlop and Selden 2009).
At least 11 species of mites are present at these two sites (Kethley et al. 1989; Shear
and Kukalová-Peck 1990; Bernini 1991). Four species of the early derivative oribatid
mite taxon, Enarthronota, are known from Gilboa (Norton et al. 1988). The two
species that have been described, Devonacarus sellnicki and Protochthonius gilboa,
are both small, ranging from 323 to 512 μm in length. Similarly, the endeostigmatid
mite Archaeacarus dubinini from Gilboa is less than 400 μm in length (Kethley
et al. 1989), and the Rhynie Chert endeostigmatid Protacaris crani is less than
500 μm long (Gordon and Olson 1995).
Fossil Mites 27

How does one distinguish fossil mites from other arachnids? Aside from being
generally tiny chelicerate arthropods with an anterior body region called either the
capitulum (‘little head’) or gnathosoma (‘jaw body’) and the absence of serial
plates on the body, this question is difficult to answer (see Chap. 3). Two distinct
lineages of mites are currently recognised: the Acariformes (=Actinotrichida)
and Parasitiformes (=Anactinotrichida) (Krantz and Walter 2009). All share the
general characteristics of mites but, as Lindquist (1984) has pointed out, many of
the characters used to define mites are also present in other chelicerate orders.
For example, most spiders also lack serial plates, while ricinuleids may have a
gnathosoma. Often, it seems, mites are most easily recognised by what they are
not – other arachnids.

Fossil Acariformes

Almost 300 species of fossil acariform mites have been formally described (Arillo
et al. 2012; Dunlop et al. 2007; Dunlop and Selden 2009; Selden 1993; Schmidt et al.
2012). All of the Devonian fossil mites belong to the Acariformes (and include
representatives of the Endeostigmata, Oribatida and Prostigmata), as do all of the
Palaeozoic and most of the Mesozoic fossils, and most can be placed in superfami-
lies or families with living members. Even the most ancient acarine plant-parasites,
eriophyoid mites from Triassic amber (~230 mya) (Schmidt et al. 2012), have body
plans little different from their modern relatives. Most ancient acariformans belong
to taxa whose descendants today feed on fungi and other microbes, dead plant matter
or green algae. Based on extensive faecal remains, it appears that these acariform
mites were major components of the detritivore system in Palaeozoic coal swamps
(Labandeira et al. 1997; Kellogg and Taylor 2004; Feng et al. 2010), just as they are
in modern soil-litter systems (Fig. 2.12).

Fossil Parasitiformes

In contrast, no fossil parasitiform mites are known from the Palaeozoic and even
their recent fossil history is poor, including only about two dozen species (Dunlop
et al. 2013). The oldest parasitiform fossils are ticks from late Cretaceous amber
(~90–100 mya). The earliest Opilioacarida and Mesostigmata are known from even
more recent Eocene amber ca. 44–49 mya and include the parasitid mite
Aclerogamasus stenocornis and uropodid deutonymphs tentatively assigned to the
modern genus Uroobovella and phoretic on a cerambycid beetle (Dunlop 2010;
Dunlop et al. 2013). Seius bdelloides (sometimes assigned to Sejidae, but it has
no characteristics of the family and is probably a member of the Dermanyssiae,
possibly Phytoseiidae) and a digamasellid mite, Dendrolaelaps fossilis, have been
described from mid-Tertiary amber (Selden 1993). Although it seems likely that the
28 2 The Origin of Mites: Fossil History and Relationships

Fig. 2.12 Late Cretaceous (Foremost Formation 78–79 mya) fossil Acariformes from the
Grassy Lake amber of Alberta (see McKeller and Wolfe 2010): (a) Prostigmata, Anystoidea;
(b) Prostigmata, Tetranychidae; (c) Prostigmata, Camerobiidae; (d) Oribatida (possibly
Licneremaeoidea); (e) Oribatida (possibly Archaeorchestidae) (Images by DE Walter)

Parasitiformes arose much earlier, perhaps in the Carboniferous (Mans et al. 2011),
no currently known fossils support this hypothesis.

Potential Arachnid Relatives of Mites

The possibility that mites have no close relatives among terrestrial chelicerates has
rarely been considered by modern arachnologists, but earlier workers were
impressed by how different mites were and often considered them as separate from
or basal to other arachnids (see Dunlop and Alberti 2007). At one time, sea spiders
(Pycnogonida) were thought to be related to mites and workers wondered at the
similarity between the hexapod protonymphon larvae of sea spiders (Fig. 2.13) and
the hexapod larvae of mites (Dunlop and Arango 2005). Fürstenberg (1861) even
included pycnogonids as a family of water mites in his book with the scratch-
inducing title “The itch mites of men and animals”. A third instar protonymphon
pycnogonid can be more or less hexapod (the third pair of legs are sack-like) and a
Potential Arachnid Relatives of Mites 29

Fig. 2.13 Sea spider larval


stage: 3rd instar
protonymphon of a
pycnogonid, venter. Note
strand of silk on viewer’s
right (SEM by HC Proctor)

capitulum-like pair of chelicerae with palp-like processes above the proboscis. The
“palp”, however, is a spinning spine and silk is produced from a pore at its tip and
probably is derived from the endite of the chelifore coxa. The next two appendages
transform into palps and ovigers during development (Bain 2003) and it is only the
sack-like blobs at the rear (bud-like in earlier protonymphon instars) that become
the first of the walking legs. Thus, the pycnogonid protonymphon does not appear
to be constructed like a mite larva.
More modern workers have accepted mites as good arachnids, but also have
tended to assume that mites are (a) monophyletic and (b) relatively derived within the
Arachnida. Extremely poor taxon sampling of mites, often restricted to one or two
species, has been a hallmark of most previous studies of arthropod, or even arach-
nid, phylogeny. As a result, mites have tended to come out in anomalous locations
or near various other poorly studied taxa. Morphology and biology of the various
contenders for a sister group to mites are discussed below (see Dunlop and Alberti
2007 for a thorough review).

Palpigradi

Palpigrades are minute, enigmatic arachnids with a rather generalised morphology


and numerous, seemingly ancestral, character states, e.g. chelicerae three seg-
mented, palps leg-like, book lungs absent, first pair of legs antenniform and telson
30 2 The Origin of Mites: Fossil History and Relationships

(flagellum) present (Kaestner 1968; Savory 1977; van der Hammen 1989). They
have often been considered the most primitive of arachnids and some authors have
suggested that that they are not completely terrestrial. For example, sand-dwelling
species of Leptokoenenia are able to swim easily in sea water (Monniot 1966).
However, this genus appears to be recently derived and most other palpigrades are
clearly terrestrial (Condé 1996). Also, palpigrades have numerous well-developed
trichobothria (sensory hairs that function only in air). Prokoenenia wheeleri has
paired invaginations on Segments X–XII (‘lung sacs’) and a pair of papilla-like
bumps called verrucae on the underside of Segment IX. These may be remnants of
ancestral book lungs or of book gills. Modern palpigrades, similar to many acari-
form mites, are thought to respire cutaneously (van der Hammen 1989).
Based on his concepts of the origin of feeding structures in Chelicerata, van der
Hammen (1989) considered the Acari to be diphyletic (in contrast to monophyletic)
and contended that Acariformes and Parasitiformes are only distantly related. Van
der Hammen’s morphology-based phylogeny implies that at least three independent
colonisations of land by arachnids occurred: by ancestral Opiliones (harvestmen),
by Scorpionida and by an ancestor of the rest of the arachnids. According to van
der Hammen, the Acariformes and Palpigradi are sister groups. However, the char-
acters that he presents to support this union are a mixture of ancestral characters
(e.g. tridactyl claws) and interpretations of development or morphology that are not
generally accepted. In fact, there seem to be few if any character states that the
Acariformes and Palpigradi share that are not found in other arachnids.
One general point that favours a palpigrade–mite relationship is that both are
small, soil-dwelling animals whose first terrestrial ancestors may have invaded the
land through soil interstices (Condé 1996). Both early derivative acariform mites
and palpigrades appear to rely on cutaneous gas exchange for respiration. Some
palpigrades have three pairs of opisthosomal sacs and adult acariform mites have three
pairs of genital papillae that represent remnants of the limbs of three opisthosomal
segments (Grandjean 1946). Although genital papillae have no respiratory function,
they are used for osmoregulation (water uptake and ion balance) (Evans 1992) and
so derivation from the gill branch of an ancient limb is reasonable hypothesis. It is
possible that these three pairs of remnant limbs are homologues in palpigrades and
acariform mites, but segmentation in the acariform mite opisthosoma appears to be
difficult to resolve (Barnett and Thomas 2012).
Generally, genital papillae are added sequentially during postembryonic ontogeny
in acariform mites: larvae have no papillae, protonymphs have one pair, deutonymphs
two pairs and tritonymphs usually have the maximum of three pairs (see Chap. 4).
Van der Hammen (1989) pointed out that the lung sacs in P. wheeleri appear to be
similarly added in sequential moults. However, there is good reason not to accept
this similarity as a shared derived condition. The trilobite larva of the horseshoe
crab has only two pairs of book gills. Additional pairs of book gills are added
sequentially with moults, until the adult complement of five pairs is reached
(Kaestner 1968). Presumably this sequential addition of opisthosomal structures
represents an ancestral chelicerate ontogeny. If the ancestor of acariform mites
Potential Arachnid Relatives of Mites 31

Fig. 2.14 Siro acaroides a very mite-like Cyphophthalmi Opiliones (scale bar = 1 mm) (SEMs by
DE Walter)

colonised interstitial sands directly from the ocean, then they may never have
evolved book lungs like those found in the larger arachnids.
Recently published phylogenetic analyses have done little to clarify what
relationship the Palpigradi and Acari may have. One molecular phylogeny of the
Arthropoda based on 62 nuclear protein-coding genes (Regier et al. 2010) did
recover a weakly supported (bootstrap values <50 %) sister group relationship
between palpigrades and acariform mites. However, the arachnid taxa included
were minimal and most of the nodes within the Euchelicerata were poorly
supported.

Opiliones

Harvestmen were early favourites in the contest for a close relative of mites but
recent views have been mixed. Cyphophthalmid opilionids (Fig. 2.14) are super-
ficially very mite-like in appearance (Savory 1977), and members of the mite
family Opilioacaridae are superficially opilionid-like. Like opilioacarans and
many acariformans, many harvestmen feed on particulate food and all harvestmen
and mites have their opisthosomas and prosomas broadly joined. However,
32 2 The Origin of Mites: Fossil History and Relationships

particulate-feeding and a broad waist are also characters of marine chelicerates


such as horseshoe crabs and eurypterids; therefore, rather than acting as synapo-
morphies for Opiliones and mites, they may be retained plesiomorphies. No
clearly derived characters suggest a close relationship between harvestmen and
mites and recent molecular phylogenies have not supported a sister group rela-
tionship of opilionids with either superorder of mites.

Ricinulei

One group of small (<10 mm long), spider-like arachnids called Ricinulei (Fig. 2.10e)
or hooded tickspiders (ricinus is Latin for tick) has long been a favourite candidate
for the mites’ closest living relatives. Only about 60 species in a single family
(Ricinoididae) are known from Africa and the New World, where they occur as far
north as Texas. Lindquist (1984) presented four possibly derived characters linking
the Acari and Ricinulei, of which two seem especially convincing. In both mites and
hooded tickspiders the first active stages that hatch from the egg are 6-legged
(hexapod). No other arachnids are known to have a 6-legged stage. Also, the first two
limb-bearing segments in mites are separated from the rest of the body by soft cuticle to
form the jaw-body (gnathosoma or capitulum) and the bases of the pedipalps are
fused to form a subcapitulum (also called infracapitulum). Ricinuleids also have
the chelicerae and pedipalps organised into a unit similar to the mite jaw-body.
Weygoldt and Paulus (1979) first proposed a sister group relationship between
the Ricinulei and the Acari, and named this taxon the Acarinomorpha. Shultz
(1993), in an analysis dominated by limb characters, also supported this relationship
and called the group Acaromorpha. In contrast, van der Hammen (1989) viewed the
Ricinulei as the sister group of just one group of mites, the Anactinotrichida
(=Parasitiformes), with acariform mites being distant relatives.
Unfortunately, the developmental and morphological support for a sister group
relationship between mites and ricinuleids is not as strong as it first seemed. Early
developmental stages are unknown for most extinct groups of arachnids; for exam-
ple, it is unknown if a hexapod stage may have been present in Trigonotarbida. Also,
why have a hexapod stage at all? Could there be a benefit to suppressing the fourth
pair of legs until after hatching in tiny chelicerates? If so, then convergence becomes
a likely explanation for this shared character.
If a hexapod larval stage is a primitive condition, or due to convergent evolution,
then only the gnathosoma remains as a strong shared derived character linking mites
and hooded tickspiders. Sharing a name, however, is not necessarily the same as being
the same structure in an evolutionary sense (homology). Hooded tickspiders have a
unique cap-like plate called the cucullus (Latin for hood) that is separated by soft
cuticle from the rest of the prosoma, covers the chelicerae, and can be drawn against the
subcapitulum. No mite has a cucullus, so this weakens any proposed similarity.
Parasitiform mites do have a subcapitulum formed from the medially fused bases of the
pedipalps, as do hooded tick spiders. The subcapitulum forms the floor of the mouth,
Potential Arachnid Relatives of Mites 33

Fig. 2.15 A pseudoscorpion (a) dorsal habitus (long hairs on chelae represent trichobothria) and
(b) close-up of chelicerae (SEMs by DE Walter)

much like our lower jaw, and is probably an important adaptation to fluid-feeding pred-
ators on land: the mass of saliva and digested prey might float in water well enough to
be sucked in, but would dribble down to the surface on land. All arachnids have evolved
one way or another to solve this problem, but in addition to parasitiform mites and
hooded tickspiders, some whipscorpions and pseudoscorpions also form a subcapitu-
lum from the median fusion of the pedipalps. Moreover, acariform mites form their
subcapitulum in a different manner than do parasitiform mites (see Chap. 3).
Thus, we are left with only weak morphological inference for mites and hooded
tickspiders being closely related (see Talarico et al. 2011). Also, support for this
relationship is poor or absent from recent molecular phylogenetic analyses (see
below). Finally, Jason Dunlop (2006) has made an interesting case for hooded
tickspiders being derived trigonotarbids and the similarities in body form between
the two groups are striking.

Pseudoscorpionida

Pseudoscorpions (Fig. 2.15) appear in the fossil record in the Devonian (392 mya),
not much later than acariform mites (410 mya), but have only rarely been thought of
as relatives of the mites (Dunlop and Alberti 2007). Pseudoscorpions long precede
parasitiform mites in the fossil record, and seem unlikely to have shared a recent
common ancestor with them. However, two recent analyses (Regier et al. 2010;
Dabert et al. 2010) have found some support for a sister group relationship between
34 2 The Origin of Mites: Fossil History and Relationships

Parasitiformes and Pseudoscorpionida. Neither of these studies was specifically


designed to determine the relationships of parasitiform mites, so neither has particu-
larly good taxon sampling, but the relationship is an interesting hypothesis.

Solifugida

Of all of the currently living groups of arachnids, the solifigids (camel spiders, wind
scorpions) might seem to be the least likely to be close relatives of mites. Camel
spiders (Fig. 2.10f) tend to be rather large, fast moving predators, with enormous
chelicerae and long pedipalps that end in sticky pads for capturing prey. Solifugid
fossils only date back to the Carboniferous (~308 mya) (Dunlop 2010) and they
have usually been thought to be close relatives of the pseudoscorpions (Dunlop and
Alberti 2007).
However, some early mite workers were struck by the similarity in form of
solifugids and acariform mites in the family Rhagidiidae – these are fairly early
derivative acariform mites with large chelicerae that are fast and furious predators.
In a posthumously published work, the prominent Russian acarologist Aleksey
A. Zachvatkin specifically proposed that acariform mites were closely related to
solifugids (Dunlop and Alberti 2007). Other prominent acarologists have also
pointed out similarities between solifugids and acariform mites (e.g. Grandjean
1954; Alberti 2000). One particularly interesting similarity is that the solifugid egg
hatches into an inactive stage that, except for having eight legs instead of six, is
similar to the inactive acariform mite prelarva. In any case, two recent molecular
phylogenetic analyses (Dabert et al. 2010; Pepato et al. 2010), both devoted to
teasing out relationships among acariform mites, have found strong support for
solifugids being the sister group of acariform mites.

Summary and Preview

Chelicerate arthropods had invaded developing terrestrial ecosystems from the


ancient oceans by at least the late Silurian period more than 400 mya. Scorpions
were once thought to be the earliest of these invaders and the founders of the
Arachnida. All arachnids, including the mites, were thought to be descended from
one such group of terrestrialized scorpions. However, it now seems likely that a
variety of different marine chelicerate lineages may have colonised land and that
extant arachnids represent their descendants. It also seems likely that the tiny arach-
nids that we call mites represent two rather distantly related lineages, the Acariformes
and Parasitiformes. These ‘mite’ lineages appear to be more closely related to other
groups of living arachnids than to each other.
Acariform mites were among the earliest terrestrial colonists and probably
invaded land directly from the ocean by way of the interstices in beach sand.
References 35

The earliest known members of this lineage are minute, have no respiratory organs
and exchange gas across their cuticles in moist soil habitats. They have retained an
ancestral marine chelicerate mode of feeding mode (ingesting particles of food) but
have terrestrial adaptations such as trichobothria, which sense air currents. By the
early Devonian period, acariformans were diverse, if tiny, members of the soil
fauna, and many fossils and fossil coprolites attributed to wood and leaf-boring
oribatids mites have been described from the Palaeozoic and Mesozoic Eras. Early
radiations filled the soil system with a bewildering array of minute microbivores and
scavengers and oribatid mites appear to have been (see Chap. 6). Later radiations in
association with plants (Chap. 8) and animals (Chaps 9 and 10) have produced a
similarly dazzling diversity of species, such that perhaps a half million species of
acariformans are alive today.
Parasitiform mites are more enigmatic. Their fossil record is sparse, beginning
only in the late Cretaceous, and known fossils all fit within living genera. Most para-
sitiform mites are predators that feed on the fluids of their prey, as do most other
arachnids, or parasitize invertebrates (e.g. Varroa on honeybees) and vertebrates
(e.g. fowl mites, ticks). Although not as successful as their distant relatives in the
Acariformes, today parasitiform mites are the dominant predators in soil systems,
useful biological control agents in agriculture, and the most important vectors of
disease after the mosquitoes (Chap. 8).

References

Alberti, G. (2000). Chelicerata. In B. G. M. Jamieson (Ed.), Progress in male gamete ultrastruc-


ture. In K. G. Adiyodi & R. G. Adiyodi (Eds.), Reproductive biology of the invertebrates,
Vol. 9B (pp. 311–388). New Delhi/New York: Oxford & IBH Publishing/Wiley.
Arillo, A., Subias, L. S., & Shtanchaeva, U. (2012). A new species of fossil oribatid mite
(Acariformes, Oribatida, Trhypochthoniidae) from the lower Cretaceous amber of San Just
(Teruel Province, Spain). Systematic & Applied Acarology, 17, 106–112.
Bain, B. A. (2003). Larval types and a summary of postembryonic development within the
pycnogonids. Invertebrate Reproduction and Development, 43, 193–222.
Barnett, A. A., & Thomas, R. H. (2012). The delineation of the fourth walking leg segment is tem-
porally linked to posterior segmentation in the mite Archegozetes longisetosus (Acari:
Oribatida, Trhypochthoniidae). Evolution & Development, 14, 383–392. doi:10.1111/j.1525-
142X.2012.00556.x.
Bergström, J. (1979). Morphology of fossil arthropods as a guide to phylogenetic relationships.
In A. P. Gupta (Ed.), Arthropod phylogeny (pp. 3–56). New York: Van Nostrand Reinhold
Company.
Bernini, F. (1991). Fossil Acarida. In A. Simonetta & S. C. Morris (Eds.), The early evolution
of metazoa and the significance of problem taxa (pp. 253–262). Cambridge: Cambridge
University Press.
Bernini, F., Carnevale, G., Bagnoli, G., & Stouge, S. (2002). An early Ordovician mite (Acari:
Oribatida) from the island of Öland, Sweden. In F. Bernini, R. Nannelli, G. Nuzzaci, & E. de
Lillo (Eds.), Acarid phylogeny and evolution. Adaptations in mites and ticks (pp. 45–47).
Dordrecht: Kluwer.
Briggs, D. E. G., & Fortey, R. A. (1989). The early radiation and relationships of the major
arthropod groups. Science, 246, 241–243.
36 2 The Origin of Mites: Fossil History and Relationships

Briggs, D. E. G., Siveter, D. J., Sutton, M. D., Garwood, R. J., & Legg, D. (2012). Silurian
horseshoe crab illuminates the evolution of arthropod limbs. Proceeding of the National
Academy of Sciences of the USA, 109, 15702–15705. doi:10.1073/pnas.1205875109.
Condé, B. (1996). Les Palpigrades, 1885–1995: Acquisitions et lacunes. Revue Suisse de Zoologie
Hors série, 1, 87–106.
Dabert, M., Witalinski, W., Kazmierski, A., Olszanowski, Z., & Dabert, J. (2010). Molecular
phylogeny of acariform mites (Acari, Arachnida): Strong conflict between phylogenetic signal
and long-branch attraction artifacts. Molecular Phylogenetics and Evolution, 56, 222–241.
Donoghue, P. C. J., & Antcliffe, J. B. (2010). Origins of multicellularity. Nature, 466, 41–42.
Dunlop, J. A. (1996). Evidence for a sister group relationship between Ricinulei and Trigonotarbida.
Bulletin of the British Arachnological Society, 10, 193–204.
Dunlop, J. A. (1997). Palaeozoic arachnids and their significance for arachnid phylogeny.
Proceedings 16th European Colloquim of Arachnology (pp. 65–82).
Dunlop, J. A. (2006). Evidence for s sister group relationship between Ricinulei and Trigonotarbida.
Bulletin of the British Arachnological Society, 10, 193–204.
Dunlop, J. A. (2010). Geological history and phylogeny of Chelicerata. Arthropod Structure &
Development, 39, 124–142.
Dunlop, J. A., & Alberti, G. (2007). The affinities of mites and ticks: A review. Journal of
Zoological Systematics and Evolutionary Research, 46, 1–18.
Dunlop, J. A., & Arango, C. P. (2005). Pycnogonid affinities: A review. Journal of Zoological
Systematics and Evolutionary Research, 43(1), 8–21. doi:10.1111/j.1439-0469.2004.00284.x.
Dunlop, J. A., & Selden, P. A. (1998). The early history and phylogeny of chelicerates. Systematics
Association Special Volume Series, 55, 221–235.
Dunlop, J. A., & Selden, P. A. (2009). Calibrating the chelicerate clock: A paleontological response
to Jeyaprakash and Hoy. Experimental & Applied Acarology, 48, 183–197.
Dunlop, J. A., Anderson, L. I., & Braddy, S. J. (2004). A redescription of Chasmataspis laurencii
Caster & Brooks, 1956 (Chelicerata: Chasmataspidida) from the Middle Ordovician of
Tennessee, USA, with remarks on chasmataspid phylogeny. Transactions of the Royal Society
of Edinburgh-Earth Sciences, 94, 207–225.
Dunlop, J. A., Penn, D., Tetlie, O. E., & Anderson, L. I. (2007). How many arachnid fossils are
there? Journal of Arachnology, 36, 267–272.
Dunlop, J. A., Kontschán, J., & Zwanzig, M. (2013). Fossil mesostigmatid mites (Mesostigmata:
Gamasina, Microgyniina, Uropodina), associated with longhorn beetles (Coleoptera:
Cerambycidae) in Baltic amber. Naturwissenschaften, 100, 337–344.
Evans, G. O. (1992). Principles of Acarology. Wallingford: CAB International.
Feng, Z., Wang, J., & Liu, L.-J. (2010). First report of oribatid mite (arthropod) borings and
coprolites in Permian woods from the Helan Mountains of northern China. Palaegeography,
Palaeoclimatology, Palaeoecology, 288, 54–61.
Fisher, D. (1979). Evidence for subaerial activity of Euproops danae (Merostomata: Xiphsuridae).
In M. H. Nitecki (Ed.), Mazon Creek fossils (pp. 379–447). New York: Academic Press.
Friend, J. A., & Richardson, A. M. M. (1986). Biology of terrestrial amphipods. Annual Review of
Entomology, 31, 25–48.
Fürstenberg, M. H. F. (1861). Die Krätzmilben der Menschen und Thiere. Leipzig: Wilhelm Engelmann.
Garwood, R. J., & Dunlop, J. A. (2011). Morphology and systematics of Anthracomartidae
(Arachnida:Trigonotarbida). Palaeontology, 54, 145–161.
Giribet, G., & Edgecomb, G. (2012). Reevaluating the arthropod tree of life. Annual Review of
Entomology, 57, 167–186.
Gordon, M. S., & Olson, E. C. (1995). Invasions of the land – The transitions of organisms from
aquatic to terrestrial life. New York: Columbia University Press.
Gould, S. (1989). Wonderful life – The Burgess Shale and the nature of history. New York: Penguin
Books.
Grandjean, F. (1946). Au sujet de l'organe Claparède, des eupathides multiples et des taenidies
mandibulaires chez les acariens actinochitineux. Archives des Science Physiques et Naturelles,
Genève, 28, 63–87.
References 37

Grandjean, F. (1954). Ètude sur les palaeacaroides (Acariens, Oribates). Mém Mus Nat Hist Nat
Paris, Sér A Zoologie, 7, 179–274.
Hanken, N. M., & Størmer, L. (1975). The trail of a large Silurian eurypterid. Fossils and Strata,
4, 255–270. Oslo.
Haug, J. T., Waloszek, D., Maas, A., Liu, Y., & Haug, C. (2012). Functional morphology,
ontogeny and evolution of mantis shrimp-like predators in the Cambrian. Paleontology, 55,
369–399.
Hjelle, J. T. (1990). Anatomy and morphology. In G. A. Polis (Ed.), The biology of scorpions
(pp. 9–63). Stanford: Stanford University Press.
Jeram, A. J., Selden, P. A., & Edwards, D. (1990). Land animals in the Silurian – arachnids and myria-
pods from Shropshire, England. Science, 250, 658–661. doi:10.1126/science.250.4981.658.
Jones, D., & Morgan, G. (1994). A field guide to crustaceans of Australian waters. Sydney: Reed
Books.
Kaestner, A. (1968). Invertebrate zoology (Vol. II). New York: Wiley.
Kamenz, C., Staude, A., & Dunlop, J. A. (2011). Sperm carriers in Silurian sea scorpions.
Naturwissenschaften, 98, 889–896.
Kellogg, D. W., & Taylor, E. L. (2004). Evidence of oribatid mite detritivory in Antarctica during
the late Paleozoic and Mesozoic. Journal of Paleontology, 78, 1146–1153.
Kenrick, P., Wellman, C. H., Schneider, H., & Edgecomb, G. D. (2012). A timeline for
terrestrialization: Consequences for the carbon cycle in the Palaeozoic. Philosophical
Transactions of the Royal Society of London, Series B, Biological Sciences, 367, 519–536.
doi:10.1098/rstb.2011.0271.
Kethley, J. B., Norton, R. A., Bonamo, P. M., & Shear, W. A. (1989). A terrestrial alicorhagiid mite
(Acari: Acariformes) from the Devonian of New York. Micropaleontology, 35, 367–373.
Krantz, G. W., & Walter, D. E. (Eds.). (2009). A manual of Acarology (3rd ed.). Texas Tech
University Press. 807 p, 338 b/w illustrations, 60 figures ISBN 978-0-89672-620-8.
http://www.ttup.ttu.edu/BookPages/9780896726208.html.
Kühl, G., Bergmann, A., Dunlop, J., Garwood, R. J., & Rust, J. (2012). Redescription and palaeo-
biology of Palaeoscorpius devonicus Lehmann, 1944 from the Lower Devonian Hunsrück
Slate of Germany. Palaeontology, 55, 775–787.
Labandeira, C. C. (2005). Invasion of the continents: Cyanobacterial crusts to tree-inhabiting
arthropods. Trends in Ecology & Evolution, 20, 253–261.
Labandeira, C. C., & Beall, B. S. (1990). Arthropod terrestriality. Short Courses in Paleontology,
3, 214–256.
Labandeira, C. C., Phillips, T. L., & Norton, R. A. (1997). Oribatid mites and the decomposition of
plant tissues in Paleozoic coal-swamp forests. Palaios, 12, 319–353.
Lamsdell, J. C., & Braddy, S. J. (2010). Cope’s Rule and Romer’s theory: Patterns of diversity and
gigantism in eurypterids and Palaeozoic vertebrates. Biology Letters, 6, 265–269. doi:10.1098/
rsbl.2009.0700.
Levi-Setti, R. (1975). Trilobites, a photographic atlas. Chicago: University of Chicago Press.
Lindquist, E. E. (1984). Current theories on the evolution of major groups of Acari and on their
relationships with other groups of Arachnida, with consequent implications for their classifica-
tion. In D. A. Griffiths & C. E. Bowman (Eds.), Acarology VI (Vol. 1, pp. 28–62). New York:
Wiley.
Manning, P. L., & Dunlop, J. A. (1995). The respiratory organs of eurypterids. Palaeontology, 38,
287–297.
Mans, B. J., de Klerk, D., Pienaar, R., & Latif, A. A. (2011). Nuttalliella namaqua: A living fossil
and closest relative to the ancestral tick lineage: Implications for the evolution of blood-feeding
in ticks. PLoS One, 6(8), e23675. doi:10.1371/journal.pone.0023675.
Maxmen, A., Browne, W. E., Martindale, M. Q., & Giribet, G. (2005). Neuroanatomy of sea
spiders implies an appendicular origin of the protocerebral segment. Nature, 437, 1144–1148.
McKeller, R. C., & Wolfe, A. P. (2010). Canadian amber. In D. Penney (Ed.), Biodiversity of fossils
in amber from the major world deposits (pp. 96–113). Manchester: Siri Scientific. ISBN
978-0-9558636-4-6.
38 2 The Origin of Mites: Fossil History and Relationships

McLaughlin, P. A. (1980). Comparative morphology of recent crustacea. San Francisco: W.H. Freeman
and Company.
Monniot, F. (1966). Un Palpigrade interstitiel: Leptokoeninia scurra n. sp. Revue Ecologie Biologie
du Sol, 3, 41–64.
Norton, R. A., Bonamo, P. M., Grierson, J. D., & Shear, W. A. (1988). Oribatid mite fossils from a
terrestrial Devonian deposit near Gilboa, New York. Journal of Paleontology, 62, 259–269.
Paterson, J. R., García-Bellido, D. C., Lee, M. S. Y., Brock, G. A., Jago, J. B., & Edgecombe, G. D.
(2011). Acute vision in the giant Cambrian predator Anomalocaris and the origin of compound
eyes. Nature, 480, 237–240.
Pepato, R., da Rocha, C. E. F., & Dunlop, J. (2010). Phylogenetic position of the acariform mites:
Sensitivity to homology assessment under total evidence. BMC Evolutionary Biology, 10, 235.
http://www.biomedcentral.com/1471-2148/10/235.
Plotnick, R. (1996). The scourge of the Silurian seas. American Paleontologist, 4, 2–3.
Raff, R. A. (1996). The shape of life: Genes, development and the evolution of animal form.
Chicago: University of Chicago Press.
Regier, J. C., Shultz, J. W., Zwick, A., Hussey, A., Ball, B., Wetzer, R., Martin, J. W., & Cunningham,
C. W. (2010). Arthropod relationships revealed by phylogenomic analysis of nuclear protein-
coding sequences. Nature, 463, 1079–1084.
Savory, T. (1977). Arachnida. New York: Academic Press.
Schaefer, I., Norton, R. A., Scheu, S., & Maraun, M. (2010). Precambrian mites colonized land and
formed parthenogenetic clusters. Molecular Phylogenetics and Evolution, 57, 113–121.
Schmidt, A. R., Jancke, S., Lindquist, E. E., Ragazzi, E., Roghi, G., Nascimbene, P. C., Schmidt, K.,
Wappler, T., & Grimaldi, D. A. (2012). Arthropods in amber from the Triassic Period.
Proceeding of the National Academy of Sciences of the USA, 109, 14796–14801. doi:10.1073/
pnas.1208464109. www.pnas.org/cgi.
Scholtz, G., & Edgecombe, G. D. (2006). The evolution of arthropod heads: Reconciling morpho-
logical, developmental and palaeontological evidence. Development Genes and Evolution, 216,
395–415.
Selden, P. A. (1993). Arthropoda (Aglaspidida, Pycnogonida and Chelicerata). In M. J. Benton
(Ed.), The fossil record 2 (pp. 297–320). New York: Chapman & Hall.
Selden, P., & Read, H. (2008). The oldest land animals: Silurian millipedes from Scotland. Bulletin
of the British Myriapod & Isopod Group, 23, 36–37.
Shear, W. A., & Kukalová-Peck, J. (1990). The ecology of Paleozoic terrestrial arthropods: The
fossil evidence. Canadian Journal of Zoology, 68, 1807–1834.
Shultz, J. W. (1993). Muscular Anatomy of the Giant Whipscorpion Mastigoproctus giganteus
(Luca) (Arachnida: Uropygida) and its Evolutionary Significance. Zoological Journal of the
Linnean Society, 108, 335–365.
Sissom, W. D. (1990). Systematics, biogeography, and paleontology. In G. A. Polis (Ed.), The biology
of scorpions (pp. 64–160). Stanford: Stanford University Press.
Snodgrass, R. E. (1952). A textbook of arthropod anatomy. Ithaca: Comstock.
Stanley, S. M. (1973). An explanation for Cope's rule. Evolution, 27, 1–26.
Talarico, G., Lipke, E., & Alberti, G. (2011). Gross morphology, histology, and ultrastructure of
the alimentary system of Ricinulei (Arachnida) with emphasis on functional and phylogenetic
implications. Journal of Morphology, 272, 89–117.
Tollerton, V. P. (1989). Morphology, taxonomy, and classification of the order Eurypterida,
Burmeister, 1843. Journal of Paleontology, 63, 642–657.
van der Hammen, L. (1989). An introduction to comparative arachnology. The Hague: SPB Academic
Publishing.
Weygoldt, P. (1998). Evolution and systematics of the chelicerata. The Third Symposium of the
European Association of Acarologists, Amsterdam. Experimental & Applied Acarology, 22, 63–79.
Weygoldt, P., & Paulus, H. F. (1979). Untersuchungen zur Morphologie, Taxonomie und
Phylogenie der Chelicerata. 2 Cladogramme und die Entfaltung der Chelicerata. Zeitschrift fur
Zoologische Systematik und Evolutionforschung, 17, 177–200.
Chapter 3
Systematic and Morphological Survey

To work with mites (Fig. 3.1), one must first understand their morphology and
classification. Other than the need for a good microscope, the major hurdle to mite
identification is the scattered nature of the literature. This disarray includes both the
variety of languages and locations of the relevant papers and the babel of jargon
used to describe mite morphology. Fortunately, in recent years mite morpho-speak
has become more homogeneous and several excellent texts exist that provide clear
and comprehensive discussions of mite morphology. For those interested in learning
to identify mites, we recommend the ‘bible’ of acarologists: A Manual of Acarology
3rd Edition (Krantz and Walter 2009). For those who speak English and who wish
to become proficient in the identification of mites, The Acarology Laboratory at The
Ohio State University runs a series of short courses during the Northern Hemisphere
summer. The Acarology Summer Program has been in existence for over 50 years
and both authors learned much of what they know about mite taxonomy there. For
an overview of acarological resources see Google or Walter and Proctor (2010).
This book is not intended as an identification manual but anyone who works with
mites needs to know something about their taxonomy and morphology. In this chap-
ter we briefly review the major orders of mites and their functional morphology. At
the end of this chapter you’ll find a dichotomous key to the major groups that uses
the characters and their states described below.

What Is ‘Acari’? The Question of Mite Monophyly

Mites have traditionally been treated as a single taxon within the Arachnida, usually
ranked at subclass, and called either Acari or Acarina. Within the Acari two or three
groups with highly divergent morphologies have long been recognised: the
Acariformes (also called Actinotrichida) with about 42,000 named species and
Parasitiformes (Anactinotrichida including Opilioacarida) with about 13,000 spe-
cies. Lindquist et al. (2009) place each of these two lineages at the rank of

D.E. Walter and H.C. Proctor, Mites: Ecology, Evolution & Behaviour: Life at a Microscale, 39
DOI 10.1007/978-94-007-7164-2_3, © Springer Science+Business Media Dordrecht 2013
40 3 Systematic and Morphological Survey

Fig. 3.1 Assorted Acari: Mites are easier to recognise than define and the few similarities
(e.g. hexapod larvae) in their diversity of forms may mask separate origins of two lineages that are
not sister groups (SEMs by DE Walter)

superorder (Table 3.1). However, as we discuss below, and has recently been indicated
by molecular phylogenetic studies (Dabert et al 2010; Pepato et al. 2010; Regier
et al. 2010), there is relatively little evidence that these two groups are closely
related. Rather, they appear to represent two independent origins of a ‘mite-grade’
of evolution, characterized by small size, presence of two tagmata not shown by
other arachnid groups (gnathosoma and idiosoma), reduced expression of abdomi-
nal segmentation, and a life-cycle with a six-legged larval stage and eight-legged
nymphs and adults. Major morphological differences between the two groups
include basic subdivisions of the body, location and structure of spiracles and tra-
cheae, the architecture of the venter of the gnathosoma (see Fig. 3.8), the presence
or absence of setae that are birefringent under polarized light (the basis for the
names Actinotrichida [birefringent setae present] and Anactinotrichida [absent]),
and striking differences in sperm ultrastructure (see Table 3.2). Dunlop and Alberti
(2007) provide detailed discussion on this topic. Given the increasingly strong support
for the diphyly of ‘Acari’, it is likely that each of these two superorders will soon be
raised to the same rank as other major groups within the Arachnida.

Parasitiformes: Ticks and Their Relatives

The origin of the name Parasitiformes is easy to understand – many species are
important parasites of vertebrates. All share a subcapitulum formed from the median
fusion of the pedipalpal coxae (Figs 3.2a, 3.3d, 3.5d, 3.8a), a tracheal system with
Parasitiformes: Ticks and Their Relatives 41

Table 3.1 A survey of the major taxa of the Acari: Superorders, orders, suborders, and infraorders
(Modified from Lindquist et al. 2009)
Superorder Acariformes
Order Sarcoptiformes – (Ordovician?) Devonian (410 mya) to recent
Suborder Endeostigmata – 10 families of early derivative, ‘segmented’ mites; fossils from 410 mya
Suborder Oribatida – beetle mites, moss mites, box mites, and Astigmata
Infraorder Enarthronotides – 18 families of early derivative oribatids; morphology highly
variable, and some are brightly coloured, but most have subdivided notogasters and
often mineralized exoskeletons; rostrum variably developed; mostly parthenogenetic;
fossils from 392 mya
Infraorder Palaeosomatides – 6 families of lightly sclerotized, atypical oribatids with fully
exposed chelicerae (rostrum absent) that retain some structures lost in others (e.g. a pair
of median ocelli in Loftacarus); mostly bisexual; fossils from 375 mya
Infraorder Parhypsomatides – 3 families of weakly sclerotized oribatid mites with
well-developed opisthosomal glands; rostrum poorly developed; parthenogenetic; fossils
from 326 mya
Infraorder Mixonomatides – 9 families including most of the box mites; rostrum and opisthosomal
glands usually well developed; bisexual and parthenogenetic; fossils from 49 mya
Infraorder Desmonomatides – about 240 families of nothrine (6 families), brachypyline
(160 families) – the typical beetle and moss mites, also including freshwater and
intertidal mites – and astigmatine (77 families) – cheese, dust, fur, nest, phytotelmata and
intertidal mites; rostrum present traditional oribatids, but reduced in Astigmata; most have
well developed opisthosomal glands; bisexual and parthenogenetic; fossils from 197 mya
Order Trombidiformes – Devonian (410 mya) to recent
Suborder Sphaerolichida – 2 families of bizarre and poorly known mites
Suborder Prostigmata – ~150 families of mostly soft-bodied mites linked by the presence of a
tracheal system (secondarily lost in some)
Infraorder Labidostommatides – 1 family of heavily sclerotized ambush predators; bisexual;
fossils from about 45 mya
Infraorder Eupodides – 20 families of mostly soft-bodied predators, parasites (insects,
slugs, birds), fungivores, plant-parasites (earth, gall, rust mites), intertidal and marine
mites; bisexual and parthenogenetic; fossils from 410 mya
Infraorder Anystides – about 87 families of mostly soft-bodied terrestrial predators and
predators/ protelean parasites (Parasitengona) (hosts include vertebrates, arthropods)
including red velvet mites, chiggers and the water mites; fossils from 115 mya
Infraorder Eleutherengonides – about 42 families of soft-bodied to well sclerotized
predators, parasites (insects, lizards, mammals, birds), fungivores, plant-parasites (spider
mites, broad mite), intertidal and marine mites; fossils from 112 mya
Superorder Parasitiformes
Order Opilioacarida – Opilioacarans; 1 family; fossils from 49 mya
Order Holothyrida – Holothyrans; 3 families; no fossil record
Order Ixodida – ticks; 3 families; fossils from 100 mya
Order Mesostigmata – mesostigmatans; ~100 families; fossils from 49 mya
Suborder Sejida – 5 families of predators and arthropod ectosymbionts, vertebrate (lizards)
parasites; no fossils record
Infraorder Sejina – 4 families of mostly free-living predators; heteromorphic deutonymphs
are phoretic on insects; bisexual
(continued)
42 3 Systematic and Morphological Survey

Table 3.1 (continued)


Infraorder Heterozerconina – 2 families associated with millipedes, centipedes, lizards;
bisexual; male Heterozerconidae have spermatodactyl-like structure
Suborder Trigynaspida – 28 families of predators, ectosymbionts and parasites; no fossil record
Infraorder Cercomegistina – 5 families of mostly tropical, free-living predators or phoretic
associates of beetles; bisexual
Infraorder Antennophorina – 23 families of mostly tropical, free-living predators,
ectostymbionts of arthropods, parasites of ants, termites, and lizards; some phoretic on
insects as adult females; bisexual
Suborder Monogynaspida – ~77 families of predators, ectosymbionts and parasites; fossils
from 49 mya
Infraorder Uropodina – ~35 families of mostly tropical, free-living or phoretic predators and
fungivores and parasites of ants; phoretic as heteromorphic deutonymphs; bisexual
Infraorder Gamasina – ~42 families of mostly free-living or phoretic predators and
fungivores and parasites of ants, bees, moths, katydids, true bugs, lizards, snakes, birds
and mammals; phoretic as homeomorphic deutonymphs or inseminated adult females;
bisexual or parthenogenetic; male Dermanyssiae have spermatodactyl, Parasitiae a
spermatotreme

Table 3.2 Some morphological and behavioural characters that differ between the Acariformes
(*absent in some derived groups) and Parasitiformes (** absent in ticks, some derived groups).
Also see Alberti (2006)
Acariformes Parasitiformes
Setae birefringent under polarized light (except for Setae not birefringent under polarized
solenidia) light
Palp without apotele Palp with (sub)terminal apotele
(= claw remnant)**
Pharynx crescent-shaped in cross-section Pharynx triradiate (Y-shaped) in
cross-section
Median eyes and tubercle (naso) present* Median eyes and naso absent
Tritosternum (mentum or labium) incorporated into Tritosternum emergent, separate from
subcapitulum (but functionally integrated with)
subcapitulum**
Leg coxae fused to venter Leg coxae freely articulated (partially
fused in ticks)
Coxae cover sternal region for legs I–IV Sternal regions of legs II–IV distinct and
usually covered by one or more plates
Anterior 4 limb-bearing segments (cheliceral, Sejugal furrow absent, carapace entire
pedipalpal, legs I–II) separated from remainder of
body by sejugal furrow
Prodorsum with trichobothria* Prodorsum without trichobothria
Larva with osmoregulatory Claparède’s organs Larva without with Claparède’s organs
(=urstigmata)*
Genital papillae added ontogenetically* Genital papillae absent
Typical Malpighian tubules absent Malpighian tubules present
(continued)
Parasitiformes: Ticks and Their Relatives 43

Table 3.2 (continued)


Acariformes Parasitiformes
Stigmatal openings and tracheae absent in basal Stigmatal openings above or behind leg
groups; multiple origins in derived lineages bases and tracheae present in all
post-larval instars
Sperm transferred on stalked spermatophores (basal), Sperm transferred by chelicerae directly
direct insemination by a variety of methods to female genital opening (basal) or to
(including an aedeagus to a secondary sperm secondary sperm induction pore
induction pore) evolved numerous times (derived)
Spermatozoa simple and never vacuolated Spermatozoa complex and vacuolated
(vacuoles absent in some derived taxa)

stigmatal openings associated with the leg bases, the absence of both median eyes
and prodorsal trichobothria, and a distinct sternal region between the mostly freely
articulating leg coxae. Additionally, most have clearly three-segmented chelicerae
(unclear in Ixodida) and at least the basal members have a protruding tritosternum
(absent in Ixodida). The four suborders of parasitiformans are the Ixodida,
Holothyrida, Opilioacarida and Mesostigmata.
The Ixodida are ticks (Fig. 3.2) – perhaps the most familiar of all mites because of
their large size and bloodthirsty habits. All feed on the blood of vertebrates with each
stage attaching to a host for a period of slow feeding, usually over a period of several
hours to several days, followed by more rapid engorgement. They have specialised
chelicerae modified for cutting into the host’s skin and a well-developed anterior
hypostomal process covered in backward-pointing (retrorse) teeth (Fig. 3.2a) that
helps them to anchor their mouthparts during feeding and to resist host grooming. The
tarsi of the first pair of legs has a pair of pits (Haller’s organ Fig. 3.2b) with sensory
setae that detect body heat, carbon dioxide, and other host chemicals. A tritosternum
is not present. Males use their chelicerae to transfer sperm in a sac-like spermatophore
directly to the female’s genital opening (see Chap. 5). Sonenshine (1991) provides an
excellent modern review of tick morphology and biology, and their taxonomy has
been summarized by Guglielmone et al. (2010). Three families of ticks are currently
recognised and fewer than 900 species have been described.
The Nutalliellidae consists of one species Nuttalliella namaqua that lives
among rock crevices in southern Africa and feeds on lizards. Its body is leathery
and stretches during the short period of engorgement, with any excess water
excreted from the anal opening. Current analysis indicates that Nutalliellidae is
the sister group to all other ticks (Mans et al. 2011). The second family, Argasidae
(‘soft ticks’), contains about 200 species of ticks with leathery, sack-like bodies,
a sclerotized dorsal plate only in the larva, and usually numerous post-larval
instars that each feed only once before moulting. Adults may feed repeatedly
between bouts of egg-laying, but a few species do not feed as adults and produce
eggs from reserves acquired as nymphs. Bird, bats, and other mammals that live
in permanent or semi-permanent shelters are especially likely to be attacked by
argasid ticks, but some have become permanent ectoparasites of large mammals
44 3 Systematic and Morphological Survey

Fig. 3.2 Ticks (Ixodida) (a) Subcapitulum of male Winter Tick showing retrorse teeth; (b) Haller’s
Organ of tarsus I (Ixodes holocyclus); (c) Dorsal habitus female ixodid tick; (d) venter of soft tick
(Argasidae, Carios sp.) (Images by DE Walter)

(Kierans 2009). Feeding and engorgement are similar to that observed in N. namaqua,
but argasid ticks have more mechanisms for excreting excess water.
More than 700 species of ticks belong to the family Ixodidae (‘hard ticks’) and
these include the most important acarine vectors of disease in wildlife, livestock,
and people. In contrast to the other two families, ixodid ticks have a sclerotized
dorsal plate (scutum) in all stages and a plicate cuticle capable of both distension
during feeding and of growth without moulting (neosomy). Only three active stages
are present: larva, nymph, and adult. Except in a few specialists, each stage feeds on
a different host individual (3-host ticks) before dropping off and moulting. As a
result, an adult ixodid tick generally has had two extra chances of picking up and
transmitting novel disease-causing microbes, in addition to any that it may have
been acquired from its mother (transovarial transmission) and maintained through
the moults (transstadial transmission).
Parasitiformes: Ticks and Their Relatives 45

Fig. 3.3 Holothyrida (Allothyridae, Allothyrus sp.) (a) Lateral view of adult; (b) Gnathosoma
with chelicerae extended; (c) Peritreme; (d) Subcapitulum (short arrow = corniculus, long
arrow = deutosternal groove); (e) Quadrigynaspid genital shields of female (pg pregenital, lg latig-
ynal, mg mesogynal) (SEMs by DE Walter)

The Holothyrida (Fig. 3.3) also contains three families but these include fewer
than 30 described species. Adult holothyrans are large (2–7 mm long), reddish to
purplish mites completely encased in armour, but the larva and three nymphal stages
are mostly soft-bodied with at most a few small plates. Larvae are hexapod, but have
vestigial legs IV (Klompen 2011). As in ticks, a maximum of one pair of lateral eyes
and one pair of stigmata are present on the idiosoma. Holothyrans are difficult to
collect, poorly studied mites, known only from Australia and New Zealand
(Allothyridae), the Neotropics (Neothyridae) and islands that are continental frag-
ments in the Indian and Pacific Oceans (Holothyridae). Their current distribution
appears to represent vicariance on a grand scale – the break-up of the ancient super-
continent Gondwana. Allothyrids scavenge on dead arthropods and, like ticks, feed
only on fluids – haemolymph and externally digested tissues (Walter and Proctor
1998a). Species of Allothyrus have a pair of elongate, seta-like tritosternal laciniae
46 3 Systematic and Morphological Survey

Fig. 3.4 Dorsal habitus of


opilioacaran. These large,
long-legged and active mites
have a leathery prodorsum,
but are otherwise soft-bodied
and very atypical
Parasitiformes (Image by DE
Walter)

originating from a common base. Sperm transfer has not been observed but may be
similar to that in ticks (Chap. 5). Phylogenetic studies usually recover the Holothyrida
as the sister group of the ticks, and within Holothryida, the Allothyridae as the sister
group of Neothyridae + Holothyridae (Klompen 2011; Mans et al. 2011).
The Opilioacarida (Fig. 3.4), as their name implies, resemble small opilionids
(harvestmen, daddy-long-legs) and could easily be taken for the most primitive of
living mites. Appearances can be deceiving, though, and molecular analyses usually
place them either right in the middle of the Parasitiformes or as the sister group of
Holothyrida + Ixodida. To date about three dozen species (including two Tertiary
fossils) have been discovered and all are placed in a single family (Opilioacaridae).
Most opilioacarans have been found under rocks in Mediterranean, monsoonal or
other seasonally dry climates in Europe, Africa, Asia and Australia, but some are
known from rainforest litter in Central America.
Opilioacaran adults are usually 2–3 mm in length with purple-striped, brownish
grey bodies and long legs banded with white. The cuticle is leathery and lacks the
sclerotised plates that are found in most other groups of mites. Two or three pairs of
lateral eyes are present but median eyes are not. No trichobothria are present on the
prodorsum or other locations on the body; however, trichobothria have been reported
on leg femora for one species. These structures may be homologous with the trichoboth-
ria of other arachnids or may be convergently derived (as in some Mesostigmata). The
opisthosoma of the adult has four pairs of antero-lateral tracheal openings – a unique
condition in arachnids. The almost dorsal placement of the spiracles is reflected in
another name sometimes used for this group: Notostigmata (noto = back).
Opilioacarans feed on pollen, fungi and small arthropods. Large three-segmented
chelicerae are used to grasp food, and large serrated rutella (hypertrophied setae on
either side of the buccal opening) saw it off into bit-sized chunks (see Fig. 13.12a, b
Parasitiformes: Ticks and Their Relatives 47

Fig. 3.5 Tritosterna (arrows), i.e. the sternal element of legs I, in some arachnids: (a) Simple
plate (labium) with setae in a spider (Araneae); (b) Plate with pair of flagellate setae in a
micro-whipscorpion (Schizomida, Apozomus sp.); (c) Plate with flagellate setae in a larval hooded
tick-spider (Ricinulei, Cryptocellus sp.); (d) Two finger-like projections in a larval opilioacaran
(Parasitiformes, Opilioacarida). c chelicera, pc pdeipalpal coxae, I coxa of leg I (Illustrations by
DE Walter)

for an analogous system). Ingested food forms boluses in the guts of the mites.
Female Neacarus texanus from Big Bend National Park in Texas lay one large egg
at a time but the reproductive biology of these animals is otherwise unknown.
So, like basal acariform mites (see below), Opilioacarans ingest solid foods and
have well-developed rutella, but other characters support a relationship to parasiti-
form mites. These include the loss of the median eyes, the retention of a well-
developed basal cheliceral segment, the lack of prodorsal trichobothria, the lack of
genital papillae and the presence of a protruding tritosternum. A tritosternum
(literally, the sternum of the third segment, i.e. the one that bears legs I) is present
in most arachnid groups. Usually it is plate-like and is then sometimes called the
labium (Fig. 3.5), but it may have seta-like processes that extend under the
48 3 Systematic and Morphological Survey

palpcoxae (Fig. 3.5b, c). In opilioacarans the tritosternum consists of two separate
finger-like processes (Fig. 3.5d). Whether these processes are homologous with the
tritosternum of other parasitiform mites is unclear, but the larvae of Australian
Allothyridae also have separate tritosternal processes (Walter, unpublished).
Opilioacarans feed on solid food, not fluids as in other Parasitiformes and most
arachnids. In predatory Mesostigmata, the biflagellate tritosternum holds the under-
side of the flow of prey juices and helps to pump the liquid food up the deutoster-
nal grove towards the filter of the internal malae and into the pharynx. Opilioacarans
ingest solid food and if this condition is derived compared to their ancestors, then
the tritosternum may be secondarily regressed.
Although opilioacarans appear to be the most “primitive” members of the
Parasitiformes, it may well be that they are a relatively high branch on the tree.
Characteristics of these very active mites, such as the large size, hypertrichy, lack of
sclerotized plates, rutellum, rudimentary tritosternum, four pairs of respiratory
openings and ability to regenerate lost limbs, may be derived from a more plodding,
fluid-feeding and armoured ancestor. As with the holothyrans, a crucial aspect of
the biology of opilioacarans is still unknown: how do they transfer sperm? If these
mites are good Parasitiformes, then they would use their chelicerae.
The Mesostigmata (Fig. 3.6) is the most diverse and broadly distributed subor-
der of Parasitiformes with about 70 families divided into 12 major lineages. About
11,000 species have been described, of which about half are free-living predators in
soil-litter, rotting wood, compost, herbivore dung, carrion, nests, house dust or simi-
lar detritus-based systems. These predators are usually abundant and voracious
enough to regulate the populations of other small invertebrates in their habitats. No
swimming mesostigmatans are known but some species live in the intertidal zone or
the margins and surface films of freshwater streams and lakes and may spend a
considerable time submerged (see Chap. 7).
A few species of mesostigmatans, in several families, have switched from preda-
tion to grazing on fungi and some species ingest fungal spores and hyphae. Some
feed on pollen (digesting the exine externally and ingesting only fluids), nectar and
other plant fluids. Pollen feeding is common in the Phytoseiidae, a family that has
successfully colonised the leaf-surface habitat and accounts for about 20 % of
described species of Mesostigmata. Phytoseiid mites are used with great success as
classical biological control agents and are the dominant family of predatory mites
on plants in most terrestrial ecosystems (see Chap. 8). The remainder of the
described mesostigmatans are parasites. Sixteen of the 17 families in the superfam-
ily Dermanyssoidea are composed entirely of parasites of mammals, birds, reptiles
and arthropods. Dermanyssoids range from facultative ectoparasites to obligate
ectoparasites to obligate internal parasites of vertebrate ear canals and respiratory
passages (see Chap. 9).
Mesostigmatans have robust, often clearly predatory (Fig. 3.7), and sometimes
very long three-segmented chelicerae. Most are fluid-feeders with membranous pro-
cesses associated with the mouthparts that filter out particulate matter. As in ticks, the
gnathosoma resembles a small head (capitulum) with a basal, sclerotised ring formed
from the fusion of the palp coxae and gnathosomal tectum. The body is often
Parasitiformes: Ticks and Their Relatives 49

Fig. 3.6 Adult female Mesostigmata. (a) Phytoseius oreillyi (Phytoseiidae) an arboreal predator;
(b–d) Soil-inhabiting predators: (b) Asca sp. (Ascidae), (c) Sejus sp. (Sejidae), (d) Stratiolaelaps
scimitus (Laelapidae); (e) a parasite of birds, the Tropical Fowl Mite Ornithonyssus bursa
(Macronyssidae); (f) Lantana Flower Mite Proctolaelaps lobatus (Melicharidae); (g) a soil mite
with unknown feeding habits (Uropodoidea, Clausiadinychus sp.) (SEMs by DE Walter)

dorso-ventrally flattened, oval to subrectangular in shape and protected by dorsal and


ventral sclerotised shields, at least in the adults. Mesostigmatans lack any indication
of eye lenses; however, pigment spots can be seen on the anterolateral margins of the
idiosoma in some lightly sclerotised, plant-inhabiting species. The distal part of the
first pair of legs has a well-developed field of sensory setae that is sometimes within
a pocket-like depression (reminiscent of Haller’s organ in ticks) and may or may not
have paired claws and an ambulacrum. No trichobothria are present, but a few spe-
cies of Veigaia (Veigaiidae) have what appear to be functionally similar modified
50 3 Systematic and Morphological Survey

Fig. 3.7 Chelicerae of some very toothsome mesostigmatans: (a) Megisthanus sp. (Trigynaspida);
(b) Epicroseius sp. (Sejida); (c) Proarctacarus oregonensis (early derivative Gamasina);
(d) Athiasiella sp. (early derivative Dermanyssiae) (SEMs by DE Walter)

setae on the legs. Sperm transfer is by means of the chelicerae. Males of early derivative
groups of Mesostigmata transfer sperm directly to the female genital opening with
unmodified chelicerae; however, a more elaborate system of chelicera-based transfer
has evolved in some groups (see Chap. 5). Eggs are produced one at a time in the
derived and most diverse infraorder, Gamasina (hyporder Dermanyssiae), but two,
four or a few dozen eggs are matured concurrently in other suborders.

Acariformes: The Mite-Like Mites

The Acariformes is by far the more diverse of the two superorders of mites, with
over 40 thousand described species and a fossil history going back at least 410 mil-
lion years. Derived character states that appear to link this vast diversity include a
subcapitulum composed of a plate-like tritosternum (mentum – Fig. 3.8b) fused to
the lobes of the pedipalpal coxae in more derived groups (e.g. Fig. 3.13d), the pres-
ence of ontogenetically added genital papillae (Fig. 3.9), the reduction of the basal
cheliceral segment so that the chelicerae appear two-segmented (although clearly
three-segmented in some Oribatida), the fusion of the coxae to the body wall cover-
ing the sternal region, and the presence of prodorsal trichobothria (Fig. 3.13e)
composed of a cup-like bothridium and a variously modified seta-like structure
called a bothridial sensillum or (in the oribatid literature) sensillus (ss in
Fig. 3.13e). Gas exchange via tracheal tubes has evolved repeatedly within derived
groups, but the ancestral acariform mite (and many extant species) appears to
have exchanged gases directly through its cuticle. Other possible synapomorphies
include the division of the body into a four-segmented head-like region composed
of the segments bearing the chelicerae, pedipalps, and first two pairs of walking legs
Acariformes: The Mite-Like Mites 51

Fig. 3.8 Basic differences in organization of ventral gnathosoma in (a) Parasitiformes (Gamasina,
Antennolaelaps sp.) and (b) Acariformes (Oribatida, Parhypochthonius aphidinus). c corniculus,
dg deutosternal gutter, pc palpcoxa, r rutellum, t tritosternum or mentum (SEMs by DE Walter)

Fig. 3.9 Genital papillae of


the box mite Eupththiracarus
flavus (left series, dissected
and numbered anterior to
posterior). Scale bar = 0.1 mm
(SEM by DE Walter)
52 3 Systematic and Morphological Survey

(proterosoma or ‘fore body’) and a hysterosoma (‘latter body’) composed of the


idiosoma and the segments bearing the last two pairs of legs. Currently, the
Acariformes is considered to contain two suborders, the Sarcoptiformes (Fig. 3.10)
and the Trombidiformes (Fig. 3.11).
Basal within the Sarcoptiformes is a paraphyletic collection of 10 families with
barely more than 100 described species (including three genera and five species
known only from fossils). These are collectively known as the Endeostigmata.
Most endeostigmatan genera have cosmopolitan distributions and are often found
in thermally or nutritionally extreme habitats, especially deserts (cold and hot),
sands and deep soils. One or two pairs of trichobothria occur on the prodorsum,
and one or two pairs of simple ocelli may be present. Some species also have a
median ocellus under a projection of the anterior prodorsum (the naso).
Chemosensory setae (solenidia) are present on the legs. Most of the setae and
other cuticular structures have a core of material that glows under polarised light
(actinopilin, and hence an alternative name for the order, Actinotrichida). The
chelicerae are two-segmented and most species have well-developed rutella that
act as shears to clip off segments of food for swallowing. The few endeostigmatans
that have been studied ingest solid foods such as fungi, algae and soft-bodied
microinvertebrates like nematodes, rotifers and tardigrades. However, the more
derived groups (Nanorchestidae, Nematalycidae) appear to be fluid-feeders.
Most other Sarcoptiformes ingest solid foods, primarily bacteria and fungi,
microinvertebrates and bits of decomposing vegetation. The most species-rich
group of sarcoptiformans is the Oribatida (a.k.a. Oribatei, Cryptostigmata).
Soil, leaf litter, lichens, mosses and forest canopies are home to more than ten
thousand known species of oribatid mites. Many of these mites are completely
encased in sclerotized cuticle as adults and, along with strongly sclerotized mites
in other groups (e.g. Uropodina, Scutacaridae), have acquired common names
referring to their resemblance to beetles, turtles, or tortoises. Some oribatid mites
also strengthen their cuticle with a variety of calcium compounds (Norton and
Behan-Pelletier 2009).
Most oribatid mites are found in the organic layers of the soil-litter system, in
mosses and lichens, or on the surface of vascular plants where they feed as micro-
bivores, fungivores, algivores and detritivores. The immature stages are especially
likely to be found burrowing in decomposing plant matter and some tunnel into
living plant tissue. Other oribatids that appear to feed mostly on fungi or algae
also are opportunistic predators or necrophages (see Chap. 6). Many oribatid
mites have colonized the trunks of trees and shrubs, and especially in humid cli-
mates, can be found in soils suspended in the canopies of forests and even on the
surface of leaves (Chap. 8). Bogs, ponds, lakes, streams, rivers and seashores have
been invaded by other lineages of oribatid mites that live a mostly or entirely
aquatic existence (Chap. 7).
A monophyletic lineage usually given a separate rank, the Astigmata, appears
to have arisen from within the Oribatida (Norton 1998). Early derivative astig-
matans somewhat resemble the immature stages of oribatid mites and have simi-
lar feeding habits. These mites also have colonized many aquatic habitats
Acariformes: The Mite-Like Mites 53

Fig. 3.10 Representative Sarcoptiformes. (a) Endeostigmata: (Spelerochestes sp.). (b–f)


Oribatida: (b) Synchthonius crenulatus, (c) Gehypochthonius sp., (d) Phthiracarus borealis, (e)
Camisia biverrucata; (f) Oribatella jacoti. (g–h) Astigmata: (g) a house dust mite Dermatophagoides
farinae); (h) male (viewer’s right) and female rosella feather mite Rhytidelasma punctata in pre-
copulo (SEMs by DE Walter)
54 3 Systematic and Morphological Survey

Fig. 3.11 Representative Trombidiformes (Prostigmata): (a–e) Soil-inhabiting predators: (a)


Rhagidiidae, (b) Labdiostommatidae, (c) Bdellidae, (d) Cunaxidae, (e) Cheyletidae); (f) Unknown
feeding habits (Cryptognathidae); (g–h) Aquatic predators (g) Homocaligidae, (h) Water mite,
Aturidae; (i) Chigger (Trombiculoidea); (j) Spider mite (Tetranychidae); (k) Gall mite
(Eriophyidae); (l) Bee associate (Scutacaridae); (m) Soil-living fungivores (Pygmephoridae)
(SEMs by DE Walter)
How Do Mites Do the Things They Do? 55

including the intertidal zone and accumulations of water on tree holes and pitcher
plants (phytotelmata). Others have invaded the nests of social insects, burrows
made by beetles and other arthropods in living and dead wood, and rich accumu-
lations of organic matter such as dung, compost, and beach wrack where they
have complex associations with their insect hosts (Chap. 9). However, members
of the astigmatan hyporder Psoroptida are primarily associated with vertebrates
and nest-building insects including bees. Scab and mange mites, as well as house
dust mites, stored product mites, feather mites and some fur mites belong to the
Psoroptida (see Chap. 10).
The Trombidiformes do everything the Sarcoptiformes do and more. Many are
parasitic or ectocommensal on vertebrates and invertebrates; however, it is the
plant-parasitic trombidiformans that cause the most economic damage. All of
the major acarine plant parasites belong to the Trombidiformes, including spider
mites, flat mites, earth mites, gall mites, erinose mites and bud mites (see Chap. 8).
These plant-parasitic mites are found in 10 families in three trombidiform lineages.
The remaining 140 or so families contain predators, fungivores, algivores and animal
parasites. The largest and most spectacular lineage, the hyporder Parasitengonina,
includes more than 11,000 described species of terrestrial and aquatic mites.
The parasitengone life cycle is complex, involving a parasitic larva, two pupa-like
resting stages and two active predatory stages (see Chap. 4). Within this group,
velvet mites and water mites are conspicuous because of their size (some > 1 cm
long) and often brilliant colours. At the opposite end of the size scale, larval chiggers
(or scrub itch mites) are pesky inhabitants of meadows and grasslands whose bite
can result in welts and a maddening itch, or worse, transmit microorganisms that
cause scrub typhus.

How Do Mites Do the Things They Do?

Morphology and behaviour are the costumes and scripts in ‘the ecological theatre
and the evolutionary play’ (apologies to G.E. Hutchinson), so some understanding
of mite morphology is a prerequisite for any discussion of their behaviour. Mites
are arthropods with external skeletons and a maximum of six pairs of jointed limbs
and, as with all arthropods, the limbs are the key to understanding behaviour.
Below we review the morphology of mites in terms of how external structures are
used to get through life.

Sensing, Feeding, Silk and Sex: The Gnathosoma

At the front end of the mite body, the chelicerae and palps are used primarily in
capturing, tasting and ingesting food. The earliest mites probably had chelate–dentate
chelicerae, i.e. shaped like pincers with toothed opposed surfaces. The pincer is
56 3 Systematic and Morphological Survey

Fig. 3.12 (a–f) Oribatida: (a–b) Anterior view of strongly chelate-dentate chelicerae and robust
rutella of Meristacarus sp., an oribatid mite that burrows in decaying wood; (c) Brachychthonius
sp. an oribatid mite with curiously modified mouthparts and unknown feeding habits; (d) Ventral
gnathosoma of a suctobelbid mite where the rutella are modified into elongate supports for the
whip-like chelicerae; (e) Bothridial sensillum of Brachychthonius sp.; (f) Lateral habitus of female
Hermannidae showing extruded ovipositor. (g) Lateral view of gnathosoma of a fluid-feeding
prostigmatan (Stereotydeus sp.). b bothridium, c corniculus, m mentum, p pedipalp, r rutellum.,
ss sensillus (bothridial sensillum) (SEMs by DE Walter)

formed from a distal segment, called the movable digit, that articulates against a
dorsal extension of the second cheliceral segment (the fixed digit).
Because most arachnids are predators that reduce their prey to fluids before
ingestion, it is thought that predation and fluid-feeding were the ancestral behav-
iours in mites (Krantz 1978). However, it is possible that the earliest acariform
mites were more inclined to scavenge and ingest solid food, as do today’s horse-
shoe crabs and sea spiders. The earliest known fossil mites from the early to mid-
Devonian are primarily from groups whose modern descendants (Sarcoptiformes)
swallow microbes, bits of decomposing organic matter and minute, soft-bodied
invertebrates, such as nematodes (Walter 1988a; Walter and Proctor 1998a). Their
chelicerae (Fig. 3.7b, 3.12a, b) are basically two-segmented and often fairly stout
– useful organs for cutting organic matter into small bits before swallowing – and
are assisted by a pair of hypertrophied seta-like structures called rutella (singlu-
lar, rutellum). Chelicereae and rutella are often modified in mysterious ways (e.g.
Fig. 3.12c, d) that have not yet been related to feeding behaviours. In contrast, the
How Do Mites Do the Things They Do? 57

Trombidiformes feed on fluids: rutella are lost in the Prostigmata and the chelicerae
are often modified into organs for stabbing prey or hosts (e.g. fungi, plants) and
withdrawing fluids.
In the Mesostigmata and Holothyrida, however, the chelicerae are three-
segmented, relatively slender, chelate–dentate (Fig. 3.7). Most of these mites are
general predators or scavengers that use their chelicerae to rip holes in the sides of
other arthropods or to grasp, mash and chew small arthropods or nematodes. The
chelicerae can be inserted into the wounds and moved rapidly back and forth as they
shred internal tissue. The corniculi (also called external malae) (Fig. 3.12) usually
support and protect the salivary stylets – the ducts that introduce digestive enzymes
into the prey body – and may be modified to support different diets. For example,
those Mesostigmata that have re-evolved the ability to ingest solid foods often have
one or more teeth on their corniculi (e.g. Fig. 3.13d), which presumably help to col-
lect fungal spores or clip-off bits of hyphae or arthropod cuticle. In the predators,
the tritosternum is usually strongly biflagellate and plumose and is used to promote
fluids up the deutosternal gutter and through the filter of the highly fimbriate inter-
nal malae. The deutosternal gutter typically has numerous rows of small teeth in
these predators (Fig. 3.13a). Mesostigmata that feed on fungi or pollen usually have
reduced numbers of teeth in the deutosternal gutter (Fig. 3.13c, d) and often have
modified corniculi.
Among modern mites, all known Opilioacarida, most other free-living Parasitiformes
and many Acariformes retain the chelate–dentate condition. However, all ticks
and most of the other parasitic lineages within the Parasitiformes have modified
cheliceral digits that pierce or slash tissue and often have membranous corniculi
(Fig. 3.13e). In the Acariformes, mites that are parasites of vertebrates, arthro-
pods and plants tend to have needle, hook or stylet-like mouthparts. Plant para-
sites, such as spider mites, have long, whip-like stylets (see Fig. 8.4) capable of
probing beneath the epidermis of a leaf and stabbing individual cells. Many suc-
cessful lineages of trombidiform predators and fungivores also have piercing-
sucking mouthparts. They often have movable digits modified into stout, sharp
spines capable of penetrating arthropod cuticle or hyphal walls. Many of these
modified cheliceral forms are distinctive and can provide a good indication of
how their owners earn a living.
Chelicerae are also used in non-feeding roles. In agonistic (sometimes cannibalistic)
encounters, the chelicerae of mesostigmatans can be used to thrust and parry
(Lindquist and Walter 1989). Male spider mites attack and even kill predatory mites
by stabbing with their cheliceral stylets (Saito 1986a, b). Chelicerae can also be
used to hold onto a larger-bodied animal during phoretic travel.
Males of four separate lineages of Mesostigmata have modifications of a chelic-
eral digit to facilitate transfer of sperm from the male to the female (see Chap. 5). In
the suborder Gamasina, males in the Dermanyssiae have an often elaborate process
on the movable digit called a spermatodactyl (‘sperm finger’ – see Fig. 5.3) that
delivers sperm to a secondary sperm pore in the female. In the Parasitinae, males
have a slit-like process in a similar position on the movable digit that holds a packet
of sperm, the spermatotreme (‘sperm hole’ – see Fig. 5.2) – a seemingly intermediate
58 3 Systematic and Morphological Survey

Fig. 3.13 In the Mesotigmata the form of the structures on the gnathosoma often reflect feeding
biology. (a–b) Soil-ihnabiting general predators: (a) Gammasellodes sp. (Ologamasidae); (b)
Lasioseius sp. (Blattisociidae); (c) Arboreal predator Neoseiulus sp. (Phytoseiidae); (d) Fungivore-
omnivore Proctolaelaps sp. (Melicharidae); (e) Blood-sucking parasite of rodents Ornithonyssus
bacoti (Macronyssidae). c corniculus, ch chelicera, dg deutosternal gutter, im internal malae,
ph opening to pharynx, s salivary stylet, t tritosternum (SEMs by DE Walter)

step on the way to a spermatodactyl. In the suborder Trigynaspida, males in the


Celaenopsoidea have a ventral clip-like process that appears to play a similar sper-
matophore-grasping role (apparently, mating has not been observed in these mites).
In the suborder Sejida, males in the Heterozerconidae have an often elaborate pro-
cess on the movable digit that resembles the spermatodactyl in the Dermanyssiae.
The palps are used to sense food and handle prey during feeding. They typically
resemble small legs but may be antenniform (e.g. Bdellidae), raptorial (e.g.
Cunaxidae, many water mites), armed with a claw-like structure (e.g. spider mites,
Trombidiodea) or reduced to stubs (e.g. Tarsonemidae). They are usually well supplied
How Do Mites Do the Things They Do? 59

with mechanoreceptive setae, and clusters of sensory setae at the tips of the palps
presumably allow mites to taste their food before and during feeding. Many meso-
stigmatans tap the substrate with their palps as they walk.
Some spider mites (Tetranychidae) have large unicellular glands that extend
from within the body to the spinneret – a hollow seta at the tip of each palp that is
used to spin silk (see Fig. 8.4). Early derivative spider mites use silk primarily for
protection of eggs and to make a protective tent for moulting, but among the true
spider mites, silk is used throughout the life cycle: egg and moulting covers are
spun, active stages live in or under silken webs, males sequester quiescent female
deutonymphs under a silken mat and silken strands may be used as droplines and to
facilitate aerial dispersal.
Parasitiform mites do not produce silk, but the production of silk from the
gnathosoma appears to be a basic ability of acariform mites, although apparently
lost in the Oribatida (except, perhaps, for signal threads leading to spermato-
phores produced by some oribatid mites – see Chap. 5) and many derived
Trombidiformes. Silk seems to have been used originally for protecting eggs and
quiescent immature stages. For example, adult females of the endeostigmatan
Alicorhagia fragilis (Alicorhagiidae) weave a silken platform on which the eggs
are laid, and larvae and nymphs spin a protective cocoon around themselves
before becoming quiescent and undergoing a moult (Walter 1988a). The use of
silken egg platforms or cases and/or cocoons to protect moulting stages also
occurs in several trombidiform lineages: Eupodina (Eupodidae, Bdellidae,
Cunaxidae, Tydeidae, Eriophyidae), Anystina (Adamystidae, Anystidae) and
Raphignathina (Cheyletidae, Harpirhynchidae, Camerobiidae, Tetranychidae).
Except in the Tetranychidae (see above) the fine silken strands originate within the
buccal opening from modified salivary gland secretions (Gerson 1985). Some
predatory prostigmatans also use silk in prey capture (see Chap. 6).

Moving, Sensing and Interacting: The Legs

In most Parasitiformes and in many Prostigmata, the first pair of legs are used like
antennae and are often slender, elongate and lack well-developed claws. Like the
palps, the first pair of legs usually have clusters of sensory setae near their tips.
When walking, many mites perform characteristic leg-waving movements.
Sometimes these movements are stereotypic enough to be diagnostic. For example,
the morphologically similar oribatid mites Oppiella nova and Oppia nitens, which
are commonly sympatric in mesic soils, are easily distinguished by the characteris-
tic leg-waving behaviour of O. nitens that is absent in O. nova (Seniczak 1975).
The first pair of legs (= legs I) are often used to size up a potential sexual partner
or prey item through tentative tapping movements. Legs may also be used in prey
capture. In some predatory mites, the first two pairs of legs are used in conjunction
with the palps and chelicerae to form a basket-like arrangement of limbs to entrap
small arthropods (Coineau 1973; Lindquist and Walter 1989; Proctor and Pritchard
60 3 Systematic and Morphological Survey

1990). In the Epicriidae (Mesostigmata), the long slender legs I are used like fishing
poles with sticky tips: setae on the tarsi carry a drop of glue at their tips that is used
to snare prey and drag them back into the grip of the chelicerae (Alberti 2010).
In some mites the first pair of legs are modified for clinging to a host during pho-
retic dispersal. But the second, third and fourth pairs of legs are the primary organs
of locomotion in mites, and, when the first pair are primarily sensory in function,
mites are functionally hexapod. However, in many acariform and some parasitiform
mites the first pair are also used in walking. Most water mites use all four pairs of
legs for crawling but the last pair may be trailed like stabilisers during swimming.
The first two acariform mite instars (prelarva, larva) are morphologically hexapod,
i.e. their fourth pair of legs are not expressed, but larvae are functionally quadripedal
when the first pair of legs are used exclusively for sensing. In a few groups of
plant-parasitic mites, some legs are absent in all stages (e.g. the fourth pair in
Larvacarus, Tenuipalpidae; the second and fourth pair in the Eriophyoidea). The most
extreme reduction in legs known occurs in some animal symbiotic mites such as
the insect-parasitic Podopolipidae. Larvae and adult males in this family are often
hexapods and adult females can be reduced to sac-like bodies with only remnants of
the first pair of legs.

Reproduction

Mites have a bewildering diversity of sperm transfer methods (see Chap. 5). In
males, every appendage that can be modified for reproduction has been. Direct
transfer of sperm via apposition of genital openings has evolved independently sev-
eral times in the Acari but this type of direct transfer is relatively uncommon in
mites, most of which use a more indirect route. As previously mentioned, derman-
yssine parasitiform males have spermatodactyls on their chelicerae to channel
sperm into the female’s sperm receiving opening. In other parasitiform species, the
male simply picks up a sperm packet (spermatophore) from his genital opening
with unmodified chelicerae and places it in the female's reproductive opening.
In water mites there are examples of species that have the first, second, third or fourth
pair of legs modified for holding and transferring sperm packets (Proctor 1992).
In the Hydracarina and many Mesostigmata, males have legs modified for holding
females during mating, or for attacking rival males. Tarsonemid mites capture quiescent
pre-adult females and carry them on a suction-cup process on their posterior,
although over-enthusiastic suitors sometimes sequester developing males, rather
than females (Garga et al. 1997). Finally, there are many terrestrial and aquatic
mites in which males and females never come into contact during sperm transfer;
rather, males leave spermatophores on a substrate and females pick these up later.
Even stranger are the many cases in which mites can do without sperm: reproduction
by virgin females (parthenogenesis) is very common in mites (see Chap. 5).
Female mites have fewer modifications for gamete – or rather zygote – extrusion.
Most mites lay eggs on a substrate without having special genital structures for
How Do Mites Do the Things They Do? 61

oviposition. However, many oribatid females have an extrusible ovipositor


(Fig. 3.13f) for probing into soil and litter niches (Krantz 1978), and females of the
water mite genus Hydrachna use their ovipositors to push eggs into the air spaces of
aquatic plant tissues. Females of another water mite, Unionicola, have long spines
on their genital valves that prise apart tissues of the sponges and mussels in which
these mites lay their eggs.

Digestion and Excretion

The basic organisation of the chelicerate digestive system is strikingly different


from that of mandibulate arthropods (Snodgrass 1948; Manton 1977). Chelicerates
do not have grinding jaws and their pincer-like chelicerae lie above a mouth opening
that is roofed by a small plate, the epistome (‘above the mouth’), that has a projecting
lip (the labrum). In horseshoe crabs, the mouth opens backwards – towards a pocket
formed by the spine-covered coxae of the pedipalps and of the second (see Fig. 2.10b)
and third pair of legs (gnathobases). Shellfish, worms and other creatures captured
by the pincers on the limbs of horseshoe crabs are crushed and killed by the spiny
coxae. The strongest pair of crushing coxae are those on the fourth pair of legs, at
the posterior end of the prosoma; these are capable of cracking open relatively
large clams (Manton 1977). As the coxae subdue and shred the food, pieces are
moved into the mouth by the pincer-like chelicerae. A muscular gizzard completes
the masticatory process and the food passes down the oesophagus into the midgut
where it is further digested before being forced into the radiating branches of the gut
(diverticula or caeca) where epithelial cells engulf the food particles and complete
the digestion.
In most arachnid orders, the coxae are not gnathobasic and the mouth points
forward into a pre-oral opening formed by the pedipalpal coxae, epistome and
labrum. The gizzard is replaced by a pharynx and only fluids enter the midgut –
sucked through filtering structures around the pre-oral opening by the dilation and
constriction of pharyngeal muscles (for an especially fine analysis of the digestive
system in Ricinulei, see Talarico et al. 2011). In this regard, most parasitiform and
trombidiform mites are typical arachnids. Digestion begins externally as enzymes
are pumped into the mangled body of a prey or a plant cell and continues in the
midgut and gastric caeca. The branching pattern of gastric caeca is often visible in
mites as digestive products accumulate in them.
Sarcoptiform mites, however, have a very different feeding mode. Chelicerae,
assisted by rutella (analogous to the spiny gnathobases in horseshoe crabs) on the
hypostome, cut off pieces of food and move them into the mouth. No gizzard is
present, so additional comminution of the food does not occur. In fact, food frag-
ments can often be identified in the gut contents of these mites. A ball of food
accumulates at the base of the oesophagus and is compacted into a food bolus that
gradually makes the transit of the midgut to the anal opening. The gastric caeca in
these mites are generally lobe-like and simple in structure. Pigments or other
62 3 Systematic and Morphological Survey

digestive products may accumulate in the lobes but the bulk of the ingested mate-
rial remains in the bolus (covered by a peritrophic membrane). Up to three boluses
in various stages of digestion are commonly present in actively feeding oribatid
mites. At the end of its journey, the bolus is expelled as a faecal pellet through a
relatively large anal opening covered by a pair of trapdoor-like valves. Opilioacarans
seem to feed in a similar manner and a few mesostigmatans have secondarily
evolved to feed on pieces of foods and also have relatively large anal openings, e.g.
some species of Proctolaelaps (‘anus hurricane’).

Keeping It All In: The Cuticle

The exoskeleton of a mite consists of numerous layers of varying composition


(see Krantz and Walter 2009 for details), but to the casual observer mites seem to
range from soft and translucent to hard and shiny. In general, immature stages are
more soft-bodied than adults, but specialized dispersal stages such as the astigmatan
hypopus tend to be better protected than free-living stages. Males are often more
heavily armoured than females and are often armed with spines and processes for
combat with other males or for holding on to females. Heavily armoured mites (see
Chap. 6) tend to be dull-coloured, often shiny browns or black, but are sometimes
bright yellow, orange or red (see Fig. 8.7). Soft-bodied mites are often translucent
(appearing white), but also come in a range of colours. Red coloration probably
denote the presence of carotenoids that protect the mites from ultraviolet rays and
may confer some protection from predators with colour vision (see Chap. 7).
The outermost layer of the cuticle is usually a layer of secretions of waxes and
proteins called the cerotegument. The primary purpose of the cerotegument appears
to be water retention, a thin sealant on the surface of the cuticle. In some mites,
however, the cerotegument is very thick and may be strongly ornamented (Fig. 3.14c, d)
or rough and deeply fissured (e.g. Fig. 3.10e). In others it may be sticky, allowing
the mite to cover itself in a layer or soil particles (see Fig. 6.12b), as in Oriflamella
lutulenta (Fig. 3.14a, b) or other detritus. In a variety of oribatid mites found in dry
microhabitats (e.g. rocks, tree trunks, tree holes), the cerotegument is covered by a
field of minute (typically 1–2 μm in diameter and height) pustules with a crystalline
microstructure (Fig. 3.15). It seems likely that these waxy crystalloids have a water
repellent effect similar to the “Lotus-effect” known for similar structures on the
leaves of some pond plants (Walter 2009).

Identifying Mite Superorders and Orders

Although this is not an identification manual, a good way to introduce the major
groups of mites is to provide a dichotomous key. Keys are logical tools for mak-
ing choices among characters to arrive at an identification. Character choices are
Identifying Mite Superorders and Orders 63

Fig. 3.14 The wax layer or cerotegument is often a thin layer that reduces water loss from the
cuticle, but is sometimes sticky, with an adherent layer of soil or detritus, or very thick and form-
less to highly ornamented. (a) The ambush predator Oriflamella lutulenta (Mesostigmata,
Ologamasidae) covered in a layer of soil. (b) Posterior dorsal shield cleaned of soil and showing
setae inserted on candlestick-like tubercles that hold the soil layer. (c) Eupterotegaeus rostratus
(Oribatida, Eutegaeoidea) with thick, ornate cerotegument in place. (d) E. rostratus with cerotegu-
ment removed. Scale bars = 100 µm (Images by DE Walter)

presented as pair of character sets (couplets). Each choice leads to another


choice until the user reaches the correct answer – or an incorrect answer or the user
gives up in frustration! More modern methods of identification include interac-
tive electronic keys based on matrices of taxa and character states (see Walter
and Winterton 2007 for a review). Although interactive keys may be easier to use
and more likely to lead to a correct identification, they cannot yet be loaded into
books or papers. Most of the literature available for identification of mites is
based on dichotomous keys.
In the following key, each couplet has two leads and each lead has a series of
character states given in parallel with the other. These characters may not be present
in all members of each group (e.g. larval mites lack any respiratory structures) and
64 3 Systematic and Morphological Survey

Fig. 3.15 Waxy crystalline


pustules on the notogaster of
a gymnodamaeid mite
(Roynortonella sp.). At the
adult moult, the cerotegument
appears wet and smooth, but
ornamentation, including
these pustules, appears as the
coating dries. The pustules
probably carry drops of water
off the mite, much as similar
structures do on the leaves of
the Water Lotus. Scale
bars = 1 µm (inset = 100 nm).
(SEMs by DE Walter)

may not be visible on each and every individual (things can be secondarily lost or
simply broken off), but they should work for most specimens. Usually the taxonomy of
a group and its keys are mostly based on a specific life stage, e.g. the adult female, adult
male, parasitic larva, or phoretic deutonymph. In the following key, adults (especially
females) will key most easily, but with luck, other post-larval stages may key out.

Key to the Superorders and Orders of the Acari

1. With one or all of the following characters: (a) filamentous tracheal system
opens from 1–4 pairs of stigmata located above the bases of the legs, often asso-
ciated with a papillate peritreme or a single stigma on sieve plate behind each
leg IV; (b) chelicerae usually chelate-dentate with 3 distinct segments includ-
ing a relatively long basal piece, but sometimes modified into a complex slicing
organ or stylet-like; (c) subcapitulum formed from the median fusion of the
pedipalpal coxae and often with a median gutter; (d) a finger-like tritosternum
usually associated with the median gutter and often ending in a pair of whip-like
laciniae; (e) pedipalpal coxae often continue dorsally to form a ring-like gna-
thosoma; (f) distinct sternal region present ventrally between the freely articu-
lating bases of the legs … Superorder Parasitiformes (go to 2)
– With one or all of the following characters: (a) tracheal system either
absent; restricted to short, blind sacks on the legs; emanating from the
genital region; or opening from a single pair of stigmata located at the
bases of the chelicerae or anterior margin of the body, in the latter case a
gutter-like peritreme may be present and sometimes elaborated into a
maze over the cheliceral bases; (b) cheliceral base absent or reduced to
Identifying Mite Superorders and Orders 65

tiny remnant so that they appear composed of only two segments, che-
late-dentate or highly modified into various stabbing and sucking organs;
(c) subcapitulum formed from the fusion of a plate-like tritosternum
(mentum) with pedipalpal coxae, never with a median gutter; (d) pedipal-
pal coxae never continue dorsally to form a ring-like gnathosoma; (e)
sternal region occupied partly or entirely by the fusion of the bases of the
legs (coxae) to the body … Superorder Acariformes (go to 5)
2 (1). Chelicerae modified into complex slicing structures with a distal pair of cut-
ting organs; tritosternum never present; subcapitulum without a median gut-
ter, but with a hypostomal process covered with backward pointing teeth …
Order Ixodida (ticks)
– Chelicerae composed of three distinct segments, usually chelate-dentate,
but sometimes stylet-like; flagellate tritosternum often present; subcapit-
ulum usually with a distinct median gutter and not prolonged into an
anterior process covered with backward pointing teeth … go to 3
3 (2). Adults usually well sclerotized with an anterior shield over the region bear-
ing the legs, which is often extended to cover most of or the entire body;
venter similarly protected by a series of shields; flagellate tritosternum
often present; one pair of tracheal openings situated above the bases of the
legs; blind or with one pair of lateral simple eyes … go to 4
– Adults without sclerotized shields; tritosternum a pair of non-flagellate,
finger-like processes; more than one pair of tracheal openings situated
above the bases of the legs (4 pairs in adults, less in other stages); 2–3
pairs of lateral eyes … Order Opilioacarida
4 (3). Adults usually with these characters: (a) small to medium-sized mites, rarely
over 1 mm in length, and usually with extensive areas of soft cuticle ven-
trally; if completely encased in shields, then female genital opening cov-
ered by 1–3 shields; (b) tritosternum usually present and associated with
median subcapitular gutter with rows of teeth; (c) ocelli never present; (d)
subcapitulum with 4 pairs of setae … Order Mesostigmata
– Adults usually with these characters: (a) large mites (often over 1 mm
long) with the dorsum covered by a dome-shaped shield and the venter
completely covered by various shields, including 4 shields over the
female genital opening; (b) tritosternum usually absent, but if present,
then weakly formed and subcapitular gutter without teeth; (c) one pair of
lateral ocelli present or absent; (d) subcapitulum with 5 or more pairs of
setae … Order Holothyrida
5 (1). Usually with all or most of these characters: (a) chelicerae chelate, usually
chelate-dentate, rarely attenuate, never fused together; (b) rutella present;
(c) tracheal system absent or composed of short sack-like invaginations on
legs or body, rarely arising from genital region, peritremes never present;
(d) prodorsum with 0–2 pairs of trichobothria, but idiosoma and legs
66 3 Systematic and Morphological Survey

without trichobothria; (e) legs ending in 1–3 simple claws, rarely median
claw ray-like or lateral claws toothed, but tenant hairs not present; (f) eyes
present in only some early derivative taxa … Order Sarcoptiformes
– Usually with some of these characters: (a) chelicerae rarely chelate-
dentate, usually highly modified into a variety of needle-like, knife-like
or stylet-like structures, often partially or completely fused into a sin-
gle functional unit; (b) rutella never present; (c) tracheal system some-
times absent, but usually produced from stigmata at base of chelicerae
or on anterior margin of body, rarely arising from genital region, perit-
remes associated with gnathosoma may be present and sometimes
highly elaborate; (d) prodorsum with 0–2 pairs of trichobothria; idio-
soma and legs sometimes with trichobothria; (e) legs ending in diverse
array of claws, pads and rayed structures, often including tenant hairs;
(f) eyes often present … Order Trombidiformes

Summary

Two major lineages of mites are alive today and the basic differences among these
‘'superorders’' are often more striking than the similarities. All mites have an ante-
rior section resembling a tiny head (the gnathosoma or capitulum) that includes
the two limbs devoted to feeding: the chelicerae that lie above the oral opening,
and the pedipalps, whose fused coxae form the floor and sides to the opening to
the digestive system. The remainder of the body is more or less fused into a sac-
like idiosoma that contains organs of cognition, digestion, excretion and repro-
duction. Except in the Sarcoptiformes, movement is usually hexapod in post-larval
stages, with the first pair of legs being used as feelers. Parasitiform and trombidi-
form mites resemble other arachnids in that they feed primarily on fluids includ-
ing blood, plant fluids or externally digested tissues. Opilioacarid mites, most
Sarcoptiformes and a few Mesostigmata, however, consume particles of food:
fungal spores, pollen grains or bitten-off bits of larger food items. The great vari-
ety of feeding and reproductive habits exhibited by the Acari will be discussed in
detail in the following chapters.

References

Alberti, G. (2006). On some fundamental characteristics in acarine morphology. – Atti della


Accademia Nazionale Italiana di Entomologia. Rendiconti LIII, 2005, 315–360.
Alberti, G. (2010). On predation in Epicriidae (Gamasida, Anactinotrichida) and fine-structural
details of their forelegs. Soil Organisms, 82, 179–192.
Coineau, Y. (1973). Les Caeculidae (Acariens Prostigmates) quelques aspects de leurs particularités
éco-éthologiques. Bulletin D’Écologie, 4, 329–337.
References 67

Dabert, M., Witalinski, W., Kazmierski, A., Olszanowski, Z., & Dabert, J. (2010). Molecular
phylogeny of acariform mites (Acari, Arachnida): Strong conflict between phylogenetic signal
and long-branch attraction artifacts. Molecular Phylogenetics and Evolution, 56, 222–241.
Dunlop, J. A., & Alberti, G. (2007). The affinities of mites and ticks: A review. Journal of
Zoological Systematics and Evolutionary Research, 46, 1–18.
Garga, N., Proctor, H., & Belczewski, R. (1997). Leg size affects mating success in Tarsonemus
confusus Ewing (Prostigmata: Tarsonemidae). Acarologia, 38, 369–375.
Gerson, U. (1985). Webbing. In A. W. Helle & M. W. Sabelis (Eds.), Spider mites, their biology,
natural enemies and control (Vol. 1, pp. 223–232). Amsterdam: Elsevier.
Guglielmone, A. A., Robbins, R. G., Apanaskevich, D. A., Petney, T. N., Estrada-Peña, A., Horak,
I. G., Shao, R., & Barker, S. C. (2010). The Argasidae, Ixodidae and Nuttalliellidae (Acari:
Ixodida) of the world: a list of valid species names. Zootaxa, 2528, 1–28.
Kierans, J. E. (2009). Order Ixodida. In G. E. Krantz & D. E. Walter (Eds.), A manual of acarology
(3rd ed., pp. 111–117). Lubbock: Texas Tech University Press.
Klompen, H. (2011). Holothyrids and ticks: new insights from larval morphology and dna sequenc-
ing, with the description of a new species of Diplothyrus (Parasitiformes: Neothyridae).
Acarologia, 50(2), 269–285. doi:10.1051/acarologia/20101970.
Krantz, G. W. (1978). A manual of acarology. Corvallis: Oregon State University Bookstores.
Krantz, G. W., & Walter, D. E. (Eds.). (2009). A Manual of Acarology (3rd ed.). Texas Tech
University Press. 807 pages, 338 b/w illustrations; 60 figures ISBN 978-0-89672-620-8. http://
www.ttup.ttu.edu/BookPages/9780896726208.html.
Lindquist, E. E., Krantz, G. W., & Walter, D. E. (2009). Classification. In G. E. Krantz & D. E. Walter
(Eds.), A manual of acarology (3rd ed., pp. 97–103). Lubbock: Texas Tech University Press.
Lindquist, E. E., & Walter, D. E. (1989). Biology and description of Antennoseius janus, new spe-
cies (Mesostigmata: Ascidae), a mesostigmatic mite exhibiting adult female dimorphism.
Canadian Journal of Zoology, 67, 1291–1310.
Mans, B. J., de Klerk, D., Pienaar, R., & Latif, A. A. (2011). Nuttalliella namaqua: A Living Fossil
and Closest Relative to the Ancestral Tick Lineage: Implications for the Evolution of Blood-
Feeding in Ticks. PLoS One, 6(8), e23675. doi:10.1371/journal.pone.0023675.
Manton, S. M. (1977). The arthropoda. Oxford: Clarendon.
Norton, R. A. (1998). Morphological evidence for the evolutionary origin of Astigmata (Acari:
Acariformes). Experimental & Applied Acarology, 22, 559–594.
Norton, R. A., & Behan-Pelletier, V. M. (2009). Suborder Oribatida. In G. E. Krantz &
D. E. Walter (Eds.), A manual of acarology (3rd ed., pp. 430–564). Lubbock: Texas Tech
University Press.
Pepato, R., da Rocha, C. E. F., & Dunlop, J. (2010). Phylogenetic position of the acariform mites:
Sensitivity to homology assessment under total evidence. BMC Evolutionary Biology. 10, 235.
http://www.biomedcentral.com/1471-2148/10/235.
Proctor, H. C. (1992). Mating and spermatophore morphology of water mites (Acari: Parasitengona).
Zoological Journal of the Linnean Society, 106, 341–384.
Proctor, H. C., & Pritchard, G. (1990). Prey detection by the water mite Unionicola crassipes
(Acari: Unionicolidae). Freshwater Biology, 23, 271–279.
Regier, J. C., Shultz, J. W., Zwick, A., Hussey, A., Ball, B., Wetzer, R., Martin, J. W., & Cunningham,
C. W. (2010). Arthropod relationships revealed by phylogenomic analysis of nuclear protein-
coding sequences. Nature, 463, 1079–1084.
Saito, Y. (1986a). Prey kills predator: Counter-attack success of a spider mite against its specific
phytoseiid predator. Experimental and Applied Acarology, 2, 47–62.
Saito, Y. (1986b). Biparental defence in a spider mite (Acari: Tetranychidae) infesting Sasa bam-
boo. Behavioural Ecology and Sociobiology, 18, 377–386.
Seniczak, S. (1975). Revision of the family Oppiidae Granjean, 1953 (Acarina, Oribatei).
Acarologia, 17, 331–345.
Snodgrass, R. E. (1948). The feeding organs of Arachnida, including mites and ticks. Smithsonian
Miscellaneous Collections, 110, 1–93.
Sonenshine, D. E. (1991). Biology of ticks (Vol. 1). New York: Oxford University Press.
68 3 Systematic and Morphological Survey

Talarico, G., Lipke, E., & Alberti, G. (2011). Gross morphology, histology, and ultrastructure of
the alimentary system of Ricinulei (Arachnida) with emphasis on functional and phylogenetic
implications. Journal of Morphology, 272, 89–117.
Walter, D. E. (1988). Predation and mycophagy by endeostigmatid mites (Acariformes:
Prostigmata). Experimental & Applied Acarology, 4, 159–166.
Walter, D. E. (2009). Genera of Gymnodamaeidae (Acari: Oribatida: Plateremaeoidea) of Canada,
with notes on some nomenclatorial problems. Zootaxa, 2206, 23–44.
Walter, D. E., & Proctor, H. C. (1998). Feeding behaviour and phylogeny: Observations on early
derivative Acari. Experimental & Applied Acarology, 22, 39–50.
Walter, D. E., & Proctor, H. C. (2010). Mites as modern models: Acarology in the 21st century.
Acarologia (Paris), 50, 131–141. doi:10.1051/acarologia/20101955.
Walter, D. E., & Winterton, S. L. (2007). Keys and the crisis in taxonomy: Extinction or reinven-
tion? Annual Review of Entomology, 52, 193–208.
Chapter 4
Life Cycles, Development and Size

As mammals, we receive a lot of help from mom in the form of milk and protection
during our early postpartum years, and of course many months of room and board
inside the womb before then. Maternal care is likewise widespread among arach-
nids and mothers often carry eggs and developing young. Some horseshoe crab
females also carry their eggs until they hatch (Shipley 1909). On land, arachnid
mothers often build a burrow, a silk-lined chamber or a silken egg sac (spiders) in
which eggs or hatchlings are protected. Many scorpions, whipscorpions, sun scorpi-
ons, spiders and pseudoscorpion females guard their young after hatching until they
are fully active and ready to begin hunting. In contrast, paternal care appears to be
much rarer among arachnids, but is known from five families of harvestmen in the
superfamily Gonyleptoidea (Proud et al. 2011). For example, males in the neotropi-
cal genus Zygopachylus construct a nest into which females place their eggs after
mating. The males clean the eggs of fungal parasites and ward off potential preda-
tors, especially other opilionids (Mora 1990). Other cases of paternal care have been
reported (Martens 1993) and some female opilionids also guard their eggs and
young juveniles (Mitchell 1971; Ramires and Giaretta 1994; Proud et al. 2011);
however, most opilionids limit maternal care to the use of ovipositors to hide eggs
in crevices or in the soil (Preston-Mafham and Preston-Mafham 1993).
From what we know of mites, parental care beyond the oviposition stage is rela-
tively rare. When it comes to relative reproductive effort, speed of development and
modifications of ontogeny, however, mites are without peer. In this chapter we
review the life cycles of mites and end with a discussion of the relationships between
life history parameters, generation times and size (Fig. 4.1).

D.E. Walter and H.C. Proctor, Mites: Ecology, Evolution & Behaviour: Life at a Microscale, 69
DOI 10.1007/978-94-007-7164-2_4, © Springer Science+Business Media Dordrecht 2013
70 4 Life Cycles, Development and Size

Fig. 4.1 (a) A predatory mesostigmatan taking a Tortoise Mite (Prostigmata, Scutacaridae) for a
ride. (b) Venter of a scutacarid mite: most are associated with ants, bees, wasps, and beetles, but
some are found on mites; (c) Details of anterior end with hooked claws on leg I (arrow) that is used
to attach to a carrier (SEMs by DE Walter)

Oviposition

Almost all mites begin their lives as eggs, laid singly or en masse. Fertilization may
occur just before egg-laying, or eggs with relatively well developed embryos or
even fully developed larvae (ovolarviparity) may be oviposited (Bergmann and
Heethoff 2012). Occasionally, mite eggs have the shape and aspect of miniature
hen’s eggs but can also be spherical, baroquely ornamented or bizarrely formed. For
example, the eggs of ornate false spider mites (Tuckerellidae) are scale-like red
flecks covered with waxy ribs and horns. The eggs of bdellid snout mites (Bdellidae)
are hairy balls that look more like fungal fruiting structures than eggs. Variations of
shape, colour and ornamentation are common enough for it to be often possible to
differentiate eggs of sympatric species (Pérez 1996). Whatever their eggs’ appear-
ance, however, mites are usually very particular about where they lay them.

Parental Care

Maternal care in mites is usually limited to hiding eggs in crevices, laying them
within silken webs, covering them with detritus particles or a jelly-like sheath, or
conveniently dying with them protected inside the mother’s body. For example, red-
legged earth mites, Halotydeus destructor (Penthaleidae), feed on pasture plants
during the autumn and winter. As spring approaches, the females die, their bodies
filled with diapausing eggs. The eggs remain within the female’s body until they
Parental Care 71

hatch the next autumn (Wallace 1970). What appears to be a similar behaviour
occurs in some mites with well-sclerotised cuticles. For example, it is relatively
common to find egg-filled cadavers of female oribatids, and larvae sometimes hatch
from these eggs, but it is not clear how easy it is for them to escape from what is
probably a maternal tomb (Norton 1994). Oribatids that lay eggs with fully devel-
oped prelarvae or larvae are also known and in Archegozetes longisetosus, the ovi-
duct appears to function as facultative brood chamber for the developing eggs
(Bergmann and Heethoff 2012). Parasitiform mites with apparently fully developed
larvae still encased within the egg shell are not uncommon and adult female
Macrocheles schaeferi have been reported to assist the larva to escape from the egg
during hatching (Walter 1988b).
Actual brooding behaviour where a female sits over her egg clutch until they
hatch is known only in argasid ticks. One bat parasite, Antricola marginatus, not
only broods her clutch, but the hatched larvae clamber on top of the females and are
carried around the bat guano habitat in which the nymphs and females live (Labruna
et al. 2012). Antricola are very unusual ticks in that only the larvae are blood feeders
and the guano-inhabiting stages apparently derive their nutrition from the faecal
matter itself. It seems likely that the females may eventually crawl up the walls of
the caves to release their larvae near the bats, but this has not yet been observed.
Another case of possible parental care in mites is the oribatid Damaeus verticil-
lipes (Oribatida, Damaeidae). In this species ovipositing females may lay their eggs
on the backs of conspecific males (Cancela da Fonseca 1975). Males of several other
species of Damaeidae have been observed with attached eggs, but none appear to have
been studied. A few species of halacarids were thought to carry eggs attached to
their posterior legs until they hatch, but these ‘eggs’ have been shown to be protozoans
of the subclass Peritichia (Krantz and Walter 2009).
Carrying or guarding eggs for longer than described above is rarely reported
among acarines, but more complex relationships have been reported in the
Cheyletidae and Tetranychidae (Saito 2010). The predatory cheyletid Cheyletus
eruditus is reported to guard eggs against intruders (Summers and Witt 1972;
Preston-Mafham and Preston-Mafham 1993). Moreover, adult females of the
parthenogenetic chelytid Hemicheyletia morii congregate around their silken
nests on leaves, thereby providing their eggs with the double protection of silk
and ring of predatory chelytids. Unwary arthropods that wander by are bitten on
a leg and dragged into a swarm of ravenous mites (Mori et al. 1999; Saito 2010).
Similar nest-guarding behaviour is probably widespread in this family because in
Australian forests the cheyletids Oudemansicheyla coprosomae and Hemicheyletia
wellsi also actively defend their eggs. A female lies in ambush at the base of a
leaf, usually along the main vein, and attacks mites and small insects that attempt
to crawl onto the leaf (Walter and Proctor 1999). Usually, 3–4 eggs are brooded
at a time, each within a silken covering. Females stand their ground when dis-
turbed and repeatedly snap their palps and stab with their chelicerae when
probed. Hatchling larvae and protonymphs can often be found near the mother
but they soon disperse out onto the leaf surface. Some subsocial spider mites in
the genus Schizotetranychus s.l. carry this protective behaviour one step further.
72 4 Life Cycles, Development and Size

Both males and females actively defend their young within silken nests from
marauding phytoseiid mites (Saito 1986a, b, 1997, 2010).
Most mites show considerable care in the choice of sites where eggs are laid.
Oribatid mites usually use an extrusible ovipositor (see Fig. 3.13f), sometimes as
long as the female’s body, to place their eggs within crevices. Some water mites
prepare oviposition sites in the stems of aquatic plants with their chelicerae (see
Chap. 7). Female dermanyssine Mesostigmata develop a single large egg at a time.
Each egg fills the opisthosoma, causes considerable distension and may weigh as
much as 40 % of the non-gravid female’s body mass (thus exceeding the kiwi
15–20 % for maternal investment in a single egg). Under favourable laboratory
conditions, females may lay 4–8 eggs, or 1–2 times their body mass, every day.
Before oviposition, the female may spend several minutes investigating cracks and
crevices. The egg exits from the genital opening, which often resembles a trapdoor
hinged posteriorly at about the level of the fourth pair of coxae and covering the
entire intercoxal width. As the egg is extruded anteriorly into the sternal region, the
female bends her gnathosoma to receive it and grasps the egg between the palps
and the dorsum of the gnathosoma. The egg is so large that females often need to
stand on fully extended legs to allow its passage. Using the chelicerae, palps and
first pair of legs, the female carefully places each egg in a crevice or other recess
and may spend several minutes tapping it into place. Particles of soil may be
scraped onto the egg and patted into place but soon the female loses interest and
wanders off, possibly because the pressure of the next swelling egg initiates search
for a new oviposition site. In contrast, engorged hard ticks often lay very large egg
masses after dropping from their hosts, and seemingly have little choice is where
to lay their eggs, but as each egg is laid the female coats it with a waxy secretion
for protection.

Egg Number and Egg Size

The complete gamut of reproductive strategies can be found in the Acari, from the
repeated production of single eggs or egg clutches (iteroparity) to a single burst of
reproduction (semelparity), and often extremes of effort can be found within a single
lineage. For example, adult female argasid ticks typically produce about a hundred
eggs after each feeding event but ixodid ticks have a single mass laying that can
exceed 20,000 eggs (Evans 1992). Very small mites often produce a single egg at a
time and, in most mites, body size and the number of eggs per clutch tend to be cor-
related. For example, Cook et al. (1989) found two strong linear relationships between
egg volume, egg number and adult female volume across 27 species of water mites of
the genus Arrenurus. Each line was explained by the type of larval host. In those spe-
cies parasitising large hosts (dragonflies and damselflies), large clutches of small eggs
were produced, whereas in those species parasitising small hosts (mosquitoes and
midges), females had small clutches with relatively large eggs. Within each group egg
number and female size showed a strong positive relationship.
Egg Number and Egg Size 73

Fig. 4.2 A female quill mite


Dermoglyphus sp. from a
Galapagos Ground Finch
Geospiza fortis that is
carrying a mature larva
(arrows) (Image by HC
Proctor)

The Mesostigmata appear to be an exception to the association between body


size and egg number. Some mesostigmatans associated with arthropods (e.g.
Heterozerconidae, Megisthanidae) may produce 12–20 eggs at a time; however,
within the suborder Dermanyssiae, only a single egg at a time is produced by a
female, irrespective of body size (Walter and Proctor 1999). Within other suborders,
there is some correlation between body size and egg number but the relationship is
not strong. In general, 1, 2 or 4 eggs seem to be the normal range for clutches in
non-dermanyssine Mesostigmata.
In many mite eggs appear to be deposited at an early stage of development.
However, laying eggs with fully developed larvae (ovoviviparity) or giving birth to
fully formed larae (larviparity) is known in a number of free-living and parasitic
mites (Kethley 1990; Evans 1992; Norton 1994; Lange and Tolstikov 1999). Many
dung (e.g. Macrocheles, Zygoseius) or plant-inhabiting (e.g. Phytoseiidae) meso-
stigmatans retain eggs with fully developed larvae within their bodies until just
before hatching and occasionally larviposition occurs. Larviposition may occur in
some feather mites, based on observations of well-developed larvae inside the bod-
ies of females (see Fig. 4.2). Some permanent vertebrate-parasitic Dermanyssoidea
give birth to protonymphs, skipping an active larva entirely (Radovsky 1994).
Several groups of heterostigmatans display physogastry in which all or part of
the idiosoma swells into a grotesquely enlarged sac that eventually accommodates
the developing offspring (Kaliszewski et al. 1995). Females attach to a host, usually
74 4 Life Cycles, Development and Size

an insect or a fungus, and the body swells to many times its original size. In some
species eggs or larvae are produced but in many heterostigmatans, sexually mature
adults emerge from the mother. Depending on the species, from fewer than a dozen
to nearly 400 new adults emerge from the mother. Sex ratios of offspring of mated
females are strongly female-biased, with males typically representing only 1–5 %
of the brood (unmated females produce only male offspring through arrhenotok-
ous parthenogenesis) (see Chap. 5). Mating takes place within the maternal body
or immediately on emergence. Generation times are typically very short and these
mites have some of the highest intrinsic rates of population growth known for mul-
ticellular animals (Bruce and Wrensch 1990).

Postembryonic Development

After hatching, mites – like all arthropods – regularly shed their external skeletons
during development and thereby may change their external appearance. The cast
skins are called exuviae (both singular and plural, as in ‘clothes’). Setae, plates or
other structures (e.g. genital papillae, lyrifissures, pores) may be added sequentially
during development, allowing relatively easy recognition of different stages. Both
Opilioacarida and early derivative Acariformes have what is considered the ances-
tral acarine developmental series: egg, two hexapod larval stages (prelarva, larva),
three ocotopod nymphal stages (protonymph, deutonymph, tritonymph) and adult
(Fig. 4.3). No prelarvae have been reported from the Holothyrida, Mesostigmata, or
Ixodida (Fig. 4.4), although a remnant may remain within the eggs of soft ticks
(Aeschlimann 1984).
When comparing developmental stages, homology is inferred from the changes
in morphological characters of the instar (i.e. the mite between moults), not from
the moult per se. In some ticks and parasitengone mites, one or more additional
moults may occur during development or after sexual maturation that do not result
in the addition of new morphological characters. These are thought to be additions
to the basic life cycle and not remnants of ancestral ontogenies. Perhaps the addi-
tional moults are a means of repairing cuticle or increasing size. In heterostigmtans,
two layers of cuticle may be formed during the moult from larva to adult. The
intermediate cuticle is generally formless and is supposed to be a remnant of one or
more ancestral nymphal instars, retained presumably because of some required
developmental sequence. This is probably the case in the Parasitengonina (see Life
cycle of the Parasitengonina, below). When determining homology among instars,
it is important to compare equivalent stages and not to be misled by the additional
(or missing) moults.
Development is usually gradual, with immature mites more or less resembling
the adults (homeomorphic). In a few mite lineages one instar differs dramati-
cally in form from the others and is said to be heteromorphic. For example,
Astigmata have heteromorphic deutonymphs referred to as hypopi (or hypopo-
des) and Parasitengonina have heteromorphic larvae. In a sense, most acarine
Postembryonic Development 75

Fig. 4.3 Developmental stages in Acariformes: (a) Full acariform life cycle (most Sarcoptiformes,
many Trombidiformes); (b) Tritonymph supressed (many Prostigmata); (c) A single active stage
(larva) between egg and adult (most Heterostigmatina); (d) In some Heterostigmatina development
occurs within the physogastric body of the female and new adult mites emerge from her body.
e egg, pl prelarva, l larva, pn protonymph, dn deutonymph, tn tritonymph, ad adult (Image by HC
Proctor and C. Harvey)

larvae are heteromorphic because they are usually six-legged (hexapod), rather
than eight-legged (octopod) as are nymphs and adults. Prelarvae, when present
and well formed, are also hexapod, and are invariably strongly heteromorphic.
Immature stages and adults of many oribatid mites are quite different in form and
can be difficult to associate without rearing, and are easily confused with astig-
matans by novices. Both are whitish, lightly sclerotised sarcoptiform mites with
sac-like bodies and single claws. Adults of higher oribatids (beetle mites), in
contrast, are strongly sclerotised, typically a shiny brown or deep black in colour,
and often have wing-like lateral processes (pteromorphs) and three claws.
Cuticular growth within an instar (neosomy) is known for chiggers
(Trombiculidae), hard ticks and some parasitic Mesostigmata (Radovsky 1994) but,
76 4 Life Cycles, Development and Size

Fig. 4.4 Developmental stages in Parasitformes: (a) The full parasitiform life cycle is known only
in the Opilioacarida, but a prelarva may be present in the Holothyrida. Soft ticks (Argasidae) some-
times have supernumery nymphal stages, but a prelarva is absent; (b) In the Mesostigmata, only
two nymphal stages are present; additions of setae during development suggest that the final
nymphal stage is suppressed; (c) In hard ticks (Ixodidae), only one nymphal stage is present. e egg,
pl prelarva, l larva, n nymph, pn protonymph, dn deutonymph, tn tritonymph, ad adult (Image by
HC Proctor)

in general, intra-moult increase in size is restricted to swelling or unfolding of previ-


ously deposited cuticle. Although such ancillary moulting is rare in the Acari, the
loss, suppression and reduction of instars is an acarine leitmotiv.

Prelarva and Larva

In most arachnids the first active instar is an eight-legged creature with imperfectly
formed limbs and little or no setation (Kaestner 1968; Foelix 1982; Polis and Sissom
1990; Canard and Stockmann 1993). This ‘larva’ (Foelix 1982), ‘incomplete juve-
nile’ (Ji, Ji1 or JP of Canard and Stockmann 1993) or ‘J1 larva’ (as we will call it)
usually can move but does not feed on external resources. In some of the larger
arachnids (scorpions, spiders, whipscorpions), the J1 larvae climb onto the back of
their mother and undergo further development fuelled by yolk reserves until
Prelarva and Larva 77

moulting into the ‘nymph’ or second instar juvenile (J2). The J2 nymph usually has
functional mouthparts and actively hunts and acquires resources not present in the
egg. In the opilionid Erginulus clavotibialis, a J1 larva with vestigial mouthparts,
eyes, setation and limb segmentation emerges from the egg and moults almost
immediately into an active J2 nymph (Goodnight and Goodnight 1976).
Many spiders and whipscorpions have 1–3 moults that occur after the end of
embryonic development but precede the formation of the J1 larva. These prelarval
or foetal instars are only partially developed and may be retained within the egg
(something that might also happen within eggs of acariform mites– see Shatrov
1999a). Those that burst from the chorion are immobile, have eight legs with incom-
plete segmentation, and lack eyes, claws and setae (Canard and Stockmann 1993).
Modern xiphosuran eggs hatch into a ‘blastodermic cuticle’ that gives rise to the
trilobite larva (Shipley 1909). Possibly the prelarval cuticles in spiders and whip-
scorpions are remnants of planktonic stages in their marine ancestors.
In most mites, the first active instar is a hexapod feeding stage called the larva.
Buds for the fourth pair of legs appear early in embryonic development but then
usually regress (Barnett and Thomas 2012), although in the parasitiform orders
Opilioacarida and Holothyrida unsegmented precursors of the fourth pair of legs are
present in the larva (Klompen 2011). In Opilioacarida and Acariformes, another
stage, the prelarva, precedes the larva. The prelarva of the opilioacarans, like the
larva, has three pairs of segmented legs and bumps where the fourth pair of legs
would arise. The gnathosoma is moderately well developed but no body setae are
expressed (Coineau and van der Hammen 1979; Klompen 2000). It is unclear if
the prelarva emerges from the egg or if it moults to the larva within the eggshell
(Klompen 2000). In endeostigmatans and basal trombidiformans, an inactive
hexapod prelarva pops out of the eggshell but other than this jack-in-the-box
action is usually immobile (but see below). As with the octopod J1 larva in spi-
ders, whipscorpions and opilionids, the hexapod prelarva of mites does not feed
on external resources (yolk may be present) and the chelicerae are poorly devel-
oped or vestigial (Fig. 4.5).
In acarology, stages of ontogenetic development are often called stases
(Grandjean 1938, 1970). In the Acariformes, the prelarva is considered by to be a
regressive stase, i.e. morphologically and behaviourally simplified from a previ-
ously more fully developed form and, in a sense, degenerate. At best, the limbs and
their setation are rudimentary and prelarvae of many species are ‘reduced’ to a
featureless layer of cuticle or are entirely absent. Most early derivative acariformans
have inactive, non-feeding prelarvae. In mite jargon, presumed regressive stases are
referred to as elattostases (‘reduced stages’) if they can move but not feed, and
as calyptostases (‘veiled stages’) if they can neither move nor feed. Calyptostases
often consist of little more than a layer of cuticle within the skin of the previous
stase (apoderm) (Shatrov 1999a, b). In most oribatids (including astigmatans) and
in the endeostigmatans Terpnacarus glebulentus and Alycus roseus, the prelarva is
little more than a layer of cuticle formed before, and within which, the larva develops
(Walter 1988a; Kethley 1992; Lange and Tolstikov 1999). Usually this cuticle is
shed with the eggshell, but in the astigmatan Bonomoia opuntiae the larva emerges
78 4 Life Cycles, Development and Size

Fig. 4.5 Prelarva of an


erythraeid (Prostigmata) mite
removed from its egg shell
(SEM by DE Walter)

from the eggshell encased in the prelarval cuticle (Wirth 2006). Other early derivative
acariformans have a calyptostatic prelarva that barely protrudes from the egg after
it splits, although it may have some hypertrophied setae that protect the protruding
dorsum (e.g. Hybalicus). In Alicorhagia fragilis, the adult female spins a small
sheet web on which she lays her eggs. When the eggs hatch, the legs of the prelarva
expand away from the body, pushing the eggshell remnants away (Walter 1988a).
Other than the, presumably, hydrostatic expansion of the legs, the exposed prelarva
of this species is completely inactive, with rudimentary chelicerae, leg segments
and setation, and resembles the immobile octopod first instar of solfugids (Grandjean
1938; Cloudsley-Thompson 1992).
A few acariformans, however, have an active, ellatostatic prelarva similar to the
J1 larva in opilionids. In the anystid mite Chaussieria venustissima, the prelarva
emerges from the eggshell and wanders around for 2–5 h before becoming quies-
cent and moulting to the larva (Otto 1996). Elattostatic prelarvae are known in other
anystid mites (Otto 1999) and the adamystid Saxidromus delamarei (Coineau 1976),
both in the Anystina, and in the endeostigmatid Speleorchestes poduroides
(Nanorchestidae) (Schuster and Pötsch 1989). All of these mites seem to be mem-
bers of relatively derived taxa within their lineages and certainly are not the most
primitive of acariform mites. Other known acariform prelarvae are calyptostatic and
most are partially or fully retained within the egg.
In contrast to prelarvae, mite larvae tend to approach life with open mouths. For
example, larval Mesostigmata from soils in Colorado typically have chelicerae that
Prelarva and Larva 79

are well developed enough to capture, puncture and feed on smaller invertebrates –
and the behavioural repertoire to acquire their prey. Larval Lasioseius
(Blattisociidae), Gamasellodes (Ascidae), Dendrolaelaps (Digamasellidae),
Rhodacarus and Rhodacarellus (Rhodacaridae) will starve within a week if food is
not available. In contrast, larval feeding is facultative in species of Protogamasellus
(Ascidae) and absent in species of Laelapidae and Macrochelidae (Walter and
Lindquist 1989; Walter and Proctor 1999). Larvae of phytoseiid mites living on
plant leaves also vary from obligate to facultative to non-feeding, with only a few
derived taxa being the latter (Zhang and Croft 1994). Larval feeding also appears
to be the norm in the Acariformes. Endeostigmata (Walter 1988a) and, as far as is
known, Oribatida (Walter and Proctor 1998) have larvae that actively feed. Most
Prostigmata have feeding larvae, although those of some Labidostommatidae and
Rhagidiidae (Eupodina) are reported to be non-feeding (Evans 1992). As in ticks,
the parasitic larvae of Parasitengonina actively seek their hosts and some, such as
Vatacarus ipoides, a trombiculid mite that lives in the nostrils of sea snakes, feed
only as larvae.
Determining the homologies of instars or stages within a group as diverse and
ancient as the arachnids may not be possible but we believe that mites will be best
understood when their relationships to other chelicerates have been fully explored.
To date, the only coherent approach to this problem within acarology has been the
school of thought founded by François Grandjean that considered regressive evolu-
tion to be the dominant force in the phylogenetic history of the Acari. For example,
in this tradition Kethley (1990) hypothesised that the prelarvae of ancestral mites
were fully formed and fully functional and that the few known ellatostatic prelarvae
are closest to this ancestral condition. However, Otto (1997) argues that modern
prelarvae that are capable of movement were derived several times from immobile
prelarvae in taxa that inhabited dry habitats, probably because movement increases
the ability of the prelarva to secure a humid microsite for moulting or to avoid being
cannibalized by siblings.
Are mite prelarvae a regressive remnant of a formerly more perfect stage? The
overall behavioural and ontogenetic trends discussed above suggest the following
hypothesis. A non-feeding instar subsisting on yolk reserves and cared for by the
mother is characteristic of most arachnid orders. Because both this stage and the
associated maternal behaviour are present in both pulmonate arachnid and scorpion
lineages, we consider it an ancestral character that was probably inherited from
marine ancestors. The yolk-dependent acarine prelarva, therefore, is homologous to
the non-feeding arachnid J1 larva (Table 4.1). The loss of extended maternal guard-
ing of newly hatched young may explain the inactive acarine prelarva. With no
mother to seek out and crawl upon, mobility in the mite prelarva no longer provided
a selective advantage (except in some modern forms characteristic of dry habitats).
Unlike the arachnid J1, the fourth pair of legs in mite prelarvae is rudimentary or
completely suppressed. This may allow the redirection of resources to other parts of
the developing embryo. In the oribatid Archegozetes longisetosus the apparent limb
bud of legs IV develops long after the expression of the other prosomal segments in
the embryo, and the expression of the segment that will eventually bear legs IV is
80 4 Life Cycles, Development and Size

Table 4.1 Possible homologies among juvenile stages in the Chelicerata


Taxon Foetal instars J1 J2 J3 J4 J5 Jx
Scorpionida (Placental) M 8 NF 8F F F F 6–9
Uropygida (Brood pouch) M 8 NF 8F F F F –
Amblypygida (Brood pouch) M 8 NF 8F F F F 6–8
Araneae (Cocoon) M 8 NF 8F F F F 5–12
Solfugida – Im 8 NF 8F F F F 6–8
Pseudoscorpionida (Brood pouch) M 8 NF 8F F – – –
Palpigradi ? ? 8F F F – –
Opiliones ? M 8 NF 8F F F F 6–7
Opilioacariformes – Im 6 NF 6F F F F –
Acariformes – Im 6 NF 6F F F F –
Parasitiformes – ? 6F F F F (Ticks)
Ricinulei ? ? 6F F F F –
M mobile, Im immobile, NF nonfeeding, F feeding, 6 hexapod, 8 octopod

tied to the development of opisthosomal segmentation (Barnett and Thomas 2012).


How this may or may not be related to the sequential addition of opisthosomal
segments during post-embryonic development (anamorphosis) is not clear.
Why the prelarva or other stages (or at least their apoderms) are retained where
they now have no obvious function may have to do with the canalisation of develop-
ment: particular developmental steps or additions of structures require the forma-
tion of the apoderm of the previous stage. However, in some mite lineages the
prelarva and other stages are apparently lost.
The first feeding stage in mites is the hexapod larva, while in other arachnids it
is the octopod J2 nymph. Assuming these two stages are homologous (a hypothesis
in need of testing), the three nymphal stages of mites may then be homologous to
the succeeding nymphal stages in other arachnids (i.e. J3, J4, J5), provided no
instars have been added to or subtracted from these developmental sequences. In
mites, additional moults are known only in a few seemingly derived mite taxa;
therefore, these are most likely to represent evolutionary innovations and not
retained or re-expressed ancestral arachnid stages.

Suppression and Skipping of Stages

Reduction in the number and degree of expression of free-living ontogenetic stages


has been a recurring theme in acarine evolution. However, one major radiation of
mites, the Sarcoptiformes, has found the expression of a full developmental sequence
no impediment to success. Most endeostigmatans express a full developmental
sequence although only two octopod immature stages may be present (e.g.
Alicorhagia) and the prelarva may be retained within the egg. In all known oribatid
mites in the traditional sense (excluding Astigmata), a full developmental sequence
is present, although prelarvae are reduced to a layer of cuticle with little surface
Suppression and Skipping of Stages 81

structure that is retained inside the egg in most groups. Exceptions occur in some
Palaeosomata where the prelarva pops out of the egg like some endeostigmatans and
in some Astigmata where the larva emerges from the egg coated in the prelarval
cuticle (Wirth 2006).
The Astigmata (a derived lineage of oribatid mites) are a special case. Astigmatans
often have an unusual deutonymph modified for dispersal – an elattostatic hetero-
morph or hypopus in mite speak. Hypopi (heteromorphic deutonymph is preferred
by some, but we prefer the shorter hypopus, hypopi) do not have functional mouth-
parts and usually have incomplete guts (Houck 1994). In most taxa the formation of
these modified deutonymphs is facultative. When it is not formed or when it is
entirely suppressed, as in parasitic and ectocommensal species, the protonymph
moults directly to the tritonymph, skipping the deutonymphal instar entirely.
Non-astigmatan sarcoptiform mites are developmentally conservative and unusu-
ally so for mites. For most lineages of Acari, the evolutionary trend has been to
reduce the number of ontogenetic stages, or at least the number of active develop-
mental stages. For example, a synapomorphy of the Mesostigmata is the reduction
of their ontogeny to a larva and two nymphal instars.
In the Mesostigmata, pharate stages form directly under the cuticle of the previ-
ous stage (even being able to move until just before the old cuticle is shed), with no
indication of the apoderm of a missing stage present. Deutonymphal and adult
mesostigmatans typically have the same setation, with differences confined to
changes in extent of sclerotized plates, the opening of the genital atrium, and the
development of other sexual characters. One idea is that mesostigmatan ontogeny
results from skipping the tritonymphal stage that is present in Holothyrida and arga-
sid ticks. Another hypothesis might be that the tritonymph becomes the reproduc-
tively competent adult and the ancestral adult is not expressed. However, adult
mesostigmatans tend to resemble adult holothyrans in degree of sclerotization and
development of plates, so this would seem to be a more complicated hypothesis .
Another parasitiform lineage carries this trend even farther: ixodid ticks have only
a single nymph between the larva and adult. Possibly, the ixodid nymph is the
protonymphal stage, and the adult either a sexually mature deutonymphs or an adult
stage reached without expression of two nymphal stages.
In the Prostigmata, the number of octopod stages may be reduced to two, one or
none. André and van Impe (2012) argue that the terminal reproductive stage in the
spider mite Tetranychus urticae is a sexually mature tritonymph, so the adult stage
is absent. In the Eriophyoidea, a vast clade of plant-parasitic gall and rust mites,
there are no octopods stages – and no hexapod stages, either: the third and fourth
pair of legs are not expressed in any Eriophyoidea and have been absent since at
least the Triassic (Schmidt et al. 2012). Gall and rust mites moult twice after hatch-
ing. Interpreting these stages is especially difficult because so many external char-
acters have been lost, but the characters that remain suggest that it is the larval and
protonymphal stages that are expressed before the adult (Lindquist 1996).
The most extreme truncations in life cycles among any animal species occur in
the prostigmatan group Heterostigmata. Early derivative heterostigmatans in the
Tarsocheylidae and Heterocheylidae have a relatively complete life cycle with only
82 4 Life Cycles, Development and Size

the prelarva and tritonymph seemingly absent (Lindquist 1986). However, most
heterostigmatans have only a single active pre-adult stage, the hexapod ‘larva’. This
trend towards reduced active stages reaches its most extreme possible condition in
genera such as Dolichocybe, Siteroptes, Pyemotes and Acarophenax where develop-
ment, and often mating, occur within the physogastric body of the mother
(Kaliszewski et al. 1995).

Life Cycle of the Parasitengona

The prostigmatic suborder Parasitengonina includes about 80 families and over


11,000 described species of terrestrial and aquatic mites. One unifying characteris-
tic of this morphologically and behaviourally diverse group is their very compli-
cated life cycle – perhaps the most complex of all acarine developmental patterns
(Fig. 4.6). Although a prelarva has been described for only a few species of water
mites (Hydracarina) (e.g. Meyer 1985) and terrestrial parasitengones (e.g. Johnston
and Wacker 1967), it is likely to be a universal stage in this group. The prelarva
bursts through the eggshell but is otherwise inactive (Meyer 1985) and the larva
emerges from the thin prelarval cuticle (Shatrov 1999a, b).
Most larval Parasitengonina seek out and parasitise a broad range of arthropods
and vertebrates. However, a few terrestrial species and several dozen aquatic ones
have dispensed with the parasitic stage, with rather drastic repercussions on other
aspects of their life history and ecology (see Chap. 9). Depending on the species, the
larva spends from days to months on the host and may be moved a considerable
distance from its point of origin. After engorging on host fluids, the swollen larva
drops off and waddles or swims to a safe site where it transforms to a calyptostatic
protonymph. Terrestrial species usually seek a humid site or bury themselves in soil,
whereas water mites attach to aquatic vegetation or seek shelter in the tissues of
freshwater sponges and bivalves (Hevers 1980).
The deutonymph develops within the skin of the protonymph, which in turn is
enclosed in the larval cuticle. The deutonymph that emerges is a very different crea-
ture from the larva. Morphologically, these two stages bear no resemblance to each
other in the structure of palps, legs or body. This radical transformation is similar to
the structural rearrangement that occurs in insect pupae, which has led some water
mite biologists to call the life cycle of parasitengones ‘essentially holometabolous’
(Smith and Cook 1991). As well as looking different from the larva, the deutonymph
behaves differently. It is an active predator of the eggs, larvae and pupae of insects,
soft-skinned mites and (in water mites) small crustaceans (see Chap. 7). The
deutonymph feeds until it is bloated and then enters a second calyptostase, the
tritonymph. In this case the transformation is less extreme and the eclosed adult is
usually morphologically similar to the deutonymph. However, in water mites the
adults are often more highly sclerotised than the deutonymphs and in sexually
dimorphic species the female resembles the deutonymph more closely than the
often bizarrely modified male. Adult parasitengones are predatory and have diets
Life Cycle of the Parasitengona 83

Fig. 4.6 The life cycle in the Parasitengonina is complex and has alternating active and inactive
cryptic stages. E egg, pl hatched prelarva, al active larva, pl parasitic larva, cpn calyptostatic
protonymph, adn active deutonymph, ctn calyptostatic tritonymph, ad adult emerging from cuticle
of tritonymph (Image by DE Walter (after Michener 1946))

similar to their deutonymphal stages, although they may feed on larger prey. Only
the first active stage of parasitengone mites is parasitic; the other stages are free-
living. Animals that are parasitic only as immatures are called protelean parasites
(Treat 1975). In mites, this developmental pattern is rare outside the Parasitengonina
but some astigmatans have deutonymphs that are protelean parasites in mammalian
hair follicles or under the skin of birds (Athias-Binche 1991).
Parasitengones are among the longest-lived of acariform mites. Species of giant
velvet mites found in desert regions of Africa and North America are thought to
live several years as an adult. This long life may be due in part to these mites being
active only immediately after rainfall: the intervening periods are spent buried and
inactive (Crawford 1990). Adults of the water mite Limnochares aquatica live for
more than 2 years in captivity (Böttger 1972). Bader (1980) reports that adult
84 4 Life Cycles, Development and Size

lifespans of 1–2 years are common for water mites that dwell in cold streams; how-
ever, mites in warm water may live only 2–6 months as adults and have no more than
one generation per ice-free season. One problem with living for a long time is wear
and tear on the cuticle. Old water mites are often covered with encrusted dirt, algae
and phoretic protozoans and may bear scars from predator attacks (Proctor, unpub-
lished data). Perhaps as a result of selection to renew the cuticle, Limnochares aquat-
ica is able to undergo supernumerary moults up to a year after laying eggs (Böttger
1972). Some individuals moulted twice during Böttger’s study and in every case the
adult entered a calyptostatic state prior to moulting. Post-adult moults and calypto-
stases also have been observed in terrestrial parasitengones. Michener (1946) states
that Microtrombidium maculatum (Microtrombidiidae) may moult up to seven times
as an adult. The amount of growth between such moults varied greatly (e.g. the
increase in length of the tarsi on the first pair of legs ranged from 0.02 to 0.19 mm).

Paedomorphosis, Progenesis and Neoteny

If female individuals that do not appear to be adults are able to reproduce, this giving
birth by ‘children’ is call paedogenesis. Aphids are notorious for being born already
pregnant with young that are themselves pregnant, in a series reminiscent of a set of
nested Russian matryoshka dolls. In mites, however, reproduction by sexually imma-
ture individuals has been reported for a few plant-parasites on citrus (Baker 1979;
Ochoa 1989). In contrast, the confusingly similar term paedomorphosis (‘child
shape’) refers to the presence of ancestral juvenile characters in later developmental
stages (often the adult). Several authors have applied this concept to mites (Kethley
1974; Atyeo et al. 1984; OConnor 1984). For example, the adult female of Metacheyletia,
a cheyletid mite living in the feathers of parrots, is hexapod and resembles ‘a large,
globular larva with functional genitalia’ (Atyeo et al. 1984). Gould (1977) proposed
that paedomorphosis results from two ecologically distinct selective regimes. According
to Gould’s hypothesis, progenesis (‘forward production’) results from accelerated
maturity in species exploiting ephemeral resources; in contrast, retarded maturation in
species competing for stable resources results in neoteny (‘extended youth’). In both
cases the result is a sexually mature developmental stage that to some degree resembles
an earlier developmental stage of a presumed ancestor.
Several groups of sarcoptiform mites have adults that are poorly sclerotised and
resemble the immatures of related taxa. However, there is no indication that these
‘larviform’ adults result from extended developmental periods (neoteny). Rather,
evolutionary change in the Acari seems to be characterised by the loss of segments,
structures and stages and by the acquisition of smaller body sizes and faster devel-
opmental rates. Gould considered Siteroptes graminum (Fig. 4.7) to be an acarine
example of progenesis. Development is faster in offspring produced physogastri-
cally and they have fewer larval cuticular structures than those that develop from
eggs laid outside the mother (Rack 1972). On a broader scale, the entire Mesostigmata
could be interpreted as an example of progenesis. Mesostigmatans are sexually
Paedomorphosis, Progenesis and Neoteny 85

Fig. 4.7 Physogastry: (a) Physogastric female of Siteroptes graminum. (b) Individual pharate
female within larval cuticle (Image by DE Walter (after Rack 1972))

mature after three moults, compared to four in holothyrans and most argasid ticks,
and they are generally smaller, less hirsute, less well armoured and develop much
faster than holothyrans (Walter and Proctor 1998). However, adult mesostigmatans
do not resemble immature holothyrans or ticks.
Many mites that exploit ephemeral resources have very rapid developmental
rates (see Developmental Rates and Generation Times) but retain the full develop-
mental sequence of more slowly developing relatives. In some of these taxa, the
reproductive stages lack ontogenetically added cuticular structures normally char-
acteristic of adults. Chant (1993) proposed that faster development within existing
stages that results in paedomorphic adults be called sequential progenesis. For
example, adult phytoseiid mites are considered to be hypotrichous (‘under hairy’),
i.e. setal complement is less than what is considered the primitive condition
(holotrichy – ‘complete hairs’) in the Mesostigmata. Chant (1993) argued that phy-
toseiid mites tend to be smaller and to develop faster than many of their more setose
relatives, and noted that within the Phytoseiidae less hairy species tended to develop
faster than more hairy species.
Chant (1993) convincingly demonstrated that the setae present in adult phyto-
seiid mites are primarily those first added in the larval stage and that setae added in
later stages of related mites are those that tend to be absent in Phytoseiidae. As a
result, adult phytoseiid mites resemble immature stages of their existing relatives.
Other mesostigmatans, however, mature even more rapidly than phytoseiid mites
86 4 Life Cycles, Development and Size

but show little or no indication of sequential progenesis, and phytoseiid mites


mature at the normal adult stage. So, neither neoteny nor progenesis appears
to apply to these mites. Perhaps post-displacement (Raff 1996), i.e. a form of
paedogenesis unrelated to maturation rates where the expression of features
(e.g. addition of setae) is delayed or eliminated, is a simpler way to account for the
loss of setae in some mesostigmatans.

Size, Developmental Rate and Generation Times

Although we think of mites as small, size is relative and the largest mite is more
than one million times the mass of the smallest (Fig. 4.8). Mass, metabolism and
growth rate are often strongly inter-related among the animals with which we are
most familiar (e.g. the mouse to elephant curve). In mites, however, these
interrelationships are not so simple.

Overview of Mite Size Patterns

Although smallness is the very essence of being a mite, some mites are noncon-
formists. Engorged adult female ticks may be the size of cherries (2–3 cm). Adults
of the African giant red velvet mites Dinothrombium tinctorium reach 13–16 mm

Fig. 4.8 Mites vary in size by many orders of magnitude (Image by HC Proctor)
Overview of Mite Size Patterns 87

in length (Thor and Willmann 1947; Newell and Tevis 1960) and other velvet,
water and holothyrid mites exceed 5 mm. Modern parasitiform mites tend to be
relatively large as adults in several lineages. For example, adult Australian allo-
thyrid mites are typically 2–3 mm in length – larger than many beetles in the same
forest litter habitats and much larger than the average parasitiform mite inhabiting
rainforest litter. In live extractions, allothyrid mites are sometimes covered with
acarid hypopi, oribatids and mesostigmatans that cling to the legs and bodies of
these lumbering giants.
Acariform mites can be among the smallest of terrestrial arthropods. Soil-litter
habitats in hot and cold deserts and other dry ecosystems are often dominated
by mites less than 250 μm in length (a quarter of a millimetre or 1/100th of an
inch) in the families Brachychthoniidae, Tydeidae, Tarsonemidae, Scutacaridae,
Terpnacaridae and Nanorchestidae. For example, Nanorchestes antarcticus is a tiny,
globular, hopping algivore with an average adult length of 240 μm and a width of
135 μm. It is common throughout Antarctica, as far south as the Horlick Mountains
(85 °32′S), and can occur at densities of >2000 per m2 (Block 1979). Many other
endeostigmatic mites are even smaller. For example, Alycosmesis corallium
(Terpnacaridae), which was collected in litter under gum trees in South Africa, has
an adult length of only 154–162 μm and Nanorchestes memelensis from pasture soil
in South Africa measures 149–157 μm from the tip of the gnathosoma to the end of
the idiosoma and is only 69–75 μm wide. Larvae of these tiny mites would be in the
80–120 μm length range – seemingly near or at the lower limit for mites. These
lower limits also apply to the smallest of parasitic mites, such as the Tarsonemoidea
that attach to insect eggs, the Demodicidae that live in the pores and ducts of mam-
mals and the Eriophyoidea that cause galls to form in plants. The latter, an extremely
diverse group with nearly 3,000 described species of tiny, worm-like plant-parasites,
are typically about 200 μm in length and 30–50 μm in diameter, with some species
reputed to be as short as 80 μm as adults (Lindquist et al. 1996).
Parasitiformans also can be very small. Phytoseiid mites, those most useful of
biological control agents, rarely exceed 500 μm in length and some efficient preda-
tors of spider mites (e.g. Galendromus occidentalis) are less than a quarter of a
millimetre long. An adult length of about 200–250 μm appears to be the lower limit
in the Mesostigmata but this is a characteristic size for many cosmopolitan soil-litter
inhabitants such as Rhodacarus roseus and Protogamasellus mica. The latter has
been collected at depths of up to 8 m along mesquite roots (Walter and Ikonen 1989)
and, in general, the deeper one samples in the mineral soil, the smaller and narrower
the bodies of the mites that live there. In fact, many habitats will have acarine size
ranges that are distinctive.
Figure 4.9 shows the body size distribution for 180 species of parasitiform mites
from subtropical rainforest in the Green Mountains section of Lamington National
Park in Queensland, Australia. The body length of the smallest species, a phytoseiid
mite (240 μm), was used to set the histogram interval. The number of species in
each interval declines steeply with increasing size. Mites 2 mm in length or greater
were rare and consisted primarily of holothyrans (2000 μm), three species of ticks
(2500–5000 μm) and a number of trigynaspid mites (to 2866 μm) associated with
88 4 Life Cycles, Development and Size

Fig. 4.9 Body size


distribution of 180 species
of parasitiform mites from
rainforests in the Green
Mountains region of
Lamington National Park,
Queensland, Australia
(Walter and Proctor 1999)

large arthropods living in logs. Mites collected from leaves averaged only about half
the length (mean = 349 ± 9, range = 240–400 μm) of mites collected from forest floor
habitats (mean = 709 ± 46, range = 250–5000 μm). Except for ticks, the guild of
mites with the largest members was that associated with arthropods, which includes
the largest of all known Mesostigmata, the Megisthanidae. Some species of
Megisthanus from the tropical rainforests in Far North Queensland have a body a
whopping 4 mm in length!

Developmental Rates and Generation Times

Given that mites are small and cold-blooded, one might expect them to have very
short generation times, at least given proper temperatures, enough food and a lack
of constraints such as synchronization with host life cycles. As noted above, some
mites have rates of population increase that are second to none among animals, in
part because of their very rapid life cycles. Many plant parasites have generation
times of less than 10 days and economically important spider mite pests typically
pass through a generation in 10–14 days during the growing season (Sabelis
1985). The broad mite Polyphagotarsonemus latus, a tiny tarsonemid
(170 × 100 μm) that can cause devastating damage to a number of economically
important plants, probably holds the record. In the summer, a generation takes
only 4–5 days and egg to adult time is only 3 days at 27 °C (Hambleton 1938;
Jeppson et al. 1975). Another tarsonemoid mite, Acarophenax mahunkai, parasit-
ises the eggs of mealworms. The female finds an egg to which she attaches, feeds
and swells to more than four times her previous diameter in about a day. In cul-
tures at 24–26 °C, active mite progeny were visible moving around inside the
physogastric mother at 72–96 h after parasitism had begun and all had left the
dying mother within 120 h (Steinkraus and Cross 1993).
Developmental Rates and Generation Times 89

Predatory Mesostigmata also can develop rapidly. At the moderate temperature


of 25 °C the nematode-predatory pachylaelapid Zygoseius furciger develops from
newly laid egg to adult female in 108 ± 4 h, i.e. it passes through more than a fifth of
its developmental time each day. A spider mite predator, the phytoseiid Phytoseiulus
longipes, can complete development in about 94 h (Takahashi and Chant 1992).
Macrocheles muscaedomesticae, a predator of flies in chicken dung, is even faster.
At its optimum temperature of 30 °C, male M. muscaedomesticae averaged 66 h
from egg to adult and females averaged 77 h (Ho 1985, 1989). Species of
Macrocheles predatory on fly eggs and nematodes on the surface of a dung pat – an
extremely ephemeral resource – have minimum developmental times of 34–51 h at
temperatures of 30 °C and above, and mean generation times of 4½ days (Cicolani
1979; Geden et al. 1990; Ruf 1996).
Daily developmental rates of 10–20 % are common among the Macrochelidae in
herbivore dung, the Phytoseiidae on plants and many soil-inhabiting Mesostigmata
(Krantz 1983; Sabelis and Janssen 1994; Walter and Ikonen 1989; Ruf 1996). These
mites show a strong response to temperature, with generation times falling rapidly
with increasing temperature (Fig. 4.10a). The lower and upper limits to this response
are related to the normal habitat of the mite. For example, the small ascid
Gamasellodes vermivorax is found in the drier soils of the Great Plains in North
America where surface soil temperatures can rise to over 50 °C in the summer
(Walter 1987). Adult mites will survive for at least a day at 45 °C in an incubator but
will not lay eggs. Eggs transferred to this temperature will not develop. The critical
upper limit to development is between 35 °C and 40 °C and the lower limit near
10 °C. However, over the 25 °C range between these two extremes, developmental
rates increase sharply with temperature (Fig. 4.10b). Zygoseius furciger from high
elevation meadows (>3,000 m) in the Rocky Mountains has a similar response;
however, the curve is shifted to the left. At 10 °C, a temperature at which G. vermi-
vorax fails to develop, Z. furciger is able to progress from egg to adult in 23 days.
At 30 °C, a suboptimum temperature for G. vermivorax, all Z. furciger died either
during development or at the adult moult, although the latter took only 3 days to
reach their final, fatal transformation.
However, not all mites exhibit such rapid developmental patterns. Many para-
sites are closely tied to their hosts’ life cycles and may pass through only a single
generation per year. Mites with complicated life cycles often have extended devel-
opmental and generation times, as for example the giant red velvet mites that prey
on desert-inhabiting termites during the insects’ irregular nuptial flights (see Life
Cycle of Parasitengona). However, even given similar ecological constraints, devel-
opmental rates show strong phylogenetic patterns, both across and within lineages.
At temperatures where species in some families or suborders can complete a
generation in less than a week, others take 2 weeks or a month and, among oribatid
mites, a month or more per instar may be more normal.
The suborder Oribatida, a paraphyletic assemblage when Astigmata is excluded,
contains over 10,000 described species distributed across more than 1,100 genera in
its traditional ‘beetle mite’ form. These mites are usually particulate-feeding micro-
bivores or saprophages that retain the full ancestral developmental sequence.
90 4 Life Cycles, Development and Size

Fig. 4.10 Effect of


temperature on (a) generation
time and (b) developmental
rates of two species of
Mesostigmata from Colorado,
USA. Zygoseius furciger
(Pachylaelapidae) from a
subalpine meadow in the
Rocky Mountains and
Gamasellodes vermivorax
(Ascidae) from the Great
Plains. Mites were reared on
nematodes in small vials in
incubators (Walter and
Proctor 1999)

However, this gustatory and developmental conservatism has not been a barrier to
evolutionary success. Even the most extreme terrestrial habitats, from frozen shores
to heat-blasted deserts, have been colonised by several lineages of oribatid mites. In
temperate to subtropical forests oribatid mites are often the most abundant arthro-
pods, from the litter to the canopy, with densities of tens of thousands to hundreds
of thousands of individuals per square metre reported (Norton 1994; Walter and
O’Dowd 1995).
Litter-soil populations of oribatid mites tend to be relatively stable, although these
mites often have long developmental times and relatively low reproductive outputs. In
the laboratory at temperatures of 20–30 °C, egg to adult developmental times of 3–50
weeks are commonly reported for oribatid mites from temperate zone forests. Many
species appear to be univoltine and those from colder climates clearly have multi-year
life cycles. Few tropical species have been studied but there is some indication that
generation times of one or two months may be more typical (Honciuc 1996). Generally,
species from more derived lineages appear to have faster developmental rates and,
Dissociation Between Body Size and Developmental Rate in Mesostigmata 91

within a lineage, there appears to be a negative relationship between body size and
developmental rate (Norton 1994). However, the Astigmata, a derived group of oriba-
tid mites that traditionally have been honoured with their own suborder, often have
very short generation times, especially among those that have exploited ephemeral
habitats (OConnor 1994).

Dissociation Between Body Size and Developmental


Rate in Mesostigmata

Scaled co-variation between the size of organisms and their metabolic processes is
a well-known phenomenon among animals. For example, when log metabolic rate
is plotted against log body mass, species tend to fit a line defined by a 3/4 power
function (the ‘mouse-to-elephant curve’, Brown 1995). Conversely, specific meta-
bolic rate (i.e. energy consumption per unit body mass) decreases with increasing
body size along a −1/4 curve, so that, per gram of body mass, an elephant uses
less energy per gram than does a mouse (Schmidt-Nielsen 1984). So, smaller organ-
isms tend to have faster paced lives than larger organisms and life-history parame-
ters dependent on numerous metabolic processes (e.g. developmental time) should
be correlated with body size. This trend should be especially obvious in related
species living under similar conditions.
In spite of the generality of the 3/4 power relationship (Brown 1995, but see
White et al. 2007), size is not the only arbiter of metabolic rate. Bats, for example,
tend to live 5–10 times as long as rodents of similar size (Guarante 1997), sug-
gesting that phylogenetic history may override physiological inertia. In even
smaller organisms, the 3/4 power relationship may be irrelevant. The regulation of
metabolic activities of the nematode Caenorhabditis elegans has been shown to
be under genetic control. Mutations in nematode ‘clock’ genes strongly affect
developmental rate, pace of behaviours and lifespan (Ewbank et al. 1997).
Mesostigmatic mites that prey on nematodes also seem to be an exception to
size–metabolism allometry.
The Mesostigmata are the dominant group of predatory microarthropods in spa-
tially and temporally continuous habitats such as soil and leaf litter. However, they
are also common in patchier habitats, such as herbivore dung, rotting fruit and rot-
ting fungal sporocarps. Some Mesostigmata are general predators that develop
equally well on a variety of small invertebrate prey but many have a preference for,
or are dependent on, nematodes as a food source (Walter et al. 1987; Walter and
Ikonen 1989).
As mentioned in Chap. 3 the subordinal name Mesostigmata is derived from the
mid-body position of the openings to the highly branched tracheal system that rami-
fies through the body of the mite (Evans 1992). West et al. (1997) proposed a model
to explain the generality of the 3/4 power relationship based on transport of essen-
tial materials through tubular systems such as blood vessels or arthropod tracheal
tubes. Because mesostigmatic mites have a highly branched tracheal system, the
92 4 Life Cycles, Development and Size

Fig. 4.11 Mean developmental


rate versus mean dry weight
for 15 species of Mesostigmata
from grassland soils in
Colorado, USA (Walter and
Proctor 1999)

West et al. model (1997) predicts that they should follow the 3/4 power relationship.
However, when developmental rates are plotted against dry weights for 15 species
of Mesostigmata from a variety of grassland soils in Colorado, no significant slope
is generated (Fig. 4.11). Although body mass varies more than an order of magni-
tude across these 15 species of mites, developmental rates are much less variable
and do not tend to increase with body mass. Within a genus, developmental rates
tend to be similar, irrespective of body size.
Possibly, the flat distribution of developmental time against body size results
from constraints on exploiting ephemeral resources, such as nematodes and other
prey whose populations flush in response to irregular desert rains or rotting organic
material. Larval Lasioseius starve in less than a week (Walter and Lindquist 1989)
but adults survive without prey for many weeks (Walter and Proctor 1999).
Therefore, the selective pressure for rapid development of mites to adults must be
very strong in temporary habitats and may result in the evolution of genetic over-
rides on normal allometric growth. Developmental stages, however, appear to be
genetically fixed. Although some Mesostigmata have non-feeding stages, they pro-
duce the full ontogenetic sequence of the order. Therefore, strong reductions in
body size may be difficult to achieve and increases in developmental rate must
occur within developmental stages, rather than by suppression of developmental
stages (Athias-Binche 1987; Chant 1993).

Dispersal, Migration and Phoresy

Movements within and between habitats are important components of any organ-
ism’s life history. Movement that maintains an organism within its usual home
range can be subdivided into a number of categories that relate to its speed, direc-
tion, and purpose; however, the ultimate result of these behaviours is that organisms
use the resources available at a particular site to promote growth and reproduction
Dispersal, Migration and Phoresy 93

(Dingle 1996). In terms of mite movement within habitats, foraging (movement in


search of resources) has been best studied in the predatory mites used in biological
control. Regular back and forth movements in response to daily changes in resource
availability, predation dangers or environmental tolerances (e.g. temperature,
humidity) are analogous to the daily movement of people to and from work or
school, and have been termed commuting. In desert soils, mites commute into sur-
face litter in response to rainfall and also have a daily commute into litter in the
early morning and back into mineral soil by mid-day (called ‘vertical migration’ in
Whitford et al. 1981).
Dispersal is a commonly used, but somewhat confusing term, for leaving one
area or habitat patch for another. We usually think of ‘dispersal’ as a kind of scat-
tering of individuals across a landscape; however, dispersal behaviours can have
a variety of results. If, as is the case with many mites, it is an immature stage of
an obligatory bisexual species that is ‘dispersing’, then scattering of the popula-
tion could lead to many Robinson Crusoes, each living out their life marooned on
some lonely habitat island. One might expect then that selection would tend to
favour behaviours that produced ordered and predictable movement from one
site to another.
Migratory behaviour occurs when an organism suspends normal feeding, devel-
opment and reproduction and begins expending energy to move from one place to
another. Migration is associated with specific starting and stopping behaviours and
is more persistent and less meandering than normal movement.

Migratory Stages

In mites, dispersal is often complex and difficult to shoehorn into standard defini-
tions. For example, spider mites spread rapidly within a plant because a proportion
of newly fertilised adult females will emigrate regardless of the population density
of the natal leaf (Kennedy and Smitley 1985). These mites may walk to their new
leaves but others literally cast their fates to the wind. Under high population levels,
deteriorating host plant condition and reduced humidity levels, most spider mites
shift from their normal sedentary behaviour, become photopositive, cluster on the
top of plants and migrate on the wind. Some species of spider mites in the genera
Oligonychus, Eotetranychus and Panonychus spin down from a silken thread and
dangle until a breeze snaps the thread and carries them off (often called ballooning,
as in spiders, but different in that the silk is not reeled out on the wind until lift is
achieved). However, species of Tetranychus have a more complicated migratory
behaviour. Females, and sometimes nymphs, cluster at the tips of plants and respond
to wind by orienting away from the source of light, raising their first pair of legs and
body surfing on the wind (Kennedy and Smitley 1985).
Most of the spider mites that have been studied in agricultural habitats are
polyphagous (consume plants of many species). Therefore aerial dispersal, although
it must involve high mortality, is probably a reasonable alternative to remaining on
94 4 Life Cycles, Development and Size

a host plant of declining quality and increasing number of predators. Inseminated


females that are able to feed on a range of plant hosts should have a fair chance of
founding new colonies. However, even the highly host-specific Eriophyoidea
migrate passively on the wind as adult females and this is less easy to understand
(Sabelis and Bruin 1996). In general, eriophyoid mites feed on perennial hosts and
cause relatively minor damage to the host plant. About half the species are able to
manipulate the physiology of the host to form galls of various sorts in which they
live relatively well protected from predators. Also, on average, eriophyoids are the
smallest of all mites and most are thought to be species- or even species-tissue spe-
cific parasites. They seem, therefore, quite unlikely to encounter a new suitable host
through passive dispersal. However, many eriophyoids orient to wind currents, use
their caudal lobe to spring well above the leaf boundary layer and increase these
behaviours in response to increasing wind (Lindquist and Oldfield 1996). They also
appear able to move vast distances in the air and survive the extreme cold and desic-
cation associated with long-distance aerial dispersal (Zhao and Amrine 1997).
The instar at which migration occurs has a strong phylogenetic component in
mites that is often correlated with ecology. Oribatid mites seem to disperse primar-
ily as adults (at least in those species where the immatures burrow in decaying
vegetation); however, many Astigmata have a deutonymphal stage especially modi-
fied for migration, the hypopus. The deutonymphal stage has been lost in parasitic
and ectocommensal Astigmata and movement from host to host is either by adult
females or, in the case of parasites with highly modified adults, by the larvae. Most
Prostigmata appear to migrate as adult females; however, parasitengone larvae also
disperse on their hosts. Mesostigmata usually migrate as either mated but prerepro-
ductive adult females (most Dermanyssiae) or as deutonymphs (Sejina, Uropodina,
Parasitina, Digamasellidae). In the Uropodina, deutonymphs may exhibit a ‘normal’
and a ‘dispersal’ form (wandernymphen). However, the larva is sometimes the dis-
persing stage in parasitic Mesostigmata (Radovsky 1994).

Phoresy

Migration is a major challenge for any mite not living in a continuous, persistent
habitat. Even within such seemingly uniform and extensive habitats as a forest floor,
desirable microhabitats may be few and far between. Since mites are tiny and wing-
less, movement between resources that are patchy in space or time may require
help. The use of one organism (the carrier, host or porter) by another (the phoretic)
for transport is termed phoresy (from the Greek ‘to carry’) and is a common behav-
iour among mites (Fig. 4.12), small insects, pseudoscorpions, molluscs, nematodes
and microbes.
The term phoresy was first proposed by French entomologist and bostrichid bee-
tle specialist Pierre Lesne in 1896 and to quote from a delightful paper by the
famous myrmecophile and ant acarologist W.M. Wheeler (1919): “In 1896 Lesne
called attention to a number of small insects that habitually ride on larger insects.
Dispersal, Migration and Phoresy 95

Fig. 4.12 Mites are usually phoretic as either deutonymphs of both sexes or as adult females.
Examples of phoretic deutonymphs: (a) Parasitid mite holding onto the mouthparts of a fly
(Anthomyiidae, Pegomyia sp.); (b) another anthomyiid fly (Eutrichota sp.) with a cluster of
Halolaelaps sp. (Mesostigmata, Halolaelapidae) posterior to the scutellum and rows of Myianoetus
sp. (Astigmata, Histiostomatidae) attached at the juncture of the dorsal tergites; (c) Venter of a
histiostomatid hypopus showing sucker plate; (d) a Uroseius sp. (Uropodoidea) attached to a
phorid fly (Images by DE Walter)

To this phenomenon he applied the term “phoresy” and showed that it is distin-
guished from ectoparasitism by the fact that the portee does not feed on the porter
and eventually dismounts and has no further relations with the latter.” Mites are
especially adept phoretics and often have highly modified phoretic stages (e.g. the
astigmatan deutonymph, which has been around for at least 50 million years,
Dunlop et al. 2012) or morphs within a particular stage (e.g. phoretomorphs in the
Heterostigmata) (Moser and Cross 1975) (see Chap. 9).
As classically defined in the mid-twentieth Century, phoresy required an animal
to stop its normal behaviours (e.g. seeking food or members of the opposite sex),
seek out a carrier, mount the carrier and refrain from doing anything other than
holding on, and then after a time to dismount from the carrier and resume normal
behaviours. The heteromorphic deutonymph of Astigmatina (aka hypopus) is a clas-
sic phoriont – it lacks functional mouthparts, has no foregut (and so feeding would
seem to be impossible), and has a ventral sucker disk formed from modified setae
for holding onto an insect or other larger arthropod (sometimes even other mites).
96 4 Life Cycles, Development and Size

Deutonymphs of Parasitus coleoptratorum (Mesostigmata, Parasitidae), found on


geotrupine scarab beetles that frequent herbivore dung, are typical phorionts. When
beetles arrive at a fresh dung pat, the phoretic deutonymphs leave their carrier, moult
to adults and reproduce. Female P. coleoptratorum can lay an egg every hour (up to 25
per day) and development to the deutonymph is rapid (Costa 1969). Normally, these
new deutonymphs would seek out a beetle, migrate to the next pat of fresh dung and
repeat the cycle. However, if the appropriate cues are not available, then 60–80 % of
P. coleoptratorum can delay the moult to adult for up to 6 months (Rapp 1959).
So, in an evolutionary ecology sense, classical phoresy can be thought of as an
ectosymbiosis in the commensal category where the mite gets a free ride and the
carrier is unaffected. Dispersal is assumed to be at the root of the evolution of these
behaviours in the mites (although phoresy may actually concentrate, rather than
disperse the mites). But are things really so simple?
Again the jargon is a bit confusing: ‘commensal’ implies eating together, and the
mite can’t eat while on the carrier or it wouldn’t be considered phoretic. Consider
the case of water mite larvae – they use adult aquatic insects to ‘disperse’, but since
they feed on their carriers while doing so, they are considered parasites, not phori-
onts. Then there are the strange hypopi of Hemisarcoptes cooremani which hitch
rides under the elytra of ladybird beetles in the genus Chilochorus. Since Chilochorus
beetles like to eat mites, this seems like a dangerous thing to do – but hypopi of
H. cooremani that do not find a beetle die. Those that do find a beetle hang on for
5–21 days and swell up before leaving the beetle. Marilyn Houck (1994), who has
studied this interaction, has suggested that beetle hemolymph (reflexive bleeding is
a defense of many ladybird beetles – pick them up and noxious yellow blood oozes
from their leg joints) may provide critical nutrients, possibly via the anal vestibule
of the mites (which connects to a hindgut).
Mites can accidentally wander onto larger organisms and be transported to new sites
but establishment of new populations would then be largely a matter of luck. Phoretic
behaviour is not accidental in this way, although it may have evolved from less formal
associations. Instead, phoretics go through a sequence of behaviours that are quite dif-
ferent from non-phoretic individuals. Many of these behaviours are analogous to those
used by parasites to find their hosts (Athias-Binche and Morand 1993). In a sense, all
phoresy can be considered an exploitation of the carrier, and therefore, parasitic.
However, species interactions must be defined in terms of their ultimate effects on the
fitness of the participants if they are to make any ecological or evolutionary sense.
Traditionally, phoresy has been differentiated from parasitism by requiring the
phoretic individual to enter a state of diapause and not feed on the host (Athias-Binche
and Morand 1993). However, any interaction between individuals of two species that
benefits one but causes a reduction in the net fitness of the other is parasitic. A phoretic
does not need to feed on a carrier to cause it harm, as has been shown for Southern
Pine Beetles that are weighed down with masses of uropodid mites (Kinn and
Witcosky 1977) or flies so encumbered with hypopi (Elzinga and Broce 1988) or
ereynetid tritonymphs (Zhang and Sanderson 1993) that they are unable to fly.
Lisa Hodgkin and her colleagues at the University of Melbourne published
(2010) a study with some interesting experimental data. They studied a bark beetle
Summary 97

(Ips grandicollis) introduced into Australia and its phoretic mites and demonstrated
that ‘phoresy’ can be both dynamic and complex. Phoretic mites were associated with
both negative (adults) and positive (larvae) effects on beetle reproduction and devel-
opment. Heavy mite loads were not good for female beetles, but having mites in the
galleries resulted in bigger and healthier offspring. Perhaps when we are considering
defining complex interactions like phoresy we should remember what Shakespeare’s
Hamlet pointed out to Horatio: ‘There are more things in heaven and earth, Horatio,
Than are dreamt of in your philosophy’.
In contrast, being fed upon does not necessarily reduce the fitness of a carrier.
For example, passalid beetles seem to suffer no ill effects from the hordes of large
mesostigmatans that forage on their bodies. More importantly, a phoretic that causes
no harm to a carrier in flight may devastate the carrier’s brood if they exit the carrier to
feed on or compete with the carrier’s offspring (Eickwort 1994). Some associations
traditionally considered phoretic are clearly parasitic or appear to be evolving
towards negative effects on the carriers (Houck 1994). These, and other mite-carrier
relationships, will be considered in more detail in Chap. 9. For now, the point that
we wish to emphasise is that the changes in behaviour associated with phoresy indi-
cate that it is a set of behaviours that allows a mite to move from one place to
another through the help of a larger organism, irrespective of the proximate or ulti-
mate effects on the fitness of the carrier.

Summary

Mites may make mediocre mothers but they are masters of ontogenetic manipulation.
The basic acarine developmental pattern of egg, prelarva, larva, protonymph,
deutonymph, tritonymph and adult has been modified repeatedly. In terrestrial
chelicerates, the planktonic instars of marine ancestors either have been lost or
appear as poorly formed foetal cuticles. The first post-egg stage subsists on rem-
nants of yolk and has incompletely formed mouthparts and limbs. In most arach-
nids this J1 nymph is octopod and dependent on protection from the mother until it
has moulted into the J2 nymph that is capable of hunting and feeding. In mites a
hexapod prelarva (which we think is homologous to the J1 nymph) is present in both
Acariformes and Parasitiformes, but ephemeral, often retained within the egg as a
simple layer of cuticle, and when expressed from the egg is usually immobile and
never feeds on external resources. Mite larvae (J2 nymphs) actively feed, except in
a derived lineages where feeding is facultative or absent. Nymphal instars beyond
the tritonymph and post-adult moults are known in only a few, derived mite taxa and
appear to represent means of repairing cuticle or increasing body size, rather than
being retained ancestral stages.
Suppression of one or more nymphal instars and accelerated development are
recurrent themes in the evolution of the Acari. Many oribatid mites, however, are
able to maintain high population levels and a ubiquitous distribution while encum-
bered with the full developmental sequence. Within the Astigmata, a derived lineage
98 4 Life Cycles, Development and Size

of oribatid mites, the deutonymph is a non-feeding dispersal stage that may be


facultatively or completely suppressed. A number of highly diverse and successful mite
lineages in both the Parasitiformes and Acariformes are characterised by terminal
truncation of nymphal instars (or perhaps, accelerated sexual maturity), very rapid
developmental rates and extraordinary rates of population increase. The Parasitengonina
have followed a third route to evolutionary success through parasitism of vertebrates
and arthropods as larvae and by alternating active, predatory stages with inactive,
hidden stages that occur within the cuticle of the previous stage. The ultimate in the
abbreviated approach to life occurs in the Heterostigmata, where grotesquely swollen
adult females give birth to fully formed and already mated adult offspring.
Although the majority of mites are minute, adult body lengths vary from about
80 μm in some plant parasites to 30,000 μm in engorged ticks and biovolumes vary
by about 6 orders of magnitude. Within a group of closely related mites with similar
ecological constraints, life- history parameters such as clutch size, developmental
rate and generation time are usually positively related to body size. However, phy-
logenetic constraints on physiological processes are obvious throughout the Acari,
with some lineages, such as the Oribatida, characterised by long generation times
and low rates of increase. Others, such as species of Macrocheles associated with
patchy, ephemeral habitats like dung, have extraordinarily short generation times
and high rates of increase. The most rapidly developing acariformans are in the
Tarsonemina (Heterotigmata) where some parasites of plants and of insects are able to
complete development in about 3 days at moderate temperatures. In the
Parasitiformes, similar developmental rates are found among the Dermanyssiae.
These mites mature only a single large egg at a time, irrespective of body size or
ecological role, and developmental rates show no linear relationship to body size.
Mites are small and wingless but they are abundant in many patchily distributed
resources. Some mites take to the wind directly but the primary mechanism that
these mites use to travel from one resource patch to another is phoresy, i.e. hitching
rides on larger organisms. Phoresy is a form of migration and, like any migrant, a
phoretic mite behaves quite differently from a non-migrant. Phoretic mites search
for a carrier rather than food, attach to some part of that carrier, enter a period of
altered behaviour (e.g. inactivity or parasitic feeding) and depart from the carrier at
the appropriate time in a new habitat. The phoretic relationship may be mutually
beneficial to the mite and the carrier, it may be neutral, it may result in a loss of fit-
ness to the carrier, or it may vary in its effects dynamically over time. In any case, a
primary benefit to the mite is movement: this is the core of the relationship, not
whether or not they feed while hitching a ride.

References

Aeschlimann, A. (1984). What is our current knowledge of acarine embryology? In D. A. Griffiths


& C. E. Bowman (Eds.), Acarology VI (Vol. 1, pp. 90–99). Chichester: Ellis Horwood.
André, H. M., & van Impe, G. (2012). The missing stase in spider mites (Acari: Tetranychidae):
When the adult is not the imago. Acarologia, 52, 3–16.
References 99

Athias-Binche, F. (1987). Signification adaptives des différents types de développements


postembryonnaires chez les Gamasides (Acariens: Anactinotriches). Canadian Journal of
Zoology, 65, 1299–1310.
Athias-Binche, F. (1991). Evolutionary ecology of dispersal in mites. In F. Dusbabek & V. Bukva
(Eds.), Modern acarology 1 (pp. 27–41). Prague: SPB Academic.
Athias-Binche, F., & Morand, S. (1993). From phoresy to parasitism: The example of mites and
nematodes. Research and Reviews in Parasitology, 53, 73–79.
Atyeo, W. T., Kethley, J. B., & Perez, T. M. (1984). Paedomorphosis in Metacheyletia (Acari:
Cheyletidae), with the Description of a New Species. Journal of Medical Entomology, 21,
125–131.
Bader, C. (1980). Some biological and ecological data on water mites, mainly some significant
data on the life-duration. International Journal of Acarology, 6, 239–243.
Baker, E. W. (1979). A note on paedogenesis in Brevipalpus sp. (Acari: Tenuipalpidae), the first
such record for a mite. International Journal of Acarology, 5, 355–379.
Barnett, A. A., & Thomas, R. H. (2012). The delineation of the fourth walking leg segment is
temporally linked to posterior segmentation in the mite Archegozetes longisetosus (Acari:
Oribatida, Trhypochthoniidae). Evolution & Development, 14, 383–392. doi:10.1111/j.
1525-142X.2012.00556.x.
Bergmann, P., & Heethoff, M. (2012). The oviduct is a brood chamber for facultative egg retention
in the parthenogenetic oribatid mite Archegozetes longisetosus Aoki (Acari, Oribatida). Tissue
& Cell, 44, 342–350.
Block, W. (1979). Nanorchestes antarcticus Strandtmann (Prostigmata) from Antarctic ice.
Acarologia, 21, 173–176.
Böttger, K. (1972). Vergleichend biologisch-ökologische Studien sum Entwicklungszyklus
Süsswassermilben (Hydrachnellae, Acari) I. Der Entwicklungszyklus von Hydrachna globosa
und Limnochares aquatica. Int Rev Ges Hydrobiol, 57, 109–152.
Brown, J. H. (1995). Macroecology. Chicago: University of Chicago Press.
Bruce, W. A., & Wrensch, D. L. (1990). Reproductive potential, sex ratio, and mating efficiency of
the straw itch mite (Acari: Pyemotidae). Journal of Economic Entomology, 83, 384–391.
Canard, A., & Stockmann, R. (1993). Comparative postembryonic development of arachnids.
Memoirs of the Queensland Museum, 33, 461–468.
Cancela da Fonseca, J. P. (1975). Notes Oribatologiques. Acarologia, 17, 320–330.
Chant, D. A. (1993). Paedomorphosis in the Family Phytoseiidae (Acari: Gamasina). Canadian
Journal of Zoology, 71, 1334–1349.
Cicolani, B. (1979). The intrinsic rate of natural increase in dung Macrochelid mites, predators of
Musca domestica eggs. Bollettino di Zoologia, 46, 171–178.
Cloudsley-Thompson, J. L. (1992). Solifugae – and keeping them in captivity. Arachnida:
Proceedings of a Symposium on Spiders and Their Allies. London.
Coineau, Y. (1976). Les pariades sexuelles des Saxidrominae Coineau 1974 (Acariens Prostigmates,
Adamystidae). Acarologia, 28, 234–240.
Coineau, Y., & van der Hammen, L. (1979). The postembryonic development of Opilioacarida,
with notes on new taxa and on a general model for the evolution. Proceedings of the 4th
International Congress of Acarology. Budapest: Akadémiai Kiadó.
Cook, W. J., Smith, B. P., & Brooks, R. J. (1989). Allocation of reproductive effort in female
Arrenurus spp. water mites (Acari: Hydrachnidia; Arrenuridae). Oecologia, 79, 184–188.
Costa, M. (1969). The association between mesostigmatic mites and coprid beetles. Acarologia,
11, 411–428.
Crawford, C. S. (1990). Scorpiones, Solifugae, and associated desert taxa. In D. L. Dindal (Ed.),
Soil biology guide (pp. 421–475). New York: Wiley.
Dingle, H. (1996). Migration: The biology of life on the move. New York: Oxford University
Press.
Dunlop, J. A., Wirth, S., Penney, D., McNeil, A., Bradley, R. S., Withers, P. J., & Preziosi, R. F.
(2012). A minute fossil phoretic mite recovered by phase-contrast X-ray computed tomogra-
phy. Biology Letters, 8, 457–460.
100 4 Life Cycles, Development and Size

Eickwort, G. C. (1994). Evolution and life-history patterns of mites associated with bees. In
M. A. Houck (Ed.), Mites: Ecological and evolutionary analyses of life-history patterns
(pp. 218–251). New York: Chapman & Hall.
Elzinga, R. J., & Broce, A. B. (1988). Hypopi (Acari: Histiostomatidae) on house flies (Diptera:
Muscidae): A case of detrimental phoresy. Journal of the Kansas Entomological Society, 61,
203–208.
Evans, G. O. (1992). Principles of acarology. Wallingford: CAB International.
Ewbank, J. J., Barnes, T., Lakowski, B., Lussier, M., Bussey, H., & Hekimi, S. (1997). Structural
and functional conservation of Caenorhabditis elegans timing gene clk -l. Science, 275,
980–983.
Foelix, R. F. (1982). Biology of spiders. Cambridge: Harvard University Press.
Geden, C. J., Stinner, R. E., et al. (1990). MACMOD: A simulation model for Macrocheles mus-
caedomesticae (Acari: Macrochelidae) population dynamics and rates of predation on imma-
ture house flies (Diptera: Muscidae). Environmental Entomology, 19, 578–586.
Goodnight, M., & Goodnight, C. J. (1976). Observations on the systematics, development and
habits of Erginulus clavotibialis (Opiliones: Cosmetidae). Transactions of the American
Microscopical Society, 95, 654–664.
Gould, S. (1977). Ontogeny and phylogeny. Cambridge: Harvard University Press.
Grandjean, F. (1938). Sur l’ontogénie des Acariens. Comptes Rendus de l’Académie des Sciences,
206, 146–150.
Grandjean, F. (1970). Stases – Actinopiline – Rappel de ma classification des Acariens en trois
groupes majeurs. Terminologie en soma. Acarologia, 11, 796–827.
Guarante, L. (1997). What makes us tick? Science, 257, 943–944.
Hambleton, E. J. (1938). A ocorrencia do acaro tropical Tarsonemus latus Banks (Acar.,
Tarsonemidae) causador da rasgadura das folhas nos algodoais de S. Paulo (Vol. 9, pp. 201–209).
Sao Paulo: Arquivos do Instituto Biologico.
Hevers, J. (1980). Biologisch-ökologische Untersuchungen zum Entwicklungszyklus der in
Deutschland auftretenden Unionicola-Arten (Hydrachnellae, Acari). Archiv für Hydrobiologie
Supplementband, 57, 324–373.
Ho, C. -C. (1985). Mass production of the predaceous mite, Macrocheles muscaedomesticae
(Scopoli) (Acarina: Macrochelidae) and its potential use as a biological agent of house fly,
Musca domestica L. (Diptera: Muscidae). Ph.D. Dissertation, p. 186. University of Florida,
Gainesville.
Ho, C.-C. (1989). Studies on the Biology of Macrocheles muscaedomesticae (Scopoli) (Acarina:
Macrochelidae). Chinese Journal of Entomology, 3, 181–187.
Hodgkin, L. A., Elgar, M. A., & Symonds, M. R. E. (2010) Positive and negative effects of phoretic
mites on the reproductive output of an invasive bark beetle. Australian Journal of Zoology, 58,
198–204. http://www.publish.csiro.au/paper/ZO10034.
Honciuc, V. (1996). Laboratory studies of the behaviour and life cycle of Archegozetes longi-
setosis Aoki 1965 (Oribatida). In R. Mitchell, D. J. Horn, G. R. Needham, &
W. C. Welbourn (Eds.), Acarology IX: Vol. 1. Proceedings (pp. 637–640). Columbus: Ohio
Biological Survey.
Houck, M. A. (1994). Adaptation and transition into parasitism from commensalism: A phoretic
model. In M. A. Houck (Ed.), Mites: Ecological and evolutionary analyses of life-history pat-
terns (pp. 252–281). New York: Chapman & Hall.
Jeppson, L. R., Keifer, H. H., & Baker, E. W. (1975). Mites injurious to economic plants. Berkeley:
University of California Press.
Johnston, D. E., & Wacker, R. R. (1967). Observations on postembryonic development in
Eutrombicula splendens (Acari – Acariformes). Journal of Medical Entomology, 4, 306–310.
Kaestner, A. (1968). Invertebrate Zoology, Volume II. New York: John Wiley & Sons Interscience
Publishers.
Kaliszewski, M., Athias-Binche, F., & Lindquist, E. E. (1995). Parasitism and parasitoidism in
Tarsonemina (Acari: Heterostigmata) and evolutionary considerations. Advances in Parasitology,
35, 335–367.
References 101

Kennedy, G. G., & Smitley, D. R. (1985). Dispersal. In W. Helle & M. W. Sabelis (Eds.), Spider
mites, their biology, natural enemies and control (Vol. 1A, pp. 233–242). New York: Elsevier.
Kethley, J. (1974). Developmental chaetotaxy of a paedomorphic celaenopsoid, Neotenogynium
malkini n.g., n.sp. (Acari: Parasitiformes: Neotenogyniidae n. fam.) associated with millipedes.
Annals of the Entomological Society of America, 67, 571–579.
Kethley, J. (1990). Acarina: Prostigmata (Actinedida). In D. L. Dindal (Ed.), Soil biology guide
(pp. 667–756). New York: Wiley.
Kethley, J. (1992). The prelarva of Alycus roseus Koch (Bimichaeliidae: Acariformes: Acari).
Canadian Journal of Zoology, 68, 1058–1061.
Kinn, D. N., & Witcosky, J. J. (1977). The life cycle and behaviour of Macrocheles boudreauxi
Krantz. Zeitschrift fuer Angewandte Entomologie, 84, 136–144.
Klompen, J. S. H. (2000). Prelarva and larva of Opilioacarus (Neocarus) texanus (Chamberlin and
Mulaik) (Acari : Opilioacarida) with notes on the patterns of setae and lyrifissures. Journal of
Natural History, 34, 1977–1992.
Klompen, H. (2011). Holothyrids and ticks: New insights from larval morphology and dna
sequencing, with the description of a new species of Diplothyrus (Parasitiformes: Neothyridae).
Acarologia, 50, 269–285. doi:10.1051/acarologia/20101970.
Krantz, G. W. (1983). Mites as biological control agents of dung-breeding flies, with special refer-
ence to the Macrochelidae. In M. A. Hoy, G. L. Cunningham, & L. Knutson (Eds.), Biological
control of pests by mites (pp. 91–98). Berkeley: University of California Agriculture Experiment
Station Special Publication 3304.
Krantz, G. W., & Walter, D. E. (Eds.). (2009). A manual of acarology (3rd ed.). Texas Tech
University Press. 807 p, 338 b/w illustrations; 60 figures ISBN 978-0-89672-620-8. http://
www.ttup.ttu.edu/BookPages/9780896726208.html.
Labruna, M. B., Nava, S., Guzmán-Cornejo, C., & Venzal, J. M. (2012). Maternal care in the soft
tick Antricola marginatus. Journal of Parasitology, 98, 876–877.
Lange, A. B., & Tolstikov, A. V. (1999). Ovoviviparity, prelarva and the peculiarities of eclosion in
fresh-water oribatid mites Thrypochthoniellus setosus (Will.) and Hydrozetes lemnae (Coggi).
Acarina, 7, 13–21.
Lindquist, E. E. (1986). The world genera of Tarsonemidae (Acari: Heterostigmata): A morpho-
logical, phylogenetic, and systematic revision, with a reclassification of the family-group taxa
in the Heterostigmata. Memoirs of the Entomological Society of Canada, 136, 1–517.
Lindquist, E. E. (1996). 1.5.2 Phylogenetic relationships. In E. E. Lindquist, M. W. Sabelis, & J. Bruin
(Eds.), Eriophyid mites. Their biology, natural enemies and control (Vol. 6, pp. 301–327).
Amsterdam: Elsevier.
Lindquist, E. E., & Oldfield, G. N. (1996). Evolution of eriophyoid mites in relation to their host
plants. In E. E. Lindquist, M. W. Sabelis, & J. Bruin (Eds.), Eriophyid mites – their biology,
natural enemies and control (Vol. 6, pp. 277–300). Amsterdam: Elsevier.
Lindquist, E. E., Sabelis, M. W., & Bruin, J. (1996). Eriophyoid mites, their biology, natural ene-
mies and control. Amsterdam: Elsevier.
Martens, J. (1993). Further cases of paternal care in Opiliones (Arachnida). Tropical Zoology, 6, 97–107.
Meyer, E. (1985). Der Entwicklungszyklus von Hydrodroma despiciens (O.F. Müller 1776) (Acari:
Hydrodromidae). Archiv für Hydrobiologie Supplementband, 66, 321–453.
Michener, C. D. (1946). The taxonomy and bionomics of some Panamanian trombidiid mites.
Annals of the Entomological Society of America, 39, 349–380.
Mitchell, R. W. (1971). Egg and young guarding by a Mexican cave-dwelling harvestman,
Hoplobunus boneti (Arachnida). The Southwestern Naturalist, 15, 392–395.
Mora, G. (1990). Paternal care in a neotropical harvestman, Zygopachylus albomarginis
(Arachnida, Opiliones: Gonyleptidae). Animal Behaviour, 39, 582–593.
Mori, H., Saito, Y., & Tho, Y. (1999). Co-operative group predation in a sit-and-wait cheyletid
mite. Experimental and Applied Acarology, 23, 643–651.
Moser, J. C., & Cross, E. A. (1975). Phoretomorph: A new phoretic phase unique to the
Pyemotidae (Acarina: Tarsonemoidea). Annals of the Entomological Society of America, 68,
820–822.
102 4 Life Cycles, Development and Size

Newell, I. M., & Tevis, L., Jr. (1960). Angelothrombium pandorae n.g., n. sp. (Acari, Trombidiidae)
and notes on the biology of the giant red velvet mites. Annals of the Entomological Society of
America, 53, 293–305.
Norton, R. A. (1994). Evolutionary aspects of oribatid mite life histories and consequences for the
origin of the Astigmata. In M. A. Houck (Ed.), Mites: Ecological and evolutionary studies of
life-history patterns (pp. 99–135). New York: Chapman & Hall.
Ochoa, R. (1989). A note on paedogenesis in Tetranychoidea. International Journal of Acarology,
2, 117–118.
OConnor, B. M. (1984). Phylogenetic relationships among higher taxa in the acariformes, with
particular reference to the Astigmata. In D. A. Griffiths & C. E. Bowman (Eds.), Acarology VI
(Vol. I, pp. 19–27). Chichester: Ellis Horwood Ltd.
OConnor, B. M. (1994). Life-history modifications in astigmatid mites. In M. A. Houck (Ed.),
Mites: ecological and evolutionary analyses of life-history patterns (pp. 136–159). New York:
Chapman & Hall.
Otto, J. (1996). Observations on prelarvae in Anystidae and Tenerifiidae. In R. Mitchell, D. J. Horn,
G. R. Needham, & W. C. Welbourn (Eds.), Acarology 9: Volume 1, proceedings (pp. 343–354).
Columbus: Ohio Biological Survey.
Otto, J. (1997). In R. Mitchell, D. J. Horn, G. R. Needham, & W. C. Welbourn (Eds.), Observations
on prelarvae in Anystidae and Tenerifiidae (pp. 343–354). Columbus: Ohio Biological Survey.
Otto, J. (1999). Systematics and natural history of the genus Chaussieria Oudemans (Acarina:
Prostigmata: Anystidae). Zoological Journal of the Linnean Society, 126, 251–306.
Pérez, T. M. (1996). The eggs of seven species of Fainalges Gaud and Berla (Xolalgidae) from the
green conure (Aves, Psittacidae). In R. Mitchell, D. J. Horn, G. R. Needham, & W. C. Welbourn
(Eds.), Acarology IX: Volume 1, proceedings (pp. 297–300). Columbus: Ohio Biological Survey.
Polis, G. A., & Sissom, W. D. (1990). Life history. In G. A. Polis (Ed.), The biology of scorpions
(pp. 161–223). Stanford: Stanford University Press.
Preston-Mafham, R. A., & Preston-Mafham, K. G. (1993). The encyclopedia of land invertebrate
behaviour. Cambridge: The MIT Press.
Proud, D. N., Víquez, C., & Townsend, V. R., Jr. (2011). Paternal care in a Neotropical harvestman
(Opiliones: Cosmetidae) from Brazil. Journal of Arachnology, 39, 497–499.
Rack, G. (1972). Pyemotiden an Gramineen in schwedischen landwirtschaftlichen Betreiben. Ein
Beitrag zur Entwicklung von Siteroptes graminum (Reuter, 1900) (Acarina: Pyemotidae).
Zoologisches Anzeiger, 188, 157–174.
Radovsky, F. J. (1994). The evolution of parasitism and the distribution of some dermanyssoid
mites (Mesostigmata) on vertebrate hosts. In M. A. Houck (Ed.), Mites, ecological and evolu-
tionary analyses of life-history patterns (pp. 186–217). New York: Chapman & Hall.
Raff, R. A. (1996). The shape of life: genes, development and the evolution of animal form.
Chicago: University of Chicago Press.
Ramires, E. N., & Giaretta, A. A. (1994). Maternal care in a neotropical harvestman Acutisoma
proximum (Opiliones, Gonyleptidae). Journal of Arachnology, 22, 179–180.
Rapp, A. (1959). Zur Biologie und Ethologie der Kafermilbe Parasitus coleoptratorum L. 1758
(Ein Beitrag zum Phoresie Problem). Zoologische Jahrbücher. Abteilung für Systematik,
Ökologie und Geographie der Tiere, 86, 303–366.
Ruf, A. (1996). Life-history patterns in soil-inhabiting mesostigmatid mites (Dermanyssina,
Parasitina). In R. Mitchell, D. J. Horn, G. R. Needham, & W. C. Welbourn (Eds.), Acarology
IX: Volume 1, proceedings (pp. 621–628). Columbus: Ohio Biological Survey.
Sabelis, M. W. (1985). Reproductive strategies. In W. Helle & M. W. Sabelis (Eds.), Spider mites,
their biology, natural enemies and control (Vol. 1A, pp. 265–278). New York: Elsevier.
Sabelis, M. W., & Bruin, J. (1996). Evolutionary ecology: life history patterns, food plant choice
and dispersal. In E. E. Lindquist, M. W. Sabelis, & J. Bruin (Eds.), Eriophyid mites – their biol-
ogy, natural enemies and control (pp. 329–366). Amsterdam: Elsevier.
Sabelis, M. W., & Janssen, A. (1994). Evolution of life history patterns in the Phytoseiidae. In
M. A. Houck (Ed.), Mites, ecological and evolutionary analyses of life-history patterns
(pp. 70–98). New York: Chapman & Hall.
References 103

Saito, Y. (1986a). Prey kills predator: Counter-attack success of a spider mite against its specific
phytoseiid predator. Experimental and Applied Acarology, 2, 47–62.
Saito, Y. (1986b). Biparental defence in a spider mite (Acari: Tetranychidae) infesting Sasa bam-
boo. Behavioural Ecology and Sociobiology, 18, 377–386.
Saito, Y. (1997). Sociality and kin selection in Acari. In J. C. Choe & B. Crespi (Eds.), Evolution of
social behaviour in insects and arachnids (pp. 443–457). Cambridge: Cambridge University Press.
Saito, Y. (2010). Plant mite and sociality – diversity and evolution (p. 187). Tokyo: Springer.
Schmidt, A. R., Jancke, S., Lindquist, E. E., Ragazzi, E., Roghi, G., Nascimbene, P. C., Schmidt,
K, Wappler, T., & Grimaldi, D. A. (2012). Arthropods in amber from the Triassic Period.
PNAS, 109, 14796–14801. www.pnas.org/cgi/doi/10.1073/pnas.1208464109.
Schmidt-Nielsen, K. (1984). Scaling – Why is animal size so important? Sydney: Cambridge
University Press.
Schuster, R., & Pötsch, H. (1989). Another record of an active prelarva in mites. In G. P. Channabasavanna
(Ed.), Acarology (Vol. 1, pp. 261–265). New Delhi: Oxford and IBH Publishing.
Shatrov, A. B. (1999a). External morphology of the quiescent instars of trombiculid mites
(Acariformes: Trombiculidae) with notes on their moulting processes. Acta Zoologica
(Stockholm), 80, 85–95.
Shatrov, A. B. (1999b). Contribution to the prelarva status: The moulting cycle of the calyptostasic
prelarva of the trombiculid mite Leptotrombidium orientale (Acariformes : Trombiculidae).
Acarologia, 40, 265–274.
Shipley, A. E. (1909). Introduction to Arachnida and king crabs. In S. F. Harmer & A. E. Shipley
(Eds.), The Cambridge natural history (Vol. IV, pp. 255–279). Melbourne: The Macmillan
Company.
Smith, I. M., & Cook, D. R. (1991). Water mites. In J. H. Thorp & A. P. Covich (Eds.), Ecology
and classification of North American freshwater invertebrates (pp. 523–592). San Diego:
Academic.
Steinkraus, D. C., & Cross, E. A. (1993). Description and life history of Acarophenax mahunkai,
n.sp. (Acari, Tarsonemina: Acarophenacidae), an egg parasite of the lesser mealworm
(Coleoptera: Tenebrionidae). Annals of the Entomological Society of America, 86, 239–249.
Summers, F. M., & Witt, R. L. (1972). Nesting behavior of Cheyletus eruditus (Acarina:
Cheyletidae). Pan Pacific Entomologist, 48, 261–269.
Takahashi, F., & Chant, D. A. (1992). Adaptive strategies in the genus Phytoseiulus Evans (Acari:
Phytoseiidae): I. Developmental Times. International Journal of Acarology, 18, 171–176.
Thor, S., & Willmann, C. (1947). Acarina Trombidiidae. Das Tierreich, 71b, 187–541.
Treat, A. E. (1975). Mites of moths and butterflies. Ithaca: Cornell University Press.
Wallace, M. M. H. (1970). Diapause in the aestivating egg of Halotydeus destructor (Acari:
Eupodidae). Australian Journal of Zoology, 18, 295–313.
Walter, D. E. (1987). Life history, trophic behavior and description of Gamasellodes vermivorax n.
sp. (Mesostigmata: Ascidae) a predator of nematodes and arthropods in semiarid grasslands.
Canadian Journal of Zoology, 65, 1689–1695.
Walter, D. E. (1988a). Predation and mycophagy by endeostigmatid mites (Acariformes:
Prostigmata). Experimental and Applied Acarology, 4, 159–166.
Walter, D. E. (1988b). Macrocheles schaeferi (Acari: Mesostigmata: Macrochelidae), a new spe-
cies in the subbadius group from grassland soils in the central United States. Annals of the
Entomological Society of America, 81, 386–394.
Walter, D. E., & Ikonen, E. K. (1989). Species, guilds and functional groups: Taxonomy and
behavior in nematophagous arthropods. Journal of Nematology, 21, 315–327.
Walter, D. E., & Lindquist, E. E. (1989). Life history and behavior of ascid mites in the genus
Lasioseius (Acari: Mesostigmata) from grassland soils in Colorado with taxonomic notes and
a description of new species. Canadian Journal of Zoology, 67, 2797–2813.
Walter, D. E., & O’Dowd, D. J. (1995). Beneath biodiversity: Factors influencing the diversity and
abundance of canopy mites. Selbyana, 16, 12–20.
Walter, D. E., & Proctor, H. C. (1998). Feeding behaviour and phylogeny: Observations on early
derivative Acari. Experimental and Applied Acarology, 22, 39–50.
104 4 Life Cycles, Development and Size

Walter, D. E., & Proctor, H. C. (1999). Mites: Ecology, evolution and behaviour (p. 322). Sydney:
University of NSW Press. ISBN 0 86840 529 9.
Walter, D. E., Hunt, H. W., & Elliott, E. T. (1987). The influence of prey type on the development
and reproduction of some predatory soil mites. Pedobiologia, 30, 419–424.
West, G. B., Brown, J. H., & Enquist, B. J. (1997). A general model for the origin of allometric
scaling laws in biology. Science, 276, 122–124.
Wheeler, W. M. (1919). The phoresy of Antherophagus. Psyche, 26, 145–152.
White, C. R., Cassey, T., & Blackburn, T. M. (2007). Allometric exponents do not support a uni-
versal metabolic allometry. Ecology, 88, 315–323.
Whitford, W. G., Freckman, D. W., Elkins, N. Z., Parker, L. W., Parmalee, R., Phillips, J., &
Tucker, S. (1981). Diurnal migration and responses to simulated rainfall in desert soil microar-
thropods and nematodes. Soil Biology and Biochemistry, 13, 417–425.
Wirth, S. (2006). Development of the prelarva and larval behaviour to open the eggshell in the
Histiostomatidae (Astigmata). Abhandlungen und Berichte des Naturkundemuseums Goerlitz,
78, 93–104.
Zhang, Z.-Q., & Croft, B. A. (1994). A Comparative Life History Study of Immature Amblyseius
fallacis, Amblyseius andersoni, Typhlodromus occidentalis and Typhlodromus pyri (Acari:
Phytoseiidae) with a Review of Larval Feeding Patterns. Experimental and Applied Acarology,
18, 631–657.
Zhang, Z.-Q., & Sanderson, J. P. (1993). Association of Ereynetes tritonymphs (Acari: Ereynetidae)
with the Fungus Gnat, Bradysia impatiens (Diptera: Sciaridae). International Journal of
Acarology, 19, 179–183.
Zhao, S., & Amrine, J. W., Jr. (1997). Investigation of snowborne mites (Acari) and relevancy to
dispersal. International Journal of Acarology, 23, 209–213.
Chapter 5
Sex and Celibacy

In our everyday experience, animals have two separate sexes (Fig. 5.1). Some animals,
and many plants, combine the two in a single individual but they still produce male
and female gametes. This is such a basic generality that we become uneasy with
the suggestion that there are many taxa (some of them mites, see Parthenogenesis)
in which males are irrelevant.
What do we mean when we say that it is ‘normal’ to have two sexes? Fundamentally,
females are individuals that produce large-sized gametes that are well-stocked with
food (= ova) and males produce lean, minimalistic gametes (= spermatozoa). Having
gametes of different sizes – anisogamy – is the rule in multicellular plants and
animals (Parker 1982). In contrast, some algae undergo sexual reproduction by
the fusion of equal-sized gametes – isogamy (Raven et al. 1981). Some protozoans
exhibit isogamy while others are anisogamous and these states may differ between
congeneric species. Although the two types of gametes are very different in size in
most plants and animals, in protozoans the size differences range from slight to very
marked (Schmidt and Roberts 1985). Anisogamy prevails in more derived taxa,
while isogamy seems restricted to more ancestral groups of organisms, and so it is
assumed that isogamy is the primitive state.
One of the basic precepts in ecological and evolutionary theory is that resources
for any endeavour are limited. If one makes larger gametes, then one must also
make fewer. If one makes smaller gametes, the result is more of them. This is
another characteristic of anisogamous organisms: females produce few (large)
gametes and males produce many (small) ones. Although some unicellular organisms
(e.g. certain Chlamydomonas) have flagellated male and female gametes (Raven
et al. 1981), in the vast majority of anisogamous taxa, male gametes are motile and
female gametes are sedentary.
So compared to spermatozoa, ova are large, rare and immobile. The small, active
sperm vastly outnumber the cells with which they desire union. As only one sperm
can successfully partner with any given egg (with rare exceptions), ova as a resource
are in short supply. The aggressive contests between male mountain sheep, elk and
elephant seals for access to females are simply the outward manifestation of their

D.E. Walter and H.C. Proctor, Mites: Ecology, Evolution & Behaviour: Life at a Microscale, 105
DOI 10.1007/978-94-007-7164-2_5, © Springer Science+Business Media Dordrecht 2013
106 5 Sex and Celibacy

Fig. 5.1 A mating pair of


feather mites (Echinozonus
sp., Pterolichidae)
(SEM by HC Proctor)

sperm battling for access to eggs. Similarly, males may compete for female attention
by displaying flashy good looks or strenuous dancing. Darwin (1871) termed the
two types of competition intrasexual and intersexual respectively and suggested
that these two types of sexual selection were responsible for many extravagant male
characteristics (see Sexual selection).
Parker (1970) pointed out that an even more basic level of intrasexual combat
occurs between male gametes competing for a limited egg resource; he termed this
sperm competition. Sperm competition is a strong force selecting for and maintain-
ing the small size of male gametes (Parker 1982, but see Immler et al. 2011) and it
has probably played a vital role in the origin of internal sperm transfer.

Modes of Sperm Transfer

Animal life originated in the sea and it is among marine animals that we find the
greatest variety of sperm-transfer mechanisms. To understand how the diversity of
mating behaviours arose in mites it is helpful to consider evolutionary trends in
sperm transfer.
There is a general belief that the ancestral state of sperm transfer in animals
involved external fertilisation with freely spawned sperm and eggs (e.g. Parker
1978, 1984; references in Rouse and Fitzhugh 1994). Free-spawning without close
contact between the sexes occurs in many extant aquatic taxa, both invertebrate (e.g.
many annelids and echinoderms) and vertebrate (many fish). In such situations,
sperm competition is the only competitive interaction between males and the only
way to ‘win’ is to produce enormous clouds of sperm. Yund and McMartney (1994)
have demonstrated that intrasexual competition for fertilisations occurs in free-
spawning ascidians and bryozoans. Another feature associated with free-spawning
is gamete dilution, which sometimes results in a high proportion of eggs going
unfertilised (e.g. Levitan et al. 1992). Parker (1970, 1984) proposed that intrasexual
Distribution of Sperm-Transfer Modes Among Non-Acarine Animals 107

selection favours males that both avoid excessive expenditure on gametes and fertil-
ise a greater proportion of a female’s clutch by shedding sperm only when in close
proximity to a female. Although Parker focussed on benefits of close association to
the male, pairing to spawn benefits both male and female, as the latter may gain a
higher proportion of fertilised eggs. Parker suggested that copulation – placement
of sperm by a male into the receiving structure of a female (Hinton 1964) – evolved
from broadcast spawning by progressive steps, driven by intrasexual selection. This
scenario is no doubt true for metazoan animals as a whole, although the primitive
nature of broadcast spawning is in doubt for some groups in which more direct
forms of transfer may be ancestral (Buckland-Nicks and Scheltema 1995; Rouse
and Fitzhugh 1994). Regardless of which came first, it is clear that some animal taxa
display a greater diversity of sperm-transfer modes than other. Mites are near the
top of the list.

Distribution of Sperm-Transfer Modes Among


Non-Acarine Animals

The ways in which gametes get together can be classified on the basis of where the
gametes are when they meet and how much contact there is between male and
female. External fertilisation involves the release of gametes by both sexes, followed
by fertilisation in the outside environment. Males and females may be paired during
this activity (e.g. frogs, many fish) or they may be distant (e.g. many marine inver-
tebrates, in which case it is termed broadcast spawning). External fertilisation
requires an aqueous or at least very moist habitat to allow sperm to reach and
penetrate eggs. Among extant cheliceriforms, external fertilisation with pairing is
restricted to the marine horseshoe crabs and pycnogonids (Proctor 1998). Some argue
that the extinct eurypterids also had external fertilisation (Boucot 1990); however,
Braddy and Dunlop (1997) and Kamenz et al. (2011) suggest that fertilisation may
have been internal, following uptake of externally deposited spermatophores and
possibly a scorpion-like mating dance.
Internal fertilisation involves fertilisation of eggs while they are still in the
female. Although internal fertilisation was probably a prerequisite for successful
invasion of dry land by animals, it has evolved many times in aquatic habitats
(Proctor 1998). For example, Kano (2008) found that internal or semi-internal
fertilisation has arisen at least six times in deep-sea lineages of the mollusc group
Vetigastropoda, and suggests that this is a response to inefficiency of broadcast
spawning for organisms with small body sizes, low population densities, and lack of
environmental cues (e.g., changes in temperature or light regime) that might
synchronize gamete release. Retention of fertilised eggs by the female, brooding,
occurs in many species scattered through otherwise externally fertilising marine
taxa (e.g. echinoderms, David and Mooi 1990). Males of these taxa seldom show
modifications for direct insemination, suggesting that females ‘suck up’ freely
spawned sperm and use them to fertilise their eggs internally. Even some barnacles,
108 5 Sex and Celibacy

famously endowed with enormous penises, have the option of supplementing their
directly transferred sperm supply with freely spawned sperm acquired from the
water (Barazandeh et al. 2013). Thus a Parkerian process of increasing proximity
of male and female at time of spawning, followed by the female’s ‘decision’ to
brood her eggs, may be the evolutionary scenario associated with the repeated
evolution of direct sperm transfer. All arachnids, including mites, have internally
fertilised eggs.
Being human, we tend to feel that our way, copulation, is the only proper way for
sperm to get inside a female; however, males and females of other internally fertilising
taxa are not so fixated on closeness. There are numerous classifications of sperm-
transfer modes based on the degree of contact between the sexes (e.g. Alexander
1964; Schaller 1971; Weygoldt 1969). Here we use the four categories outlined by
Proctor (1998). Direct transfer or copulation involves the greatest degree of contact
and occurs when the male places his sperm on or in the sperm-receiving structure of
a female. One reason we consider copulation to be ‘normal’ is that the land animals
we are most familiar with all transfer sperm directly. All amniote vertebrates (reptiles,
birds, mammals) transfer sperm directly, as do all spiders and all winged insects,
and a large number of other groups including onychophorans, leeches, flatworms and
nematodes. In these taxa, sperm may be transferred in a liquid medium (e.g. amniotes,
most spiders and insects) or bound up together in a cohesive packet called a
spermatophore (e.g. many insects, a few spiders) (Mann 1984). Among arachnids,
all species of the Solifugae, Ricinulei, Opiliones and Araneae transfer sperm
through copulation.
Males need not place sperm directly in the female’s genital opening. Depending
on the species, male onychophorans may place semen or spermatophores in the
genital opening of their mate, or somewhere else on her body where lytic enzymes
in the spermatophore break down her epidermis to release sperm into her haemocoel
(Tait and Briscoe 1990). Another sort of traumatic, or hypodermic, insemination
occurs in bedbugs and some plant bugs (Hemiptera) in which the male pierces the
female’s integument with his sharp aedeagus (Tatarnic and Cassis 2010). Although
traumatic insemination has not been reported in mites, it seems likely that ancestors
of some acarine taxa did engage in this rather rough form of sperm transfer
( see Intersexual conflict below).
Pairing with indirect transfer occurs when a male pairs with a particular female
but does not introduce sperm into her. Rather, he deposits a spermatophore on a
nearby substrate and attempts to induce her to take the sperm from it. The most
famous of paired-indirect taxa are scorpions, which engage in a pas-de-deux or
mating dance prior to sperm deposition. All Amblypygi, Uropygi and Schizomida
pair, dance or promenade, and indirectly transfer sperm via often very complex
spermatophores, as do some species of pseudoscorpions (references in Proctor 1998).
Salamanders and newts (Amphibia: Urodela) also typically show pairing with
indirect transfer of spermatophores (e.g. Arnold 1976), as do some hexapods and
myriapods (Proctor 1998).
Unpaired, or dissociated, sperm transfer occurs when males do not court a par-
ticular female. Males of taxa showing incomplete dissociation must chemically or
Diversity of Sperm-Transfer Behaviours in Mites 109

tactilely detect a female prior to depositing a spermatophore but do not direct their
intentions toward any one female. Complete dissociation appears the most unlikely
form of transfer to us, as males and females need never contact each other, even
chemically, for successful deposition and uptake of sperm. Attraction of females to
the ejaculates is mediated by a variety of chemical and structural modifications of
the spermatophores (Proctor 1998). Dissociation is the mode of choice for many
pseudoscorpions and springtails (Collembola).

Diversity of Sperm-Transfer Behaviours in Mites

So where do mites fit in this scheme of transfer modes? They fit everywhere. In fact,
mites show the greatest diversity of sperm transfer behaviours of any group of
arthropods. In the following section we take you on a brief tour through the acarine
Kama Sutra.

Reproductive Anatomy

Before surveying sperm-transfer behaviour in mites, it is useful to know something


of their reproductive anatomy. More detailed explanations can be found in Evans
(1992), Krantz and Walter (2009), and the papers by Gerd Alberti and his colleagues
(see below).
Primary genital openings are those from which semen or spermatophores
(in males) or eggs (in females) exit the body of the individual. In many mite taxa,
females have secondary genital openings (variously termed sperm induction pores,
copulatory pores, solenostomes, etc.) into which males introduce sperm. These will
be discussed in the appropriate taxonomic sections below. In females, there may be
a temporary sperm-storage structure (a spermatheca or seminal receptacle) near
the copulatory opening (Alberti and Crooker 1985; Nuzzaci and Alberti 1996).
Spermathecae are not as common in mites as they are in insects.
In insects, fertilisation of eggs typically occurs when eggs pass by the opening of
the spermatheca and the female squeezes out a small dollop of sperm. Although
female insects in some taxa (e.g. social Hymenoptera) can store sperm for long
periods, oviposition usually occurs quite soon after sperm transfer, often while the
male mate is still present. In most mites, fertilisation appears to occur in the ovary
after sperm has migrated there (Feiertag-Koppen and Pijnacker 1985); however,
some taxa with spermathecae may fertilise eggs in an insect-like fashion (Nuzzaci
and Alberti 1996).
Oviposition in mites seldom occurs immediately after sperm transfer; it may be
days, weeks or months afterwards. The general lack of a spermatheca and the
temporal separation between sperm transfer and oviposition in mites means that
110 5 Sex and Celibacy

certain phenomena common in insects (e.g. internal manipulation of stored sperm


and post-copulatory guarding by males, Thornhill and Alcock 1983) are seldom
observed in the Acari.
The ovaries of female mites may be single, paired or a grape-like cluster of lobes.
There are typically paired oviducts regardless of the type of ovary, suggesting that
the original state is paired ovaries. Eggs may be matured one at a time, as small
clutches or as very large clutches of thousands of ova (see Chap. 4). Eversible ovipositors
are present in some mites.
In male mites, testes, like ovaries, may be single, paired or form grape-like
clusters. Vasa deferentia are usually paired. Acarine spermatozoa lack the propelling
flagellum we tend to consider the hallmark of sperm. The flagellum has been lost
from the spermatozoon independently in more than 30 animal and non-animal
taxa (Morrow 2004). Within the Arachnida, tail-less sperm also occur in the
Opiliones, Palpigradi and Solifugae, while sperm in the remaining groups of arachnids
are flagellated.
Alberti (1995, 2002a, b; Alberti et al. 2000, 2007) summarises the huge morpho-
logical variety in arachnid, and especially acarine, sperm cells. Two major subdivisions
of mites can be recognised on the basis of sperm structure: the Opilioacarida,
Holothyrida, Ixodida and Mesostigmata have elongate spermatozoa; the Prostigmata,
Astigmata and Oribatida have more globular sperm cells. The size range of
sperm cells is enormous, from less than 2 μm in the prostigmatan Labidostoma
(=Nicoletiella) lutea (Labidostomatidae) to 1,000 μm in the tick Ornithodoros
tholozani (Argasidae) (Alberti 1995). Radwan (1996) found that size of spermatozoa
varies between individual males in the astigmatan Rhizoglyphus robini (Acaridae).
Sperm cells are often transferred to the female in an inactivated state, but once in
the milieu of the female reproductive tract, they become capacitated. Capacitation
involves changes in the morphology, physiology and behaviour of spermatozoa and
is widespread among animals. However, this process is particularly spectacular in
ticks in which the sperm cells double in length during capacitation, undergo
complete reorganisation and become capable of a creeping locomotion (Oliver 1982;
Sonenshine 1991). In the soft tick Ornithodoros moubata, sperm become highly
motile only if the female receiving them has had a recent blood meal (Resler et al.
2009); this may mean that only well-nourished females are likely to have their
eggs fertilized.
Males of most species of mites have accessory glands near the junction of the
vasa deferentia that secrete various products that are incorporated into the ejaculate.
Many acariformans produce spermatophores and the ejaculatory duct is often
modified into an array of internal sclerites and muscles termed the ejaculatory
complex that plays a role in shaping the spermatophore. Barr (1972) considered the
morphology of the ejaculatory complex to be phylogenetically informative within
water mites. Among mites that copulate, it is the rare species that transfers sperm
directly from the male’s primary genital opening to the female’s. Rather, various
parts of the male’s body may be modified for sperm transfer; these are described
below for particular taxa.
Diversity of Sperm-Transfer Behaviours in Mites 111

The Parasitiformes: Elaborations on a Theme

As far as is known, all bisexual parasitiformans copulate, with males using their
chelicerae to transfer sperm to the female’s copulatory opening. Ticks, some meso-
stigmatans (e.g. Epicriidae, Zerconidae, Uropodina, Sejoidea and Trigynaspida),
and presumably opiliacarans and holothyrans share a similar method of direct
transfer. In these groups, vacuolated sperm are placed by the male in the primary
female genital opening that leads to a pair of oviducts (Evans 1992; Walter and
Proctor 1999; Alberti 2002b; Krantz and Walter 2009). For those species in which
transfer has been observed, males extrude flask-shaped or globular spermatophores
from their genital openings between or behind the coxae of their second pair of legs.
The male then flexes his gnathosoma ventrally and extends one of his chelate
chelicerae towards the ‘neck’ of the spermatophore. Depending on the taxon, the palps
and/or the first pair of legs are used to pull the spermatophore from the genital
opening. In the Epicriidae, Zerconidae, Uropodina and Sejoidea, there are no obvious
structures specialized for spermatophore manipulation, but some male Trigynaspida
have modified sternal setae or ridges that may aid spermatophore transfer, and male
Celaenopsoidea have a clasp-like structure on the movable digit of the chelicerae.
In the Ixodida and many early derivative mesostigmatans, the male transfers
sperm into the female’s primary genital opening. This is termed tocospermy, where
toketos refers to ‘bringing forth young’, alluding to the role of the primary genital
opening in oviposition. In Mesostigmata, the variation in genetic systems and
behaviours is such that Alberti (2002b) has suggested the terms archispermy
(hypothetically ancestral vacuolated sperm delivered to the primary female genital
opening and thence through paired oviducts to reach the eggs) and neospermy
(modified sperm delivered to a primary or secondary opening leading to a single
oviduct) as in the Parasitidae and Dermanyssiae. In ticks, although the site of insem-
ination on the female is invariant, the location where mating takes place differs
between the two major families. In the Argasidae, which tend to feed quickly and
leave the host, sperm transfer takes place in resting sites away from the host; in the
Ixodidae, which have more lengthy associations with their food sources, mating
usually occurs on the host’s body (Yuval 1994), though Ixodes spp. may copulate
both on or off a host (Kiszewski et al. 2001). Assembly pheromones are produced
in some tick species to facilitate congregation for mating and may attract potential
mates to a currently parasitised host (e.g. Rechav et al. 1997).
Mating behaviour in the uropodine mite Caminella peraphora is particularly
interesting and mysterious. Females of this semi-aquatic mesostigmatan often bear
a ring of amorphous material around their opisthosomas, connected to a sac of similar
material resting on their dorsums. The origin and function of this ‘signet ring’ was
unclear until Compton and Krantz (1978) observed their reproductive behaviour.
Prior to sperm transfer, the male mounts the female and faces the opposite way.
The female produces a mass of ‘ring’ material from under her genital shield. The
male then moves posteriorly and ventrally so that he is beneath her facing the same
direction. He then presses a spermatophore directly from his genital opening into
112 5 Sex and Celibacy

Fig. 5.2 Chelicerae of a


male parasitid mite showing
slit-like spermatotreme
(arrow) (SEM by DE Walter)

the amorphous material near her anal opening. The spermatophore is enveloped in
the material and somehow a hollow tube forms in the material leading from the
spermatophore to the female’s genital shield. Compton & Krantz suggest that
sperm travels from the spermatophore to the female’s genital opening at this point.
The male returns to his original position and uses his legs to manipulate the
still- extruding ring material along the sides of the female up onto her dorsum.
The material eventually hardens to form a double-chambered cell on the female’s
back. Compton and Krantz felt that the embedding of the spermatophore in the ring
material is an adaptation to prevent water damage to the ejaculate. However, they
could offer no function for the double-chambered dorsal sac, nor for the male’s
involvement in its construction.
In the Parasitidae, males place the sack-like spermatophore into the female’s
endogynum (or endogynium), an often elaborate structure just internal to the primary
genital opening. From the endogynum, the sperm apparently enter the hemocoel to
reach the eggs (Alberti 2002b). Male parasitids have chelicerae that are modified to
hold the spermatophore. On the movable digit of each chelicera a simple slit or
more complex fine duct forms the spermatotreme (Di Palma et al. 2009) (Fig. 5.2).
At the very beginning of spermatophore extrusion, the male places a chelicera over
his genital opening. The spermatophore casing is extruded through the spermatotreme
opening in its collapsed state, and once through is filled with sperm and accessory
gland products so that the spermatotreme ends up holding the balloon-like
spermatophore by a narrow neck. Male parasitines have their genital openings
moved forward to just behind the tritosternum. This allows uptake of the ejaculate
without the extreme gnathosomal gymnastics required of males in taxa with more
posterior openings. All male Dermanyssiae share this derived spermatophore
opening and ribbon-shaped sperm and females have one large ovary and oviduct
(Alberti 2002b). Convergent movement of the male genital opening to the base of
the tritosternum is found in some Sejida and Antennophorina with vacuolated
sperm. The sperm cells of Parasitidae and Dermanyssiae are modified into elongate
structures and called ribbon sperm. Alberti (1995, 2002b) hypothesises that the
Diversity of Sperm-Transfer Behaviours in Mites 113

Fig. 5.3 Chelicerae of two


male dermanyssine mites
showing some rather simple
spermatodactyls (arrows)
(SEMs by DE Walter)

altered shape is an adaptation to the long journey to the ovary that the sperm make
once inside a female.
In the Dermanyssiae a complex structure on the chelicera called a spermatodactyl
(= sperm finger) (Fig. 5.3) is used to transfer ribbon sperm much like a hypodermic
needle (Di Palma et al. 2009). In Veigaiidae, the spermatodactyl may be a simple
tube (Di Palma et al. 2006) and is possibly intermediate between the slit-like
spermatotreme and more morphologically complex spermatodactyls (Di Palma et al.
2012). In some taxa, males produce spermatophores with distinct casings (e.g.
Rhodacaridae, Lee 1974), while others extrude seminal droplets from their genital
openings (e.g. Macrochelidae, Krantz and Wernz 1979). Some members of the
Dermanyssiae show the ancestral mode of sperm uptake by males, in which the
chelicerae are flexed ventrally to contact the ejaculate (e.g. Ologamasidae, Krantz
1978). Males of the Macrochelidae show a more derived behaviour. The male
extrudes a tubular extension of the genital chamber towards the chelicerae and
produces a droplet of seminal fluid (Krantz and Wernz 1979). The male dips the
arthrodial brush (a membranous tuft located at the junction of the fixed and movable
digits) of one chelicera into the fluid, presumably mopping it up. He then apposes
the arthrodial brush of that chelicera to the groove in the spermatodactyl of the
opposite chelicera, where the seminal fluid is transferred to the cheliceral tip by
capillary action. Some non-dermanyssines have independently evolved similar
sperm-transfer structures. Among the Heterozerconoidea, the Discozerconidae have
a cheliceral extension that arises from the fixed rather than movable digit and resem-
bles a spermatodactyl, while the Heterozerconidae have a spiral canal originating
from the base of the fixed digit (Di Palma et al. 2008). In males of the Celaenopsoidea
(Antennophorina), the movable digit has a ventral process that looks as if it may be
used to hold the neck of a spermatophore against the digit, but mating has not been
described in this group.
The Dermanyssiae have another unusual feature in their reproductive systems, in
this case associated with the female rather than the male. In this group sperm is
not transferred to the primary genital opening but rather to a sperm access system
via a secondary genital opening in the female that is sometimes far removed
from the primary opening. Female dermanyssines typically have a secondary
genital opening in the unsclerotised membrane at the base of their third pair of legs.
114 5 Sex and Celibacy

Fig. 5.4 Generalized venter


of female ologamasid mite
showing the variety of
placements of sperm-
receiving openings in the
Mesostigmata. The pore-like
openings are all bilaterally
symmetrical (After Lee 1974)

In the Rhodacaroidea, however, these openings may be near the bases of their fourth
pair of legs, on ventral shields posterior to their fifth pair, or even on the trochanters
and femora of their third pair (Fig. 5.4). The proximity of the secondary openings to
legs has resulted in this form of transfer being termed podospermy. A more general
term for insemination via a secondary genital opening, regardless of its location
on the female, is porospermy (Evans 1992). Although sperm transfer has not been
observed in the Heterozerconoidea, females of at least one species have what appear
to be secondary genital openings at the bases of their ventral sucker disks (Gerdeman
and Klompen 2003).
Dermanyssine females have two types of podospermal systems. In the laelapid
type, tubes from the left and right openings meet internally at a sac in which sperm
appears to be stored. In the phytoseiid type, there is no median merger of tubes and
sperm is stored in two separate sacs, one at the base of each secondary genital
opening. Whereas in ticks the spermatozoa move up the oviducts into the ovary
where fertilisation takes place, in Mesostigmata with ribbon sperm it appears that
spermatozoa migrate through the walls of the sperm-storage sacs, navigate through
the female’s haemocoel, and penetrate the walls of the ovary to gain access to the
ova within. Alberti (1995) suggests that the wall-penetrating ability of ribbon sperm
allowed the evolution of the strange podospermal insemination system of derman-
yssines from the more ‘traditional’ insemination site used by the Parasitidae.
The sperm-transfer behaviour of the Phytoseiidae, the best-studied dermanyssine
family, includes two basic pre-copulatory behaviours (Amano and Chant 1978;
Schulten 1985). In one type, shown by Amblyseius and Typhlodromus species, the
male climbs onto the back of the female so that both mites are facing in the same
direction. The male then turns 180° and moves posteriorly and under the female,
ending up venter-to-venter with both facing in the same direction. The other type is
shown by Phytoseiulus species where the male first contacts the female ‘head-to-head’
and then flips over on his dorsum and crawls beneath her, ending up in the same
final position as in the first type.
Diversity of Sperm-Transfer Behaviours in Mites 115

Insemination occurs when the male phytoseiid introduces the spermatodactyl of


a chelicera into one of the female’s copulatory pores (located between the third and
fourth coxae). It appears that a small, flask-shaped spermatophore is held between
the chelicerae and squeezed so that its contents are channelled into the copulatory
pore along a groove in the spermatodactyl. The spermatheca associated with that
pore swells and, within a few minutes, a wall-like structure appears to form around
the transferred ejaculate. Once formed, this wall remains for a few days. Multiple
inseminations are detectable by the presence of more than one of these structures in
the spermatheca. These objects are often referred to as ‘endospermatophores’ and
the flask-like hull that remains in the male’s chelicerae, the ‘ectospermatophore’.
However, as the membrane surrounding the sperm mass in the spermatheca appears
to be constructed after transfer, perhaps through some chemical interaction between
male and female fluids, ‘endospermatophore’ is a misleading name. True endo- and
ectospermatophores are present in the ejaculates of the Ixodida (see Exploding
sperm packets).
Lee (1974) describes and illustrates mating in a number of genera of Rhodacaroidea.
As these mites also belong to the Dermanyssiae, males have spermatodactyls and
females have secondary copulatory pores. Lee observed a peculiar behaviour in
males of Euepicrius filamentosus, Gamasellus traghardi, Antennolaelaps aremenae
and Athiasella dentata. Prior to extrusion of the spermatophore from the male’s
genital opening, he appeared to insert his chelicerae beneath the female’s epigyneal
shield (which covers the primary genital opening) and move them ‘in and out
alternately as if feeding on her genital region’. This is reminiscent of the behaviour
of ticks, in which males probe the female gonopore with their mouthparts (Feldman-
Muhsam 1986). Males may probe the genital opening of an allospecific female but
they do not go on to extrude a spermatophore. Presumably in both the rhodacarids
and ticks there are species-specific chemicals in the genital opening that males
sense with chemoreceptors on their mouthparts.
Lee states that after ‘tasting’ the female’s primary genitalia, the male extrudes a
large spermatophore about the size of the male’s sternum, hence approximately 1/4
to 1/3 of the male’s body length. Lee separated a male and female Athiasella dentata
before mating was completed and noted that the neck of the flask-like spermato-
phore was held between the digits of one chelicera. In Euepicrius filamentosus, the
male inserted the tips of both of his spermatodactyls simultaneously into one of
the female’s copulatory pores (at the base of the fourth coxa). The spermatophore
shrank as the contents appeared to be directed into the pore. Afterwards the male
consumed the remainder of the spermatophore’s contents and discarded the casing.
This sequence of behaviours was similar in the other species that Lee observed, with
males inseminating only one of the two copulatory pores. In contrast, Rath et al. (1991)
suggest that male Tropilaelaps clarae (Laelapidae) inseminate both left and right
copulatory pores during a single mating. Woyke (1994) also noted that male T. clarae
move from the left to the right side of the female during mating and added the
observation that this movement is interspersed with bouts of stroking, in which the
male rapidly runs his second leg over the epigynial area several hundred times.
Perhaps the stroking stimulates the female to take in sperm from the spermatophore.
116 5 Sex and Celibacy

In Phytoseioidea, the frequency of copulations often appears to be a species-level


characteristic. Inseminated females of some species have only a single endosper-
matophore, while others may carry eight or more in each spermatheca (Walter and
Proctor 1999).

The Adventurous Acariformes

Diversity of sperm-transfer modes is not uniformly distributed among acariform mites.


Within the Sarcoptiformes, astigmatans invariably copulate, while oribatids display
the opposite tendency and most have dissociated transfer. It is in the Prostigmata
that the greatest diversity of transfer modes lies, although again, some groups are
monotypic (e.g. the Tetranychoidea, Raphignathoidea, Cheyletoidea and Tarsonemina
all transfer sperm directly with sclerotised intromittent organs). In those groups
that do show variation (e.g. the Parasitengona), it is possible to test hypotheses about
evolutionary factors that lead to changes in sperm-transfer mode. For example,
Witte (1984) suggested that the evolution of swimming by several lineages of water
mites (Hydracarina) selected for direct sperm transfer because of the reduced
probability of females contacting spermatophores deposited on a substrate. Proctor
(1991a) tested this idea using a comparative approach and found that copulation had
evolved more often than expected in swimming water mites than in crawling ones;
however, this correlation did not explain all independent origins of copulation.
The Tydeidae also exhibit variation in transfer modes (dissociated, Schuster and
Schuster 1970; direct, Knop 1985) and may also be an appropriate taxon for
comparative analysis. Below we survey the sperm-transfer modes of the Acariformes
by suborder.

Typical Oribatida

Many hundreds of species of oribatids are all-female parthenogens (see


Parthenogenesis), and never produce functional male offspring. Sexual species
typically show dissociated sperm transfer in which males readily deposit stalked
spermatophores in the absence of females. Males have a complex of sclerites associ-
ated with the ejaculatory duct (the ‘spermatophoric organ’, analogous to the ejacu-
latory complex of water mites) used to shape the spermatophore. A fairly common
form of sexual dimorphism in oribatids involves modifications to the system of
dermal glands, in which one or more pairs in the male are enlarged or associated
with protuberances (Behan-Pelletier and Eamer 2005, 2010). These glands presum-
ably produce sex pheromones that allow males and females to find each other, or
females to find deposited spermatophores. Norton and Alberti (1997) suggest that
such highly modified glands may be particularly important in oribatids that live
in low-humidity environments, in which spermatophores may lose their viability
quickly.
Diversity of Sperm-Transfer Behaviours in Mites 117

Fig. 5.5 Dorsal view of the


male of a sexually dimorphic
arboreal oribatid mite
(Symbioribates sp.). The
blade-like rostral setae are
found only on the male
(Illustration by HC Proctor)

There are a few taxa in which there appears to be a closer association between
males and females but in which sperm transfer has not been observed. The male
Erogalumna zeucta (Galumnidae) walks behind the female with his prodorsum
beneath her opisthosoma and first pair of legs on her posterior (Evans 1992), as does
the male Collohmannia gigantea (Collohmanniidae) (Schuster 1962). In the latter
species, the small male occasionally places his rostrum against the anogenital region
of the huge female and shoves so violently as to turn her over. This is strangely similar
to the behaviour of male springtails in the genus Heterosminthurus (Collembola:
Sminthuridae) in which the male inserts and rubs his forehead setae in the genital
opening of the female (Bretfeld 1970). The rostral spines present on males of a
Symbioribates species (Symbioribatidae) from Queensland rainforests (Fig. 5.5)
may indicate that this species also engages in such genital stimulation.
No other interactions were observed for E. zeucta but Schuster described additional
odd behaviour for C. gigantea. After the pairing march proceeds for some time, the
male moves from behind the female to in front of her. There he leans forward to
balance on his first three pairs of legs with his fourth pair and his opisthosoma raised
high off the substrate. He extrudes his spermatophoric organ and deposits a droplet
(possible of semen, or some nutritive gift) on his apposed fourth legs. He then
presents his raised fourth legs to the female, who nibbles at them. During this the
male taps his hind legs up to 48 times with his spermatophoric organ. Schuster
supposed that sperm transfer occurred sometime after this display, but did not
observe it. As he saw no free-standing spermatophores in the mites’ containers, it
seems likely that a form of direct transfer occurs in C. gigantea.
The best evidence of (almost) direct transfer in an oribatid mite has been observed
in a species of Pilogalumna (Galumnidae) (Estrada-Venegas et al. 1996). Like
C. gigantea and E. zeucta, a male of this species initiates interactions with the larger
118 5 Sex and Celibacy

female by pushing his rostrum beneath her body, but in Pilogalumna, the male
deliberately tips the female on her side and then apposes his venter against hers.
The male extrudes a mass of material from his genital opening that accrues on the
female’s venter. The mass is about 40 % of the female’s body length and so must
represent a large expenditure of energy on the male’s part. The female moves the
ejaculate into her genital opening with her legs. As the authors point out, deposition
of spermatophores on the bodies of females is a rare behaviour among arthropods,
although it occurs in some crustaceans, and is common in onychophorans and
leeches. The behaviour of the water mite Physolimnesia australis (Limnesiidae)
may be similar in that the male appears to deposit a sperm mass on the female’s
dorsum (Proctor 1997). Because the males of these mite species do not put sperm in
the female’s sperm-receiving structure, this type of behaviour falls between the
direct and paired-indirect categories defined above.

Astigmata

There is little doubt that the Astigmata evolved within the Oribatida (Norton et al.
1993; Dabert et al. 2010), so the ‘Oribatida’ in the broad sense approach water mites
in diversity of transfer modes. Although oribatids sensu stricto almost always show
dissociated sperm transfer with a few examples of pairing, members of the Astigmata
invariably copulate. A copulating pair of astigmatans found in Baltic amber
indicates that this has been true for at least 40 million years (Klimov and Sidorchuk
2011). Not only do male astigmatans transfer sperm directly, they do so with a
sclerotised intromittent organ typically termed an aedeagus (morphology described
by Walzl 1991). Presumably the astigmatans evolved from a group of oribatids
that had close pairing between the sexes during sperm transfer; however, it has
been suggested that they are derived from a parthenogenetic lineage of oribatids
(see Immaculate conception below). In either case, the aedeagus is an autapomorphy
of the Astigmata.
Female astigmatans also display novel genitalic characters. Rather than accepting
the aedeagus into a true genital opening, the female receives it in a secondary genital
opening (copulatory pore, bursa copulatrix) located posterior to the true opening.
This is also apparent in the 40 MY old fossil mentioned above, in which the female’s
copulatory opening appears to be on a dorso-posterior protuberance (Klimov and
Sidorchuk 2011). In some feather mites, the female’s copulatory pore is located at
the end of a long sclerotised tube (spermaduct) extending beyond the terminus of
her opisthosoma. This external spermaduct is often longer than the male’s aedeagus
(Gaud and Atyeo 1996) (Fig. 5.6). In most Astigmata, the male’s aedeagus is situ-
ated near the coxae of the fourth pair of legs. To copulate, astigmatans assume a
variety of positions. In the Glycyphagidae, the male mounts the female so that both
mites face in the same direction (Evans 1992). In the Acaridae, male and female
face in opposite directions with the male’s venter apposed to the female’s dorsum.
Male acarids grip the smooth-skinned females using suckers on the fourth pair of
legs and near the anus (Fig. 5.7).
Diversity of Sperm-Transfer Behaviours in Mites 119

Fig. 5.6 Females of many


feather mite species have
external spermaducts that
extend outside their bodies,
such as that of this
Trouessartia sp. female
(Trouessartiidae) from the
black-faced bunting Emberiza
spodocephala (Aves:
Emberizidae) (SEMs by HC
Proctor)

Fig. 5.7 Sancassania


berlesei (Acaridae) in copula:
the male is inserting his
ventrally located aedeagus
into the dorsally located
copulatory opening of the
female (SEM by HC Proctor)

Many feather mites are strongly dorso-ventrally flattened. The male lies on top
of the female, facing in the opposite direction, often clasping her with specialised
hind legs and setae, and maintaining his grip with adanal suckers. In some species
the spermaduct of the female is inserted inside the male’s genital apparatus during
120 5 Sex and Celibacy

copulation (e.g., Trouessartia, Gaud and Atyeo 1996, pers. obs. HP). Male feather
mites often exhibit remarkable intraspecific polymorphisms, with a single species
presenting small female-like males (homeomorphs) and large, elaborately modi-
fied males (heteromorphs). Male polymorphism occurs in a number of other
mite taxa but it is especially common in the Astigmata (see Intrasexual competi-
tion below). Additionally, feather mites are unusual because asymmetrical males
with hypertrophied setae and legs on one side of the body occur in several taxa.
They may be the sole type of male (i.e. in species with monomorphic males) or only
the heteromorphs may show asymmetry. Right and left asymmetry appear to be
expressed equally. The utility of this asymmetry is mysterious (Gaud and Atyeo
1996), but Dubinin (1951) hypothesized that asymmetrical males are better able to
brace themselves in the slanted feather corridors while holding onto females and
thereby withstand aerodynamic stress (and hydrodynamic, for aquatic hosts) (for
illustrations see Chap. 9, Mites on and in feathers).
Boczek and Griffiths (1979) studied the mating behaviour and capacity of males of
two common stored-product pests, Acarus siro and Lardoglyphus konoi (both Acaridae).
After a male and virgin female mated, a single sac-like mass would appear in the
female’s seminal receptacle. The authors suggested that an elastic spermatophore
was transferred through the aedeagus into the receptacle; however, Evans (1992)
feels it more likely that a liquid ejaculate develops a wall once inside the spermatheca,
much the way that the ‘endospermatophore’ of phytoseiids arises. When single pairs
were kept together, after 8 days the average number was six spermatophores per
female (1–13) for L. konoi and 15 (9–23) for A. siro. When single males were main-
tained with 1, 2, 5 (and 10 for A. siro) females, Boczek and Griffiths found that extra
females induced male L. konoi to produce more spermatophores (four times more in
the 5-female vs. 1-female situation), whereas male A. siro consistently produced
about 25 spermatophores regardless of the number of females. For both species, the
female may initiate mating. The authors observed that a female would follow the path
of a male around the observation chamber but it was not until the opisthosoma of the
female came close to the male’s gnathosoma that he would show interest. This suggests
a localised release of pheromone on the female’s part. Similarly, putting a female in
a crowded chamber often caused males to attempt to mate with each other.

Prostigmata

The task of summarising mating behaviour in the Prostigmata is daunting because


of the diversity of sperm-transfer modes. The group most thoroughly surveyed, and
most diverse, is the cohort Parasitengona. Witte (1984, 1991) has reviewed sperm
transfer in terrestrial parasitengones and Proctor (1992) that in the Hydracarina.
Although some terrestrial parasitengones show pairing behaviour, none transfer
sperm directly and most are dissociated to some degree (e.g. males may require an
initial contact with a female to start depositing spermatophores but need no subse-
quent contacts). Male Allothrombium lerouxi (Trombidiidae) show lek-like
behaviour in which they defend small spermatophore-deposition territories from
Diversity of Sperm-Transfer Behaviours in Mites 121

Fig. 5.8 Courtship and sperm transfer in the water mite Neumannia papillator (Unionicolidae):
(a) Male trembles the forelegs in the direction of the female who is in a hunting stance; (b) The
male deposits a network of spermatophores and fans his hind legs over them (Illustrations by HC
Proctor)

other males (Moss 1960). Moss observed one male tapping a female that had entered
his territory but did not witness spermatophore uptake. Mutual tapping and moving
in a circular ‘dance’ are common to a number of trombidiid and erythraeid species
that show paired-indirect transfer.
Water mites (Hydracarina) display all possible types of spermatophore transfer.
Some species of Eylais (Eylaidae) show the (probably) ancestral behaviour of paired-
indirect transfer, complete with a circular dance. Many Neumania (Unionicolidae)
have a very complex courtship involving leg-trembling, foot-stomping and leg-fanning
(Proctor 1991b) (Fig. 5.8). Neumania species feed on copepods. It appears that the
leg vibrations of the courting male ‘deceive’ the female into striking at the male as
if he were a copepod. The male responds to being grasped by depositing a large and
complex mass of spermatophores and then fanning his hind legs over them. The female,
if virgin, responds to the fanning by running towards the male and hence over the
spermatophores. The sperm packets are sticky and become attached to the female’s
venter, and she later brushes them into her genital opening. Why would the male
want to induce predatory behaviour in the female? Males respond to contacting
the bodies of females and contacting substrates recently occupied by females by
trembling, but only if he trembles in front of a female will he be grasped. So the
male likely benefits by using the female’s response to ascertain that he is indeed
courting a real female rather than the chemical residues of one, and thereby avoids
wasting his spermatophores.
From this paired, indirect mode of sperm transfer, water mites have diversified in
both directions. Many taxa display dissociated transfer of spermatophores (Proctor
1992), including some Neumania and species in the related genus Unionicola.
122 5 Sex and Celibacy

Dissociation appears to be a risky business because a male cannot be sure that a


female will find his spermatophores. Methods employed by males of dissociated
species to increase the probability of spermatophore encounter by females are
discussed in Fields of Fragrant Spermatophores below.
Direct sperm transfer has evolved repeatedly in different lineages of water mites.
Proctor (1991a) estimated that there have been 91 independent evolutions of copu-
lation in the approximately 6,000 species of Hydracarina. A robust phylogeny of
water mites is needed to determine whether this rather high estimate is justified.
Species whose males are modified for copulation occur in all major superfamilies
except for the monogeneric Hydrachnidoidea. In some cases, the presumably ancestral
shape of a stalked spermatophore with a sperm-filled apical capsule is retained with
almost no change, while in others the outward appearance is greatly altered. A male
Piona (Pionidae) transfers jelly-like blobs from his genital opening to the female’s
using his modified third pair of legs; however, examination of the internal structure
of these blobs reveals them to be a bundle of miniature spermatophores (14–18 in
P. carnea) embedded in a gelatinous matrix (Leimann 1991). Proctor (1991a) found
evidence that lineages composed of swimming mites were more likely to evolve
copulation than those made up of crawling mites; however, the correlation was far
from perfect. There are many species of water mites that copulate but do not swim,
or that swim but do not copulate. Factors responsible for changes in sperm-transfer
mode are likely to be specific to different taxa. For example, Unionicola species
endoparasitic in mussels are much more likely to transfer sperm directly than are
free-living congeners. Proctor (1992) suggested that the slimy substrate inside the
mantle cavity selects against deposition of spermatophores.
Kirchner (1967) described the sperm transfer of the marine mite Halacarellus
basteri (Halacaridae). Males do not deposit spermatophores unless females are
present. When a male and female meet, the female touches the male with her palps
and then climbs on top of him. The male then raises the fore part of his body and
lowers his genital plates to ‘scour’ the substrate. In this way, the female is lifted
rhythmically up to a dozen times. Kirchner felt that this interaction was essential to
induce the males to deposit their elaborate spermatophores. He did not describe
sperm uptake by females.
Among other prostigmatan mites, there are no known examples of indirect
transfer with pairing outside of the Parasitengona, Anystina and Eupodina. However,
many groups show dissociated transfer, including Bdellidae (Alberti 1974 ),
Labidostommatidae (Vistorin 1978), Nanorchestidae (Schuster and Schuster 1977),
Rhagidiidae (Ehrnsberger 1977), and some Tydeidae (Schuster and Schuster 1970)
and Eriophyoidea (Oldfield and Michalska 1996). Rhagidiid males do not produce
stalked spermatophores; rather, they extrude thin threads from their genital openings
that bear one (Rhagidia arenaria, R. clavicrinita) or several (R. halophila) droplets
of liquid ejaculate (Ehrnsberger 1977). Males of the other families mentioned
produce stalked spermatophores.
Oldfield and Michalska (1996) review sperm-transfer behaviour in gall mites
(Eriophyoidea). Spermatophores resemble lily pads, with a short thin stalk and a broad,
flattened, leaf-like head that is horizontal relative to the substrate. The spermatozoa
Diversity of Sperm-Transfer Behaviours in Mites 123

are contained in a central reservoir in the head. Only about 50 spermatozoa are held
by each spermatophore. When pre-emergent (pharate) females are present, males of
some species deposit their spermatophores near her body, sometimes entirely
surrounding her with a fence of ejaculates. When females are not nearby, males tend
to deposit dense clusters of up to 100 spermatophores near their normal feeding
sites. Aculus fockeui (Eriophyidae) males deposit spermatophores at a rate of about
30 per day for two weeks, for a total of 420 spermatophores in a male’s lifetime.
Insemination occurs when a recently emerged female encounters a spermatophore,
mounts it and takes sperm into her genital opening from the central reservoir, leaving
the stalk and deflated head behind. Michalska (2011) compared spermatophore
output between male A. fockeui, which don’t deposit preferentially near females,
and Aculops allotrichus, which deposit spermatophores near the bodies of quiescent
female tritonymphs. The expectation from sperm-allocation theory would be that
A. fockeui should produce many small spermatophores (increasing the probability
of encounter with females) and A. allotrichus should produce fewer larger ones.
The first prediction was upheld, with A. fockeui depositing 5x more spermatophores
on average per day than A. allotrichus; however, contrary to expectations, A. fockeui’s
spermatophores were also larger and contained slightly more spermatozoa those
of A. allotrichus. A confounding factor is that both male and female A. fockeui are
larger than the respective sexes of A. allotrichus, so a better comparison might
involve standardization of ejaculate output per unit body mass.
Female Aculus robiniae search intensively in the area where they moult, pre-
sumably because males usually encircle the pharate females with spermatophores.
Spermatozoa are stored in the spermathecae. Oddly, there appears to be both a
systematic and an ecological difference in whether one or both spermathecae are
used. Diptilomiopid females and eriophyid females that feed on dicotyledonous
plants store sperm in only one spermatheca, while phytoptid females and eriophyids
from monocots store sperm in both spermathecae.
Among the Prostigmata, direct sperm transfer is exhibited by many Hydracarina
(Proctor 1992), Adamystidae (Coineau 1976), Tetranychidae (Cone 1985), Tarsonemina
(Evans 1992), Cheyletidae (Summers and Witt 1973), Pterygosomatidae (Mostafa
1974) and some Tydeidae (Knop 1985). The adamystid Saxidromus delamarei
(sometimes placed in a different family, the Saxidromidae), like many species of
Arrenurus water mites (e.g. Proctor and Smith 1994), has a form of direct transfer
that appears to have recently evolved from indirect transfer. When a sexually
interested male and female meet, the female crawls on top of the male, facing the
opposite direction (Coineau 1976). The male’s first pair or legs, with which he
holds the female’s third and pair of legs, are more robust and presumably stronger
than those of the female. The male presses his genital aperture to the substrate and
lifts his posterior opisthosoma, drawing out a large, thick-stalked spermatophore.
The male then rotates 180° so that the female’s genital opening is above the
spermatophore tip. He then bows down and lowers the female so that she is impaled
on the spermatophore. The top half of the spermatophore is broken off inside the
female. This action requires the help of the male, and if he is disturbed at this point
the female may be left to die in this uncomfortable position. Thus the male is
124 5 Sex and Celibacy

responsible for the placement of the sperm but does not insert it with an intromittent
organ. In two other genera of adamystids, Bovidromus and Rhinodromus, the initial
capture of the female is similar, but spermatophore morphology is very different
(Alberti et al. 2010). Males produce tiny thin-stalked spermatophores bearing
simple liquid droplets. After pressing the female onto the spermatophores, male
Bovidromus and Rhinodromus can turn around, add another droplet of ejaculate, and
then repeat the process up to a total of five times. This economy of stalk use may
energetically benefit males of these taxa, and it may also benefit the females, who
do not run the risk of being abandoned to die in an impaled state. A very unusual
aspect of adamystid reproductive biology is the male’s production of synspermia –
aggregates of four sperm cells that do not separate during meiosis. The number of
synspermia per spermatophore varies from several (for Saxidromus) to only a single
one (for the other two genera). Whether a synspermium is capable of fertilizing
more than one egg is unclear. The only other animals that transfer synspermia
during mating are a few species of spiders (Alberti et al. 2010).
Male Pterygosoma mutabilis (Pterygosomatidae) demonstrate behaviour inter-
mediate between that of Saxidromus delamarei and that of taxa with intromittent
structures. Pterygosma mutabilis is an ectoparasite of agamid lizards that spends
most of its life beneath the host’s scales, feeding on blood (Mostafa 1974). Males
have a dorsal genital opening. Mostafa observed one male wandering over the
surface of his host, apparently in search of a female. The male had extruded a long,
sac-like spermatophore from his genital opening and was carrying it on his back.
Then the male discovered a female resting nearby and attempted to insert the
spermatophore into her genital opening by walking beneath her, lowering his
anterior end and raising his posterior. The female refused to open her genital
valves and the male eventually gave up. Mostafa noted that the male then tried to
draw his spermatophore back into his genital pouch, presumably for use on a more
fortuitous occasion.

Spermatophore Structure and Function

External spermatophore morphology in mites ranges from the stripped-down and


purely functional droplet spermatophores of oribatids to the baroquely decorated
ejaculates of halacarids and bdellids. A sampler of spermatophores from just
one clade, the Hydracarina, illustrates this great diversity in morphology (Fig. 5.9).
Although some of this structure may be designed to internally court females, as
suggested by Eberhard (1985), other structural aspects are likely to reflect selection
for an increase in the stability of spermatophores or for an increase in the encounter
probability of spermatophores and females in dissociated species (Proctor et al. 1995).
Internally, even oribatid spermatophores display a surprising complexity of layers
(Alberti et al. 1991). The most intricate internal morphology is shown by the
spermatophores of ticks: their complex and mysterious structure is discussed below.
Diversity of Sperm-Transfer Behaviours in Mites 125

Fig. 5.9 Diversity of spermatophore structure in the Hydracarina (Illustrations by HC Proctor)

Exploding Sperm Packets

Sonenshine (1991) summarises the structure and behaviour of tick spermatophores.


After a male tick finds a suitable and interested female, he produces a flask-shaped
spermatophore that is composed of two basic layers, the ecto- and endospermato-
phore. First he exudes a clear droplet from his genital opening that is composed of
a viscous mixture of proteins and mucopolysaccharides. This liquid congeals at
the surface to form a membranous skin. Into this ectospermatophore flow opaque
seminal fluids and immature sperm cells (prospermia). Then the endospermatophore
is introduced into the ectospermatophore, sealing it at the male’s genital opening.
Argasids introduce two endospermatophores into each ectospermatophore but ixodids
126 5 Sex and Celibacy

produce only one. In both groups, the endospermatophore is bilobed and filled
with several proteinaceous substances, one of which may inhibit the proliferation
of bacteria.
The strangest ingredient found in the spermatophores of some species is the
symbiotic yeast-like fungus Adlerocystis. Adlerocysts may help to maintain viability
of the spermatozoa until the female is ready to oviposit. Evidence for this is that in
ticks that oviposit soon after copulation, adlerocysts do not attach to sperm cells,
while in those species that may experience a long delay between copulation and
oviposition – up to years in some Argasidae – the symbionts do adhere to the
spermatozoa (Feldman-Muhsam 1991).
When the spermatophore is complete, a process taking only about 30 s, the
male grasps it with his chelicerae and inserts it into the female’s genital opening.
In argasids, the male reaches down to his gonopore to grasp the emergent sper-
matophore but in ixodids the male maintains his chelicerae in the female genital
opening, where he introduced them prior to spermatophore production, and moves
the spermatophore to his mouthparts via contortions of his body. Sonenshine
(1991) states that the spermatophore is not a passive tool in sperm transfer but
rather “is an active partner in copulation and itself accomplishes the process of
sperm transfer” (p. 321). Once the spermatophore is placed on the female’s genital
opening, a CO2 ‘bomb’ goes off inside the ectospermatophore, causing the neck
of the ejaculate to extend into the female’s gonopore, forcing the contents – sperm,
seminal fluid and Adlerocystis – into the endospermatophore and shooting the
endospermatophore into the female’s reproductive tract. This extrusive explosion
continues even if the spermatophore is separated from the ticks. In argasids, the
paired endospermatophores head off into the paired uteri, while in ixodids the single
endospermatophore is held in the female’s seminal receptacle. The ectospermato-
phore, its role complete, drops off the female. The sperm cells complete their
capacitation within the female (see Reproductive anatomy) and in argasids may
be stored in the uterus for many years, allowing the female to fertilise her eggs
without seeking further copulations.

Fields of Fragrant Spermatophores

To humans, with our reliance on direct sperm transfer, the feature of dissociated
transfer that seems most precarious is the apparently low probability that a small
female mite will stumble across a much smaller droplet of sperm in a vast volume
of soil or water. Males of dissociated species, however, have evolved an array of
devices that increase the chances of females finding their ejaculate. Most obvious
among these are signalling threads and tracks, which are produced by oribatids and
many terrestrial and aquatic prostigmatans, as well as other dissociated arthropods
(millipedes, pauropods, pseudoscorpions) (Proctor 1998). These filaments mechan-
ically increase the radius of interception of a spermatophore. They can also help
Diversity of Sperm-Transfer Behaviours in Mites 127

guide the female to the location of the ejaculate, which may be at the centre of
radiating threads (Oppedisano et al. 1995) or at the end of a zigzag track of decreasing
width (e.g. Witte 1991).
Terrestrial parasitengones have the most elaborate and diverse repertoire of
signalling devices. For example, Johnstoniana errans (Johnstonianidae) deposits a
linear zigzag track at the midpoint of which is the spermatophore; at each change of
angle is a ‘signalling stalk’ that looks like a scaled-down spermatophore minus
the sperm cells (Witte 1991). Some dissociated species of water mites also have
such signalling devices (e.g. Hydrachna, Limnochares) but in comparison to their
terrestrial relatives, their tracks are crude and haphazard.
Less readily observed but likely common to all species with indirect sperm transfer
(both paired and dissociated) is the use of spermatophore-associated pheromones to
attract females. Behavioural evidence for pheromones include observations of mites
moving towards spermatophores from a distance, males touching their genital openings
on the substrate before or after depositing spermatophores, males fanning their legs
over spermatophores to direct water currents towards females and the loss of attrac-
tiveness of spermatophores over time (implying dissipation of a chemical) (Proctor
1998). Witte (1984) suggested that attraction of females through spermatophore-
associated pheromones is more important in swimming water mites, where females
may move through the water column rather than being restricted to a two-dimensional
substrate. In such situations, deposition of huge fields of fragrant spermatophores
may allow females to detect the presence of ejaculates even when swimming
some distance from the source. Extensive, dense ‘lawns’ of spermatophores are
produced by male Hydrachna (Hydrachnidae) and clumps of dozens to hundreds of
spermatophores are also commonly produced by male Limnesia (Limnesiidae) and
Hydrodroma (Hydrodromidae) (Proctor 1992).
As well as attractants that act at a distance, there are undoubtedly contact
chemicals that allow females to recognise spermatophores of their own species.
This would be particularly important for dissociated species that share deposition
sites with closely related taxa. To our knowledge, no studies have investigated
whether females of dissociated mite taxa differentiate between spermatophores of
conspecifics and related congeners that share the same habitat. However, Oldfield
and Michalska (1996) describe spermatophore recognition by female gall mites
from different host plants and note that, for some species, proper discrimination
requires that the female be on her own host.
A third way to increase the probability of contact between females and spermato-
phores involves the placement of ejaculates where females are likely to encounter
them in the course of their daily activities. For example, the water mite
Hydrodroma despiciens (Hydrodromidae) feeds on the eggs of chironomid midges.
Male H. despiciens deposit spermatophores preferentially on these egg masses
(Proctor 1992). Females in search of food will thus also encounter a supply of
sperm. Males of some gall mites encircle the bodies of quiescent pre-adult females
with spermatophores so that when they emerge as adults, the sperm will be conve-
niently within reach (Oldfield and Michalska 1996; Michalska 2011).
128 5 Sex and Celibacy

Sexual Selection

When Darwin (1871) split off sexual from natural selection, he considered only two
categories: conflict between members of one sex (intrasexual competition) and
attraction between members of the opposite sexes (intersexual selection, typically
female choice of males). However, a third category has recently been identified:
conflicts of interest between the sexes (intersexual conflict). Most examples of sexual
selection in mites fall into the first category; however, evidence for female choice
and intersexual conflict is growing.

Intrasexual Competition: Male Modifications


for Mate Monopolisation

Competition between males for access to the eggs of females has resulted in the
evolution of a range of secondary sexual characters, both behavioural and morpho-
logical. This is as true for mites as it is for all other organisms with separate
sexes. One major difference between mites and the best studied group of arthropods,
insects, is that sequestering of a female after insemination is widespread in Insecta
(Alcock 1994) but is extremely uncommon in the Acari. The only documented
examples in mites are Rhizoglyphus robini (Radwan and Siva-Jothy 1996), the
special cases of biparental defence against predators in spider mites (Tetranychidae)
(Saito 1986a, b) and of harems in mussel-inhabiting water mites (e.g. Davids et al. 1988,
see Mate-guarding and aggression below) and spider mites (Saito 1990).
In contrast, pre-copulatory (or more precisely, pre-inseminatory) guarding is
very common. This may be because: female mites often become unreceptive soon
after mating (e.g. Tetranychidae, Cone 1985, and likely also Histiostomatidae,
Fashing 2008); because the period between mating and oviposition is very long and
so a male would gain more by seeking other mates than by sequestering his current
one; or because first-male sperm precedence is very strong. In Arrenurus water
mites, male and female often remain physically attached for hours after sperm
transfer has occurred (Proctor and Smith 1994). This may be a type of mate-guarding
(Alcock 1994), or it may be a subtle form of post-copulatory courtship that is neces-
sary to convince a female to take up sperm (see Eberhard 1985); this would place
the behaviour in the realm of intersexual selection through female choice, although
the distinction between the two types of sexual selection is often subtle.

Mate-Guarding and Aggression

Guarding of pre-imaginal females occurs in many groups of mites with direct sperm
transfer. When it involves close physical contact between the male and pharate
female it is termed precopula. In the Mesostigmata, pre-copulatory guarding has
been observed in the Phytoseiidae, Blattisociidae, Laelapidae, Ameroseiidae and
Sexual Selection 129

Fig. 5.10 Males of many Proctophyllodes species (as this one from the red-breasted nuthatch Sitta
canadensis) use their tube-like anal suckers (a, c) to attach to finger-like extensions from the
posterior dorsum of female tritonymphs (b, d) (Pictures by K. Byers (with permission))

Macrochelidae (Yasui 1988, Walter pers. obs.). In some ascid and laelapid mites,
the males may even ride around on the backs of female deutonymphs and adults.
In the Astigmata, precopula has been observed in many taxa including the
Histiostomatidae (Fashing 2008), Psorpotidae, Chirodiscidae (Evans 1992) and
many families of feather mites (Witalinski et al. 1992). Males of these groups use
modified limbs and/or suckers on the legs or next to the anal opening to hold onto
the female tritonymphs, and the tritonymphs in turn may have special tubercles,
disks or lobes that the males attach to, holding on until the adult female emerges
(Fig. 5.10). Witalinski (1990) and Witalinski et al. (1992) discuss the fine structure
130 5 Sex and Celibacy

Fig. 5.11 Male Broad Mite


Polyphagotarsonemus latus
(Tarsonemidae) carrying a
pharate female (Illustration
by DE Walter and C Harvey)

of adanal suckers in male astigmatans. Fashing (2008) describes how males of the
unusual histiostomatid genus Creutzeria, which live in the water inside pitcher
plants, deal with the complexities of carrying around a tritonymph while swimming.
The first legs of males are greatly elongated relative to those of the females and
nymphs, and also have modified tibial setae that together with the tibiae form a
deep groove into which the trochanters of the tritonymph’s legs II fit snugly. A male
Creutzeria wraps his legs I around the pre-adult female, clicks the tibial grooves
into place, and can continue swimming with his cargo safely locked into his
undercarriage. The Astigmata also demonstrate a rare occurrence of post-copulatory
contact that appears to serve as mate-guarding. Radwan and Siva-Jothy (1996) have
demonstrated that in Rhizoglyphus robini (Acaridae), a species in which individuals
of both sexes may mate several times a day, a male increases his fertilisation of a
clutch by prolonging his attachment to a female for up to 6 h after copulation. Since
many astigmatans stay in copula for extended periods, this phenomenon may be
of more general distribution.
In the Prostigmata, precopula occurs in the Tarsonemina, Tydeidae, Cheyletidae,
Eriophyidae and Tetranychidae (Michalska and Boczek 1991; Evans 1992). In the
Tarsonemina, males use their modified fourth pair of legs (when present) and a
sucker-like device termed a genital capsule to hold onto the pharate female (within
the larval skin) (Fig. 5.11). Garga et al. (1997) observed that the fourth pair of legs
of Tarsonemus confusus (Tarsonemidae) are also used in aggressive contests
between males for possession of pharate females and that the size of these was
a better predictor of which male would win than was male body size. A male
Tarsonemus may trundle about with an attached pharate female for up to 24 h before
the adult female emerges; however, the act of mating takes only a few minutes at
most, after which the pair separates (Evans 1992). In the Podapolipidae, another
group in the Tarsonemina, males seem to have grown impatient and copulate directly
Sexual Selection 131

with the skin of the larval female (Evans 1992), presumably fertilising the pharate
adult female beneath. In the Eriophyidae, male Vasates robiniae will guard quiescent
deutonymphal females by lying beside their still forms until the adult female
emerges, grappling with other males that attempt to take over (Michalska and
Boczek 1991). If the first male fails to chase away the second, the latter may become
a co-guarder. In the late summer, when pharate females are scarce, there may be as
many as eight males guarding each deutonymph. Sperm transfer in eriophyids is
indirect, so this is an example of pre-inseminatory rather than pre-copulatory
guarding.
Pre-copulatory associations and male aggression have been very well studied in
the Tetranychidae. Mites of this family lack the tritonymphal stage, moulting
directly from deutonymph to adult. Deutonymphal females close to transformation
construct a silken pad on which they settle to await their moult (Cone 1985). The
closer a female is to eclosion the more attractive she is to males, presumably through
increased release of arrestant pheromones. When a male encounters an attractive
pharate female, he settles down beside her, ready to defend his nubile booty. The
male lays down several strands of silk over the quiescent female; these allow him to
monitor both her movement prior to moulting and the arrival of any other male.
Females typically are interested in mating for only a short while immediately after
eclosion, whereas males are able to mate many times; so there is an oversupply of
amorous males, which leads to great competition for pharate females. Males vie for
possession of females by sparring with their front legs, by spitting silk and by
stabbing at each other with extended chelicerae (Potter et al. 1976; Enders 1993).
Occasionally, one male pierces the other, resulting in the rapid exsanguination and
death of the wounded male. Larger males are usually successful at defending their
females and in some spider mites the males are much larger than females (see Fig. 8.5).
If neither male can overpower the other they may become co-guarders of the female;
Potter et al. (1976) found that the co-guarding male that assumes a position on top
of the female tends to be the first to copulate.
The greatest extremes of mate-guarding are shown by males that keep harems.
This behaviour has been best studied in mammals such as elk, horses and elephant
seals, in which the harem-master has greatly elevated reproductive success relative
to those males without harems. For a male to sequester a group of females and guard
them from other suitors, the females must be defensible (Emlen and Oring 1977).
The tendency of most ungulates to form herds may allow stallions and stags to
guard females and the limited space on breeding beaches makes calving females
defensible for male elephant seals.
Harem-keeping has evolved in two groups of mites with very different lifestyles.
Many species in the water mite genus Unionicola (Unionicolidae) dwell inside the
mantle cavities of freshwater mussels where they feed on host tissues (see Chap. 7).
Sex ratios of mites inside mussels are often extremely female-biased. One explana-
tion for such biases is female-skewed primary sex ratios such as are found in many
mites and insects inhabiting small patches of habitat (see Sex ratio manipulation).
But, in such cases, it is typically the closely related descendants of a single founding
female that occupy such patches. In contrast, Unionicola in mussels are very
132 5 Sex and Celibacy

unlikely to be related to each other because larvae disperse from natal mussels
in search of midges to parasitise and new mites entering mussels are likely to be
derived from various sources.
Dimock (1983) studied the behaviour of male Unionicola formosa from mussels
in North Carolina. Most mussels (86 %) harboured only a single male, whereas
females ranged from 6 to 78 per mussel. Dimock found only 23 of 360 mussels to
have more than one male and in these cases it was usually a smaller, perhaps recently
emerged, male paired with a larger one. To determine whether the sex ratio could be
a product of intermale aggression, he compared the survivorship of male–male
pairs, male–female pairs and female–female pairs left together for 3 days.
Survivorship in pairs with one or two females was almost 100 % but all male–male
pairs were reduced to a single living male. Combat between male U. formosa
involved grappling until one male grasped the other’s leg and pierced his rival with
his chelicerae. To determine whether mortal combat also took place within the
mussels, Dimock introduced additional males into the mantle cavities of freshly
collected bivalves. After three days he removed the inhabitants and found in 23 of 24
cases only a single living male was present. In the 24th mussel, two dead males
were found locked together. Edwards and Dimock (1991) found that larger male
U. formosa were much more likely to emerge victorious from paired combat. They
also observed that the palps of males are larger than those of females, presumably
the result of intrasexual selection for stronger weapons.
Not all mussel-dwelling Unionicola show such aggression. Davids et al. (1988)
compared the behaviour of males of the European species U. ypsilophora and
U. intermedia. Like U. formosa, males of U. ypsilophora fight to the death. In contrast,
male U. intermedia show no aggression towards potential rivals. The European
species also differ in male phenology (U. ypsilophora males are present throughout
the year, U. intermedia only in the summer months) and in mode of sperm transfer
(U. ypsilophora males are not modified for direct transfer, U. intermedia are);
however, it is not clear whether one or both of these differences contributes to the
contrast in intrasexual aggression. Differences in aggression may also explain why
U. ypsilophora are able to exclude U. intermedia from the mussel Anodonta cygnea,
a host that the latter species will use in the absence of U. ypsilophora.
Similarly, Saito (1990) observed variation in intermale aggression between two
congeneric species of spider mite (Tetranychidae). Male Schizotetranychus longus
tolerate the presence of other males, whereas those of S. miscanthi are extremely
aggressive, climbing on rivals and stabbing them to death with their cheliceral stylets.
Both Schizotetranychus species live in communal nests composed of numerous
adults, juveniles and eggs. Saito surveyed the inhabitants of 110 S. longus and
86 S. miscanthi nests. For S. longus there were 0–32 males per nest, while for
S. miscanthi the range was 0–3. For the first species, numbers of males per nest
increased with overall nest size, while for the latter there was no such correlation:
most nests contained only a single male monopolising a dozen or so females. Saito
could think of no demographic or behavioural factors that could explain why one
species shows harem polygyny and the other a scramble-type polygyny.
Sexual Selection 133

Aggressive behaviour may also vary intraspecifically. As mentioned previously,


male polymorphism (andropolymorphism) occurs in a number of mite taxa, including
the Dermanyssidae, Digamasellidae, Laelapidae, Macrochelidae, Parasitidae
(Mesostigmata), Cheyletidae and Anystidae (Prostigmata), Acaridae and numerous
families of feather mites (Astigmata) (Gaud and Atyeo 1996; Gwiazdowicz 2004;
Timms et al. 1981). In general, the highly modified morph is termed a heteromorph,
and the one that is more similar to the female is a homeomorph (see below for a
more detailed scheme for the Acaridae). Male feather mites of many species show
an additional twist to polymorphism in that heteromorphs may display asymmetry
of the body, legs or setae. According to Gaud and Atyeo (1996), this asymmetry is
equally expressed on left and right sides; the adaptive value of such asymmetry,
however, is unknown.
The causes and consequences of andropolymorphism have been best studied in
the Acaridae, where the existence of multiple male forms has been known for more
than a century. Polymorphism occurs only in the subfamily Acarinae, where males
are represented by up to four morphs (Timms et al. 1981). The homeomorph has a
third pair of legs that are similar in size and shape to those of the female and has
body setae of a similar length to the female’s. The bimorph has a longer body and
longer dorsal setae than the female. The heteromorph has a very much thickened
pair of legs, typically the third pair, but has setae are similar to the female.
The pleomorph has a pair of thickened legs and has a longer body and longer setae
than the female, i.e. it has the characteristics of the bimorph and heteromorph
combined. Occasionally malformed males with only a single thickened leg occur.
Not all species exhibit all possible morphs and the presence and ratio of different
morphs may also vary between populations of the same species. For those taxa in
which behaviour has been observed, pleomorphs are aggressive fighters that attack
other males, while other morphs are comparatively placid. Morph type may be at
least partially inherited (e.g. in Rhizoglyphus robini, Radwan 1995; Smallegange
2011) or determined by environment, as Timms et al. (1981) showed for Sancassania
(=Caloglyphus) berlesei. Regardless of paternal type, male S. berlesei reared in
low-density situations became pleomorphs (well armed), while those raised in
low-density chambers grew up to be bimorphs (poorly armed). Timms et al. (1980a)
determined that a volatile chemical (or mixture of chemicals) was produced by large
colonies of S. berlesei that suppressed the production of well-armed pleomorphic
males. In a companion study (Timms et al. 1980b) they found that pleomorphic
males attack and kill newly emerged bimorphs and pleomorphs by grasping them
firmly with their modified third pair of legs and puncturing the integument with the
fourth pair. Pleomorphs killed bimorphs regardless of whether females were
present. Older males of both types were more heavily sclerotised and better able to
survive attacks. If two older, well-armed pleomorphs encountered each other ‘they
would back off and advance in different directions’. Bimorphs were non-aggressive,
both among each other and towards pleomorphs. Some species of harvestmen
show similar morphological and behavioural dimorphism in males, with aggressive
large-legged ‘majors’ and peaceful small-legged ‘minors’ (Zatz et al. 2011).
134 5 Sex and Celibacy

Fig. 5.12 Male Falculifer rostratus (Falculiferidae) from the rock pigeon (Columba livia) come in
two morphs: a female-like homeomorph and a highly modified heteromorph with enlarged front
legs and greatly extended lower digits of the chelicerae (Pictures by HC Proctor)

Radwan (1993) investigated the trade-off between being an aggressive pleomorph


(‘fighter’) versus a passive bimorph (‘scrambler’) in C. berlesei. He constructed
colonies of two sizes: small, with two individuals each of the fighter and non-fighter
morphs and six females; and large, with 30 individuals of each morph and 80
females. He found that in small colonies, aggressive pleomorphs had higher repro-
ductive success because a single male could kill many, often all, of his rivals. In large
colonies, fighters had lower reproductive success because they spent more time
fighting, less time copulating and were much more likely to be killed than were non-
fighters. Radwan suggests that pleomorph males will do best in recently founded
(and so small) colonies where individual males can readily do away with their rivals,
while peaceful scramblers do best in large colonies where the rate of encounter
between males would mean that pleomorphs would spend all of their time fighting
and little time mating. This hypothesis matches well with the observations of Timms
et al. (1980a) that development of pleomorphs is suppressed in large colonies.
In contrast, Proctor et al. (2009) did not find a similar negative correlation between
population density of the feather mite Falculifer rostratus (Falculiferidae) on rock
pigeons and the proportion of the highly modified heteromorphic males (Fig. 5.12).
Sexual Selection 135

Fig. 5.13 Pre-copulatory


guarding. This large male
feather mite, Dicamaralges
dasybrachius (Psoroptoididae),
is holding on to a tritonymphal
female with the aid of tarsal
suckers on his fourth legs.
He will mate with her when
she moults into an adult
(SEM by HC Proctor)

In a very well-designed experiment, Smallegange (2011) tested whether fighter


and scrambler morphs in male bulb mites, Rhizoglyphus robini, were determined by
the morphology of their fathers, the nutritional status of their mothers, or the quality
of food consumed during their own development. She found that females mated to
fighters that were fed a high-calorie diet tended to lay larger eggs, which were more
likely to develop into fighter males. A rich diet during juvenile instars also increased
the probability that a male would end up a fighter. Paternal genotype did have
an effect, but it was only apparent when juvenile males were raised on a poor diet,
in which case sons of fighters were more likely to grow up to be fighters.
Sexual size-dimorphism, in which one sex is clearly larger than the other, tends
to be female-biased in most mites. In feather mites, however, both male- and female-
biased dimorphism occur in different taxa. Male:female body lengths vary from 0.51
(Epidermoptidae: Microlichus sp.) to 1.22 (Psoroptoididae: Dicamaralges dasybrachius)
(Fig. 5.13) (measurements from Gaud and Atyeo 1996). Larger-bodied males tend
to also show strong modifications for mate-guarding, such as massively enlarged legs.
Colwell (2000) observed similar variation in male vs female size in hummingbird-
flower mites (Mesostigmata: Melicharidae). In these mites, but not in feather mites
(K. Byers, unpublished data), the pattern follows Rensch’s Rule, which is that in
large-bodied species, males tend to be larger than females, and in smaller-bodied
species, females are larger than males. Colwell argues that this pattern could be the
136 5 Sex and Celibacy

result of sexual selection for larger and more competitive males when resources
are sufficient (and hence species are large-bodied). Smaller males may have an
advantage when resources are few (and hence species are small-bodied). Female
body size showed less variation overall than male, suggesting that there was
stabilizing selection on size in this sex.

Sperm Competition

Parker (1970) was the first to note that competition between ejaculates of different
males was likely to exert strong selection on male morphology and behaviour. Being
‘first in’ is very important for male spider mites because tetranychids exhibit
first-male sperm precedence. That is, the first male to copulate with a female
fertilises the majority of her eggs (Cone 1985), particularly if he is allowed to mate
with her without interruption (Potter and Wrensch 1978). The structure of the female’s
internal reproductive system and the path of spermatozoa after insemination are
thought to play important roles in determining the patterns of fertilisation priority in
arthropods (Kaster and Jakob 1997). In spider mites the spermatozoa appear to
migrate from the seminal receptacle, through the haemocoel of the female, into the
ovary (Feiertag-Koppen and Pijnacker 1985). So sperm from the first copulation get
a head start on those from subsequent matings.
The occurrence of pre-copulatory mate-guarding in the mesostigmatan
Macrocheles muscaedomesticae (Macrochelidae) would seem to predict a similar
phenomenon. In this species, the male rides about on the back of the deutonymphal
female and mates with the adult female immediately after her emergence. Males of
this species are heavily armed with spines on their second and fourth pair of legs,
which they use to attack rivals (Yasui 1988), and also to hold the female during
sperm transfer. Yasui (1988) used double matings of electrophoretically different
strains of M. muscaedomesticae to determine whether there was first-male precedence.
The evidence was overwhelming: only one of 410 offspring were produced with sperm
from a second mating. If the path of sperm inside the female in the Macrochelidae
involves travelling through the haemocoel, as it does in the Tetranychidae, then one
would predict that other dermanyssines also show first-male precedence.
In striking contrast, double-mating experiments with irradiated male deer ticks
(Ixodes dammini) revealed that the second male fertilised the most eggs (Yuval and
Spielman 1990). Yuval and Spielman (1990) suggest that the prolonged probing of
the female’s genital opening by the male may allow him to remove the ejaculate of
the previous mate, so that when he introduces his own spermatophore there is a
reduced number of competing sperm. This is an intriguing suggestion and a
plausible one considering that female ixodids store sperm in a seminal receptacle
(Sonenshine 1991). Argasid females store sperm in their uteri, possibly out of reach
of the male’s mouthparts; therefore, one would predict that last-male sperm
precedence is unlikely to occur in this group of ticks.
Sperm competition in some Acaridae (Astigmata) also involves last-male prece-
dence. Radwan (1991) found in Sancassania berlesei that the second male to mate
Sexual Selection 137

fertilised on average 74–86 % of a female’s eggs, and that last-male precedence was
maintained even when a female mated with three males. This contrasts with the
findings of Zeh and Zeh (1994) for pseudoscorpions, in which last-male advantage
breaks down if there are more than two matings. The degree of precedence decreased
with the interval between first and second matings; after 6 h, there was no signifi-
cant difference between fertilisations from first and second males. Radwan and
Witalinski (1991) examined the spermathecae of female S. berlesei after completing
mating with a single male and after interrupted mating with several males. TEM
sections indicated that when matings were interrupted, males had not transferred
sperm cells; rather, they had pumped the female’s spermatheca full of an amorphous
fluid. The authors conclude that S. berlesei males first transfer a substance that pushes
previously deposited sperm away from the openings of the efferent ducts leading to
the ovaries and then deposit their own spermatozoa in this region. Thus the last male
to mate positions his sperm where the female will use it as soon as possible.
In another acarid, Rhizoglyphus robini, there is no consistent last-male advantage;
rather, the paternity of second males ranges from 0 to 93 % (Radwan and Siva-Jothy
1996). Radwan (1996) found that the strongest predictor of male fertilisation success
was the size of his spermatozoa. Males that produced larger sperm cells were more
likely to fertilise eggs, a phenomenon independent of the male’s body size. Radwan
suggests that larger spermatozoa may survive longer in the spermatheca, or perhaps
move faster through the efferent sperm ducts leading to the ovaries. A later study on
this species revealed that there was a heritable component to competitiveness of
spermatozoa within the female (Konior et al. 2005).
When Parker (1970) developed the idea of internal sperm competition as a
selective agent, he suggested that this gametic battle was simply a continuation of the
ancestral war engaged in by externally fertilising animals. In such free-spawners,
winners were (and are) likely to be determined primarily on the sperm output of the
male. The more tickets in the lottery, the more likely you are to win the prize –
fertilisation of an egg, in this case. For arthropods with dissociated sperm transfer,
external sperm competition of this sort still holds (reviewed in Proctor 1998). Males
that deposit more spermatophores will (ceteris paribus) be more likely to have their
ejaculates intercept a female. But unlike free-spawners, these arthropods have an
additional tactic in their arsenal: physical destruction of a rival’s spermatophores. In
most dissociated taxa, when a male contacts a previously deposited spermatophore
he will trample, crush or snap it off and typically deposits a new spermatophore near
or even on the crushed one (Schuster and Schuster 1970, 1977; Alberti 1974; Proctor
1992, 1998). This behaviour often results in ‘trees’ of dozens of spermatophores
piled one on top of the other. Although earlier authors interpreted these behaviours
as maintenance of a fresh sperm supply for females, it is clearly a type of male
competition. Arnold (1976) described the remarkably similar behaviour of sper-
matophore destruction by male salamanders as intrasexual competition.
How do male mites benefit from these behaviours? Most obviously, they reduce
the number of competing spermatophores. It is also possible that by depositing near
or on the remnants of other spermatophores they gain by ‘stealing’ the pheromonal
effort of previous males (Proctor 1998). Evidence for pheromone theft is that males
138 5 Sex and Celibacy

of a number of species begin depositing spermatophores when they encounter those


of other males but do not destroy the previously deposited ejaculates (Moss 1960;
Davids and Belier 1974; Ehrnsberger 1977). It may also be that many males deposit
at the same area because of some special characteristic of that site, for example,
appropriate substrate or humidity.

Intersexual Selection as an Agent of Morphological


and Behavioural Change

Although Darwin’s contemporaries were quite willing to accept the idea that
aggression between males could result in the evolution of elaborate masculine
characters, they were sceptical that female animals could select males on the basis
of ‘beauty’. It was not until the 1980s, almost a century after Darwin’s book on
sexual selection, that scientists began to investigate the role of female choice in the
evolution of male secondary sexual characters.

Female Choice

Early efforts exploring the role of female choice in sexual selection focussed on taxa
in which females select one male from a simultaneous array of potential mates.
Evidence rapidly accumulated that the most extravagantly decorated, loudest or
most vigorous of males tend to be chosen (Ryan and Keddy-Hector 1992). This has
been best documented for animals that court visually (e.g. birds, fish, salticid
spiders, many insects) or vocally (e.g. birds, frogs, many insects). There has been
little work on those that rely primarily on chemicals for mate attraction (e.g. many
mites and insects), in part because it is more difficult for humans to ascribe a
female’s decision to mate with a given male to the attractiveness of his ‘aroma’
(for examples of females selecting a male based on his pheromones, see Thornhill
1992; Rantala et al. 2002).
This type of choice may be described as an ‘active’ display of preference by the
female and is closest to Darwin’s idea, but the definition of female choice has been
stretched so that it encompasses a broad range of female characteristics. Females
may also exert passive choice, an unfortunately oxymoronic phrase, if some aspect
of their sensory system or morphology results in them being more likely to mate
with males with certain attributes (e.g. if females see red very well, then reddish
males may have higher mating success). Females may also exert cryptic choice if
they manipulate males’ ejaculates internally to favour one male over another, or
mate again if their first mate did not meet their standards. Males may be selected to
court females both externally, where their actions are visible to human observers,
and internally during copulation. Eberhard (1985, 1996) has accumulated a huge
body of evidence suggesting that cryptic female choice and copulatory courtship are
widespread, and that female choice is responsible for much of the elaboration and
Sexual Selection 139

species-specificity of male genitalia. Is there any evidence that female mites choose
their partners either before or after sperm transfer?
There are very few observations that support the idea of active female choice in
Acari. Edwards and Dimock (1991) found that large-bodied male Unionicola
formosa tend to have larger harems. Although this may be the result of females
choosing to reside in mussels that are controlled by large males, it is also possible
that large harems are more eagerly fought for by males so that only the strongest
and largest can maintain control. In a complex series of breeding and mating experi-
ments, Lesna and Sabelis (1999) demonstrated that the mating preferences of
females of the predatory laelapid Hypoaspis aculeifer depended both on the female’s
own genetic strain and on the nature of the prey present at the time of mating.
A female tended to mate with males whose genotype would best complement that
of the female’s in their mutual offspring, given the prey assemblage. However, this
mate choice is at the level of strain rather than of individual males.
Passive female choice is a much more likely phenomenon in the Acari. The
copepod-mimicking courtship of male Neumania water mites is a good example of
males taking advantage of sensory biases in females. Is female choice possible in
dissociated species in which females never meet their partners? Conspicuous
secondary sexual characters are typically absent in taxa in which males and females
do not contact each other during sperm transfer (e.g. water mites, Proctor 1991a),
implying that female choice does not act on male morphology in these species.
Eberhard (1985) and Proctor et al. (1995) found that morphology of spermato-
phores was also less elaborate in dissociated species than it was in species with
paired-indirect transfer, suggesting that female choice is not acting as strongly on
spermatophore morphology either. It seems most likely that female choice in such
species acts on the attractiveness of chemicals associated with spermatophores.
However, although female mites can differentiate among spermatophores based
on the age of the ejaculates (e.g. in oribatids, Woodring and Cook 1962), to our
knowledge there is no study of whether they differentiate among males based on
spermatophores. Work by Hedlund et al. (1990) showed that female Orchesella
cincta (Collembola) picked up spermatophores produced by males of their own
‘rearing’ group more often than those produced by males from a different group,
proving that mate choice in dissociated species is possible.
A particularly tenuous type of female choice is that associated with the ejaculate
output of males. Particularly for males with indirect transfer, in which female
encounter with spermatophores is not guaranteed, number of spermatophores
deposited may be the best predictor of male mating success. Spermatophore output
in pseudoscorpions and mites has been shown to be affected by male age and nutri-
tional status (Proctor 1998). Females, by passively encountering the spermatophores
of well-fed males more often than those of starving ones, may exert some selection
for males with better abilities to garner resources. Female mites may also exert
cryptic choice by remating if unsatisfied with their first mate. However, there is little
evidence for this. Proctor (2003) manipulated the spermatophore output and vigour
of male Arrenurus manubriator (Arrenuridae) to test whether females were more
likely to remate if their first mate was of ‘poor quality’. Although many females
140 5 Sex and Celibacy

mated with a second male when given the opportunity, there was no correlation with
the first male’s ‘quality’ and the likelihood of the female remating.
Mating in copulating species of mites is characterised in many cases by a
remarkable lack of female choice. In many taxa, males sequester females prior to
adulthood and mate with them immediately upon eclosion (see Mate-guarding and
aggression, above). In such species, females tend to mate only once or have long
refractory periods before remating. In the Tarsonemina, brothers may copulate with
sisters inside the body of their mother; here there seems even less potential for
choice (Kaliszewski and Wrensch 1993). In taxa that mate repeatedly, there is
often a predictable pattern of first- or second-male sperm precedence. Consistent
patterns of sperm precedence are problematic for advocates of cryptic female
choice (Schneider and Elgar 1998). Unless, for example, the second male is invariably
the higher quality one, consistent precedence is evidence that selection on males
to fertilise eggs is much stronger than it is on females to choose which male gets
the honour.
Eberhard (1985) argued that elaborate male genitalia were the result of female
choice. Complex, species-specific intromittent organs are common in many taxa of
mites that transfer sperm by copulation (e.g. water mites, spider mites, dermanyssine
mesostigmatans). However, it is difficult to determine whether such modifications
are the result of selection to titillate the female or to sabotage another male’s
ejaculate through internal manipulation of sperm (e.g. Radwan and Witalinski
1991) or through intersexual conflict (below). In summary, female choice of the
kind evidenced by birds, fish and many insects has yet to be convincingly shown
in the Acari.

Intersexual Conflict

Students of sexual selection seem particularly prone to fads. Over the past century,
paradigms for the interactions between males and females have changed from that
of cooperation for the good of the species, to males frantically trying to attract the
attention of females that exert their choice actively, passively and cryptically, to the
emerging belief that intersexual conflict is the prevailing force behind the evolution
of both male and female traits. Adherents of the ‘conflict’ school feel that there is a
battle between males and females for control over fertilisation. The courtship dances
and elaborate genitalia of males are interpreted not as offerings to please choosy
females but as armaments to coerce or deceive females into giving up their eggs.
Females in turn raise their thresholds of sensitivity to such male subterfuge, resulting
in cycles of deception, disbelief and stronger deception. This sexual arms race is
thought to be a driving force in the evolution of diversity (Parker and Partridge
1998; Arnqvist and Rowe 2005), just as the battles between hosts and parasites
and plants and pollinators are considered engines of diversification (Brooks and
McLennan 1991; Mitter et al. 1991; Pellmyr 1992).
Several predictions derive from the intersexual conflict hypothesis. The first is
that females are often induced into mating when it is not in their best interests. There
Sexual Selection 141

is good evidence for this in drosophilid flies. Female Drosophila melanogaster that
mate repeatedly die earlier than singly mating females (Fowler and Partridge 1989).
Rice (1996) showed that when female lineages are kept from coevolving with male
lineages, the lifespans of such females are even more reduced from mating than are
those of coevolved females. This implies that males and females are in a constant
battle in which toxic chemicals in male ejaculates usurp control over fertilisation
and oviposition. Another prediction of the conflict hypothesis is that morphological
and behavioural tactics of each sex will change over time, as modes of attack by males
are rendered ineffective by evolutionary changes in female defences. This pattern
may be recoverable through phylogenetic analysis (Holland and Rice 1998). One might
predict that male courtship would accumulate embellishments over evolutionary
time, while female coyness would increase. So, male genitalia may become
more complex in derived species and female sperm receptacles may evolve more
tortuous paths.
This appears to be a likely explanation for the secondary female genital openings
common to many taxa of copulating mites. The logical site of entry for sperm, and
the one used by all non-copulating mites, is through the primary genital opening but
in at least some Heterozerconidae, dermanyssine mesostigmatans, Astigmata and
many prostigmatans (e.g. Tetranychidae), females have a secondary opening used
solely to receive the male intromittent organ (see taxon-specific sections above).
Why have these openings been relocated?
A clue may lie in the mating behaviour of some nematodes, velvet worms, tardi-
grades, leeches and bedbugs. Species in these taxa transfer sperm by hypodermic
insemination (also known as haemocoelic or traumatic insemination). In this
form of copulation the male introduces sperm into the female’s body cavity directly
through her integument. This may be via spermatophores spiked with lytic enzymes
(e.g. in velvet worms) or needle-like intromittent organs (e.g. in bedbugs and some
of their relatives). Chapman (1971) described a scenario for the evolution of hypo-
dermic insemination in the Cimicoidea (Hemiptera), the group to which bed bugs
(Cimicidae) belong. In Alloeorhynchus flavipes, the male’s aedeagus enters the
primary genital opening of the female but pierces the wall of the female’s vagina
with a sharp apical spine so that sperm are injected into her haemocoel. They move
through the female’s body cavity, eventually congregating around the ovarioles
where they penetrate the membrane surrounding the eggs and fertilise them.
In Primicimex the sharp aedeagus penetrates an intersegmental membrane on the
female’s back to inject sperm. In this taxon, sperm end up being stored in a pair of
pouches near the oviducts. In Xylocoris galactinus, the female has a sclerotised
cuticular pouch (the ectospermalege) between the dorsal sclerites of abdominal seg-
ments 2 and 3 that receives the male’s aedeagus. Sperm is then conducted through
a cord of cells that leads from the ectospermalege into the haemocoel, and they
then accumulate in a seminal receptacle at the junction of the ovarioles. In
Anthocoris and Orius (Anthocoridae) insemination is no longer haemocoelic.
Rather, a copulatory tube opens between the ventral plates on abdominal segments
7 and 8 and are stored in a sperm pouch. From here a conducting cord leads to the
oviducts so that sperm is never free in the haemocoel.
142 5 Sex and Celibacy

The situation in Anthocoris and Orius is remarkably similar to that in dermanyssine


mites, where secondary genital openings situated on the female’s ventral plates or
on her legs receive sperm from the male’s chelicerae. Can one discern a pattern in
parasitiform mites parallel to that of the Cimicoidea? In the Ixodida, males introduce
their mouthparts into the primary genital opening of the female where they fiddle
around doing mysterious things, sometimes for hours (Feldman-Muhsam 1986).
It is only after the mouthparts are withdrawn that the spermatophore is applied to
the genital opening. In the Dermanyssiae, rhodacaroid males do likewise prior to
transferring sperm into the female’s secondary genital opening. Is there a stage in
primitive mesostigmatans, as yet undiscovered, in which males pierce the female’s
vaginal wall to provide their sperm with a shortcut to the ovaries? And if one were
to plot the movement of sperm-conducting tissues and sites of insemination in the
Mesostigmata, would one observe a pattern of males shifting penetration sites
and females attempting to regain control of the movement of sperm? Have male
spermatodactyls become sharper and more penetrating over time? It is surprising
that no one has performed such an analysis, but then, to do so would require a good
phylogeny, such as that which allowed Tatarnic and Cassis (2010) to trace the
coevolutionary battles between male and female genitalia in Coridromius plant
bugs. A similar study might involve investigating behavioural and morphological
patterns in the water mite family Arrenuridae, in which sexual dimorphism ranges
from almost zero (e.g. Wuria), where sperm transfer is presumably indirect, to
copulation without intromission (e.g. Arrenurus (Megaluracarus) spp.), to transfer
via complex intromittent structures (e.g. Arrenurus (Arrenurus) spp.) (see Proctor
and Wilkinson 2001). The huge diversity of transfer modes and male modifications
for sperm transfer displayed by the Acari is an unexploited treasure-trove for
students of sexual selection.

Parthenogenesis

Parthenogenesis literally means ‘virgin birth’, and refers to production of viable


offspring in the absence of sex. Although prokaryotes have thrived for perhaps
4,000 million years without sex, we expect organisms with chromosomes encased
in a nucleus to undergo meiosis and sexual recombination. With the exception of
domesticated poultry under stringent circumstances, no birds or mammals or birds do
without sex and few reptiles or amphibians are even facultatively parthenogenetic in
the wild (Booth et al. 2012). Mites, however, don’t feel obliged to follow these rules.

Why Have Sex?

Despite the temptation to answer flippantly, this is one of the most profound
questions in evolutionary biology (Bell 1982; Stearns 1987). Sex – the merging of
genomes from two individuals to produce offspring – is the overwhelming rule
Parthenogenesis 143

in multicellular animals. Norton and Palmer (1991) note that obligate parthenogens
make up less than 1 % of the animal kingdom. Yet there is a basic inefficiency
to producing both male and female offspring. If one placed two females in com-
petition, one that produced only female offspring and one that produced a 50:50
ratio of males and females, the all-female lineage would rapidly outproduce the
bisexual one. Also, from the founding female’s point of view, she would transmit
a much higher proportion of her genes if she and her daughters reproduced
clonally rather than having to bring in male genomes every generation. This waste
of resources on unproductive males has been termed the ‘twofold cost of sex’
(Maynard Smith 1978).
So why bother with sexuality? Many hypotheses have been advanced (Stearns 1987),
although empirical evidence for most is rather scanty. At the species-selection level,
sexuality may reduce the probability of extinction by generating greater genetic
diversity (through recombination), thereby allowing more rapid adaptation to
changing environments. Its parallel at the individual level is bet-hedging: in variable
environments, one should not put all of one’s eggs in a single genetic basket.
Changing environments may be biotic as well as abiotic. Parasites and disease-
causing agents typically have an advantage in that their more rapid life cycles
allow them to adapt quickly to the host. If hosts are invariant in their genetic makeup,
this allows diseases to rapidly adapt and sweep through populations. It is clear
in this case that recombination allows host organisms to stay in the arms race.
The opposite side of the coin is that sex helps to prevent undesirable variation via
DNA-repair, where chromosomal damage from one parent can be repaired by
undamaged information in the homologous chromosome from the other parent.
The DNA-repair hypothesis applies to all eukaryotes regardless of the vacillations
of their environments.
Some combination of these selective agents, acting both at and above the level of
the individual, may explain the overwhelming dominance of sexual reproduction in
animals. Most vertebrate parthenogens seem to be recently derived from sexual
species, suggesting that loss of sex shunts a lineage into an evolutionary dead end
(e.g. Bell 1982). However, there are a few niggling taxa that seem to flout these
rules. The most celebrated of these are the bdelloid rotifers, a group of several
hundred species in which males have never been discovered. Bdelloids are often
represented as ‘the chief counterexample’ to the evidence that asexual lineages are
short-lived (Stearns 1987). But they are not alone. Based on fossilized evidence of
brood pouches, the ostracod family Darwinulidae appears to have been functionally
asexual for at least 200 MY (Martens et al. 2003) with males in living taxa being
both exceedingly rare and sterile (Birky 2010). Several families of oribatid mites,
comprising hundreds of species, are obligatorily parthenogenetic, as are numerous
taxa in the Prostigmata and Mesostigmata (Norton et al. 1993).
Despite the wealth of examples in the Acari, they have been virtually ignored by
most evolutionary biologists interested in the problem of all-female parthenogenesis.
But in contrast, mites have played a starring role in the development of sex-ratio theory
(Hamilton 1967). Below we outline some of the outstanding features of asexual
reproduction in mites in the hope that someone will take the ball and run with it.
144 5 Sex and Celibacy

Distribution of Parthenogenesis in Mites

Norton et al. (1993) catalogue the distribution of genetic modes of reproduction in


the Acari. These can be categorised as: diplodiploid, in which both sons and daugh-
ters are produced from fertilised eggs and have one set of chromosomes from each
parent; arrhenotoky, in which sons are produced from unfertilised eggs and are
haploid while daughters are produced from fertilised eggs and are diploid; and
thelytoky, in which daughters are produced asexually and retain diploidy. Examples
of amphitoky (or deuteroky) in mites, in which both sons and daughters are
produced asexually, appear to be mistaken instances of arrhenotoky (Norton et al.
1993). Parahaploidy, also known as pseudoarrhenotoky, involves the production
of haploid sons from fertilised eggs that lose the paternal genome early in develop-
ment. One more recently discovered, and very rare, mode is that of all-female haploid
parthenogenesis, which to date appears to occur only in one animal species, the
tenuipalpid mite Brevipalpus phoenicis (Weeks et al. 2001).
Nothing is known of the genetic systems of the Opilioacariformes or Holothyrida
but they appear most likely to be diplodiploid. In support of this, Allothyrus
species tend to have about equal sex ratios in the field and males and females are
similar in size. All ticks are diplodiploid with the exception of the thelytokous
Amblyomma rotundatum and Haemaphysalis longicornis, the latter of which is triploid.
The Mesostigmata contain diplodiploids, arrhenotokes, thelytokes and parahaploids
(Walter and Proctor 2007). Of the last, the Phytoseiidae are the best-studied and,
except for a few thelytokous species, most appear to be parahaploid.
Among the Acariformes, the early-derivative families collectively known as the
Endeostigmata are about evenly divided between thelytokous species and bisexual
ones, although it is not clear whether males of the latter taxa are produced sexually
or via arrhenotoky. Among the Prostigmata, all Parasitengona appear diplodiploid,
although there is one suspected case of thelytoky in a species of Balaustium
(Erythraeidae). In other acariform taxa, arrhenotoky is widespread and thelytoky
crops up in numerous lineages. Many species are known to be haplodiploid but the
relevant genetic mechanism (arrhenotoky or parahaploidy) has not been determined.
As mentioned above, oribatids include a large number of thelytokous species (about
10 % of the ca 7,000 described spp.) (Fig. 5.14). Most of the remaining species
appear to be diplodiploid, with a smattering of haplodiploids of uncertain mecha-
nism. In the Astigmata most species are either diplodiploid or arrhenotokous, with
very few examples of thelytoky.
One of the most remarkable things about Norton et al.’s (1993) survey of genetic
mechanisms in mites is how evolutionarily flexible they appear. Variation in repro-
ductive mode occurs within families, genera and even within species. For example,
the cyclamen mite Phytonemus pallidus (Tarsonemidae) appears to have thelytokous
and sexual populations. Occasionally, non-functional males appear in otherwise
thelytokous populations. These males often produce few, spermless spermatophores
or exhibit hermaphroditic structures. This suggests that the genetic machinery for
constructing males is still present in the genomes of many overtly asexual mites
Parthenogenesis 145

Fig. 5.14 Oppiella nova is


an obligately thelytokous
oribatid mite (i.e. an
all-female parthenogenetic
‘species’). Cosmopolitan in
distribution, less than a
quarter of a millimetre in
length and, contrary to
theory, abundant in both
disturbed and virgin habitats
(SEM by DE Walter)

(see also Weeks et al. 2001). Another point of interest is that asexuality, particularly
in the Oribatida, is clearly not an evolutionary dead end. Thelytokous lineages have
radiated to produce hundreds of species, much as they have in the bdelloid rotifers
and darwinulid ostracods. The prediction from ‘normal’ genetics would be that
asexual lineages would soon exhaust variation and become dangerously homozygotic.
Palmer and Norton (1992) did indeed find that, relative to sexual oribatids, thelytokous
species tend to show fewer multilocus genotypes. However, mean heterozygosity
was still high. Wrensch et al. (1994) suggest that the peculiar nature of mite
chromosomes – they are holokinetic and undergo inverted meiosis – may allow
retention of diversity and even genetic repair. The model organism Caenorhabditis
elegans also has holokinetic (= holocentric) chromosomes, and geneticists have
made good progress towards understanding meiosis in this species (Durnberg 2001),
perhaps we will soon be able to test Wrensch et al.’s hypothesis.
Given the wide occurrence and repeated evolution of thelytoky, do patterns in the
Acari support any of the hypotheses about the value of sex? Norton and Palmer
(1991) surveyed the distributions of thelytokous species of oribatids. They found
that thelytokous taxa were common in recently disturbed soil habitats, including
tilled and frequently flooded land. They were also more common in deep soil than
in the surface litter layer. The authors note that under these situations, the biotic
diversity of the habitat is rather low, and so presumably is the need to ‘keep up with
the Joneses’ via sexual reproduction. This supports the idea that sex is useful in
biotically complex habitats. Other evidence in favour of this idea is that thelytokous
endostigmatans are also very common in deep soil. However, many of these
thelytokes are small and the average size of soil mites tends to decrease with depth
146 5 Sex and Celibacy

in the soil (Walter, unpublished). Thus, it may be that that mate-finding is difficult
in deep soil, thereby selecting against sexual reproduction (Norton et al. 1993).
Cianciolo and Norton (2006) explicitly tested whether diversity of competitors and
predators correlated positively with sexuality in oribatids using data from literature
and field surveys, and found no evidence for the expected relationship.
In dermanyssine Mesostigmata, thelytoky is common in soil-litter inhabiting
species but of sporadic occurrence taxonomically (Walter and Lindquist 1995).
Most thelytokes appear to have close bisexual relatives and only a few taxa (e.g. the
genus Protogamasellus, the Asca aphidioides species group) appear to have radiated
as asexual lineages. Again, small thelytokous species appear to be relatively more
common in deep soil. Unlike in some studies of oribatid mites, however (e.g. Norton
and Palmer 1991), the distribution of all-female parthenogenesis in soil Mesostigmata
appears to be independent of disturbance and the degree of biotic complexity of the
habitat: both disturbed habitats (e.g. agricultural fields) and pristine forests have
many thelytokous species (Walter and Lindquist 1995).
Thelytoky is also very common among oribatid mites that inhabit stable habitats.
For example, freshwater and bog-dwelling oribatids are often thelytokes. For these
mites Norton and Palmer (1991) suggest that osmotic damage to spermatophores
may select for reproduction without males. This is plausible, although water mites
appear to have adapted very well to indirect sperm transfer under water. Norton
et al. (1993) note that thelytoky is remarkably rare among mites that parasitise ani-
mals (only two ticks and one suspected parasitengone), while it is quite common in
other groups of animal parasites. They offer no explanations for this relationship.
Another explanation for the evolution of parthenogenesis has nothing to with the
benefits of asexuality to the individual, but with the benefits to symbionts of that
individual. In insects and some other arthropods, Wolbachia and other bacterial
endosymbionts are known to cause feminization of males, thelytoky and female-biased
sex ratios (O’Neill et al. 1997 ). Wolbachia has been found in spider mites
(Tetranychidae) where in a few, but nowhere near all, populations, they are associated
with sex-ratio changes (Gotoh et al. 2003). By treating mites with antibiotics, Weeks
et al. (2001) determined that haploid-female parthenogenesis in Brevipalpis phoenicis
was likely caused by a bacterial endosymbiont. The symbionts somehow converted
unfertilized eggs, which would normally grow up to be males in other members of
the Tetranychoidea, to become haploid females. The bacterium has since been
named Cardinium and has been found in a wide diversity of arthropods, includ-
ing insects, spiders, harvestmen, and mites, where it sometimes co-occurs with
Wolbachia (Ros et al. 2012 and references therein).

Sex-Ratio Manipulation

It has long been noted that for sexually reproducing animals, the most common ratio
of male to female offspring is 1:1. The selective forces behind this are quite simple
(Charnov 1982). Suppose that in a population, male offspring are rarer than females.
Sex-Ratio Manipulation 147

Because sex is required for the production of offspring, each male mates with more
than one female, thereby achieving a greater individual reproductive success than
the females. Mothers that tend to produce male offspring, therefore, have more
grand-offspring than those without such tendencies. In subsequent generations,
male offspring become more common and the advantage to producing excess males
weakens, eventually resulting in an equal production of males and females. While
this ratio is the norm in vertebrates, there are many examples of invertebrates in
which the sex ratio is very female biased.
Hamilton (1967) noted that this expected 1:1 sex ratio was dependent on there
being a large, panmictic population. He observed that biased ratios were often found
among arthropods living in reproductively patchy habitats and where brother-sister
matings were common. How might these characteristics select for female-biased
sex ratios? Consider the following population structure. Habitat consists of ‘islands’
that are each colonised by a single fertilised female. The female produces offspring
that mate on the islands and then the fertilised daughters disperse to colonise new
patches. If a mutation arises such that a female produces fewer sons and more
daughters, then that female will produce a larger number of colonists than females
with 1:1 sex ratios in their offspring. This assumes, of course, that the smaller
number of sons are able to fertilise all daughters. This advantage of female-biased
progeny is reduced as the number of founding females per patch increases, for sons
become more valuable when unrelated daughters are available to be fertilised.
The predicted offspring sex ratio when there is such local mate competition (LMC)
is simply: proportion males = (number of foundresses – 1)/(number of foundresses × 2)
(Charnov 1982). When there is a single foundress, proportion males = 0, which is
interpreted to mean that a female should produce only enough sons to be certain that
her daughters are fully fertilised. When there are two foundresses, sex ratio is
predicted to be 25 % males and 75 % females. When three, then it is 33 % males and
66 % females, and so on until a 1:1 ratio is approached.
Bulmer and Taylor (1980) and Wilson and Colwell (1981) considered what
would happen if multiple generations were spent in the original patch, and then a
dispersal event took place in which mating occurred randomly between offspring
from different patches. Their models predict that the more generations spent in the
original patch, the more female-biased the sex ratio should be, because high-bias
patches produce many more colonists than low-bias ones. This is a process of group
selection, because it is only through the structured nature of the population that
female-biased sex ratios are selected (see Sober and Wilson 1998).
Hamilton (1967) noted that there was ‘extreme economy in the production of
males’ among parasitoid wasps, some bark beetles and thrips, and numerous taxa of
mites in which sex ratios were as low as 2–4 % male. In these species, mating tends
to occur between siblings, either in a host insect, the nest of the parent or, for
Tarsonemina, inside or near the body of the dead mother. The rather gruesome
incestuous mating of the latter has been popularised by Gould (1977). Clearly, these
examples fit the idea of localised mating prior to dispersal. Hamilton also observed
that many of these species exhibited arrhenotoky and felt that this mode of repro-
duction readily allowed the female to control the sex ratio of her offspring.
148 5 Sex and Celibacy

There is a huge body of literature dealing with sex ratios in mites, far too much
for us to cover adequately. We therefore recommend interested readers to peruse the
volume edited by Wrensch and Ebbert (1993), in which patterns and causes of
biased sex ratio in mites and insects are thoroughly reviewed. Here we simply skim
the surface of this sea of information and highlight evidence that it hard to explain
using conventional sex-ratio theory. Wrensch (1993) points out that sex ratio is not
fixed within a species, or within an individual for that matter. She notes that female
mites appear to control the fraction of eggs that become females on a daily basis,
rather than over their lifetimes. So the ‘ideal’ sex ratio may change, depending on
the stage of colonization, the quality of food and the overall density of individuals
in a patch (these factors are usually intertwined).
Kaliszewski and Wrensch (1993) surveyed sex-ratio evolution within the
Tarsonemina. Four families exhibit physogastric reproduction (see Chap. 4), and the
degree and location of incest varies among them. Males of Pyemotes tritici
(Pyemotidae) emerge from their mother’s body first, help their sisters to be ‘born’
and then mate with them. In Luciaphorus hauseri (Pygmephoridae), mating takes
place within the body of the mother. However, males will also mate with non-sibling
females that have ‘escaped’ from their ruptured parent before their brothers got
to them. Within the Tarsonemina, sex ratio ranges from 3:1 to 98:1 in favour of
daughters. Species with temporally patchy food supplies and those that are parasitoids
of insects tend to produce very female-biased ratios. However, true parasites,
such as the tracheal mite of honeybees, tend to produce only slightly biased broods
(2.4–4.4 females:1 male). Phytophagous species also tend to produce less biased
sex ratios (although thelytoky is common in some groups of plant-parasites).
Kethley (1971) discusses the ecology of the quill-dwelling mite, Syringophiloides
minor (Syringophilidae). Fertilised females invade the open umbilicus of developing
quills of the house sparrow, Passer domesticus. As the feather matures, the umbilicus
closes. Typically, only one female colonises each feather but occasionally as many
as ten enter a quill. The female lays 11–13 eggs, the first to develop always being a
male and the remainder being female. The foundress dies as the first generation
matures. The offspring mate among themselves and these females lay 10–12 eggs,
with only one male per mother. When the second generation matures the calmus is
packed with mites (usually 121 females and 10 males). Thus, one female ‘produces’
enough mites in two generations to fill the feathers. There is little support for the
LMC prediction of more males in multiply founded feathers, as Kethley found that
“multiple founders each perform similarly to isolated females”.
Hamilton (1967) made special note of the correlation between arrhenotoky
and extreme sex ratios because it provided a mechanism whereby females could
control the sex of her offspring. Although this generally holds true for mites
(e.g. Tetranychidae, Tarsonemina), it is not true for all acarine taxa. In the Phytoseiidae,
which almost invariably have female-biased ratios (sometimes up to 90 % females),
both sons and daughters come from fertilised eggs. However, the paternal genome
is destroyed early in embryogenesis so that male phytoseiids are effectively haploid.
Are sex ratios in the Phytoseiidae fixed or can the female manipulate them? Sabelis
and Nagelkerke (1993) suggest that they can to some extent, as females at the start
Immaculate Conception: Did Sexual Astigmatans Arise from Asexual Oribatids? 149

of their reproductive life typically produce a son first. They suggest that this may
be insurance so that if no other males are present, the female’s daughters may be
fertilised by her son. As well, the daily sex ratio produced by female Galendromus
occidentalis was very precise, as if controlled by the mother rather than being a
probabilistic ratio. Exactly how females might select which eggs are ‘cleansed’ of
their father’s genome is not clear. Even in truly haplodiploid spider mites, there
are additional complications as eggs differ in the probability that they will
receive sperm; large eggs are more likely to be fertilized than are small ones, and
females in situations of high local mate competition produce more large eggs
(Macke et al. 2012).
Finally, there are some puzzling examples of biased sex ratios in mites that lack
a clear mechanism by which to bias the sexes of their offspring. Water mites
(Hydracarina) often exhibit sex ratios in the field that are strongly biased either to
males or to females; however, they are all apparently diplodiploid. Some of this can
be explained by the sexes having different lifespans or phenologies, so that for
example, a sample in spring may show female bias while one in late summer may
be male biased. However, rearing experiments with Neumania and Unionicola
(Unionicolidae) suggest there is a subtle but significant 2:1 bias towards females,
while different female Arrenurus manubriator (Arrenuridae) produce families that
are either very female biased, male biased or unbiased (Proctor 1996). Arrenurus
chromosomes appear to behave oddly during spermatogenesis (Sokolov 1954).
Proctor (1996) suggested that sex-determination in Arrenurus (and perhaps in
the unionicolids) may be similar to that in sciarid flies, in which a parahaploid
mechanism can result in the production of all-male, all-female and mixed-sex
broods (Metz 1938).

Immaculate Conception: Did Sexual Astigmatans


Arise from Asexual Oribatids?

The belief that asexuality is a 'dead end' is so ingrained in the general evolutionary
literature, that it is hard to believe that any asexual lineages contribute much to
biological diversity. So the concept of an asexual group giving rise to a species-rich
sexual lineage seems particularly unlikely. In various papers (Norton and Palmer
1991; Palmer and Norton 1992; Norton et al. 1993; Norton 1998), however, Norton
has proposed that a sexual lineage, the Astigmata, has arisen from an asexual group
of oribatids (Malaconothroidea).
This hypothesis has two major implications. The first is phylogenetic: the
Oribatida as typically defined excludes the Astigmata and is, therefore, paraphyletic.
This is now well supported by molecular phylogenetic evidence (e.g., Dabert et al.
2010). The morphological similarities between astigmatans and juvenile oribatids
overwhelmingly support the idea of the former arising from the latter. However, the
fact that the oribatids that seem morphologically similar to early-derivative astigmatans
belong to an entirely thelytokous group has much greater implications.
150 5 Sex and Celibacy

How can sex ‘re-evolve’ from an asexual taxon with a long evolutionary history?
One possible mechanism is cyclical parthenogenesis. Although this alternation of
asexual and sexual generations is common in aphids, it has never been demonstrated
in mites. Perhaps the answer lies in the occasional reproductive ‘accident’.
Spanandry refers to the appearance of a very few males in an otherwise extremely
female-biased population. Spanandric males pop up among the offspring of many
otherwise thelytokous taxa (Walter and Kaplan 1990 c, Norton et al. 1993; Birky
2010) and they do not appear to function as sexually active males. The oribatid
ancestor of the Astigmata may have produced a rare functional spanandric male that
somehow encouraged a female to take up his sperm. One difficulty with this
scenario is that all astigmatans transfer sperm via an aedeagus, while there is only
one documented case of semi-direct transfer amongst oribatids. This species is from
the Galumnidae, a member of the derived group Brachypylina. In the phylogenetic
hypothesis diagrammed in Norton et al. (1993), the Astigmata and Brachypylina are
shown issuing from a common ancestral group, the Desmonomata. No pairing species
are known from this taxon. However, another ancestral group, the Mixonomata,
does contain a species observed to engage in a sort of pairing dance, Collohmannia
gigantea. In their broad phylogenetic analysis of the Acariformes, Dabert et al. (2010)
support the argument of Astigmata arising from within the Desmonomata; however,
they are resolved as having the sexual family Hermanniidae as a sister taxon rather
than arising within the asexual Malaconothroidea.

Summary

As they are in all things, mites are extremely diverse in their sexual habits. They
display a mind-boggling variety of sperm-transfer behaviours. This ranges from
complete dissociation between male and female, to pairing and substrate-deposited
spermatophores, to direct transfer of sperm via an intromittent organ. Sperm-transfer
behaviour is unknown in the Opilioacariformes and Holothyrida but in the remaining
Parasitiformes it is uniformly via copulation. The Acariformes are more imaginative
and display the entire range of possible modes. Copulation has evolved dozens of times
and water mites appear particular prone to evolving direct sperm transfer. In both
copulating and non-copulating taxa, spermatophores are seldom passive pedestals
for sperm but rather they take an active part in attracting and inseminating females.
In particular, the spermatophores of ticks, with their explosive sperm-delivery
mechanism and helpful symbionts, are astounding in their complex structure.
There are many examples of intrasexually selected characteristics in male mites.
Males compete with each other for access to females through fighting (e.g. harem-
keeping water mites), sequestering females before they have emerged as adults
(e.g. Tarsonemidae) or through trampling of other males’ spermatophores (many
dissociated parasitengones). They may also do battle internally through sperm
competition which, although not as common as in insects, appears to exist in some
ticks and astigmatans. In contrast, overt female choice seems rare or nonexistent in
References 151

the Acari. In copulating species, females are often impregnated after being
‘kidnapped’ (e.g. in Tarsonemidae) or while still in the bodies of their mothers
(e.g. in some Pygmephoridae). Among dissociated species, it is not clear whether
females can differentiate between spermatophores of different males, a prerequisite
for active mate choice. ‘Passive’ choice, which selects for those males best adapted
to the sensory biases and morphology of females, seems a more likely scenario.
The idea of passive mate choice merges with that of intersexual conflict as a driving
force behind the evolution of male and female sexual adaptations. Males may benefit
by inseminating females that would rather not be mated, which should lead to an
arms race between the sexes for control over fertilisation. Intersexual conflict should
be strongest in copulating species, where males have the possibility of forcing
unwilling females to accept sperm. Secondary genital openings in female derman-
yssine and astigmatic mites may have evolved through intersexual conflict over the
site of penetration, as it appears to have done in the hemipteran group Cimicoidea.
Mites also excel in the opposite of sex. Asexuality has arisen many times in the
Acari, both as arrhenotoky (production of males from unfertilised eggs) and thelytoky
(asexual production of females). No clear patterns exist to explain why some taxa
reproduce entirely parthenogenetically while related groups are sexual; however,
there may be a trend towards greater occurrence of thelytoky in abiotically constant
habitats with low biotic diversity (e.g. deep soil). Sex-ratio biases in mites often follow
the predictions of local mate competition theory but include numerous counter-
examples as well. The huge number of thelytokous oribatid species presents a
particular problem for evolutionary biology. Not only have they speciated and
diversified – thought almost impossible in asexual lineages of eukaryotes – but they
may have produced a very successful group of entirely sexual species, the Astigmata.

References

Alberti, G. (1974). Fortpflanzungsverhalten und Fortpflanzungsorgane der Schnabelmilben (Acarina:


Bdellidae, Trombidiformes). Zeitschrift für Morphologie und Ökologie der Tiere, 78, 111–157.
Alberti, G. (1995). Comparative spermatology of the Chelicerata: Review and perspective.
Mémoires du Muséum National d’Histoire Naturelle, 166, 203–230.
Alberti, G. (2002a). Reproductive systems of gamasid mites (Acari, Anactinotrichida) reconsidered.
In F. Bernini, R. Nannelli, G. Nuzzaci, & E. de Lillo (Eds.), Acarid phylogeny and evolution
adaptations in mites and ticks (pp. 125–139). Dordrecht: Kluwer.
Alberti, G. (2002b). Ultrastructural investigations of sperm and genital systems in Gamasida
(Acari: Anactinotrichida) current state and perspectives for future research. Acarologia, 42,
107–126.
Alberti, G., & Crooker, A. R. (1985). Internal anatomy. In W. Helle & M. W. Sabelis (Eds.), Spider
mites: Their biology, natural enemies and control (Vol. 1A, pp. 29–62). Amsterdam: Elsevier.
Alberti, G., Fernandez, N. A., & Kümmel, G. (1991). Spermatophores and spermatozoa of oribatid
mites (Acari: Oribatida). Part II: Functional and systematical considerations. Acarologia, 32,
435–449.
Alberti, G., Gegner, A., & Witalinski, W. (2000). Fine structure of the spermatophore and sperma-
tozoa in inseminated females of Pergamasus mites (Acari: Gamasida: Pergamasidae). Journal
of Morphology, 245, 1–18.
152 5 Sex and Celibacy

Alberti, G., Gerdeman, B., & Klompen, H. (2007). Fine structure of spermiogenesis and sperm in
a heterozerconid mite (Heterozerconina; Gamasida). In J. B. Morales-Malacara, V. Beham
Pelletier, E. Ueckermann, T. M. Pe´ rez, E. G. Estrada-Venegas, M. Badii (Eds.), Acarology XI:
Proceedings International Congress (pp. 557–560).
Alberti, C. G. Y., Fernandez, N. A., & Théron, P. D. (2010). Fine structure of the male genital syst
ems, spermatophores and unusual sperm cells of Saxidromidae (Acari, Actinotrichida).
Acarologia, 50, 243–256.
Alcock, J. (1994). Postinsemination associations between males and females in insects: The
mate-guarding hypothesis. Annual Review of Entomology, 39, 1–21.
Alexander, R. D. (1964). The evolution of mating behaviour in arthropods. Symposia of the Royal
Entomological Society of London, 2, 78–94.
Amano, H., & Chant, D. A. (1978). Mating behaviour and reproductive mechanisms of two species
of predaceous mites, Phytoseiulus persimilis Athias-Henriot and Amblyseius andersoni (Chant)
(Acarina: Phytoseiidae). Acarologia, 20, 196–213.
Arnold, S. J. (1976). Sexual behavior, sexual interference and sexual defense in the salamanders
Ambystoma maculatum, Ambystoma tigrinum and Plethodon jordani. Zeitschrift für
Tierpsychologie, 42, 247–300.
Arnqvist, G., & Rowe, L. (2005). Sexual conflict. Princeton: Princeton University Press.
Barazandeh, M., Davis, C. S., Neufeld, C. J., Coltman, D. W., Palmer, A. R. (2013). Something Darwin
didn’t know about barnacles: spermcast mating in a common stalked species. Proceedings
of the Royal Society B, 280, 20122919. http://dx.doi.org/10.1098/rspb.2012.2919.
Barr, D. (1972). The ejaculatory complex of water mites (Acari: Parasitengona): Morphology and
potential value for systematics. Life sciences contributions/Royal Ontario Museum, 81, 1–87.
Behan-Pelletier, V. M., & Eamer, B. (2005). Zachvatkinibates (Acari: Oribatida: Mycobatidae) of
North America, with descriptions of sexually dimorphic species. Canadian Entomologist, 137,
631–647.
Behan-Pelletier, V. M., & Eamer, B. (2010). The first sexually dimorphic species of Oribatella
(Acari, Oribatida, Oribatellidae) and a review of sexual dimorphism in the Brachypylina.
Zootaxa, 2332, 1–20.
Bell, G. (1982). The masterpiece of nature. Berkeley: University of California Press.
Birky, C. W., Jr. (2010). Positively negative evidence for asexuality. Journal of Heredity,
101(Supplement 1), S42–S45.
Boczek, J., & Griffiths, D. A. (1979). Spermatophore production and mating behaviour in the
stored product mites Acarus siro and Lardoglyphus konoi. In J. G. Rodriguez (Ed.), Recent
advances in acarology I (pp. 279–284). New York: Academic.
Booth, W., Smith, C. F., Eskridge, P. H., Hoss, S. K., Mendelson, J. R., & Schuett, G. W. (2012).
Facultative parthenogenesis discovered in wild vertebrates. Biology Letters. doi:10.1098/
rsbl.2012.0666.
Boucot, A. J. (1990). Evolutionary paleobiology of behavior and coevolution. Amsterdam:
Elsevier.
Braddy, S. J., & Dunlop, J. A. (1997). The functional morphology of mating in the Silurian
eurypterid, Baltoeurypterus tetragonophthalmus (Fischer, 1839). Zoological Journal of the
Linnean Society, 120, 435–461.
Bretfeld, G. (1970). Grundzüge des Paarungsverhalten europäischer Bourletiellini (Collembola,
Sminthuridae) und darous abgeleitete taxonomisch-nomenklatorische Folgerungen. Zeitschrift
fur zoologische Systematik und Evolutionsforschung, 8, 259–273.
Brooks, D. R., & McLennan, D. A. (1991). Phylogeny, ecology, and behavior. Chicago: University
of Chicago Press.
Buckland-Nicks, J., & Scheltema, A. (1995). Was internal fertilization an innovation of early
Bilateria? Evidence from sperm structure of a mollusc. Proceedings of the Royal Society of
London B, 261, 11–18.
Bulmer, M. G., & Taylor, P. D. (1980). Sex ratio under the haystack model. Journal of Theoretical
Biology Biology, 86, 83–89.
References 153

Chapman, R. F. (1971). The insects: Structure and function (2nd ed.). London: Edward Arnold Ltd.
Charnov, E. L. (1982). The theory of sex allocation. Princeton: Princeton University Press.
Cianciolo, J. M., & Norton, R. A. (2006). The ecological distribution of reproductive mode in
oribatid mites, as related to biological complexity. Experimental and Applied Acarology, 40,
1–25.
Coineau, Y. (1976). Les pariades sexuelles des Saxidrominae Coineau 1974 (Acariens Prostigmates,
Adamystidae). Acarologia, 28, 234–240.
Colwell, R. K. (2000). Rensch’s Rule crosses the line: convergent allometry of sexual size
dimorphism in hummingbirds and flower mites. The American Naturalist, 156, 495–510.
Compton, G. L., & Krantz, G. W. (1978). Mating behavior and related morphological specialization
in the uropodine mite, Caminella peraphora. Science, 200, 1300–1301.
Cone, W. W. (1985). Mating and chemical communication. In W. Helle & M. W. Sabelis (Eds.),
Spider mites: Their biology, natural enemies and control (Vol. 1A, pp. 243–251). Amsterdam:
Elsevier.
Dabert, M., Witalinski, W., Kazmierski, A., Olszanowski, Z., & Dabert, J. (2010). Molecular
phylogeny of acariform mites (Acari, Arachnida): Strong conflict between phylogenetic signal
and long-branch attraction artifacts. Molecular Phylogenetics and Evolution, 56, 222–241.
Darwin, C. (1871). The descent of man and selection in relation to sex. London: John Murray.
David, B., & Mooi, R. (1990). An echinoid that “gives birth”: Morphology and systematics of a
new Antarctic species, Urechinus mortenseni (Echinodermata, Holasteroida). Zoomorphology,
110, 75–79.
Davids, C., & Belier, R. (1974). The spermatophores of Hydrachna conjecta Koenike and the life
history of the land-living ancestors of this water mite. Proceedings of the 4th International
Congress Acarology (pp. 147–151). Budapest: Akademiai Kiado.
Davids, C., Holtslag, J., & Dimock, R. V., Jr. (1988). Competitive exclusion, harem behaviour
and host specificity of the water mite Unionicola ypsilophora (Hydrachnellae, Acari) inhabit-
ing Anodonta cygnea (Unionidae). Internationale Revue der Gesamten Hydrobiologie, 73,
651–657.
Di Palma, A., Alberti, G., Nuzzaci, G., & Krantz, G. W. (2006). Fine structure and functional mor-
phology of the mouthparts of a male Veigaia sp. (Gamasida: Veigaiidae) with remarks on the
spermatodactyl and related sensory structures. Journal of Morphology, 267, 208–220.
Di Palma, A., Gerdeman, B. S., & Alberti, G. (2008). Fine structure and functional morphology
of the spermatodactyl in males of Heterozerconidae (Gamasida). International Journal of
Acarology, 34, 359–366.
Di Palma, A., Wegener, A., & Alberti, G. (2009). On the ultrastructure and functional morphology
of the male chelicerae (gonopods) in Parasitidae and Dermanyssina mites (Acari: Gamasida).
Arthropod Structure & Development, 38, 329–338.
Di Palma, A., Alberti, G., Blazak, C., & Krantz, G. W. (2012). Morphological and functional
adaptations of the female reproductive system in Veigaiidae (Acari: Gamasida) and implications
regarding the systematic position of the family. Zoologischer Anzeiger, 251, 49–70.
Dimock, R. V., Jr. (1983). In defense of the harem: Intraspecific aggression by male water mites
(Acari: Unionicolidae). Annals of the Entomological Society of America, 76, 463–465.
Dubinin, V. B. (1951). Feather mites (Analgesoidea). Part I. Introduction to their study. Fauna
U.S.S.R., 6, 1–363.
Durnberg, A. (2001). Here, there, and everywhere: Kinetochore function on holocentric chromo-
somes. The Journal of Cell Biology, 6, F33–F38.
Eberhard, W. G. (1985). Sexual selection and animal genitalia . Cambridge: Harvard
University Press.
Eberhard, W. G. (1996). Female control: Sexual selection by cryptic female choice. Princeton:
Princeton University Press.
Edwards, D. D., & Dimock, R. V., Jr. (1991). Relative importance of size versus territorial
residency in intraspecific aggression by symbiotic male water mites (Acari: Unionicolidae).
Experimental & Applied Acarology, 12, 61–65.
154 5 Sex and Celibacy

Ehrnsberger, R. (1977). Fortpflanzungsverhalten der Rhagidiidae (Acarina: Trombidiformes).


Acarologia, 19, 67–73.
Emlen, S. T., & Oring, L. W. (1977). Ecology, sexual selection, and the evolution of mating
systems. Science, 197, 215–223.
Enders, M. M. (1993). The effect of male size and operational sex ratio on male mating success in
the common spider mite, Tetranychus urticae Koch (Acari: Tetranychidae). Animal Behaviour,
46, 835–846.
Estrada-Venegas, E., Norton, R.A., & Moldenke, A. R. (1996). Unusual sperm transfer in
Pilogalumna sp. (Galumnidae). In R. Mitchell, D. J. Horn, G. R. Needham, & W. C. Welbourn (Eds.),
Acarology IX: Volume 1, Proceedings (pp. 565–567). Columbus: Ohio Biological Survey.
Evans, G. O. (1992). Principles of acarology. Wallingford: CAB International.
Fashing, N. F. (2008). Mate-guarding in the genus Creutzeria (Astigmata: Histiostomatidae), an
aquatic mite genus inhabiting the fluid-filled pitchers of Nepenthes plants (Nepentheaceae).
Systematic & Applied Acarology, 13, 163–171.
Feiertag-Koppen, C. C. M., & Pijnacker, L. P. (1985). Oogenesis. In W. Helle & M. W. Sabelis
(Eds.), Spider mites: Their biology, natural enemies and control (Vol. 1A, pp. 117–127).
Amsterdam: Elsevier.
Feldman-Muhsam, B. (1986). Observations on the mating behaviour of ticks. In J. R. Sauer &
J. A. Hair (Eds.), Morphology, physiology, and behavioral biology of ticks (pp. 217–232).
Chichester: Ellis Horwood Ltd.
Feldman-Muhsam, B. (1991). The role of Adlerocystis sp. in the reproduction of argasid ticks.
In R. Schuster & P. W. Murphy (Eds.), The Acari: Reproduction, development and life-history
strategies (pp. 179–190). New York: Chapman & Hall.
Fowler, K., & Partridge, L. (1989). A cost of mating in female fruitflies. Nature, 338, 760–761.
Garga, N., Proctor, H., & Belczewski, R. (1997). Leg size affects mating success in Tarsonemus
confusus Ewing (Prostigmata: Tarsonemidae). Acarologia, 38, 369–375.
Gaud, J., & Atyeo, W. T. (1996). Feather mites of the world (Acarina, Astigmata): The supraspe-
cific taxa. Part I. Annalen Zoologische Wetenschappen, 277, 1–193.
Gerdeman, B. S., & Klompen, H. (2003). A new North American heterozerconid, Narceoheterozercon
ohioensis n. g., n. sp., with first description of immatures of Heterozerconidae (Acari:
Mesostigmata). International Journal of Acarology, 29, 351–370.
Gotoh, T., Noda, H., & Hong, X.-Y. (2003). Wolbachia distribution and cytoplasmic incompatibility
based on a survey of 42 spider mite species (Acari: Tetranychidae) in Japan. Heredity, 91,
208–216.
Gould, S. (1977). Ontogeny and phylogeny. Cambridge: Harvard University Press.
Gwiazdowicz, D. J. (2004). Record of heteromorphic males of Hypoaspis (Cosmolaelaps) vacua
(Michael, 1891) (Acari, Mesostigmata, Laelapidae) from Poland. Journal of the Acarological
Society of Japan, 13, 181–184.
Hamilton, W. D. (1967). Extraordinary sex ratios. Science, 156, 477–488.
Hedlund, K., Ek, H., Gunnarsson, T., & Svegborn, C. (1990). Mate choice and male competition
in Orchesella cincta (Collembola). Experientia, 46, 524–526.
Hinton, H. E. (1964). Sperm transfer in insects and the evolution of haemocoelic insemination.
Symposia of the Royal Entomological Society of London, 2, 95–107.
Holland, B., & Rice, W. R. (1998). Perspective: Chase-away sexual selection: Antagonistic seduction
versus resistance. Evolution, 52, 1–7.
Immler, S., Pitnick, S., Parker, G. A., Durrant, K. L., Lüpold, S., Calhim, S., & Birkhead, T. R. (2011).
Resolving variation in the reproductive tradeoff between sperm size and number. Proceedings
of the National Academy of Sciences, 108, 5325–5330.
Kaliszewski, M., & Wrensch, D. L. (1993). Evolution of sex determination and sex ratio within
the mite cohort Tarsonemina (Acari: Heterostigmata). In D. L. Wrensch & M. A. Ebbert
(Eds.), Evolution and diversity of sex ratio in insects and mites (pp. 192–213). New York:
Chapman & Hall.
Kamenz, C., Staude, A., & Dunlop, J. A. (2011). Sperm carriers in Silurian sea scorpions.
Naturwissenschaften, 98, 889–896.
References 155

Kano, Y. (2008). Vetigastropod phylogeny and a new concept of Seguenzioidea: Independent


evolution of copulatory organs in the deep-sea habitats. Zoologica Scripta, 37, 1–21.
Kaster, J. L., & Jakob, E. M. (1997). Last-male sperm priority in a haplogyne spider (Aranea:
Pholcidae): Correlations between female morphology and patterns of sperm usage. Annals of
the Entomological Society of America, 90, 254–259.
Kethley, J. (1971). Population regulation in quill mites (Acarina: Syringophilidae). Ecology, 52,
113–118.
Kirchner, W. P. (1967). Spermatophoren bei Halacariden. Naturwissenschaften, 54, 345–346.
Kiszewski, A. E., Matuschka, F.-R., & Spielman, A. (2001). Mating strategies and spermiogenesis
in ixodid ticks. Annual Review of Entomology, 46, 167–182.
Klimov, P. B., & Sidorchuk, E. A. (2011). An enigmatic lineage of mites from Baltic amber
shows a unique, possibly female-controlled, mating. Biological Journal of the Linnean Society,
102, 661–668.
Knop, N. F. (1985). Mating behaviour in the tydeid mite Homeopronematus anconi (Acari: Tydeidae).
Experimental and Applied Acarology, 1, 115–125.
Konior, M., Keller, L., & Radwan, J. (2005). Effect of inbreeding and heritability of sperm
competition success in the bulb mite Rhizoglyphus robini. Heredity, 94, 577–581.
Krantz, G. W. (1978). A manual of acarology. Corvallis: Oregon State University Bookstores.
Krantz, G. W., & Walter, D. E. (Eds.). (2009). A manual of acarology (3rd ed.). Lubbock: Texas
Tech University Press.
Krantz, G. W., & Wernz, J. G. (1979). Sperm transfer in Glyptholaspis americana. In J. G. Rodriguez
(Ed.), Recent advances in acarology (Vol. 2, pp. 441–446). New York: Academic.
Lee, D. C. (1974). Rhodacaridae (Acari: Mesostigmata) from near Adelaide, Australia III.
Behaviour and development. Acarologia, 16, 21–44.
Leimann, J. (1991). Structure and formation of the sperm package of Piona carnea (Koch, 1836)
(Prostigmata, Hydrachnidia), a copulating water mite. In F. Dusbábek & V. Bukva (Eds.),
Modern acarology (Vol. 2, pp. 449–454). The Hague: Academia, Prague and SPB Academic
Publishing.
Lesna, I., & Sabelis, M. W. (1999). Diet-dependent female choice for males with ‘good genes’ in
a soil predatory mite. Nature, 401, 581–584.
Levitan, D. R., Sewell, M. A., & Chia, F.-S. (1992). How distribution and abundance influence
fertilization success in the sea urchin Strongylocentrotus franciscanus. Ecology, 73, 248–254.
Macke, E., Magalhaes, S., Bach, F., & Olivieri, I. (2012). Sex-ratio adjustment in response to local
mate competition is achieved through an alteration of egg size in a haplodiploid spider mite.
Proceedings of the Royal Society B. doi:10.1098/rspb.2012.1598.
Mann, T. (1984). Spermatophores: development, structure, biochemical attributes, and role in the
transfer of spermatozoa. Berlin: Springer.
Martens, K., Rossetti, G., & Horne, D. J. (2003). How ancient are ancient asexuals? Proceedings
of the Royal Society of London B, 270, 723–729.
Maynard Smith, J. (1978). The evolution of sex. Cambridge: Cambridge University Press.
Metz, C. W. (1938). Chromosome behavior, inheritance and sex determination in Sciara. American
Naturalist, 72, 485–520.
Michalska, K. (2011). Daily production of spermatophores, sperm number and spermatophore size
in two eriophyoid mites. Experimental and Applied Acarology, 55, 349–359.
Michalska, K., & Boczek, J. (1991). Sexual behavior of males attracted to quiescent deutonymphs
in the Eriophyoidea (Acari). In F. Dusbábek & V. Bukva (Eds.), Modern acarology (Vol. 2,
pp. 549–553). The Hague: Academia, Prague and SPB Academic Publishing.
Mitter, C., Farrell, B., & Futuyma, D. J. (1991). Phylogenetic studies of insect–plant interactions:
Insights into the genesis of diversity. Trends in Ecology & Evolution, 6, 290–293.
Morrow, E. H. (2004). How the sperm lost its tail: The evolution of aflagellate sperm. Biological
Reviews, 79, 795–814.
Moss, W. W. (1960). Description and mating behaviour of Allothrombium lerouxi, new species
(Acarina: Trombidiidae), a predator of small arthropods in Quebec apple orchards. Canadian
Entomologist, 92, 848–905.
156 5 Sex and Celibacy

Mostafa, A. R. (1974). Biological and behavioral aspects of the lizard mite Pterygosoma mutabilis
Jack, 1961 (Acarina: Pterygosomidae). Acarologia, 16, 100–105.
Norton, R. A. (1998). Morphological evidence for the evolutionary origin of Astigmata (Acari:
Acariformes). Experimental & Applied Acarology, 22, 559–594.
Norton, R. A., & Alberti, G. (1997). Porose integumental organs of oribatid mites (Acari,
Oribatida). 3. Evolutionary and ecological aspects. Zoologica, 146, 115–143.
Norton, R. A., & Palmer, S. C. (1991). The distribution, mechanisms and evolutionary
significance of parthenogenesis in oribatid mites. In R. Schuster & P. W. Murphy (Eds.),
The Acari: Reproduction, development and life-history strategies (pp. 107–136). New York:
Chapman & Hall.
Norton, R. A., Kethley, J. B., Johnston, D. E., & OConnor, B. M. (1993). Phylogenetic perspectives
on genetic systems and reproductive modes of mites. In D. L. Wrensch & M. A. Ebbert (Eds.),
Evolution and diversity of sex ratio in insects and mites (pp. 8–99). New York: Chapman & Hall.
Nuzzaci, G., & Alberti, G. (1996). Internal anatomy and physiology. In E. E. Lindquist, M. W. Sabelis, &
J. Bruin (Eds.), Eriophyoid mites – their biology, natural enemies and control (pp. 101–167).
Amsterdam: Elsevier Science B.V.
O’Neill, S. L., Hoffmann, A. A., & Werren, J. H. (Eds.). (1997). Influential passengers inherited
microorganisms and arthropod reproduction. NY: Oxford University Press.
Oldfield, G. N., & Michalska, K. (1996). Spermatophore deposition, mating behavior and population
mating structure. In E. E. Lindquist, M. W. Sabelis, & J. Bruin (Eds.), Eriophyoid mites – their
biology, natural enemies and control (pp. 185–198). Amsterdam: Elsevier Science B.V.
Oliver, J. H. (1982). Tick reproduction: Sperm development and cytogenetics. In F. D. Obenchain
& R. Galun (Eds.), Physiology of ticks (pp. 245–275). Oxford: Pergamon Press.
Oppedisano, M., Eguaras, M., & Pernandez, N. (1995). Dépôt de spermatophores et structures de
signalisation chez Pergalumna sp. (Acari: Oribatida). Acarolgia, 36, 347–353.
Palmer, S. C., & Norton, R. A. (1992). Genetic diversity in thelytokous oribatid mites (Acari;
Acariformes: Desmonomata). Biochemical Systematicals and Ecology, 20, 219–231.
Parker, G. A. (1970). Sperm competition and its evolutionary consequences in the insects.
Biological Review, 45, 525–567.
Parker, G. A. (1978). Evolution of competitive mate searching. Annual Review of Entomology,
23, 173–196.
Parker, G. A. (1982). Why are there so many tiny sperm? Sperm competition and the maintenance
of two sexes. Journal of Theoretical Biology, 96, 281–294.
Parker, G. A. (1984). Sperm competition and the evolution of animal mating strategies. In
R. L. Smith (Ed.), Sperm competition and the evolution of animal mating systems (pp. 2–61).
New York: Academic.
Parker, G. A., & Partridge, L. (1998). Sexual conflict and speciation. Philosophical Transactions
of the Royal Society B, 353, 261–274.
Pellmyr, O. (1992). Evolution of insect pollination and angiosperm diversification. Trends in
Ecology & Evolution, 7, 46–49.
Potter, D. A., & Wrensch, D. L. (1978). Interrupted matings and the effectiveness of second
inseminations in the twospotted spider mite. Annals of the Entomological Society of America,
71, 882–885.
Potter, D. A., Wrensch, D. L., & Johnston, D. E. (1976). Guarding, aggressive behavior, and mating
succes in male twospotted spider mites. Annals of the Entomological Society of America,
60, 707–711.
Proctor, H. C. (1991a). The evolution of copulation in water mites: a comparative test for
nonreversing characters. Evolution, 45, 558–567.
Proctor, H. C. (1991b). Courtship in the water mite Neumania papillator: Males capitalize on
female adaptations for predation. Animal Behavior, 42, 589–598.
Proctor, H. C. (1992). Mating and spermatophore morphology of water mites (Acari: Parasitengona).
Zoological Journal of the Linnean Society, 106, 341–384.
Proctor, H. C. (1996). Sex-ratios and chromosomes in water mites (Hydracarina). In R. Mitchell,
D. J. Horn, G. R. Needham, & W. C. Welbourn (Eds.), Acarology IX: Volume 1, Proceedings
(pp. 441–445). Columbus: Ohio Biological Survey.
References 157

Proctor, H. C. (1997). Mating behaviour of Physolimnesia australis (Acari: Limnesiidae),


a non-parasitic, rotifer-eating water mite from Australia. Journal of Arachnology, 25, 321–325.
Proctor, H. C. (1998). Indirect sperm transfer in arthropods: Behavioral and evolutionary trends.
Annual Review of Entomology, 43, 153–174.
Proctor, H. C. (2003). Feather mites (Acari: Astigmata): Ecology, behavior and evolution. Annual
Review of Entomology, 48, 185–209.
Proctor, H. C., & Smith, B. P. (1994). Mating behaviour of the water mite Arrenurus manubriator
(Acari: Arrenuridae). Journal of Zoology Lond., 232, 473–483.
Proctor, H. C., & Wilkinson, K. (2001). Coercion and deceit: water mites (Acari: Hydracarina) and
the study of intersexual conflict. In R. B. Halliday, D. E. Walter, H. C. Proctor, R. A. Norton, &
M. J. Colloff (Eds.), Acarology: Proceedings of the 10th International Congress (pp. 155–169).
Melbourne: CSIRO Publishing.
Proctor, H. C., Baker, R. D., & Gwynne, D. T. (1995). Mating behaviour and spermatophore
morphology: A comparative test of the female choice hypothesis. Canadian Journal of Zoology,
73, 2010–2020.
Proctor, H. C., Williams G., Clayton, D. H. (2009). Population density and male polymorphism in
the feather mite Falculifer rostratus (Acari: Falculiferidae). In M. W. Sabelis & J. Bruin (Eds.),
Trends in acarology – Proceedings of the 12th International Congress (pp. 299–302). Springer.
Radwan, J. (1991). Sperm competition in the mite Caloglyphus berlesei. Behavioural Ecology and
Sociobiology, 29, 291–296.
Radwan, J. (1995). Male morph determination in two species of acarid mites. Heredity, 74, 669–673.
Radwan, J. (1996). Intraspecific variation in sperm competition success in the bulb mite: A role for
sperm size. Proceedings of the Royal Society of London B, 263, 855–859.
Radwan, J., & Siva-Jothy, M. T. (1996). The function of post-insemination mate association in the
bulb mite, Rhizoglyphus robini. Animal Behavior, 52, 651–657.
Radwan, J., & Witalinski, W. (1991). Sperm competition. Nature, 352, 671–672.
Rantala, M. J., Jokinen, I., Kortet, R., Vainikka, A., & Suhonen, J. (2002). Do pheromones reveal
male immunocompetence? Proceedings of the Royal Society of London B, 269, 1681–1685.
Rath, W., Delfinato-Baker, M., & Drescher, W. (1991). Observations on the mating behavior, sex
ratio, phoresy and dispersal of Tropilaelaps clarae (Acari: Laelapidae). International Journal
of Acarology, 17, 201–207.
Raven, P. H., Ebert, R. F., & Curtis, H. (1981). Biology of plants (3rd ed.). New York: Worth
Publishers, Inc.
Rechav, Y., Goldberg, M., & Fielden, L. J. (1997). Evidence for attachment pheromones in the
Cayenne tick (Acari: Ixodidae). Journal of Medical Entomology, 34, 234–237.
Resler, J. H., Frazier, J. L., Shepherd, J. G., & Modaffer, J. D. (2009). Migration and motility of
spermatozoa in the female reproductive tract of the soft tick Ornithodoros moubata (Acari,
Argasidae). Parasitology, 136, 511–521.
Rice, W. R. (1996). Sexually antagonistic male adaptation triggered by experimental arrest of
female evolution. Nature, 381, 232–234.
Ros, V. I. D., Fleming, V. M., Feil, E. J., & Breeuwer, J. A. J. (2012). Diversity and recombination
in Wolbachia and Cardinium from Bryobia spider mites. BMC Microbiology, 12(Suppl 1), S13.
Rouse, G., & Fitzhugh, K. (1994). Broadcasting fables: Is external fertilization really primitive?
Sex, size, and larvae in sabellid polychaetes. Zoologica Scripta, 23, 271–312.
Ryan, M. J., & Keddy-Hector, A. (1992). Directional patterns of female mate choice and the role
of sensory biases. American Naturalist, 139, S4–S35.
Sabelis, M. W., & Nagelkerke, K. (1993). Sex allocation and pseudoarrhenotoky in phytoseiid
mites. In D. L. Wrensch & M. A. Ebbert (Eds.), Evolution and diversity of sex ratio in insects
and mites (pp. 512–541). New York: Chapman & Hall.
Saito, Y. (1986a). Prey kills predator: Counter-attack success of a spider mite against its specific
phytoseiid predator. Experimental and Applied Acarology, 2, 47–62.
Saito, Y. (1986b). Biparental defence in a spider mite (Acari: Tetranychidae) infesting Sasa bamboo.
Behavioural Ecology and Sociobiology, 18, 377–386.
Saito, Y. (1990). ‘Harem’ and ‘non-harem’ type mating systems in two species of subsocial spider
mites (Acari, Tetranychidae). Researches on Population Ecology, 32, 263–278.
158 5 Sex and Celibacy

Schaller, F. (1971). Indirect sperm transfer by soil arthropods. Annual Review of Entomology,
16, 407–446.
Schmidt, G. D., & Roberts, L. S. (1985). Foundations of parasitology (3rd ed.). St. Louis: Times
Mirror/Mosby College Publishing.
Schneider, J. M., & Elgar, M. A. (1998). Spiders hedge genetic bets. Trends in Ecology & Evolution,
13, 218–219.
Schulten, G. G. M. (1985). Mating. In W. Helle & M. W. Sabelis (Eds.), Spider mites: Their
biology, natural enemies and control (Vol. 1A, pp. 5–65). Amsterdam: Elsevier.
Schuster, R. (1962). Nachweis enes Paarungszeremoniells be den Hornmilben (Oribatei, Acari).
Naturwissenschaften, 49, 502.
Schuster, I. J., & Schuster, R. (1970). Indirekte Spermaübertragung bei Tydeidae (Acari,
Trombidiformes). Naturwissenschaften, 57, 256.
Schuster, R., & Schuster, I. J. (1977). Ernährungs- und fortpflanzungs-biologische Studien an der
Milbenfamilie Nanorchestidae (Acari, Trombidiformes). Zoologischer Anzeiger, 199, 89–94.
Smallegange, I. M. (2011). Complex environmental effects on the expression of alternative
reproductive phenotypes in the bulb mite. Evolutionary Ecology, 25, 857–873.
Sober, E., & Wilson, D. S. (1998). Unto others, the evolution and psychology of unselfish
behaviour. Cambridge: Harvard University Press.
Sokolov, I. I. (1954). The chromosome complex of mites and its importance for systematics and
phylogeny. Trud. Leningr. Obshch. Estesvois, Otd. Zool., 72, 124–159 [in Russian].
Sonenshine, D. E. (1991). Biology of ticks (Vol. 1). New York: Oxford University Press.
Stearns, S. C. (1987). Why sex evolved and the difference it makes. In S. C. Stearns (Ed.),
The evolution of sex and its consequences (pp. 15–31). Basel: Birkahäuser Verlag.
Summers, F. M., & Witt, R. L. (1973). Oviposition and mating tendencies of Cheyletus malaccensis
(Acarina: Cheyletidae). Florida Entomologist, 56, 277–285.
Tait, N. N., & Briscoe, D. A. (1990). Sexual head structures in the Onychophora: Unique modifications
for sperm transfer. Journal of Natural History, 24, 1517–1527.
Tatarnic, N. J., & Cassis, G. (2010). Sexual coevolution in the traumatically inseminating plant bug
genus Coridromius. Journal of Evolutionary Biology, 23, 1321–1326.
Thornhill, R. (1992). Female preference for the pheromone of males with low fluctuating
asymmetry in the Japanese scorpionfly (Panorpa japonica: Mecoptera). Behavioral Ecology,
3, 277–283.
Thornhill, R., & Alcock, J. (1983). The evolution of insect mating systems. Cambridge: Harvard
University Press.
Timms, S., Ferro, D. N., & Waller, J. B. (1980a). Suppression of production of pleomorphic
males in Sancassania berlesei (Michael)(Acari: Acaridae). International Journal of Acarology,
6, 91–96.
Timms, S., Ferro, D. N., & Emberson, R. M. (1980b). Selective advantage of pleomorphic male
Sancassania berlesei (Michael) (Acari: Acaridae). International Journal of Acarology, 6, 97–102.
Timms, S., Ferro, D. N., & Emberson, R. M. (1981). Andropolymorphism and its heritability in
Sancassania berlesei (Michael) (Acari: Acaridae). Acarologia, 22, 391–398.
Vistorin, H. E. (1978). Fortpflanzung und Entwicklung der Nicoletiellidae (Labidostomatidae);
Acari, Trombidiformes. Zool. Jb. Syst., 105, 462–473.
Walter, D. E., & Kaplan, D. T. (1990). A guild of thelytokous mites associated with citrus roots in
Florida. Environmental Entomology, 19, 1338–1343.
Walter, D. E., & Lindquist, E. E. (1995). The distributions of parthenogenetic ascid mites (Acari:
Parasitiformes) do not support the biotic uncertainty hypothesis. Experimental and Applied
Acarology, 19, 423–442.
Walter, D. E., & Proctor, H. C. (1999). Mites: Ecology, Evolution and Behaviour (p. 322). Sydney/
Wallingford: University of NSW Press/CABI. ISBN 0 86840 529 9.
Walter, D. E., & Proctor, H. C. (2007). Using your genetic system to win friends and obfuscate
theorists: patterns in the distribution of genetic sex determination in the Parasitiformes. In
J. B. Morales-Malacara, V. Behan-Pelletier, E. Ueckermann, T. M. Pérez, E. G. Estrada-Venegas,
& M. Badii (Eds.), Acarology XI: Proceedings of the International Congress (pp. 637–643).
References 159

México: Instituto de Biololgía/Facultad de Ciencias/Universidad Nacional Autónoma de


México/Sociedad Latinoamericana de Acarología.
Walzl, M. G. (1991). Comparison of the sclerotized structures of Acaridae and Glycyphagidae
used for copulation. In F. Dusbábek & V. Bukva (Eds.), Modern acarology (Vol. 2, pp. 283–296).
The Hague: Academia, Prague and SPB Academic Publishing.
Weeks, A. R., Marec, F., & Breeuwer, J. A. J. (2001). A mite species that consists entirely of
haploid females. Science, 292, 2479–2482.
Weygoldt, P. (1969). The biology of pseudoscorpions. Cambridge: Harvard University Press.
Wilson, D. S., & Colwell, R. K. (1981). The evolution of sex ratio in structured demes. Evolution,
35, 882–897.
Witalinski, W. (1990). Adanal suckers in acarid mites (Acari, Acaridida): Structure and function.
International Journal of Acarology, 16, 205–212.
Witalinski, W., Dabert, J., & Walzl, M. G. (1992). Morphological adaptation for precopulatory
guarding in astigmatic mites (Acari: Acaridida). International Journal of Acarology, 18, 49–54.
Witte, H. (1984). The evolution of the mechanisms of reproduction in the Parasitengonae (Acarina:
Prostigmata). In D. A. Griffiths & C. E. Bowman (Eds.), Acarology 6 (Vol. 1, pp. 470–478).
Chichester: Ellis Horwood.
Witte, H. (1991). Indirect sperm transfer in prostigmatic mites from a phylogenetic viewpoint. In
R. Schuster & P. W. Murphy (Eds.), The Acari: Reproduction, development and life-history
strategies (pp. 137–176). London: Chapman & Hall.
Woodring, J. P., & Cook, E. F. (1962). The biology of Ceratozetes cisalpinus Berlese, Scheloribates
laevigatus Koch, and Oppia neerlandica Oudemans (Oribatei), with a description of all stages.
Acarologia, 4, 101–137.
Woyke, J. (1994). Mating behavior of the parasitic honeybee mite Tropilaelaps clareae.
Experimental and Applied Acarology, 18, 723–733.
Wrensch, D. L. (1993). Evolutionary flexibility through haploid males or how chance favors the
prepared genome. In D. L. Wrensch & M. A. Ebbert (Eds.), Evolution and diversity of sex ratio
in insects and mites (pp. 118–149). New York: Chapman & Hall.
Wrensch, D. L., & Ebbert, M. A. (1993). Evolution and diversity of sex ratio in insects and mites.
New York: Chapman & Hall.
Wrensch, D. L., Kethley, J. B., & Norton, R. A. (1994). Cytogenetics of holokinetic chromosomes
and inverted meiosis: Keys to the evolutionary success of mites, with generalizations on
eukaryotes. In M. A. Houck (Ed.), Mites: Ecological and evolutionary analyses of life-history
patterns (pp. 282–343). New York: Chapman and Hall.
Yasui, Y. (1988). Sperm competition of Macrocheles muscaedomesticae (Scopoli) (Acarina:
Mesostigmata: Macrochelidae), with special reference to precopulatory mate guarding behavior.
Journal of Ethology, 6, 83–90.
Yund, P. O., & McCartney, M. A. (1994). Male reproductive success in sessile invertebrates:
Competition for fertilizations. Ecology, 75, 2151–2167.
Yuval, B. (1994). The vertebrate host as mating encounter site for its ectoparasites: Ecological and
evolutionary considerations. Bulletin of the Society of Vector Ecology, 19, 115–120.
Yuval, B., & Spielman, A. (1990). Sperm precedence in the deer tick Ixodes dammini. Physiological
Entomology, 15, 123–128.
Zatz, C., Werneck, R. M., Macías-Ordóñez, R., & Machado, G. (2011). Alternative mating tactics
in dimorphic males of the harvestman Longiperna concolor (Arachnida: Opiliones). Behavioral
Ecology and Sociobiology, 65, 995–1005.
Zeh, J. A., & Zeh, D. W. (1994). Last-male sperm precedence breaks down when females mate
with three males. Proceedings of the Royal Society of London B, 257, 287–292.
Chapter 6
Mites in Soil and Litter Systems

The soil ecosystem is the poor man’s tropical rainforest.


(Usher et al. 1982)

People, like other large terrestrial mammals, tread the surface of the Earth and we
know best the aboveground half of the ecosystems we inhabit. Forests, grasslands,
meadows, fields, deserts and cities squat on the surface of the earth, but each has its
shadow existence belowground. Less than half of the energy fixed by the sun is
respired by the plants, animals and microbes that live aboveground: most falls into
the living system we call soil (Macfadyen 1963). Soils are well known for their
extraordinary biological diversity (Wardle 2006) and, more than any other habitat,
this largely belowground system is the empire of mites.
Soil interstices and the sheltering accumulations of decaying organic matter
probably represent the ancestral home of mites and these are still the habitats where
their greatest diversity and abundance of can be found. For example, a single 0.3 ha
site in the mountains of western North Carolina, USA, contains 170 species of litter-
inhabiting oribatid mites of which up to 40 species may be recovered from a single
litter core (Hansen 2000). Densities of 50,000–250,000 mites per square metre in
the upper 10 cm of soil are commonly reported (Petersen 1982) and estimates of a
million or more mites in a square metre of forest soil are not unusual.

The Enigma of Soil Biodiversity

At the macro-level, a handful of humus from a temperate forest looks like a handful
of dead and decomposing vegetation, perhaps with a worm or two intermingled, but
at the micro-level the diversity of microbes, microfauna and mesofauna it contains
is astonishing. Even ignoring all of this diversity except the mites, a handful of
humus may easily contain many of acarine species and hundreds of individuals

D.E. Walter and H.C. Proctor, Mites: Ecology, Evolution & Behaviour: Life at a Microscale, 161
DOI 10.1007/978-94-007-7164-2_6, © Springer Science+Business Media Dordrecht 2013
162 6 Mites in Soil and Litter Systems

Fig. 6.1 Berlese funnel extraction of a handful aspen litter under the snow cover in a frigid Alberta
winter contains at least 25 species of mites, 10 species of insects, 7 species of springtails and
3 species of spiders (Photo by DE Walter)

(Fig. 6.1). The ancient origin of the mite–soil association (Chap. 2) may help to
explain why so many soil-litter mite species exist, but how these species coexist in
a seemingly uniform layer of decomposing litter is still unclear.
Traditional ecological theory assumes that communities are usually near
equilibrium and that most available resources are used to support respiration and
population growth. Since litter does not accumulate to great depths on the surface
of the soil, the logical conclusion is that the decomposer system must be food-
limited and its members should habitually compete for resources (Hairston et al.
1960). The competitive exclusion principle argues that complete competitors
cannot coexist (Gause 1934). However, microbivore-detritivores, the most
species-rich functional group of the soil-mite fauna, appear to have a high level of
trophic overlap – they seem to eat the same foods. Many also tend to have stable
population of long-lived adults.
For example, oribatid mites are among the best-studied soil mites. Develo-
pmental times are long, adults long-lived, species diversity high, species often
occur in aggregations, and population growth tends to be within aggregations,
rather than from increasing numbers of aggregations (Usher 1975, 1976; Norton
1985; Norton and Behan-Pelletier 2009). Gut contents of most oribatids are
largely composed of fungal hyphae and spores or fungal material and plant detri-
tus. Soil fungi appear to be resource limited and, at least at the level of a given
substrate, their assemblages are structured by competitive interactions (Wardle
2006). Thus, oribatid mites would seem likely to compete for and partition their
use of fungal resources. However, although feeding studies typically show that
The Enigma of Soil Biodiversity 163

some fungi are tastier than others, co-occurring oribatids significantly overlap in
their food choices (e.g. Schneider and Maraun 2005; Schneider et al. 2005).
Therefore, there is little indication of temporal or trophic partitioning by oribatid
mites (but see Schneider et al. 2004) on two of the major niche axes assumed to
be important in the coexistence of functionally similar species in equilibrium
communities (Anderson 1975; Giller 1996; Wardle 2006).
Since adult microbivores-detritivore populations tend to be stable, there is little
evidence that predation-dependent cycles (Paine 1966), intermediate levels of
resource availability or disturbance (Connell 1978), or competition are important in
structuring communities (Wardle 2006). Similarly, predator populations also appear
to be composed of many general predators that take similar prey; yet, their adult
populations also are relatively stable. Predators may undergo seasonal bouts of
recruitment or respond to fluctuations in prey populations but, at any one time, a
diverse community of general predators is usually present in undisturbed systems
(e.g. Luxton 1966).
Soil is a matrix infiltrated by a maze of tunnels, pockets and pores. These
passageways and blind alleys provide a large amount of habitat space for small
organisms. This has been interpreted, in terms of fractal geometry, as implying that
a diversity of soil microhabitats also exists (Lawton 1984; Giller 1996). Although a
swath of litter may seem a fairly homogeneous habitat, on closer inspection even
seemingly uniform areas of the forest floor contain a variety of smaller habitats,
e.g. clumps of cryptogams (mosses, liverworts and lichens), rocks, tree trunks,
stumps, and other woody debris, inhabited by restricted subsets of the species pool
(Hammer 1972; Lindquist 1975; Wallwork 1983). Logs, for example, have been
shown to harbour high diversities and species specific assemblages of both Oribatida
and Mesostigmata (Déchêne and Buddle 2010; Skubała and Maslak 2010;
Gwiazdowicz et al. 2011; Beaulieu 2011). Even within the soil organic horizons the
structural diversity of microhabitats has been shown to have a strong positive effect
on the diversity of oribatid mites (Anderson 1978a; Abbott et al. 1980; Hansen
2000). Some level of trophic partitioning related to microhabitat differences also
has been shown (e.g. Anderson 1978b; Siepel 1990).
Increasing the diversity of the litter sources appears to have a positive influence
on oribatid mite diversity, probably because single litter types are more uniform,
produce less structural diversity during decomposition, and limit root penetration
compared to mixed litters (Hansen and Coleman 1998; Kaneko and Salamanca
1999; Hansen 2000). Litter thickness is also positively correlated with soil mite
diversity: thick organic layers have more individuals and species than thinner layers
(Gill 1969; Stanton 1979; Curry 1994). Generally, species diversity of soil arthro-
pods increases with stage of decomposition (Whitford 1996) and stage of plant
succession (Scheu and Schulz 1996; De Deyn et al. 2003; Hodkinson et al. 2004).
Finally, aboveground and belowground processes are not independent of each other,
but intimately connected (Wardle et al. 2004; Wardle 2006). Below we will review
how mites fit into soil systems and try to tease out some of the factors that may help
explain their diversity, how they function, and how they relate to other organisms.
First, though, we need to have some understanding of soil itself.
164 6 Mites in Soil and Litter Systems

What Is Soil?

Mesic forests with well-developed surface organic layers and a predominance of


fungal over bacterial decomposition are home to the highest diversities of mites but
even the driest, hottest and coldest of soils are inhabited by mites. So are the smelli-
est: compost, herbivore dung, bat guano and carrion contain specific suites of mite
taxa. Even the humus that hangs in the canopies of forests swarms with an extraor-
dinary diversity of mites (Walter and Behan-Pelletier 1999; Lindo and Winchester
2008; Beaulieu et al. 2010 and see Chap. 8). Mites thrive in parched, sandy soils
(André et al. 1994), deep soil and caves (Ducarme et al. 2004a, b). They are among
the earliest colonizers of soil exposed by retreating glaciers (Hågvar et al. 2009), are
surprisingly diverse in the High Arctic (Young et al. 2012), and dominate the ter-
restrial fauna of Antarctica (see below: Soil mites in a simple system: Antarctica).
This belowground system starts with the decaying vegetation, dung and animal
remains that accumulate on the surface (the litter layer) and continues down as
deep as roots can penetrate and nutrients percolate into the soil profile. If organic
litter did not decompose, then plant growth would soon grind to a halt for want of
essential nutrients. The soil itself is a complex of the living (soil organisms), the
dead (detritus), and the weathered remains of inorganic rocks (mineral soil).
Microbes slowly burn off detritus, stoked and dampened by myriad other soil organ-
isms. Earthworms, millipedes, termites, beetles, crustaceans, springtails, maggots,
mites and others chop, grind and burrow through plant and animal remains. As they
live and die, their droppings and their bodies are incorporated into the soil and make
the nutrients they contain available to others. The larger arthropod and worm
composters, and those that form large colonies like ants and termites, are the soil
fauna we tend to see, but in numbers of individuals and of species, it is the
minute that dominate and these are mostly mites or mite food (Wallwork 1976;
Crossley 1977; Swift et al. 1979; Seastedt 1984; Moore et al. 1988; Lavelle 1997;
Young et al. 2012).

Forest Floor Habitats

In temperate forests with a low fire frequency, leaf litter and other organic debris
accumulate on the forest floor. Coarse surface litter tends to be rather sparsely
inhabited, probably because of the large pore spaces and tendency to dry out, but
the organic layers below the surface have high diversities and abundances of
mites. Many oribatids actively burrow in leaves, small stems, cones and other
structural components of litter – actively increasing the surface area for microbial
decomposition and inoculating detritus with microbial spores. In coniferous
forests, especially under acidic conditions, a mor soil profile usually develops.
In these soils, layers of needles accumulate and slowly degrade forming a thick
covering of organic matter on the forest floor. The resulting organic horizons are
dominated by small pore spaces – ideal for mites – that decrease in size range,
What Is Soil? 165

median size and complexity with depth. Freshly fallen litter overlays a thick
fermentation layer of partially decayed humus that often contains extraordinary
densities of mites. Under the fermentation layer is a thinner layer of fine humus
particles and arthropod faeces, especially those of oribatid mites. The organic
horizons overlay mineral soil layers variously affected by leaching and deposition
from above. In mor profiles, the average body size of mites often decreases sharply
at the organic–mineral soil interface, as species composition changes and the
reduced average pore size restricts the access of larger mite species (see Swift
et al. 1979; Wallwork 1976; Eisenbeis and Wichard 1987). The deeper mineral
soils in coniferous forests are rarely sampled but mites undoubtedly occur at least
as deep as the roots penetrate.
In neutral to alkaline soils, often associated with deciduous forests, relatively
rapid decomposition and high levels of activity by soil macrofauna (especially
earthworms and millipedes) tend to result in a mull profile with a fairly well-mixed
layer of soil and organic material under the litter. This relatively sparse litter layer
has a laminar arrangement of freshly fallen and partially decomposed leaves that
provide fewer hiding places for mites than a mor, especially if composed of leaves
from a single species of tree (Hansen 2000). However, burrows made by the macro-
fauna or left by rotting roots provide major highways for mite traffic through a mull
profile, and mites follow smaller passageways along the rhizosphere into deeper
soils. Typically, mite species assemblages are drastically different in mor and mull
profiles, even when the forests are adjacent. The major input of organic matter in
mull soils is usually with litterfall in the autumn (or at the beginning of the dry
season), and much of the initial decomposition of litter may occur under a blanket
of snow.
Tropical rainforests can be similar to temperate deciduous forests in that litter
accumulation may be highly seasonal, although associated with dry, rather than
cold, periods. In seasonally dry tropical forests, litter arthropod populations may
flush in the early wet season as accumulated leaf fall is degraded (Levings and
Windsor 1982). Other aspects of soil structure and dynamics, however, are strik-
ingly different. When sufficient moisture is available decomposition is very
rapid, little litter accumulates and nutrients are quickly cycled. With little litter
available on the surface, mite densities tend to be low and mineral soil is so
densely packed with roots that sampling to any depth is difficult. Litter pockets
can be found in the buttresses of trees and on the uphill side of trees and logs but
more humus can be found hanging in the canopy – in treeholes, deep crotches
and among the roots and rhizoids of large epiphytes (Beaulieu et al. 2010).
Similar accumulations of suspended soil can be found in the cool coniferous
rainforests in the Pacific Northwest of North America (Lindo and Winchester
2008; Lindo et al. 2008). Stanton (1979) found a higher rate of spatial turnover
among tropical compared to temperate litter mites, and suggested that tropical
mites tended to be more habitat-restricted and had higher proportions of rare spe-
cies. Based on the few studies that exist, soil mite diversities can be very high in
tropical and subtropical systems (Lasebikan 1974; Chiba et al. 1975; Stanton
1979; Walter et al. 1998; Basset et al. 2012).
166 6 Mites in Soil and Litter Systems

Ephemeral Versus Stable Soil-Litter Habitats

The organic layers of mor soils, rotting wood (Fager 1968) and nests (Lindquist
1975) are relatively stable, persistent habitats for soil mites. Other components of
the soil-litter system, however, can be far less predictable in time and space. Patchy,
ephemeral habitats, such as dung, decaying fruit, rotting fungi and carrion, have
characteristic mite faunas that depend on hitch-hiking rides (phoresy) to reach these
scattered habitats, usually on winged insects. For example, phoretic species of
Macrocheles, Holocelaeno, Holostaspella, Neopodocinum and Glyptholaspis
(all Macrochelidae) and Parasitinae (Parasitidae) are the most common and abundant
mites in herbivore dung (Halffter and Matthews 1971; Krantz 1983); in contrast,
soil-inhabiting Parasitidae typically belong to the subfamily Pergamisinae and soil-
inhabiting Macrochelidae to genera (Geholaspis, Macrolaspis) or species-groups of
Macrocheles that are non-phoretic (Manning and Halliday 1994). However, genera
and species-groups characteristic of herbivore dung tend to have close relatives in
other ephemeral habitats. For example, phoretic species of Macrocheles are com-
mon in carrion (Wilson 1983), rotting cacti (Polak and Markow 1995), under the
bark of beetle-killed pines (Kinn and Witcosky 1977) and in rotting fungi (Walter
and Proctor 1999). Fungal sporocarps, from fleshy to woody, are home to an extraor-
dinarily diversity of mites (Pielou and Verma 1968; Matthewman and Pielou 1971;
Okabe and Amano 1992, 1993; O’Connell and Bolger 1997; Walter and Proctor
1998b; Walter et al. 1998).

Mites, the Rhizosphere and Mycorrhizae

Half or more of all primary production (the energy and nutrients sequestered
by plants from the sun, atmosphere and soil) is out of sight, below ground.
Roots, tubers, bulbs and rhizomes are resources for microbes, worms and arthro-
pods. In terms of plant health, mites are generally considered beneficial in
belowground systems. They are important fungivores, major predators of root
pests, especially nematodes, and only some feed directly on the living belowground
parts of vascular plants.
Mites and other members of the soil fauna tend to be concentrated where
resources are in greatest supply. In systems where organic litter accumulates on the
surface, much of the mite diversity will be located there and, in general, the deeper
the litter layer the higher the abundance and diversity of soil mites. However, mites
will penetrate the underlying mineral soils at least as deep as the roots of plants. The
soil directly influenced by roots is called the rhizosphere (Curl and Truelove 1986).
Roots tend to leak nutrients, especially carbohydrates, and these root exudates stim-
ulate microbial growth. Carbon-isotope studies have shown that the carbon that
leaks from roots is quickly taken up by mites and other microbial grazers (Garrett
et al. 2001; Ostle et al. 2007).
What Is Soil? 167

The rhizosphere appears to be an especially rich habitat for soil mites if the
plant roots have been colonised by a diversity of fungi. Many of these fungi are
mycorrhizae (from the Greek for fungus + root), generally dark-pigmented fungi
that have a complex but usually mutually beneficial relationship with plant roots
(Klironomos and Kendrick 1995). Oribatid mites are known to feed on mycorrhizal
fungi (Schneider et al. 2005), but none are known to be mycorrhizal specialists.
However, when mycorrhizal fruiting bodies are available to oribatid mites they will
eat the spores, pass viable spores in their fecal pellets, and carry viable spores on
their bodies; and thus, may be important in the local distribution of mycorrhizal
propagules (Lilleskov and Bruns 2005). Lichens are home to many mites and,
although they may consume the thallus (the ‘body’ of the lichen) and thus nega-
tively affect fitness of the lichen, they are also capable of passing viable fungal
spores and algal symbionts in their fecal pellets (Meier et al. 2002). Some mites
(especially in the Prostigmata) are likely to be specialized on certain fungal species
or growth forms, while others (e.g. many Oribatida) may be ‘choosy generalists’
(Schneider and Maraun 2005) that will feed on preferred species when available,
but make do with less tasty choices when they must. Earlier attempts to understand
oribatid mite diets using gut enzymes were difficult to interprete (Siepel and
De Ruiter-Dijkman 1993). Newer analytical techniques focused on stable isotopes
(Schneider et al. 2004; Maraun et al. 2011), fatty acids (Ruess et al. 2002) and
careful PCR analysis of dissected gut contents (Remen et al. 2010) seem to offer
more hope for increasing our currently poor knowledge of soil mite feeding
(Coleman 2008).
Aboveground herbivory is known to influence a plant’s belowground allocation
of resources (e.g. Holland et al. 1996) and to affect soil fauna (e.g. Clapperton et al.
2002) including the plant-parasitic Linotetranidae (Prostigmata: Tetranychoidea)
(Leetham and Milchunas 1985). Root-feeders obviously have the potential to
weaken plants and make them more susceptible to attack by aboveground herbi-
vores, but the relationship between plant health and susceptibility is complex. For
example, water-stressed plants have long been known to be more susceptible to spider
mite damage, because nutrients in plant cells become more concentrated and spider mite
reproduction is enhanced.
Root-symbionts such as mycorrhizal fungi may enhance the nutrient status of
plants and thus potentially make them better resources for herbivores. For example,
Hoffmann et al. (2011) found that bean plants inoculated with the arbuscular mycor-
rhizal fungus Glomus mosseae had increased production of beans (a good measure
of fitness), but also supported higher densities of spider mites and suffered more
damage to leaves. Spider mites on beans with mycorrhizal roots were apparently a
better prey resource for a third trophic level and supported higher densities of the
predatory mite Phytoseiulus persimilis. The increased predation may have allowed
the bean plants to increase the amount of carbohydrates allocated to mycorrhizal
growth, a kind of feedback loop. The plant volatiles that attract predators to spider
mite infested plants may also be positively affected by mycorrhizae (Schausberger
et al. 2012).
168 6 Mites in Soil and Litter Systems

Fig. 6.2 Anterior end of worm-like (insert with scale bar) Nematalycidae (Gordialycus sp.), a
mite known from deep in sandy soils. Legs III-IV are minute and displaced well posterior to legs
I-II. The body is encircled with rings of tubercles that may help the mite move through soil pores
(Image by DE Walter)

How Deep Is Soil?

Mites usually escape notice because they are minute and hidden (cryptozoic) in
litter and soil. They may escape even the notice of soil ecologists when they are
deep beneath our feet. Soil is rarely sampled below 15 cm (Kethley 1990) but mites
descend at least as deeply as roots penetrate and in some soils a fifth or more of mite
abundance occurs below 2 m depth (Price and Benham 1977). The fauna of deeper
soils has some similarity to that found in caves, but is distinct (Ducarme et al.
2004a). Though few acarologists dig themselves deep into the soil profile; when
they do, they find strange and wonderful mites, such as the worm-like (Fig. 6.2)
Nematalycidae (Schubart 1973; Coineau et al. 1978) or the enigmatic Pomerantziidae
(Price and Benham 1976).
Traditional extraction techniques for soil mites are based on the Berlese Funnel.
Antonio Berlese, the ‘Father of Acarology’ used a small radiator (hot water run
through pipes) placed over a sample suspended on a screen in a large funnel. A later
worker substituted an incandescent light bulb (the Tullgren-modified Berlese
Funnel) for the water pipes. In either case, the result is the same: as soil dries from
the top down, the arthropods follow the humidity gradient to the bottom of the
sample and then tumble into the funnel and collecting vial below (see Krantz and
Walter 2009). Berlese funnels take advantage of the instinct to conserve moisture by
mites and other small animals with large surface-to-volume ratios that inhabit humid
environments. This technique works best in organic soil layers and only for active
animals that track humidity levels. Quiescent arthropods or those that are adapted to
dry soils are not extracted except by chance. To investigate the diversity of mites
What Is Soil? 169

present in deep mineral soils or more shallow soils in dry environments flotation
techniques are usually required (Walter et al. 1987b; André et al. 1994; Ducarme
et al. 2004a, b).
Although some mites move up and down a soil profile in response to humidity
changes (e.g. diurnal migrations in desert litter), they also can be involuntarily
moved down the soil profile. A spectacular example was the explosion of Mt. St.
Helens in the north-western United States that knocked down entire forests and
buried humus layers under 10 cm of volcanic ash. A year later, this blanket of ash
itself contained no mites but the layers of humus it had sealed off from above con-
tained numerous oribatid mites going about their business seemingly unaffected by
the blast and entombment (Walter and Proctor 1999). The alternating bands of dark
organic matter and lighter ash layers that one can see in road cuts in that region
testify to its long volcanic history – and probably also to layers of fossilised mites.

Mites and Decomposition

In detrital habitats like dung, carrion, and rotting fruit or fungal sporocarps that are
patchy in time and space, predatory mites are a significant component of the com-
munity, but larger animals, especially insects, are the primary consumers in the
early stages of decomposition. Microbivore-detritivore mites with phoretic associa-
tions with flies and beetles do colonize these habitat patches, but where sufficient
insect consumers are present, have little chance to exploit them. When such
resources are superabundant (e.g. after snowmelt in boreal forests when moose and
deer dropping accumulated over the winter litter the forest floor) mites may be more
important in decomposition, but most of the acarine interactions in these patchy
habitats are with the insect consumers (see Chap. 9). Once these concentrated
resources become more like soil, mites and other microarthropods become more
prominent.
In soil, mites contribute to the regulation of decomposition and nutrient cycling
by preying on decomposer worms and arthropods, comminuting organic matter,
feeding on microbes, and vectoring microbial propagules (Seastedt 1984; Moore
et al. 1988; Bal 1982). This latter function may be of major importance. Although
most soil respiration is microbial, microbes have limited abilities to move from one
resource patch to another. Once the energy in a particular patch has been expended,
microbial biomass shuts down and remains dormant until new resources become
available. This has been referred to as the ‘Sleeping Beauty Paradox’ (Lavelle et al.
1994; Lavelle 1997). The dormant microbes require the 'kiss' of arthropod ‘Prince
Charmings’ to awaken. This rather deep kiss (i.e. ingestion) results in microbial
propagules being mixed with fresh resources in the fecal pellet, and often their
transportation to new sites in a manner that is directly analogous to the dispersal of
seeds by fruit-feeding vertebrates.
Rates of decomposition will vary with litter input and abiotic factors such as
rainfall and temperature, as will the distributions and abundances of the litter fauna.
170 6 Mites in Soil and Litter Systems

In the warmer areas of the world litter input can be associated with the onset of a dry
season or be more or less continuous through the year. Temperate deciduous forests
typically have a pulse of litter at the end of the growing season. In tundra, boreal
forests and temperate forests with significant winter snow cover, decomposition
does not necessarily stop at the onset of winter, but may continue over the winter
under an insulating snow pack. Snow cover is an effective buffer against surface
swings in temperature and, even in extreme cold, a subnivean (‘under snow’) layer
may develop with temperatures in the -5C to +5C range, warm enough for cold-
adapted species to function. Mites and other microarthropods are active and con-
tinue to feed in such conditions (see Fig. 6.1) and contribute to the significant
decomposition that occurs over the long winters and snow-covered springs
(Aitchison 1979; Addington and Seastedt 1999; Hågvar and Hågvar 2011). In the
aspen parklands of Alberta, the subnivean mite fauna appears to include most of the
species normally found during the summer. However, in the very early spring,
Astigmata (e.g. species of Acarus, Sancassania and Tyrophagus) are far more
abundant than usual, apparently taking advantage of the flush of fungal growth asso-
ciated with the extra moisture from the melting snow pack (Walter & Proctor pers.
obs.). Collembola also appear to be especially abundant at this time and may be an
important resource for predatory mesostigmatans, some of which are known to fol-
low the ‘snow fleas’ up through the snow pack as they forage (Leinaas 1981).
Although relationships among the components of the decomposer system in the
soil are only poorly understood, several interesting patterns have emerged in recent
research. Vegetation cover and soil properties are not independent of each other in
their effects on the soil fauna, but are also strongly affected by scale at which they
are investigated (Wardle 2006; Nielsen et al. 2012). Although structurally similar
litter types may be colonized by similar decomposer species (e.g. Walter 1985),
structurally different litters, especially in combination, may have strong effects on
decomposer diversity (e.g. Hansen 2000). Therefore, as assemblages of plant spe-
cies replace each other during succession, one might expect soil invertebrate assem-
blages to change as well. More surprisingly, it appears that soil invertebrates may
help suppress dominant early successional plant species, thus increasing local plant
diversity and contributing to successional change (e.g. De Deyn et al. 2003).

Soil Mites in a Simple System: Antarctica

The southern polar region is a huge expanse of ocean with numerous small islands
and one continent, Antarctica. More than 520 species of mites have been reported
from this area, mostly from subantarctic islands, marine collections (water column
and benthos) and as parasites of birds and marine mammals (Pugh 1993, 1994).
Although the acarine diversity of the subantarctic region is respectable, the fauna of
the continent itself is much less diverse. Except for narrow rocky coastal regions
where the snow cover melts in summer, windswept parts of mountain ranges and
extreme deserts where precipitation of any kind is rare, <95 % of the Antarctic
Soil Mites in a Simple System: Antarctica 171

continent is covered with permanent ice and snow. The lowest temperatures known
on Earth have been recorded in Antarctica (−89.6 °C) but the lack of available water
and organic matter are probably more important to the soil fauna (Cannon and
Block 1988; Barrett et al. 2006).
Excluding parasites, synanthropes and spurious records, over 50 species of
mites, 20 of springtails and one chironomid midge (Belgica antarctica) manage to
survive among growths of cyanobacteria, green algae, lichens, mosses and liver-
worts in the tussocks formed by two species of vascular plants (a grass and a pearl-
wort), in accumulations of detritus around penguin rookeries and seal wallows, in
beach wrack and in the thin mineral soils (Block 1992; Pugh 1993, 1994). Of the
33 or so mite species known from inland Antarctica, most are endemics (Block and
Stary 1996; Marshall and Pugh 1996; Sanyal et al. 2002) and appear to be of pre-
glacial origin (Pugh and Convey 2008; Mortimer et al. 2011). Antarctic mites are
interesting in themselves, but as major components in the Earth’s simplest food
webs and most extreme terrestrial ecosystems, they offer us a means of addressing
basic questions about food-web structure.

Antarctic Mites

The oribatid Alaskozetes antarcticus is probably the best studied Antarctic mite
(see Block and Convey 1995). Adults are among the largest of Antarctic arthro-
pods (>1 mm in length), are heavily sclerotised and weigh 200–330 μg. They roam
the ice-free patches of the Antarctic Peninsula feeding on algae, lichens, fungi and
detritus. Herds of several hundred thousand Alaskozetes of all stages have been
found grazing on the foliose alga Prasiola crispa but they also are abundant in
intertidal debris, penguin guano, seal wallows, bird nests and lichen-encrusted
rocks and bones.
Alaskozetes can survive being frozen into a block of ice but will die if their tis-
sues freeze. They tolerate cold by depressing their supercooling point below the
minimum winter temperature they experience. Adults have an average supercooling
point of about −30 °C and that of immature stages averages several degrees lower
(Cannon and Block 1988). As might be expected in a cold climate with only short
periods above freezing, development is slow. Females usually accumulate 4–6 eggs
in their bodies and lay them as a batch. Many females may use the same oviposition
site, resulting in large masses of eggs under rocks and in other sheltered spots. Eggs
hatch in the spring, 1–2 months after being laid. At the end of summer freeze-thaw
cycles, increasing frosts stimulate the immature mites to clear their guts, secrete
anti-freeze agents into their haemolymph and form moulting clusters. The moult to
the next instar usually occurs the following spring and one moult per year is normal.
Adult females do not produce eggs until their second year but they may survive for
several years and can produce several batches of eggs. Generation time is at least 5
years. Mortality is primarily from freezing, desiccation, starvation and predation,
with the latter concentrated in the earlier developmental stages.
172 6 Mites in Soil and Litter Systems

About a dozen other species of oribatid mites in the genera Antarcticola,


Antarctozetes, Edwardzetes, Halozetes, Liochthonius, Magellozetes, Maudheimia
and Oppia are found on the Antarctic mainland, grazing on algae, lichens, fungi and
detritus, along with about twice as many species of Prostigmata. The endeostig-
matan Nanorchestes antarcticus has the southernmost distribution of any arthropod
(85°32′S) and also has the widest geographic and altitudinal distribution (3–2,245 m)
of any Antarctic mite (Block 1979, 1984). Hundreds of thousands of these tiny
(adult live weight = 3–4 μg) hopping mites can be found grazing on microalgae in
each square metre of some barren rocky areas.
In less extreme coastal sites, thick turfs formed by mosses and other bryophytes
are dominated by species of Nanorchestes and small prostigmatans with generation
times of 2–3 years (Usher and Booth 1986). As usual, the larger the animal, the bet-
ter known its feeding habits but small mites (e.g. Nanorchestes, Eupodes,
Apotriophtydeus) are thought to feed primarily on microalgae and fungi. However,
Ereynetes species are probably predators (Krantz 1978; Zhang and Sanderson
1993). Grazing mites are most abundant in rich sites, such as the algae growing on
feathers of long dead penguins (Gressitt and Shoup 1967). Species of Stereotydeus
are the largest of the prostigmatan herbivores, at 600–700 μm in length (30–40 μg),
and are especially abundant (20–2,000 m−2) in mosses and lichens.
Compared to the interior, the Antarctic Peninsula is relatively diverse and well
studied. The Peninsula’s fauna includes the only species of free-living Mesostigmata
known from the Antarctic mainland: Gamasellus racovitzai and Hydrogamasellus
antarcticus (Pugh 1993; Marshall and Pugh 1996). The latter has not been well
studied but Gamasellus racovitzai is the top predator in maritime Antarctica and
feeds on any animal it can overpower (Lister 1984; Block 1984; Usher and Bowring
1984; Lister et al. 1987, 1988). Other predators include the prostigmatans Rhagidia
gerlachei and Tuberostoma leechi in the Rhagidiidae, and Ereynetes macquariensis
in the Ereynetidae.
The inland acarofauna of the frozen continent is composed of fewer than 30 spe-
cies of mites (Marshall and Pugh 1996). Two species of Tuberostoma (formerly
Coccorhagidia) are the only predators (Gless 1967). Many of the barren rocky sites
have only microalgae and cyanobacteria as primary producers at the base of the
food web and a few prostigmatans and nematodes as top grazers. The ice-free Dry
Valleys of Southern Victoria Land often lack mites and have an impoverished fauna
of nematodes and the occasional tardigrade or rotifer (Freckman and Virginia 1997).

An Antarctic Food Web

A simple pictorial way to express feeding relationships is a food web, i.e. a map of
who eats whom or what: a picture of an organism’s place in its trophic universe.
Like all art, a food web is strongly influenced by the perceptions and abilities of its
creator. Food webs can emphasise the consumers of a particular species or resource
(source webs) or of a particular consumer or group of consumers and their foods
Soil Mites in a Simple System: Antarctica 173

(sink webs), or they can attempt to show all the feeding links of the organisms at a
particular geographical locality (community webs) (Cohen 1978). On a more for-
mal level, food webs also are used to model the movement of nutrients and the
transfer of energy among the components of an ecosystem (Odum 1991).
Invariably, the construction of food webs reflects the theory of trophic levels.
Indeed, food webs are usually constructed in trophic steps from the basal species
(usually primary producers or detritus) to the top carnivore and the species within a
trophic level are often lumped together, to one degree or another, to form ‘trophic
species’, guilds, and the like (Moore et al. 1988; Hunt et al. 1989). Typically, the larger
the organism, the more likely it is to be treated as an individual species in a food web.
Simple food webs tend to resemble food chains, i.e. the plant–herbivore–carnivore
sequence that reflects one of our most basic views of ecosystems (and of our own
place at the top of these systems). Upon closer scrutiny, however, even very simple
systems become ensnared in a dense web of feeding relationships.
Food webs have been a central theme of the ecological literature but they have
been constructed with little consistency. The vast majority of published food webs
were produced by field biologists as simplified illustrations of the feeding relation-
ships considered important for their study, not as standardised descriptions of com-
munity trophic dynamics. These diagrams, however, became the grist for a guild of
mathematical modellers, with predictably unlikely and illogical results. The impli-
cations of honouring mathematical models over reality for our Antarctic foodweb
was discussed in Walter and Proctor (1999), but here we will limit ourselves to an
overview as an introduction to the trophic roles of mites in soil food webs.
Food webs can be considered to be composed of one or more food chains. The
length of each chain is the maximum number of times nutrients and energy are
transferred from a non-consuming basal species to an apex consumer, and it is
equal to one less than the number of species in the chain (Cohen 1978). For exam-
ple, Fig. 6.3a shows a simple sink web for the maritime Antarctic predator
Gamasellus racovitzai on a mat of the foliose alga Prasiola crispa that has been
colonised by the isotomid springtail Cryptopygus antarcticus (Usher and Bowring
1984). This system has a chain length of 2, i.e. two feeding links unite the three
trophic species (one at each of three trophic levels). Primary production at the base
of the web is represented by the alga that captures energy from sunlight and nutri-
ents from rocks or penguin guano to support its growth. Grazers, represented by
the springtail, transform plant-captured solar energy and nutrients into animal bio-
mass that supports the top predator G. racovitzai. We could expand this web by
adding some of the other grazers, e.g. Alaskozetes and Stereotydeus villosus, to the
middle trophic level; however, this would simply add two more chains of two links
and not affect the mean chain length. We could also partition Alaskozetes into two
trophic species: armoured adults and soft-bodied immature stages. Since
Gamasellus preys on the nymphs and larvae of Alaskozetes, but not the adults
(Block 1984), the average chain length would drop to 1.75 (i.e. three chains of
length 2 and one of length 1).
If all we were interested in, e.g. the effect of feeding by G. racovitzai on C. ant-
arcticus populations, then Fig. 6.2a would be an adequate ‘picture’ of our study.
174 6 Mites in Soil and Litter Systems

Fig. 6.3 A simple Maritime Antarctic foodweb supported by primary producers (moss turf and
the foliose alga Prasiola crispa). Arrows represent the flow of energy and nutrients. (a) Sink web
for predatory mite Gamasellus racovitzai that feeds on the springtail Cryptopygus antarcticus
which grazes on algae. (b) A more comprehensive community web (see Walter and Proctor 1999
for supporting references). Shaded boxes represent a compartment fuelled by heterotrophic bacte-
ria. Dotted lines indicate ‘incidental’ predation of nematodes by grazing springtails (Image DE
Walter & HC Proctor)

However, these two species do not live alone. Even a relatively simple maritime
Antarctic food web, a diversity of mites and other invertebrates live in the surround-
ing primitive soils, lichen-encrusted rocks and moss turfs (Block 1984). The appear-
ance of these animals on a particular algal mat will depend on distance from source
populations, weather conditions, the presence of appropriate cues (e.g. prey must
colonise the algae before a searching predator is likely to remain) and other stochas-
tic factors (Kitching and Pimm 1985; Kitching 1987), but any full consideration of
the trophic ecology of our top predator would necessarily need to include as many
of these interactions as possible.
Soil Mites in a Simple System: Antarctica 175

Figure 6.3b shows a more comprehensive set of feeding relationships based on


the available literature (e.g. Fitzsimons 1971; Block 1984; Lister 1984; Lister et al.
1987, 1988; Cannon and Block 1988), our own experience with soil animals and
consultations with Antarctic researchers. In this representation, 13 trophic species,
30 feeding links and 7 trophic levels are shown. Two differences from Fig. 6.2a are
immediately obvious: (1) mean chain length is much more difficult to calculate and
(2) omnivory, i.e. feeding at more than one trophic level in its special food web
definition, is common. The top predator Gamasellus, for example, feeds on all of
the trophic levels beneath it except directly on the alga. Most of the grazers in this
web feed on at least two trophic levels. Omnivory appears to be an important
feeding strategy in free-living Antarctic arthropods – as it is for many animals,
including ourselves (Walter 1987a, b; Sprules and Bowerman 1988; Pimm 1991).
This is especially important to remember when forcing mites into trophic species,
guilds or functional groups (see below).
The longest of the traditional (unshaded) food chains in Fig. 6.3b result from
three simple changes. The first is including another predatory mite in the system,
the prostigmatan Rhagidia gerlachei, a monster of 1,400 μm length (Strandtmann
1967). The second is partitioning the arthropod grazers into six trophic species.
Even this level of resolution, however, is a simplification. Only the five largest
grazers (three springtails and two mites) are treated individually: small prostig-
matans and oribatids are lumped into a single trophic species – 'small mites'. The
final change is including another resource/consumer in the system: fungi. An
algal mat is actually a complex of algae and other microorganisms that feed on
the algae and on each other, and in turn are eaten by larger organisms. Fungi are
eaten along with algae by most larger consumers and are preferentially grazed by
small mites. The new feeding relationships account for six chains of length 2
(alga-consumer–top predator), five of length 3 (alga–fungi-grazer–top predator)
and four chains of length 4 (alga–fungi-consumer–Rhagidia–top predator) for an
average chain length of 2.87. Therefore, by adding more realism to the food web,
we have been able to raise a simple Antarctic system, completely free of verte-
brates, to the upper end of the chain-length averages reported for all food webs
(Polis 1991).
Additionally, Fig. 6.2b shows a simplified bacteria-based compartment (shaded)
in the food web. Although often ignored, nematodes, protozoa and bacteria are
invariably present in any soil food web and the Antarctic is no exception (Block
1984; Vincent 1988). In even the simplest of Antarctic food webs – in the Dry
Valley deserts – microbivore and predator–omnivore nematodes can be found
(Freckman and Virginia 1997).
The longest chains result from the pivotal position and the broad diet of
Gamasellus. Like many predatory soil mites (Walter et al. 1988), Gamasellus is a
generalist that attacks most of the invertebrates of appropriate size that it encoun-
ters (Block 1984). Effective specialisation may result from habitat preferences of
the predator, size changes in the predator during ontogeny, interactions with other
predators, change in prey defences over its ontogeny or variation in relative abun-
dances of prey. However, predators such as Gamasellus that actively search for
176 6 Mites in Soil and Litter Systems

prey will encounter a variety of acceptable prey types over their lifetime and
success at prey capture and feeding will be primarily a function of prey defences.
Such opportunistic predation resulting in broad diets is not limited to soil mites and
is not rare (Polis 1991).
Grazers also have feeding behaviours that defy simple trophic classification. For
example, we credit our nematode ‘trophic species’ with swallowing only bacteria
and protozoa but, in reality, any organic bit of the appropriate size would be swal-
lowed, including microalgae and fungi (yeasts, small conidia). Depending on the
speices, maritime Antarctic nematodes may feed only on fungi, prey on other nema-
todes or eat just about anything (Spaull 1973). The only arthropod known to be a
predator of nematodes in this system is Gamasellus (Block 1984) but the isotomid
springtails undoubtedly eat nematodes as they graze algae and detritus (dotted lines
in Fig. 6.2b). There is good evidence that grazing mites and springtails actively
supplement their diet with nematodes competing for the same microbial resource
(Walter 1987b).
Although the information content of Fig. 6.2b far exceeds that of 2a, it is still
a gross simplification of the maritime Antarctic food web. For example, two
species of rhagidiid mites inhabit the Antarctic Peninsula (Strandtmann 1967;
Zacharda 1980) and their presence can affect prey selection by Gamasellus (Usher
et al. 1989). Rhagidiids may feed on larvae or nymphs of Gamasellus, adding a
loop to the web (Polis 1991). The other maritime Antarctic mesostigmatan,
Hydrogamasellus antarcticus, occurs in habitats similar to those of Gamasellus
(Pugh 1993) but none of the possible interactions between these species seem to
have been explored. We present Stereotydeus as a strict herbivore, although it also
may feed on fungi (Fitzsimons 1971). The small mite ‘trophic species’ may con-
tain both fungal and algal specialists and probably contains one predator, Ereynetes
macquariensis, which may feed on small mites (including immature stages of the
larger arthropods), eggs or nematodes. Alaskozetes has no way to avoid eating
protozoa and bacteria that live in the algae and may ‘accidentally’ or deliberately
graze on nematodes. None of the feedback loops from higher to lower trophic
levels are shown (e.g. faeces, exuviae and death-providing resources for microbes,
algae and arthropods), and disease and cannibalism are ignored. Bacteria, fungi,
protozoa, nematodes and small mites are complexes of species but all are treated
as if they are single ‘trophic species’ and other groups are entirely ignored (e.g.
viruses, tardigrades, rotifers).
The use of species or trophic species as black-boxed system variables may be
useful for modelling energy and nutrient dynamics and understanding constraints
on ecosystem-level processes. However, it is important to remember that it is energy,
not organisms, that flows through ecosystems, temporarily residing at different tro-
phic levels. Organisms use energy from many levels within their ecosystems and
selection favours survival and reproduction, not stable mathematical models. To
paraphrase Robert Paine’s (1996) commandments of food-web research: avoid jar-
gon; let nature guide mathematics, not vice versa; honour species; and pay attention
to small organisms.
Feeding Guilds and Functional Groups 177

Feeding Guilds and Functional Groups

Ecological interactions occur among individuals and can change with life stage,
season, time of day, physiological need or in response to many other variables.
Studying all of the species plus all of their interactions is a daunting task, even in
a simple ecosystem like maritime Antarctica. In more complex systems it is virtu-
ally impossible. Therefore, community ecologists must reduce that complexity to
a manageable size. Two common, if intellectually vacuous, approaches to dealing
with the diversity and complexity of soil animals have been to sort them into coarse
taxonomic or size categories. Not surprisingly, counting the total number of mites,
or mites per suborder, in various samples provides plenty of tedium, but little
insight. Size categories, e.g. microfauna (<100 μm), mesofauna (100–2,000 μm)
and megafauna (>20 mm) (Swift et al. 1979; Beare et al. 1995), at least cut across
taxonomic boundaries – soil mites fall primarily into the mesofaunal category,
along with springtails and arcane groups like pauropods and symphylans. However,
size categories tend to be used as convenient excuses for ignoring organisms that
are larger or smaller, rather than for studying the relationships between body size
and relevant ecological factors such as soil-pore size.
An alternative to arbitrary size or taxonomic classification is based on the hypoth-
esis that communities are composed of clusters of functionally similar species. The
most popular term for such a cluster is guild. Root (1967) originally proposed the
term to describe the birds on a California hillside that obtained most of their food by
gleaning insect prey from oak foliage. He argued that studies of assemblages of
organisms that used similar resources in a similar manner were ecologically rele-
vant alternatives to studies based on taxonomy. Strangely, though, only birds (Class
Aves) were members of this guild, not the insects and spiders that stalk insects on
foliage or hawk from perches in the oak canopy. To Root, guilds reflected ecological
relationships among species much the way that genera reflected phylogenetic rela-
tionships. He also explicitly stated that competition for resources was inherent in
guild membership and was the primary reason that guilds were relevant for study.
Although soil fungal communities may be structured by competition for resources,
there is no indication that soil microarthropod communities are, the reverse may in
fact be true: the high diversity of soil mites may in part result from low levels of
resource competition (Wardle 2006).
The word ‘guild’ has been a clear success but a shadow has fallen between defini-
tion and use (Jaksic 1981; Walter et al. 1988; Hawkins and MacMahon 1989;
Simberloff and Dayan 1991). Rarely has competition been shown to be limiting, or
even strong, within guilds; and typically, guilds are defined with little concern as to the
full range of species using similar resources in a similar manner. Almost invariably,
and usually implicitly, the first criterion of guild membership has been belonging to
the taxonomic group of interest, whether it is vertebrates, birds, mites or the
like. Some authors have proposed terms for more finely resolved guilds, e.g. feeding
guilds, leagues or ensembles (e.g. Walter et al. 1988; Faber 1991; Fauth et al. 1996).
178 6 Mites in Soil and Litter Systems

When predators feed on other predators, this has been called intraguild predation and
when they feed on non-predators, extraguild predation (e.g. Buitenhuis et al. 2010).
In the ecological literature, ‘guild’ has been used to enhance the ecological
importance of almost any level of ecological association from congeneric pairs of
species to mass collections of unrelated animals. Acarological studies are well
within the zoological mainstream with guilds defined as assemblages of congeneric
parasites using similar hosts (e.g. Downes 1986), astigmatans migrating as hypopi
from ephemeral habitats (Knülle 1995), predatory mesostigmatans of similar size
and reproductive mode (Walter and Kaplan 1990b), predators using differing pur-
suit and ingestion modes (Walter et al. 1988) or oribatid mites with similar feeding
capabilities (Siepel and De Ruiter-Dijkman 1993; Kaneko 1995).
A useful alternative to guild was proposed by Cummins (1974), who divided
stream-dwelling animals into functional groups based on the size of the food par-
ticles that they captured. Although its subsequent definition has converged on guild
by including the manner in which the resources are used, and some consider ‘func-
tional group’ a less elegant synonym of guild (Simberloff and Dayan 1991), in prac-
tice functional groupings have been more consistently applied and lack the
assumption about the importance of competition among members. This also avoids
the pitfall of assuming all interactions are negative. Facilitation (i.e. the action of
one organism making resources available to another) is likely to be an important
component of the interactions of soil organisms (Wardle 2006). By concentrating on
resource use first, taxonomy second and competition not at all, the functional group
concept offers a first approximation to identifying clusters of species that may be
ecologically linked. Demonstrating where and when these links may influence the
maintenance of soil biodiversity, decomposition rates, nutrient dynamics or other
ecosystem functions should be a useful goal of soil ecology.
Below, we present our view of the major functional groups of soil organisms that
have mites as members. Each group is defined by the resources used, modified by
what we believe are ecologically important differences in use and taking into account
how the use of resources by one group may change their availability to other
organisms (Jones et al. 1997). Where relevant, life history tactics are discussed as
well (see Siepel 1994). Functional groups are useful heuristic devices for considering
the trophic roles of mites in particular habitats – but they are hypotheses, or working
models, not facts. Since this is a book about mites, we will concentrate on the acarine
members and only briefly consider other organisms that may process resources in a
similar manner. Also, membership in a functional group is not necessarily exclusive –
an organism may belong to more than one functional group over its lifetime.

Comminuting Microbivore–Detritivores: Grazers and Browsers

Most soil-inhabiting sarcoptiform mites (Oribatida, Astigmata) feed by biting off


fungal spores, chunks of hyphae or strands of algae and swallowing them. Some of
the larger and more heavily sclerotised oribatids carry this feeding one step further
by eating recently dead animals (necrophagy) or dead vegetation and animal
Feeding Guilds and Functional Groups 179

material that support microbial growth (saprophagy), and so they directly contribute
to the reduction of detritus (dead organic matter). During feeding, a bolus of food
particles accumulates at the base of the oesophagus and as it passes through the
ventriculus it is covered by a thick peritrophic membrane. As digestion continues,
the pellet passes along the gut until it is eventually defecated. Oribatid dropping are
compact and generally smooth-surfaced particles, up to 200 × 140 μm in size, with-
out the mineral inclusions often found in the fecal pellets of other soil animals
(Rusek 1985). Microbial spores within these pellets are often viable.
Saprophagy probably developed as a means of capturing subsurface microbial
tissues (Norton 1985) – eating the piecrust to get at the filling. ‘Saprophages’ feed
on both detritus and on living microbes, and sometimes exclusively on the latter
(Behan and Hill 1978; Behan-Pelletier and Hill 1983) but either diet leads to the
formation of small particles that contribute to soil structure. Oribatid fecal pellets
have been major components of soils at least since the Carboniferous. In coal beds
formed by plants in Pennsylvanian swamps, both aerial and root remains of club
mosses, sphenopsids, ferns, seed ferns and cordaites contain oribatid borings and
faeces (Labandeira et al. 1997). The reduction of detritus and patches of microbial bio-
mass to small pellets is a direct engineering of the habitat and results in the production
of microhabitat spaces that can be used by smaller mites, nematodes, protozoa,
fungi and bacteria. By burrowing through internal tissues of decaying vegetation,
sarcoptiform mites both create pore spaces and disperse microbial propagules.
Unlike earthworms and larger-bodied soil arthropods like ants and termites,
mites appear to have limited abilities to directly modify pore structure in mineral
soil (Norton 1985); but, at least in dry soils, some of the larger species are able to
burrow among soil aggregates (Tevis and Newell 1962; Coineau 1973).
When a microbivore–detritivore feeds, it may obtain nutrition from ingested mate-
rial in a variety of ways. For example, to some mites, fungal spores may be much like
fruit is to many vertebrates: a tasty and digestible outer coating with an indigestible
core. However, when living cell walls are ruptured during feeding, the cell contents
become a rich source of food. For sucking-microbivores (see below), these cell con-
tents may be the only source of nutrients. Microbivore–detritivores, however, ingest
cell walls, as well as cell contents. Mites that are able to digest the structural elements
of cell walls have extra nutrients available. Additionally, dead cells (microbial as well
as those of higher plants) then become a viable food. Coprophagy is a simple, and
probably common, means of obtaining nutrition from cell-wall materials by reingest-
ing faeces after microbial digestion has taken place in the ‘external rumen’ of the soil
(Lavelle 1997). However, some mites may be able to produce the appropriate enzymes
themselves or through a mutualistic association with gut microbes (see discussions in
Norton 1985; Siepel and De Ruiter-Dijkman 1993).
Earlier attempts to classify oribatid feeding behaviour were based on the relative
proportion of various types of food in gut contents and emphasised the role these
mites play in the comminution of organic matter (Siepel and Maaskamp 1994).
Although the jargon that developed tends to obscure rather than clarify feeding
habits, certain patterns are clear: large species consume more detritus, medium-
sized species consume more fungal hyphae and small species tend to feed on fungal
spores or ‘amorphous’ material (Schuster 1956; Anderson 1975, 1978a; Luxton
180 6 Mites in Soil and Litter Systems

1972, 1979). Species feeding on decaying plant matter tend to have massive
chelicerae (Kaneko 1988). Preferences for particular resources can be related to
digestive capabilities and can vary between adults and juveniles of the same species
(Siepel 1990). The distributions of some oribatids are positively correlated to the
amount of fungal hyphae and the diversity of darkly pigmented fungi present in the
soil (Klironomos and Kendrick 1995). However, these are general trends and most
oribatids appear to be highly opportunistic in their choice of foods (Behan and Hill
1978; Behan-Pelletier and Hill 1983; Usher et al. 1982; Norton 1994; Siepel 1995).
One especially relevant form of opportunism is facultative predation. Many oth-
erwise decent fungivores will take small, defenceless organisms such as protozoa,
rotifers, nematodes and collembolans when the opportunity arises (Rockett and
Woodring 1966; Rockett 1980; Walter et al. 1986; Walter 1987b; Walter 1988b;
Walter et al. 1988; Walter and Kaplan 1990a; Norton et al. 1996). This behaviour
can result in effects much farther up the food web when oribatid mites ingest the
infective stages of tapeworms, become intermediate hosts and are ingested by mam-
mals (see above). Feeding on algae is also very common, although from an ecosys-
tem point of view, live algae capture energy from the sun and are primary producers
and their consumers are herbivores.
So, microbivore–detritivores are especially difficult animals to assign to a simple
trophic category. Depending on the habitat, time of year or stage, a particular indi-
vidual may be a predator, herbivore, fungivore, saprophage or omnivore. In all
cases, however, this feeding contributes directly to soil structure through comminu-
tion (breaking large particles into smaller ones), inoculation (many ingested micro-
bial propagules pass through digestive tracts without damage) and the formation
of fecal pellets. Grazing may also have a strong stimulatory effect on the growth of
fungal hyphae (Visser 1985; Moore et al. 1988) and remove significant numbers
of fungal propagules (Gochenaur 1987).
Mites that function as microbivore–detritivores disperse primarily as adults, have
long-lived adults, relatively long generation times, are rarely phoretic and have
repeated bouts of reproduction (iteroparity). These trends are characteristics of orib-
atid mites, which are the primary mite constituents of this group and these life-
history traits may say more about phylogenetic inertia than about adaptations to the
grazing role (Norton 1985, 1994). For example, astigmatans in the genus Tyrophagus
are common in some soils but have relatively short generation times (Walter et al.
1986), and other soil-inhabiting Astigmata use phoresy to disperse. Similarly, most
grazers are bisexual and diplo-diploid but thelytoky is common and arrhenotoky
known in some taxa (Norton et al. 1993).
In addition to the sarcoptiformans discussed above, many collembolans, insects,
crustaceans and myriapods function similarly in soil litter. In Australian forests,
talitrid amphipods, ostracods, isopods, millipedes, the caterpillars of oecophorid
moths and the grubs of cryptocephaline chrysomelid beetles are important proces-
sors of surface litter. Their fecal pellets provide an additional resource for the mostly
smaller mites and the larger of these arthropods no doubt act as incidental predators
when they consume immature stages of oribatids and other small animals that
burrow in litter (Fig. 6.4).
Feeding Guilds and Functional Groups 181

Fig. 6.4 Some soil mite microbivores: (a–d) Oribatida: (a) Rostrozetes sp. feeds on decaying
vegetation and microbes, (b) Robust chelicerae of same; (c) Scheloribates sp. a fungivores that also
takes soft-bodied prey such as nematodes; (d) Eupelops sp. a specialized fungivores with narrow,
elongate chelicerae used to collect fungal hyphae. (e–h) Prostigmata: (e) Tydeidae, (f) Chelicera
with stylet-like digits (arrow) used to puncture hyphae; (g) Scutacaridae; (h) Microdipsidae.
C = chelicera, p = palp (SEMs by DE Walter)

Piercing-Sucking Microbivores

Prostigmatans feed only on fluids – that is they pierce their host with chelicerae
and drain the fluids and externally digested remains of their food. Mycophagous
Prostigmata (e.g. Eupodidae, Tydeidae, Tarsonemidae, Scutacaridae, Pygmephoridae)
use stylet-like cheliceral digits to puncture fungi. These mites are often very abundant
in soil-litter habitats, especially in deserts, grasslands and dry forests (Kethley
1990; Curry 1994). Some small oribatid mites with attenuated mouthparts (e.g.
Suctobelboidea) appear to feed primarily on fluids (they do not form gut boluses).
182 6 Mites in Soil and Litter Systems

This is also true of a number of families of Endeostigmata (Walter 1988b). Sucking


microbivores tend to have relatively short generation times and to be more variable
in time and space compared to grazers.
Piercing-sucking is functionally different from particle feeding for a number of
reasons. Most obviously, prostigmatans directly contribute little to soil structure.
They do not comminute litter and their faeces are liquid, so only their exuviae and
bodies are added to the physical structure of soils. In some sense, their feeding is
more like parasitism than grazing, especially in physogastric heterostigmatans that
attach to a hyphal mass and become engorged (Kaliszewski et al. 1995). Fungal
spores are not ingested and any movement of microbial propagules must be of those
adhering to their cuticle or sequestered by the mites in special mycangial pouches
or sporothecae (Lindquist 1985). Also, suckers cannot use their mouthparts to bore
into detritus to obtain microbes beneath the surface. Algae and nematodes are fed
on by some sucking ‘fungivores’ and, as in grazers, a simple trophic classification
can be misleading. Aradid bugs, neanurine collembolans, tylenchid nematodes and
proturans feed on fungi in a similar manner.
Piercing-sucking microbivores, especially members of the Heterostigmata, can
be devastating pests in mushroom cultivation. For example, Microdispus lambi
(Microdispsidae) suddenly appeared in Spanish mushroom crops in 1996 and
quickly became a primary pest causing significant economic losses (Navarro et al.
2010). Several other heterostigmatans are significant pests of commercial mush-
rooms and are particularly difficult to control because of their ability to recolonise
via phoresy on insects.

Filter-Feeding Microbivores

Histiostomatid astigmatans are the only mites known to filter bacteria, yeasts,
protozoans and very fine particulate organic matter for a living (Walter and Kaplan
1990a). Their chelicerae and palps are modified into brushes that sweep fine parti-
cles towards the narrow (5 μm diameter) buccal opening. These mites are common
and often extremely abundant in any kind of wet, decomposing organic matter, e.g.
carrion, rotten fruit, rotting fungi, sap fluxes and wet litter. Generation times are
relatively short and dispersal is by phoresy of the heteromorphic deutonymph, usu-
ally on an insect. Rhabditid nematodes, some maggots and copepods filter feed on
similar resources.

Direct Plant Parasites

Root-grazing or -sucking insects, plant-parasitic nematodes and fossorial mammals


are well-known belowground herbivores (Andersen 1987) but mites that feed on
belowground parts of higher plants are poorly studied. Some tetranychoid mites
Feeding Guilds and Functional Groups 183

Fig. 6.5 Berlese extraction of soil from a pot with a dying bulb: R the Bulb Mite Rhizoglyphus
robini, G the predatory mite Gaeolaelaps aculeifer, P a male Parasitid mite, also a predator,
H hypogasturid springtail, E enchytraeid pot worm (Photo by DE Walter)

(Tuckerella, Linotetranus) are common in soil samples from grasslands and open
forests (Smith-Meyer and Ueckermann 1997) and probably feed on the crowns and
roots of grasses. The aboveground parts of grasses, herbs and ferns are attacked by a
variety of eriophyoid and tetranychoid (e.g. Bryobia, Oligonychus, Tetranychus) mites
that sometimes seek refuge in the soil, especially to overwinter in cold climates. To
the degree that these obligate plant parasites are host-specific, the diversity of these
mites will reflect the diversity of plants present in a system. Mites that feed on vascu-
lar plants are best known in agricultural systems where they are often significant pests
(see Chap. 8), but even there the belowground feeders are poorly known.
Oribatid mites are occasionally reported as boring in live roots or tubers but these
attacks are probably secondary after damage and invasion of plant tissues by fungi.
Few records exist of oribatids burrowing into healthy plant tissue except in aquatic
habitats (Krantz and Lindquist 1979). For example, ducked (Lemna app.) is often
rife with burrows of juvenile Hydrozetes (Oribatida), although it is unclear if these
mites are eating healthy or senescent tissue (see Chap. 8). However, some astig-
matans, especially species of Rhizoglyphus (Fig. 6.5), can damage tubers or other
underground storage structures or seeds (e.g. peanuts). These mites seem to feed at
the juncture of healthy and fungus-infected tissues, and avoid decayed tissues, but
are significant pests of onions and other bulbs including ornamental flowers.
Many oribatid mites feed directly on algae and lichens, and so act as herbivores
in an ecological sense. Free-living green algae growing rocks and bark (e.g.
Protococcus species) make excellent food for culturing many oribatid mites. Algae
and lichens do occur on the soil surface, but they are best studied on rocks, bark, and
184 6 Mites in Soil and Litter Systems

tree canopies (see Chap. 8). In these habitats, oribatid mites are importance
lichenivores and immature stages often burrow within the lichen thallus (Seyd and
Seaward 1984; Behan-Pelletier and Walter 2000; Fisher et al. 2011). The complex
of bacteria, algae, fungi, lichens, and mosses that form crusts on the soil in open
semiarid and arid systems is also home to numerous mites that feed on the micro-
flora (Neher et al. 2009).

Mites and Moss

Bryophytes (mosses, hornworts and liverworts), lichens, and free-living algae lack
the grandeur of trees, but can be found in even the driest and coldest ecosystems.
This is not by chance: they are well adapted to tolerate desiccation and freezing and
opportunistically take advantage of even brief pulses of water and warmth. Mosses
(Bryophyta), especially, are common on the ground and often first colonists on bare
patches of soil. Although the bryosphere (Lindo and Gonzalez 2010), living and
dead mosses with their associated microbes and arthropods, extends into aquatic
and arboreal systems, it is also intimately linked to the soil and decomposition,
especially in cool-temperate and boreal forests.
In closed boreal forests, feather mosses (e.g. Hylocomium splendens) form a
blanket-like cover over the surface of the ground and are an important habitat for
predatory Mesostigmata (Salmane and Brumelis 2008). Oribatid mites are often
called moss mites, and mats of bryophytes support a surprising diversity of these
mites. As well as the living bryophytes, these mats accumulate detritus and senes-
cent bryophyte material that support fungal growth. Moss mats usually consist of an
upper green living layer and lower layer of senescent to dead and decaying moss
leaves, rhizoids, and other detritus (Lindo and Gonzalez 2010). Remarkably, there
is little evidence that oribatid mites do more than incidental grazing on living mosses
(Schuster 1956; Woodring 1963; Lindo and Gonzalez 2010; Perdomo et al. 2012). Only
one possibly obligate relationship between a bryophyte, in this case a liverwort, and an
oribatid mite is known (Colloff and Cairns 2011). Numerous fungi and bacteria, includ-
ing nitrogen-fixing cyanobacteria, nematodes and other small soft-bodied inverte-
brates are associated with the moss mats, and it seems that these symbiotes provide
most of the nutritional resources for oribatid mites living in moss.
However, recent evidence suggests mites may play an important role in the moss
life cycle. About half of all mosses have male and female structures on different
individuals and male sperm requires a continuous layer of water to reach female.
Cronberg et al. (2006) have demonstrated that when oribatid mites (and springtails)
are present in microcosms with separated unisexual Bryum argenteum, female
mosses are fertilized. Moreover, the mites prefer fertile female moss shoots, pre-
sumably because they secrete sugars and other nutrients on which the mites feed.
Bryophytes on the surface of the soil and on logs and rocks also harbour a diver-
sity of prostigmatans whose darkly pigmented internal body contents probably rep-
resent chlorophyll and its breakdown products. Species of Eustigmaeus, an aberrant
Feeding Guilds and Functional Groups 185

genus of the mostly predatory Stigmaeidae, feed on mosses (Gerson 1972). The
Penthalodidae (Stereotydeus, Penthalodes) are thought to feed on moss leaves. The
closely related earth mites (Penthaleidae: Halotydeus, Chromotydaeus, Penthaleus,
Linopenthaleus and Linopenthalodes) are parasites of low-growing vascular plants
and, as their common name implies, are often found in soil, as are the large, long-
legged Eriorhynchidae (Eriorhynchus) in Australian forests (Qin and Halliday
1997). Earth mites appear to feed on soil mosses and algae during development and
only acquire cheliceral stylets long enough to puncture leaf cells as adults.

Indirect Plant Parasites

Mycorrhizal fungi are ‘digestive mutualists’ that colonise roots, infiltrate the rhizo-
sphere soil and sequester and transport phosphorus, nitrogen and other nutrients into
the roots of higher plants in return for carbohydrates (Lavelle et al. 1994). So, these
fungi are mutualistic extensions of the roots of plants. When mites and other soil
animals feed on mycorrhizae they are, in effect, also feeding on the plants. By remov-
ing senescent mycorrhizal biomass, dispersing the propagules of mycorrhizae or
attacking plant-parasitic fungi in the rhizosphere (Wiggins and Curl 1979), this graz-
ing may be beneficial to plants. Also, some mycorrhizae may preferentially form
their spores within the shells of dead oribatids mites (Rabatin and Rhodes 1982).
However, high densities of mycophagous mites, especially of sucking fungivores or
other fungal-feeding soil arthropods and nematodes, may reduce the effectiveness of
mycorrhizae and limit plant growth (Edwards and Stinner 1988; Lussenhop 1992;
Klironomos and Kendrick 1995).

The Worm-Eaters: Nematophages

Once long-neglected aspect of soil ecology is the relationship between nematodes


and mites. This neglect is not especially surprising, since the serious study of
free-living and plant-parasitic nematodes is an even younger discipline than aca-
rology. Historically (and bizarrely) the study of soil-inhabiting nematodes – plant
parasites, bacteriovores, fungivores and predators – is a discipline within plant
pathology, rather than zoology. As a result, few soil zoologists know much about
nematodes and, when plant pathologists look for natural enemies of nematodes,
they tend to concentrate on microbial diseases, nematode-trapping fungi and
predatory nematodes. However, understanding the population dynamics of both
soil mites and soil nematodes requires understanding that for many mites, espe-
cially Mesostigmata, nematodes are a preferred prey (Karg 1961, 1983; Muraoka
and Ishibashi 1976; Sharma 1971; Stirling 1991; Van de Bund 1972; Walter
1988b; Walter et al. 1987a; Walter and Ikonen 1989; Walter and Kaplan 1991;
Beaulieu and Walter 2007).
186 6 Mites in Soil and Litter Systems

Fig. 6.6 Soil Mesostigmata that take both small arthropods and nematodes as prey. (a) Ologamasid
with chelicerae extended. (b) Subcapitulum of zerconid with chelicerae partially withdrawn; (c)
Chelicera of Asca sp.; (Ascidae) (d) Chelicera of Lasioseius sp. (Blattisociidae) (SEMs by DE
Walter)

In addition to the disciplinary disjunction between soil acarology and nematology,


two mistaken assumptions have impeded understanding of the strong interaction
between mites and nematodes. First, soil-pore spaces can exist in air-filled or
water-filled states and mites have been assumed to be restricted to the former and
nematodes to the latter. Although some nematodes (e.g. dauer larvae with double
cuticles) are able to inch their way – and even to jump – through air-filled spaces,
most nematodes do best when free water is available and many require a film of
water for movement (Nicholas 1984). Even the thinnest film of water, however, may
be enough and mites readily extend their chelicerae into water to extract nematodes.
Soil mites probably do best in air-filled soil pores but they survive periodic flooding
and many are able to forage for nematodes, and even protozoans, within a water layer
(Walter and Kaplan 1990a). So, the faunal dichotomy between air and water in soil
pores is not as strong as has been assumed.
The second mistaken assumption is that nematodes have extensive size-based
refugia where they are safe from harvest by predatory mites. Nematodes are very
slender and soft-bodied, and they can enter very small soil-pore spaces unavailable
to the mostly broader and less flexible mites. However, many nematophagous mites
are elongate, flexible and/or capable of extruding their chelicerae considerable
distances (see Fig. 6.6) into small soil pores. Except in highly compacted soils, it is
likely that much of the pore space available to nematodes is also accessible to prob-
ing mite chelicerae and refugia must be small indeed to protect the worms. Most
food resources will not be restricted to those refugia, so nematodes must eventually
Feeding Guilds and Functional Groups 187

come out – and run a gauntlet of hungry mites. Nematodes are better protected when
they dehydrate during cryptobiosis but, even then, some sarcoptiform mites will eat
them like so many dried snacks (Walter et al. 1988).
Many, if not most, Mesostigmata that live in soil, decomposing vegetation,
compost, dung and carrion will feed on nematodes. Often these mites show their
most rapid development, lowest mortality and highest fecundity when fed nema-
todes (Walter et al. 1987a; Walter and Ikonen 1989). Species that prefer nematodes
or that feed only on them are especially common in the families Ascidae,
Eviphididae, Macrochelidae, Uropodidae and Zerconidae. Some genera of laelapid
mites that tend to be associated with ant nests (e.g. Laelaspis, Cosmolaelaps) also
seem to be specialised nematophages (Walter 1988a). More general predators that
readily feed on nematodes are found in the Laelapidae, Parasitidae, Rhodacaridae
and Ologamasidae.
Nematophagy in the Acariformes is less common but some Endeostigmata are
primarily nematophagous (e.g. Alycus and Alicorhagia) (Walter 1988b). A few
predatory Prostigmata will take nematodes, e.g. some Bdellidae (Pugh and King
1985) and Cunaxidae (Walter and Kaplan 1991). However, some of the most impor-
tant reports of significant levels of nematophagy involve the Tydeidae, mites often
thought to be primarily fungivorous (Santos and Whitford 1981; Santos et al. 1981).
In Chihuahuan Desert soils, tydeid mites have a keystone position – their elimina-
tion is associated with exponential increases in rhabditid nematodes which then
overgraze soil bacteria and reduce decomposition rates (Whitford 1996). Predation
by many oribatid mites on nematodes has been observed in the lab and confirmed in
the field using molecular methods (Heidemann et al. 2011). Nematophagy may be
common among other soil arthropods, both predators and grazers, but it has scarcely
been studied. However, many tardigrades, nematodes, fungi, protozoa and bacteria
eat nematodes (Sayre and Walter 1991; Stirling 1991).

Predators of Arthropods

Eating mites and springtails is the traditional role assigned to predatory Acari and
they carry it off well. Early derivative acariform mites graze or suck on microbes but
arthropod predation has arisen in a number of lineages, especially within the
Prostigmata. The snout mites in the Bdelloidea (Bdellidae, Cunaxidae) are entirely
predatory, as are the fast-moving, pale Rhagidiidae and the slower, heavily sclero-
tised yellow to green mites in the Labidostommatidae. Seven of the eight families in
the Anystina (Anystidae, Caeculidae, Teneriffiidae, Adamystidae, Paratydeidae,
Pseudocheylidae and Pomerantziidae) prey on arthropods, as do the red velvet mites
(Erythraeoidea, Trombidiioidea). Red velvet mites are often large and brightly
coloured and, like the closely related water mites (see Chap. 7), the larva is usually
a protelean parasite of an insect (although arachnids and vertebrates are also
attacked: see Chap. 10). One subfamily of cheyletid mites, the Cheyletinae, are
primarily predators of mites and other small arthropods, and often seem to prefer
188 6 Mites in Soil and Litter Systems

Fig. 6.7 Some Prostigmata that take primarily arthropod prey: (a, b) Cunaxa spp. (Cunaxidae) –
reflexed palps (arrow in a) act as a snap-trap for springtails and mites and sometimes have addi-
tional spines or bosses (arrow in b) to stun prey. (c) Bellidae – like Cunaxidae are called snout
mites because of their elongate chelicerae, but the palps are sensory, rather than raptorial. (d)
Rhagidiidae – fast moving predators that run down their prey and overcome them with their
enlarged chelicerae. (e) Labidostommatidae – heavily armour ambush predators of springtails,
mites and small insects. C chelicera, p palp (SEMs by DE Walter)

astigmatan prey. The Raphignathoidea consists of nine families of predatory mites


or presumed predatory mites. The well- studied taxa in the Stigmaeidae and
Eupalopsellidae are mostly predatory on mites and small insects but other families
are more enigmatic. Raphignathoidea are often major components of dryland soils
(e.g. dry grasslands, deserts, eucalypt woodlands) but little is known about their
biology (Fig. 6.7).
Mesostigmatans in the family Laelapidae are among the more aggressive preda-
tors of arthropods found in soil. Some of these species have proven useful for control
of pests in glasshouses and several species are now mass reared in commercially in
North America. The most commercially important of these predators are the hypoas-
pidines (i.e. Hypoaspis sensu lato), especially species in the genera (subgenera of
some authors) Gaeolaelaps (= Gaeolaelaps) and Stratiolaelaps. Thrips are vulnera-
ble to these mites when the penultimate instars enter the soil to pupate, fungus gnats
Feeding Guilds and Functional Groups 189

Fig. 6.8 Predation by various Mesostigmata species and populations against penultimate instar
Western Flower Thrips (WFT) in (a) open arenas and (b) soil microcosms. Arena tests are good
indicators of potentially useful predators, but soil tests are a more realistic assessment of what may
happen in the field or greenhouse (Image by DE Walter, data from Vänninen and Walter 2003)

are attacked in the larval and egg stages and bulb mites are attacked at all stages
(Barker 1969; Hamlen and Mead 1979; Lobbes and Schotten 1980; Gillespie and
Quiring 1990; Chambers et al. 1993; Wright and Chambers 1994; Lesna et al. 1996;
Ali et al. 1997). The level of attack against target prey can vary with the field popula-
tion of the predator (Fig. 6.8) and the method used to evaluate predation rates (e.g.
simple arenas vs soil microcosms), but a number of soil-inhabiting Mesostigmata
show promise as biological control agents (Vänninen and Walter 2003).
Most free-living Mesostigmata are predators, predator–scavengers or predator–
fungivores (e.g. Proctolaelaps, Protogamasellus) and only a few lineages appear to
be primarily fungivores (e.g. Ameroseiidae) (Walter and Proctor 1998a). Some
Uropodina can be reared on fungi (Woodring and Galbraith 1976; Hutu 1991) but
others are known to be predators of fly larvae and nematodes (O'Donnell and Axtell
190 6 Mites in Soil and Litter Systems

1965; O'Donnell and Nelson 1967). Many mesostigmatans that capture arthropods
also feed on nematodes and other soft-bodied invertebrates, and can be considered
general predators (Walter et al. 1988; Walter and Oliver 1990). The predatory role
of mites is important for understanding their influence in soil systems and especially
in peripheral soil-litter systems such as dung or carrion. So, in the next section we
will consider predatory behaviour in soil mites in greater detail.

Predation in the Soil

The dominant acarine soil predators are in the Mesostigmata, especially the subor-
ders Dermanyssina and Parasitina, and in numerous families in the Prostigmata.
Predatory prostigmatans are highly variable in their search behaviours but most
dermanyssine mesostigmatans are restless animals, constantly on the move, search-
ing through litter or soil pores for prey. This is reflected in the Greek-derived end-
ing of many generic names, e.g. -seius (‘to move to and fro, shake’) and -laelaps
(‘a hurricane’). Being constantly on the prowl means that they can readily locate
patches of sedentary or less mobile prey (e.g. nematodes, aphids, eggs, pupating
thrips) and this is one factor in their importance as commercial biocontrol agents.
In artificial soil habitats, Mesostigmata are especially useful biological control
agents even against active insects. For example, Stratiolaelaps scimitus (some-
times masquerading under the name Hypoaspis miles) can be used to control the
larvae of fungus gnats (Sciaridae) and in mushroom culture (Castilho et al. 2009;
Stephen and Schweizer 2009), and soil-inhabiting crambid caterpillars
(Duponchelia fovealis) in greenhouses (Messelink and Van Wensveen 2003);
Parasitus consanguineous will control both fungus gnats and the larger larvae of
the phorid fly Megaselia halterata (Szlendak and Lewandowski 2009).
Many mesostigmatans are extremely aggressive and will worry at the leg joint of
a much larger arthropod until they draw blood (Walter and Oliver 1990). It has even
been suggested that some of these animals will hunt in packs (Usher and Davis
1983). Group feeding on a larger worm or arthropod is common (Blaszak et al.
1990; Seeman and Walter 1997) but cooperative hunting is unlikely. Typically, the
initial attack is by a single mite and then either the thrashing of the prey attracts the
others (Fig. 6.9) or they stumble across the dying prey and commence feeding.

Cruise and Pursuit Predators

Mesostigmata characteristically search rapidly and pursue their prey. Similar behav-
iours also occur among the predatory Prostigmata. For example, whirligig mites
(Anystidae) are relatively large, often bright red and speedily spin across ground,
rocks and vegetation hunting for small insects and larger mites. Some cunaxid mites
are also constantly on the move. Coleoscirus simplex travels 20–23 cm per minute
Predation in the Soil 191

Fig. 6.9 Three Macrocheles


superbus (Mesostigmata:
Macrochelidae) share an
oligochaete worm (after
Blaszak et al. 1990)

at 25 °C (Walter and Kaplan 1991). Since Coleoscirus has a snout–vent length of


580 μm, this means that the mite runs almost 400 body lengths in a minute. In
human terms, this would be equivalent to sprinting 800 m every minute – for hours
on end. Prey are rapidly circled and subdued by a series of rapid jabs with the
chelicerae and finally grasped between the palps and chelicerae, elevated above the
substrate, and drained of their fluids. The crumpled cuticle is then discarded.

Ambush or Sit-and-Wait Predators

Coleoscirus rarely stops moving except to moult, mate, feed or lay an egg. However,
species of Neoscirula, another genus in the same subfamily and living in the same
habitat (Walter and Kaplan 1991), has a much more sedate style, often pausing to
rest and covering only about 1 cm per minute (about 30 body lengths). Even less
energetic are cunaxids in the subfamily Cunaxinae (e.g. species of Armascirus,
Cunaxa, Dactyloscirus and Rubriscirus). These animals spend most of the time
standing in one spot with their large, spiny palps re-curved over their bodies
(Fig. 6.10). When a springtail or mite walks past them, they pounce. The palps are
whipped forward and the victim is impaled on the snout-like chelicerae.
Ambush predation is common in the Prostigmata. For example, Caeculidae are
related to the whirligig mites discussed above but have a very different strategy for
ambushing springtails and small insects like bark lice (Coineau 1973). We have
observed an Australian representative of this group, Neocaeculus sp., from the
Wollemi region of New South Wales. This circular mite is covered with armoured
plates and knobby setae, and its first pair of legs are very stout and long with thick
192 6 Mites in Soil and Litter Systems

Fig. 6.10 Cunaxid mite feeding on a springtail. In ambush stance (and during feeding) the raptorial
palps are reflexed over the body like a set mouse trap. When a passing springtail stimulates the
elongate trichobothria on the prodorsum, the palps snap forward and implae the prey on the snout-
like chelicerae (Image by DE Walter & HC Proctor)

spines on their interior faces (hence the common name of rake-legged mites for this
family). It will stand for days in one spot with these legs elevated at a 45° angle to
the ground, completely relying on the movement of its prey to bring about an
encounter. When a springtail wanders under the upraised legs, they come crashing
down, scooping the victim into the hook-like chelicerae. A number of water mites
use similar behaviours (see Chap. 7).
A very unusual form of ambush behaviour (‘fatal flatulence’) has been reported
for nymphs of the trombiculid velvet mite Womersia strandtmanni (Huber 1978).
Springtails are attracted to the posterior of these mites, which they tap with their
antennae. In response, the mite elevates its rear end and secretes a drop of fluid. The
springtails imbibe the drop and then begin a zigzag run which ends when they fall
paralysed. A trombiculid may poison a series of springtails before encountering a
paralysed victim and feeding.

Saltatory Search

The amount of time and energy spent searching for prey can be viewed as a gradient
(O’Brien et al. 1990). At one end are the speed demons, like most mesostigmatans,
anystids and Coleoscirus. At the other end are the patient and deadly, like Neocaeculus.
However, most predatory prostigmatans and some mesostigmatans fall in between the
two extremes of constant movement or days of patient anticipation, and instead show a
series of starting and stopping behaviours called saltatory search (O’Brien et al. 1990).
Predation in the Soil 193

Many snout mites (Bdellidae and Cunaxidae) search an area by walking from
spot to spot, probing crevices with their antenna-like palps and elongate chelic-
erae – often looking much like red bags on stilts with a hypodermic needle at
one end. An exception in colour (purple) and cheliceral form (more like can
openers than needles), Cyta latirostris is a ‘snout mite’ that is able to prey on
well-armoured oribatid mites. When a beetle mite is encountered, Cyta squirts
out a strand of sticky silk to entangle its victim and then uses its short, stout
cheliceral hooks to pry open a hole in the mite's armour (Alberti 1973). Other
snout mites feed on soft-bodied mites or springtails, speedily encircling the
victim while entangling its legs and body with strands of silk (Alberti and
Ehrnsberger 1977; Walter et al. 1988). Most snout mites are mostly soft-bodied
and red to orange in colour, but Trachymolgus (= ‘rough hide’) purpureus is, as
the name suggests, purple and protected by pitted plates (see Fisher et al. 2011
for some spectacular images).
Mesostigmatans in the primitive uropodine genus Polyaspis and in the more
derived dermanyssine genus Veigaia are also rather deliberate in their searching,
and often pause in wait. Species in both genera have very elongate cheliceral digits
used to snap-trap their prey. Like many of the soil mites on the ambush side of the
gradient, these mesostigmatans are specialised predators of springtails.
Some mesostigmatans look more like they are placidly walking along a trout
stream with a pair of fishing rods than hunters. Species of Podocinum are slow-
moving, stately mites with very long front legs that they dangle in front of them.
When a springtail is encountered by the light touch of the long distal setae at the
tips of the tarsi, the mite pauses, and then quickly scoops their prey in towards
their chelicerae. The superficially similar Epicrius species carry this long-
legged fishing one step further. The long ventral setae on the tips of the front
legs have blobs of glue with which they ensnare their collembolan prey (Alberti
2010).

Constraints and Variations

Cruise predators are more likely to take sessile or clumped prey and, in general, they
will capture relatively smaller and younger prey compared to ambush predators
(Greene 1986). For example, nematodes are small and tend to occur in high density
patches. Nematodes are not eaten by cunaxid ambush predators but are taken by
cruise predators in the family, such as Coleoscirus and Neoscirula (Walter and
Kaplan 1991). Red velvet mites (Erythraeoidea, Trombidioidea, Anystidae) are
great wanderers and capable of rapid movement. Some species capture springtails
and insects (e.g. Young & Welbourn 1987); however, many specialise on non-motile,
clumped prey like the eggs of springtails or insects (Lipovsky 1954; Sharma et al.
1983; Wohltmann et al. 1996). Similarly, macrochelid mites rapidly search dung for
clusters of fly eggs (Krantz 1983).
194 6 Mites in Soil and Litter Systems

Ambush predators are often capable of capturing relatively large prey (Enders
1975). The energetics of ambush predation, however, demands that prey be highly
mobile and predictably present (Huey and Pianka 1981). Springtails are common
and usually abundant soil animals but many are fast moving and have the ability to
jump away from danger. The sudden pounce, raptorial limbs and squirting of silk
used by ambushing mites represent a variety of solutions to springtail escape
behaviour.
However, some cruising mites are efficient predators of springtails. Coleoscirus
is so fast that it simply runs over its prey, impaling small springtails before their
escape behaviour can be activated. In contrast, some cruising mesostigmatans
alter their behaviour when a relatively large prey item like a springtail is encoun-
tered. When Antennoseius janus comes upon a springtail, rapid movement stops
and the victim is stealthily approached. When close, the predator pounces, using
the antenniform first pair of legs, the second pair of legs and the gnathosoma to
form a basket-like arrangement that scoops the victim into the snapping chelic-
erae (Lindquist and Walter 1989).

Intraguild Predation

Predators are animals that prey on other animals, and we tend to think of these inter-
acting organisms as falling into two classes: predators and prey. However, the vic-
tim may itself be a predator and, if the two also compete for prey, then the ecological
and evolutionary implications of this interaction (intraguild predation) are quite
different from ‘normal’ predation (Polis et al. 1989). Intraguild predators not only
gain the immediate benefits of the energy and nutrients in the victim’s body; they
may also eliminate a potential predator of their young and may husband resources
by reducing the level of exploitation competition for other prey. Humans, of course,
are rather deliberate intraguild predators, although in Western societies the use of
the eliminated carnivores as trophies is more highly valued than their use as meat.
Other animals tend to be less deliberate in their hunting of potential competitors but
within-guild prey are often a high proportion of the harvest (Polis et al. 1981, 1989).
This is especially true of relatively large predatory mesostigmatans in soil food
webs (e.g. Walter and Oliver 1990).
Degree of prey specialisation, relative size and effectiveness of defence are the
three main arbiters of intraguild predation in mites. Specialised nematophages do
not attack other predatory mites but they will feed on predatory nematodes, thereby
reducing competition for prey (Walter et al. 1988; Walter and Ikonen 1989).
Uropodid mites are heavily sclerotised as adults and generally well protected from
predation. Those uropodids that feed on nematodes, worms and maggots are often
relatively large but do not attack even much smaller and weakly sclerotised mites.
Similarly, ambushing caeculid mites ignore many prey of seemingly suitable size
until a springtail or barklouse walks by (Coineau 1973). Species of rhagidiid mites
in the maritime Antarctic appear to be primarily predators of Paraisotoma and
Predation in the Soil 195

Fig. 6.11 Intraguild predation simply means predators that eat each other as well as a common
prey on a different trophic level (in this case rootknot nematodes). In this web, solid arrows indicate
that all stages of a predator are eaten and open arrows that only the small, lightly sclerotized
immatures are eaten (Image by DE Walter, data from Walter and Kaplan 1991; Walter and
Lindquist 1995; Walter and Proctor 1999)

Cryptopygus springtails and have not been shown to be intraguild predators.


However, remains of the predatory mite Rhagidia can represent a significant propor-
tion of the items detected in the guts of its competitor and intraguild predator,
Gamasellus racovitzai (Lister et al. 1987).
Many soil-inhabiting mesostigmatans have feeding behaviours similar to that of
G. racovitzai (see Fig. 6.2b), i.e. they are general predators that may attack any
arthropod or worm that they encounter, at least within a given size range, and the
success of the attack depends primarily on chance and the efficacy of any host
defence. For example, Pugh and King (1985) used field observations, supplemented
by protein electrophoresis and laboratory tests, to study the feeding habits of the
ologamasid Hydrogamasus salinus in the British intertidal zone. In the field, H. sali-
nus was observed preying on nematodes, oligochaetes, springtails and larval and
adult flies, and scavenging on damaged barnacles. Electrophoretic analysis sup-
ported the field observations and also found indications of intraguild predation on
another ologamasid mite. Other intertidal mesostigmatans and bdellid snout mites
had similar feeding habits, including occasional feeding on other predators (Pugh
and King 1985).
Intraguild predation can be intense enough to severely reduce the populations of
some predators. For example (see Fig. 6.11), in research glasshouses in Orlando,
196 6 Mites in Soil and Litter Systems

Florida, the ascid mite Lasioseius subterraneus colonises the soil around the roots
of potted plants infested by rootknot nematodes where it can reach densities of over
400 mites per litre of soil (Walter et al. 1993). Lasioseius is larger than most of the
other half-dozen species of predatory mites in this system and readily feeds on
them. However, when the larger and more aggressive laelapid mite, Gaeolaelaps
sp., colonises pots, Lasioseius populations decline precipitously (Walter and Lindquist
1995). In laboratory arenas, Gaeolaelaps sp. quickly captures and eats Lasioseius, and
presumably the same behaviour contributes to the rapid decline in Lasioseius popu-
lations in pots.
Although Gaeolaelaps sp. is the top predator in this system and can capture and
eat any of the other predatory mites, mycophagous mites, springtails and nema-
todes present, it is not immune to attack. Nymphal and larval Gaeolaelaps are
poorly sclerotised, relatively small and susceptible to predation by the cunaxid
snout mite Coleoscirus simplex (Walter and Kaplan 1991). Coleoscirus also feeds
on the immature stages and relatively small males of the second largest predator in
the system, L. subterraneus, but Lasioseius in turn is able to capture nymphs of
Coleoscirus (Fig. 6.11). Only active nymphs are captured. When preparing to
moult, immature Coleoscirus enter a crevice and spin a web around themselves
which provides good protection from other predators (including other Coleoscirus).
Because of their speed and size, adult Coleoscirus are immune to predation from
all except the largest predator, Gaeolaelaps, but smaller cunaxids (e.g. Neoscirula)
and mesostigmatans (e.g. Gamasellodes) in this system are constantly at risk of
being eaten by larger predators and by each other (Walter and Kaplan 1991; Walter
et al. 1993).

Cannibalism

Some scorpions are their own worst enemies (Polis et al. 1981) and this may be true
of some predatory mesostigmatans as well (Lindquist and Walter 1989). Within the
Phytoseiidae, cannibalistic interactions are typically adult females consuming eggs
or larvae (Schausberger 2003). In one species of phytoseiid mite, the rate of canni-
balism has been found to vary geographically (Croft and McMurtry 1971).
Generally, however, our observations indicate that predatory prostigmatans are
more likely to eat their own kind, including their own offspring, than are predatory
mesostigmatans.
The reasons for this dichotomy are not clear but may relate to the prevalence of
inbreeding and parthenogenesis in the Mesostigmata. When members of a popula-
tion are closely related, then a reluctance to eat one’s relatives makes long-term
evolutionary sense (Saito 1997). For example, adult females of Lasioseius dentatus
eat conspecific nymphs only when starved (Walter and Lindquist 1995). Since
L. dentatus is an all-female species that reproduces by thelytokous parthenogenesis
(see Chap. 5), in many populations a conspecific is likely to be part of an identical,
Avoiding Predation: Defences of Mites and Mite Prey 197

or at least very similar, genetic lineage (e.g. sister, daughter, daughter of a sister).
However, its bisexual relative, Lasioseius subterraneus, will eat conspecific nymphs
even when alternative prey is available (Walter and Lindquist 1995). Females of at
least one species of phytoseiid are able to recognize their own offspring and prefer-
entially eat the young of other females (Schausberger 2003). Jerking behaviour of
young phytoseiids when contacted is thought to potentially be a ‘please-don’t-eat-me’
signal to conspecifics (Blackwood et al. 2001).
Bisexual species of Lasioseius are probably parahaploid (Norton et al. 1993) and
therefore the only functional chromosomes in males are maternal (see Chap. 5).
Migration in L. subterraneus is by inseminated adult females, which produce broods
with highly skewed sex ratios (80 % female). Therefore, sib-matings and a high
level of inbreeding are probable characteristics of populations of this mite. Although
they are cannibalistic on conspecific nymphs, attack rates are relatively low (13 %)
compared to other bisexual ascid mites with more even sex ratios and probable
diplo-diploid reproductive modes (Walter and Lindquist 1995). For example, in
laboratory populations of Antennoseius janus (57 % female), both adults and
nymphs are aggressively cannibalistic (Lindquist and Walter 1989). Other probable
diplo-diploid mesostigmatans with approximately 50:50 sex ratios also readily
attack and eat their immatures (Walter, unpublished). Cannibalism among adult
mesostigmatans, however, is rare – probably because of the protection offered by
sclerotised plates and equivalent size.

Avoiding Predation: Defences of Mites and Mite Prey

Soil and litter often contain both high densities and high diversities of predators,
and the predation pressure on arthropods and nematodes often must be intense.
Not surprisingly, a variety of ways of avoiding being eaten have evolved. Escape
via large size is a ‘defence’ common to both arthropods and nematodes that are
stalked by predatory mites. The thrashing of large nematodes can dislodge and
fling off attacking mites (Walter and Ikonen 1989); however, most soil nematodes
are relatively small, soft-bodied and the vermiform stages have no obvious
defence. Adult female rootknot and cyst nematodes become embedded in root tis-
sue, where their bodies swell and become covered with a gelatinous cap into
which they lay their eggs. This stage is well protected from predators, although
some mites do attack the egg masses (Inserra and Davis 1983). Many rhabiditid
nematodes produce a resistant dauer larva by retaining, and sometimes modify-
ing, the skin of the previous moult (Nicholas 1984). This stage is more resistant to
environmental stress and the thicker cuticle may provide protection against preda-
tion. Dauer larvae of entomogenous nematodes are able to avoid predatory mites
by climbing or jumping onto their backs. Although they rarely infect mites, they
do use them as phoretic carriers and are capable of leaving the mites to attack
insect larvae (Epsky et al. 1988).
198 6 Mites in Soil and Litter Systems

Jumping

Rapidly leaving the vicinity of a larger, potentially dangerous animal is always a


good idea and many mites can produce quick spurts of speed, either forward or
backward. Jumping, the ultimate in rapid egress, is a common escape mechanism of
collembolans (springtails), especially those that live on the soil surface and in the
litter layer. However, jumping also has evolved independently a number of times in
soil-inhabiting Acari. Species of Nanorchestes (Endeostigmata) are tiny (<300 μm),
spherical mites that bound away when approached by a predator. Species of Eupodes
(Prostigmata) are very active runners, both forward and backward, that are able to
leap when close-pressed. Species of Terpnacarus (Endeostigmata) graze on fungi in
dry soils and also readily jump, as do some Scutacaridae (Ebermann 1995). Oribatids
in the genera Indotritia (Schubart 1967) and Zetorchestes (Krantz 1978; Krisper
1990, 1991) are also able to jump. Microtrombidiid larvae have highly modified
tarsi on their hind legs and are suspected of being able to jump, both to escape and
to latch onto passing insects, as do some hypopal astigmatans. Jumping parasitiform
mites, however, are rare. To date, jumping in Mesostigmata is known from only two
genera of beetle associates, Micromegistus and Saltiseius (Walter 2000). The latter
is sometimes found on passalid beetles and, as in hopping hypopi, jumping may
function as both a defence and to facilitate migration.
The jumping mechanisms used by these mites are only partially understood.
Species of Eupodes have characteristically enlarged hind femora that presumably
give them their lift. Zetorchestes have a cuticular catapult mechanism at the base of
their fourth pair of legs that allows them to spring away (Krisper 1990, 1991) in a
manner analogous to springtails. However, the box mite Indotritia cf heterotrichia
uses hydraulic extension of all four legs to achieve extraordinarily rapid takeoff
(0.5 ms), a leap it is able to repeat two or three times in a row (Wauthy et al. 1997).

Chemical Defence

Springtails are classic jumpers but some species rely on a coating of easily shed
scales or chemical defences to foil predators. For example, the springless springtail
Tullbergia granulata produces small droplets on its body when attacked. Predators
encounter these drops when they attempt to bite the springtails – and then quickly
back away and can be seen vigorously cleaning their mouthparts. Subsequent
attacks are often successful, suggesting that the chemical reserves are quickly
depleted (Walter et al. 1988). Many astigmatans produce alarm pheromones that
may have a deterrent effect on predators (Matsumoto et al. 1979; Kuwahara 1990).
Some metastriate ticks produce gland secretions that protect them from ants (Yoder
1993) and some oribatid mites produce a sticky substance that deters ant attacks
(Masuko 1994). Nymphs of Allothyrus mites (Holothyrida: Allothyridae) produce
droplets from gland openings when annoyed and these presumably represent defen-
sive chemicals (Walter and Proctor 1998a).
Avoiding Predation: Defences of Mites and Mite Prey 199

Oribatid mites can be divided into two groups based on the presence of a pair
of opisthonotal oil glands that open to the dorsolateral surface of the notogaster.
All but the most ‘primitive’ groups of Oribatida, the Enarthronotides and
Palaeosomatides (both known from Devonian fossils), have these glands. In glandu-
late oribatids these openings are usually modest, but sometimes lead to a pair of
large spouts that may have an associated ‘trigger seta’. If the cuticle is lightly sclero-
tized, a well developed reservoir may be seen under the glands, often a purplish
colour (e.g. Astigmata and juveniles of traditional Oribatida).
A large number of potentially noxious aromatic and chain hydrocarbons and
alkaloids are associated with the opisthonotal glands. Some of these chemicals func-
tion as sex pheromones (Mizoguchi et al. 2005), but most appear to act as alarm
pheromones, especially in the Astigmata (Shimizu et al. 2004). Astigmata tend to
exploit concentrated resources and achieve high local population densities, probably
often founded by a few related colonists, so alarm pheromones may be particularly
useful to them. Soil oribatids, although sometimes concentrated on a resource (e.g. a
fungal sporocarp), tend to have more dispersed lives and may be less likely to benefit
from an alarm signal. However, a similar alarm function has been demonstrated in
some of the non-astigmatic Oribatida (Shimano et al. 2002; Raspotnig 2006). The
soil, however, has a diversity of aggressive predators that will attack oribatid mites
including mites, spiders, insects, and small vertebrates (see Acarophagy below).
Recent studies (Heethoff et al. 2011 and its associated videos, Heethoff and Raspotnig
2012) have convincingly demonstrated that the opisthonotal gland secretions also act
as effective deterrents to predatory beetles.

Autotomy, Armour, Hairs, Dirt and Thanatosis

When a predator attacks an arthropod, it must first grab hold of its potential victim,
and protruding from most arthropods are a variety of tempting handles – the legs.
Opilioacarans are able to shed legs (autotomy) to escape predators and subadults
are able to regrow them at later moults (Walter and Proctor 1998a). Many long-
legged prostigmatans (e.g. Eupodidae) are able to lose legs and survive, although
they appear unable to regenerate the missing limbs.
Another response to a looming predator is to curl up the legs against the body
and play dead (thanatosis). For example, adult Allothyrus rely primarily on heavy
armour and thanatosis for protection from arthropod predators (Walter and Proctor
1998a). When attacked by ants or beetles, these relatively large mites (2–3 mm in
length) pull the legs against the body, fall over and remain motionless. After a few
snaps at the mite, the aggressor (possibly also influenced by mite secretions) usually
gives up and leaves.
Armour is by far the most common defence mechanism in soil mites and it is often
combined with thanatosis and curling of legs against the body or withdrawing append-
ages into protected recesses. Adults of many major taxa are encased in thick, sclero-
tised plates and shields. Adult Mesostigmata usually have one or two shields covering
200 6 Mites in Soil and Litter Systems

most of the dorsum and several plates covering the venter to varying extents. Adult
males are often completely encased in armour but females must be able to expand
their opisthosomas to accommodate the large eggs and generally have extensive areas
of soft cuticle laterally and ventrally. However, both sexes of most uropodid mites are
completely encased in armour. Many uropodids also have recessed areas in the shields
into which they can withdraw their legs (pedofossae – ‘leg ditches’, see Plate 29 in
Eisenbeis and Wichard 1987 for an excellent picture), and are often called turtle mites
(the same common name has been applied to the Scutacaridae). Adult holothyrans are
also completely encased in armour, as are adult Labidostommatidae and Penthalodidae
in the Prostigmata. Other prostigmatans (e.g. Adamystidae, Caeculidae, Stigmaeidae)
have extensive development of sclerotised shields.
Many oribatid mites have thick, waxy ceroteguments that may supplement their
armour. In Australian rainforest soils, some species of Ologamasidae, Uropodidae
and Oribatida have sticky ceroteguments that accumulate thick encrustations of
strongly bound soil particles. These ‘dirt’ mites (e.g. Fig. 6.12b) are typically slow-
moving and stop when touched. The soil may simply increase the effectiveness of
the armour coating but it may also act as a kind of camouflage by hiding the taste
or feel of the mite from its predators. This crypsis may also be used offensively.
In Queensland, Oriflammella lutulenta (Ologamasidae – see Fig. 3.13a, b) com-
pletely encases itself in a fine coat of soil particles soon after a moult – even the leg
segments are covered with soil. The dorsal tactile setae extend above the dirt armour
on long pedicles (Halliday 2008). As with other Ologamasidae these mites are pred-
ators of arthropods and nematodes, but slow-moving ambush predators that easily
catch fast-moving springtails (Beaulieu and Walter 2007 – as ‘new genus near
Gamasellus’). Some oribatids with sticky cerotegument accumulate a cap of detri-
tus (see Plate 34 in Eisenbeis and Wichard 1987 for excellent pictures) and euphe-
rederm oribatids carry ‘scalps’, i.e. the top part of the exuviae from previous moults.
Early derivative oribatids have sclerotised, leathery or mineralised plates protecting
their bodies, and some (e.g. Cosmochthonius, Beklemishevia) have hypertrophied
setae (Fig. 6.12c) that can be erected to physically keep a predator away (Walter
et al. 1988). Several lineages have independently developed all-encasing body
armour. However, the legs of these mites, although themselves armoured, are weak
points in their defence. A solution to these Achilles’ heels has been derived indepen-
dently by several lineages of ‘box’ mites with ptychoid (‘folded’) arrangements
between the anterior and posterior body regions (Fig. 6.12a, b). For a box mite to
close itself off from predators, its outer ‘shell’ must be solid and the base of the legs
must be articulated with surrounding sclerotized plates and capable of deformation
so that the legs can be folded internally and the volume thus displaced accommo-
dated (Schmelzle et al. 2009). Box mites look like strange Swiss army knives with
the legs protruding blade-like from a cavity between an anterior carapace and a
posterior, heavily mineralised body (Norton and Behan-Pelletier 1991). When
attacked, box mites withdraw their legs into the cavity and seal it off by lowering the
carapace, thus presenting a smooth, featureless ball to a predator. Rolling into a ball,
unfortunately, is no protection from politicians. The late Australian $50 paper note
carried a representation of a box mite, strategically placed below a microscope, on
Avoiding Predation: Defences of Mites and Mite Prey 201

Fig. 6.12 Armoured mites don’t just rely on a hard cuticle for protection: most also have protec-
tive ledges to protect their limbs (usually the weakest spot in their defences) or other more elabo-
rate defences. Box mites (a, b) have soft cuticle around the bases of the legs and are able to
withdraw their legs into their body and close the aspis over the opening protecting both legs and
mouthparts. This ability to jacknife their legs (or close the box) is called ptychoidy. Some species
(b) have the additional protection of gluing a protective covering of soil over their bodies. Species
of Cosmochthonius (c), however, have rigid leg insertions – but long, plumose erectile setae that
provides a pincushion of protection when attacked (but the setae relax when they need to move
through soil pores). SEMs by DE Walter

the side featuring Clunies Ross (presumably to represent his interest in mites as
intermediate hosts of sheep cestodes). Perhaps the only currency in history to feature
a mite, alas, neither scientist nor mite has survived the conversion to plastic bills.
The so-called ‘higher oribatids’ include the majority of the more than ten thou-
sand described species. Adults are almost completely protected by a cap-like shell
(notogaster) immovably attached to a prodorsal shield and a solid ventral plate
underneath. A narrow band of soft cuticle allows some expansion during feeding or
egg production. Higher oribatids also have tecta (tectum, tecta – from the Latin for
‘roof’) that protect their legs. Most tecta are simple shelves or ridges under which
legs can be withdrawn, but some tecta are wing-like (pteromorphs) and some
pteromorphs are hinged and connected to the notogaster by large muscles that can
202 6 Mites in Soil and Litter Systems

be pulled down to cover the legs. Oribatids, and other similarly well-defended
mites, are often called beetle or armoured mites in the literature (‘panzer’ mites in
Russian). For consistency, we will use beetle mites as the common name for
Oribatida (less the Astigmata) and armoured mites for any mites with well-developed
sclerotised or mineralised plating.
Specialised predatory mites such as the bdellid mite Cyta latirostris (Prostigmata)
are able to penetrate the cuticle of oribatid mites but most other predators are
thought to be limited to feeding on the more lightly sclerotised immature stages.
Certainly, feeding on adult oribatids is a tough proposition. Predatory mesostig-
matans can be observed attempting to feed on adult oribatids but usually give up
after a few minutes. However, some ologamasid mites are able to feed on oribatids,
as are adult Labidostommatidae. The smaller phthiracaroid, oppioid and haploze-
toid species that are relatively lightly sclerotised seem to be most at risk but even
large galumnoids are common victims. However, most armoured mites that end up
as food do so for animals larger than most mites.

Acarophagy: Mites as Food for Larger Animals

Pieces of mites can be seen in the gut contents of centipedes, scolopendrellid sym-
phylans and campodeid diplurans (Walter et al. 1989; Walter and Proctor 1999), and
many other common soil-litter predators, e.g. spiders, pseudoscorpions, opilionids,
ground beetles, rove beetles and ants, also are likely to include a high proportion of
mites in their diets (see Eisenbeis and Wichard 1987). Even the very unusual glacier-
inhabiting Northern Rock Crawlers (Grylloblatta campodeiformis, Grylloblattodea)
eat oribatid mites (Pritchard and Scholefield 1978). Unique-headed bugs
(Enicocephalidae) and larval dusky wings (Neuroptera, Coniopterygidae) take weakly
armoured and soft-bodied mites (Molleman and Walter 2001). Other small, predatory
insects that inhabit litter should also be considered potential acarophages. Although
these feeding relationships have rarely been studied, the abundance of mites in soil
and litter suggest that they should be important prey. Certainly, ants readily eat larger
mites. For example, ticks of all stages, including engorged females, are captured and
killed by ants (Wilkinson 1970; Butler et al. 1979; Dawes-Gromadzki and Bull 1997).
The dacetine ant Pyramica mazu feeds on predatory soil Mesostigmata by grabbing
them by the first legs or chelicerae before stinging them and carrying them off to feed
their larvae (Masuko 2009). While many predators take lightly armoured or soft-bodied
species or stages, the more heavily armoured mites are not immune.

Eating Armoured Mites

Although armour is the most common form of defence in soil mites, it is not neces-
sarily impregnable. Some groups of small insects appear to specialise on well
sclerotised mites. For example, armoured mites (especially Oribatida, Uropodida)
Acarophagy: Mites as Food for Larger Animals 203

Fig. 6.13 Remains of armoured mitmites (Uropodidae on left, Galumnidae on right) that have
fallen prey to ant-like stone beetles (SEMs by DE Walter)

are reported as prey of beetles in the Staphylinidae subfamilies Pselaphinae (Park


1947) and the Ant-like Stone Beetles of the Scydmaeninae (Schuster 1966; Schmid
1988), and Ptiliidae (Riha 1951; Cancela da Fonseca 1975). Schmid (1988) found
two basic types of attack in adult ant-like stone beetles. In the can-opener technique,
beetles capture mites, turn them over, hold them with their legs and use their man-
dibles to puncture the anal valves and ream open the ventral shield. Other beetles
hold their prey with labial suckers and use their mandibles to plane and saw through
the dorsal cuticle of the mite. Beetle larvae curl around their prey to subdue them
but use techniques similar to adults to puncture their cuticle. Each type of armoured
mite requires specific capturing and handling methods, which are reflected in vari-
ous modifications of the mouthparts and legs of the specialists. For example, the
planer–cutters, can capture only oribatid mites with smooth cuticles: those with a
thick, waxy cerotegument or rough surface cuticle cannot be held by labial suckers.
Some beetle species are generalists, however, and require relatively long handling
times for some armoured prey (up to 4 h).
The ant-like stone beetles beetles in Australian rainforests that have been inves-
tigated to date are more general in their prey preferences (Walter and Proctor 1999;
Molleman and Walter 2001). Individual beetles often capture a variety of armoured
adult oribatid, uropodid and mesostigmatic mites (Fig. 6.13). Modified hairs on the
tarsi of the beetles allow them to hold the mites as they flip them over, prune off the
legs and puncture the body through the gnathosoma (often forcing the chelicerae
and hypostome deep into the body). Although many different families of oribatid
mites are attacked, those in the Oppioidea are the most common victims. The pri-
mary weakness of these mites is the long legs that are not able to be withdrawn into
protected recesses of the body. So, even the well-armoured species are easily cap-
tured and eventually penetrated.
204 6 Mites in Soil and Litter Systems

Based on observations of live extractions from Australian rainforests, ants,


especially the smaller species, are important mortality agents of soil mites (Walter
and Proctor 1999). Ants often attack, maim or kill the larger mites in live litter
extractions, but ignore those that are relatively small. Two small (2–3 mm) species
in the ant genus Myrmecina take a high proportion of armoured mite prey including
oribatids and mesostigmatans (Masuko 1994). Mites are captured by workers and
carried back to the nest, where they adopt a curled posture as they manipulate the
mite with their legs and mandibles before finally tearing a hole through the anterior
portion of the ventral shield to one side or the other of the midline. The mite, resem-
bling an opened can of food, is then given to an ant larva whose head fits neatly into
the opening. Larvae of a species of Adelomyrmex in the neotropics have similar
elongate heads and also appear to feed on armoured mites (Masuko 1994). Some of
the smaller Neotropical species of Pheidole also appear to have a predilection for
oribatid mites and no difficulty in smashing through their armour (Wilson 2005).

Vertebrates That Eat Mites

Oribatid mites are intermediate hosts of anoplocephalid tapeworms (Denegri 1993)


because they are able to ingest small packets of food. Although fungal spores and
hyphae are more common in gut contents, small invertebrates also show up – many
oribatids are facultative predators of soft-bodied invertebrates such as nematodes
and tardigrades, and even slow moving or dead mites and springtails are eaten on
occasion. When the invertebrate swallowed by an oribatid is the egg of a tapeworm,
and the egg develops into the infective cysticercoid stage in the mite, the mite
becomes an intermediate host (Schuster and Cootzee 2012). All that remains is for
the mite with its internal passenger to get into the definitive host’s gut. Since many
oribatids that live in open habitats will clamber up on the grasses, low herbs, and
shrubs to graze on fungi, the stage is set. Animals that graze or browse vegetation
do not stop to flick off the mites, but ingest the lot, and mammals – from rabbits to
elephants (McAloon 2004) – may get the tapeworm along with the mite. Similarly,
grooming behaviour by vertebrates usually consists of licking or pecking off dirt
and itchy bits. Since mites are common in most habitats, they must often meet with
grooming accidents. Anyone who has paid close attention to their morning bowl of
cereal, fresh fruit or dinner salad would have to come to the somewhat disturbing
conclusion that we also eat mites more than we might like (Skubała et al. 2006).
Some vertebrates are exceptional in that they seem to like to eat mites. Oxpeckers
or Tickbirds (Buphagus spp.) are one such example – they pick blood-filled ticks
and other ectoparasites off the backs of large mammals in Sub-Saharan Africa.
Although they may be more blood-thirsty than previously thought (Weeks 2000), a
large part of their diet is ticks. Similarly, the Black-billed Magpie (Pica hudsonia)
will feed on the large female Winter Ticks (Dermacentor albipictus) on Moose
(Alces alces) and cache others in the litter under the snow where, if the birds forget
the location, the ticks’ survival may be enhanced (Samuel 2004). Small birds that
Acarophagy: Mites as Food for Larger Animals 205

forage on the ground eat at least some of the larger mites encountered (e.g. Kristin
1993; Savage et al. 2010), but the difficulty of getting gut contents from birds
without harming them seems to have limited the number of studies. Small, ground-
dwelling lizards also eat mites (Valdecantos et al. 2012), but apparently not as commonly
as the small frogs with which they co-occur (Lima et al. 2000). Amphibians, the
mite-gluttons of the vertebrate world, make up for any distaste for mites among lit-
ter dwelling vertebrates. Frogs (Pengilley 1971; Simon and Toft 1991; Donnelly
1991; Stewart and Woolbright 1996), toads (Flowers and Graves 1995) and sala-
manders (Norton and MacNamara 1976; Maiorana 1978; Lynch 1985; Utzeri et al.
2004; Walton and Steckler 2005; Walton et al. 2006; Mead and Boback 2006) all
feast on mites. Commonly reported patterns among litter amphibians are that mites
are the most abundant prey in amphibian stomach contents or second only to ants.
Soil-inhabiting mites commonly approach densities of a million individuals in a
square metre of forest floor. Although many of these may be too small to attract the
attention of animals hunting by sight, others will be within the size ranges taken by
small vertebrates foraging on the ground, low vegetation or tree trunks. Lungless
salamanders (Plethodontidae) are partly or fully terrestrial, often small enough that
mites are appropriate-sized prey, and often have mites in their gut contents (Norton
and MacNamara 1976; Lynch 1985). For example, Maiorana (1978) found that
oribatid mites accounted for about 27 % of the potential prey of California Slender
Salamanders (Batrachoceps attenuates) living under logs. Since oribatid mites
accounted for 35 % of the stomach contents of these salamanders, a slight prefer-
ence for oribatids mites. However, laboratory tests showed that the salamanders
dropped oribatid mites from their diet more quickly than they did soft-bodied smin-
thurid springtails of similar body size (Maiorana 1978). Presumably, salamanders
balance their preferences with what is available and with dietary requirements.
Oribatid mites often have high levels of minerals such as calcium and phosphorus
(Seastedt 1984) that could be important to small vertebrates. In any case, the leaf-
litter inhabiting salamander Plethodon cinereus can both positively affect the num-
ber of mites (Walton and Steckler 2005), presumably by reducing the number of
larger invertebrate predators, and also reduce the density of oribatid mites in micro-
cosms by about 2/3rds (Walton et al. 2006). Since these salamanders prefer larger
over smaller prey, their effects on mites are likely to vary with the availability of
larger prey. Similarly, the Eastern Newt (Notophthalmus viridescens, Salamandridae)
also seems to prefer the larger litter mites that it encounters (Norton and MacNamara
1976). In any case, although armour is obviously little protection against being
swallowed whole by a much larger animal, it does resist digestion and allows these
mites to be readily identified by researchers where more soft-bodied species may
have quickly disappeared.
Small frogs are even more likely than salamanders to feed on mites. For example,
Pengilley (1971) found that four of five species of frogs <35 mm long in the southern
highlands of New South Wales had fed on mites (4–100 % of the gut contents). All
feeding subadults, but only 54 % of feeding adults, of the Australian Corroboree
Frog (Pseudophryne corroboree Moore) had eaten mites. Similarly, all 109 females
of the Costa Rican Strawberry Poison Frog (Dendrobates pumilio) sampled by
206 6 Mites in Soil and Litter Systems

Donnelly (1991) contained mites, which accounted for 38 % of their diet. In Hawai’i,
the introduced Coqui (Eleutherodactylus coqui) has been associated with the reduc-
tion in the number of litter-inhabiting mites by more than one-third (Choi and Beard
2012). Why small anurans are so fond of armoured mites, however, is less easy to
understand. Because of their small size, mites have a very high surface-to-volume
ratio, i.e. they have more shell and less meat than do larger arthropods. Each feeding
event takes time and energy, but brings relatively less reward for each armoured
mite than for similar sized soft-bodied prey. However, these small and often brightly
coloured frogs may have a special reason for preferring oribatid mites.

Poison Frogs and Cleptotoxins

Poison Dart Frogs (Dendrobatidae) are extraordinarily colourful – and poisonous!


As a general rule, brightly coloured animals (including insects and mites) are not
good things to eat and often it is best to leave them entirely alone. Advertising a
poisonous, venomous, or otherwise obnoxious nature with bright colours is called
aposematism (loosely translated = ‘go-away-from signal’). However, Poison-Dart
Frogs don’t make their own poisons, but ‘steal’ them from what they eat (i.e. clep-
totoxins) – mostly oribatid mites, ants and millipedes (Toft 1995; Saporito et al.
2009, 2011; Raspotnig et al. 2011). Even within the Dendrobatidae, aposematic
colouration and its concomitant toxicity have evolved several times (Santos et al.
2003) and at least three other lineages of poisonous frogs are known (Saporito et al.
2011). Small frogs (and most of these frogs are less than 6 cm long as full grown
adults) living in the litter layer tend to eat mites at least in the proportion that they
encounter them (Toft 1995). Apparently, they have not only evolved to deal with the
toxins with which these mites defend themselves (Heethoff & Raspotnig 2011), but
to use them to their own advantage. One wonders if terrestrial salamanders, often
brightly coloured and usually with noxious skin secretions (Brodie et al. 1991), also
sequester toxins from their prey.

Body Size Patterns

We must take it for granted that a large part of the mite fauna of the world will remain
unsampled, unnamed, and unclassified [not to mention unwept, unhonoured, and unsung]
for decades to come. (Earnst Mayr 1969, interpolated by Robert May 1978)

G. Evelyn Hutchinson and his students were responsible for generating many
hypotheses to explain general patterns in community ecology from trophic levels to
body size differences among congeners (Hutchinson 1965). Two of Hutchinson’s
ideas about the importance of body size, both published in The American Naturalist
in 1959, are especially relevant to soil biodiversity. First, Hutchinson and MacArthur
(1959) suggested that an animal’s perception of the habitats in an area was basically
Body Size Patterns 207

two-dimensional and size dependent. The number of available habitats, and species
occupying those habitats, should climb rapidly from the smallest size class to the
modal class, and then decline less rapidly with increasing size. At the large end of
the size spectrum, the number of species (S) should decrease with the characteristic
length (L) of the animal (S ~ L−2). This model was the basis of May’s (1978) influ-
ential speculation that insects were highly diverse because they were small and that
the diversity–size shortfall below the modal size class was due to taxonomic igno-
rance of very small species such as mites.
May’s speculation, which ended a symposium on insect diversity, has been
honoured with a pyramid of publications on the relationships between size and
diversity. Although those who try to climb that monument may be in for a very
rapid descent, more than 30 years later the quest for the perfect slope continues
(e.g., Yamanaka et al. 2012). However, Hutchinson’s second paper (Hutchinson
1959), his presidential address to the American Society of Naturalists, has been
treated with anything but reverence (e.g. Simberloff and Boecklen 1981).
Although much challenged, Hutchinson’s basic observation is all too commonly
true. He noted that when two closely related species, typically congeners, are
found living together in the same habitat, one species is usually twice the mass
of the other. If the linear dimensions of the animals are considered, either body
length or the length of an appendage used in feeding, the ratio of the larger to the
smaller species is typically about 1.3 (range = 1.1–1.4, mean = 1.28), which in
animals of similar body shape implies a two-fold difference in body mass
(Hutchinson 1959, 1965).
Hutchinsonian size differences are a common, but difficult to explain, phenom-
enon. Although Hutchinson interpreted the size differences under the prevailing
dogma of character displacement through competition, this seems inadequate. For
example, Hutchinsonian size differences are generally present between instars in
arthropods (i.e. Dyar’s Constant) and among series of musical instruments and other
objects of human manufacture (Horn and May 1977; Maiorana 1978). Eadie et al.
(1987) have suggested that Hutchinson’s rule is an artefact of arbitrary sampling
from animals or objects with a log normal size distribution and a low variance, and
that no biological cause needs to be invoked.
Body size displacement conforming to Hutchinson’s rule is common among
some assemblages of soil mites (Fig. 6.14). For example, Walter and Norton (1985)
analysed the size distribution of the oribatid mites in Pinus radiata litter at three
sites in California. When 46 species at one of the sites were arranged in a size
sequence, the ratio of the larger to the smaller mite ranged from 1.00 to 1.14, with
83 % of the values falling below 1.05. Therefore, as an assemblage, these 46 species
filled the available size space with no sign of regular two-fold displacement in mass.
However, 13 congeneric species pairs occurred at the three sites and their within
genus body size ratios ranged from 1.20 to 1.39 and averaged 1.28, exactly as
observed by Hutchinson (1959). Whatever the cause, when congeneric pairs of orib-
atid mite species are sympatric, the adult female of one tends to be twice the mass
of the other and the doubling of mass between instars during ontogeny suggests that
these size differences are maintained among equivalent life stages.
208 6 Mites in Soil and Litter Systems

Fig. 6.14 Range of body lengths for four species of oribatid mites in the genus Scheloribates
sympatric in the litter of Monterrey Pine (Pinus radiata). Closely related soil mites that occupy the
same habitat tend to a mean difference in length of 30 %, i.e. each larger species is twice the mass
of the next smaller (Image by DE Walter; see Walter & Norton 1984 for details)

Sensitivity and Diversity: Soil Mites as Environmental


Indicators

The small size of most soil animals and the difficulty of observing rhizosphere
interactions in situ have impeded the understanding of their importance in soil
systems. Most studies have been restricted to comparing the abundances and
diversities of soil mites to rates of decomposition and nutrient cycling, the effects
of disturbance, vegetational cover and the concept of sustainable agriculture (see
Anderson and MacFayden 1976; Wallwork 1976; Swift et al. 1979; Petersen and
Luxton 1982; Seastedt 1984; Moore et al. 1988; Behan-Pelletier et al. 1985;
Hendricks et al. 1990; Wolters 1991; Wardle 2006). In most analyses, counts at the
subordinal level are the limits to resolution but a few studies have attempted to tease
out finer taxonomic patterns.
Also, most studies have concentrated on the easily sampled upper 5–10 cm of
soil, ignoring both vertical migrations by the fauna in the upper soils and the deeper
soil fauna. In dry soils, or those exposed to environmental stress, commuting to a
deeper soil level may be a means of avoiding desiccation. Soil compaction and
vegetation removal by cattle can cause a decrease in small oribatids that usually
avoid drought by moving deeper into the soil profile. Larger drought-intolerant
oribatid species that are less able to move into the soil profile can be eliminated by
Sensitivity and Diversity: Soil Mites as Environmental Indicators 209

even a single cutting in grassland (Siepel 1996). Soil mite diversity also may be
negatively affected by burning (Yeates and Lee 1997; Camann et al. 2008;
Malmstrom et al. 2008), clear-felling (Abbott et al. 1980; Lindo and Visser 2004),
fertilization practices (Cao et al. 2011) and tillage regimes in agricultural systems
(Kolesnikov 2010).
Mites in freshwater have long been recognised as useful bioindicators (see Chap. 7)
but until recently less so in soil systems. Predatory soil mites used in biological
control are sometimes used as specific indicators of toxic effects of chemicals used
in plant protection (e.g. Smit et al. 2012). Taxa that are primarily predatory, e.g.
Mesostigmata, are sometimes assessed at greater or lesser degrees of taxonomic.
Other studies have shown species-specific responses by mesostigmatans to soil
compaction (Heisler 1995) and pesticide contamination (Koehler 1992). Or to
Mesostigmata as a group being adversely affected by pentachlorophenol (PCP) pol-
lution (Salminen and Haimi 1996).
However, it is oribatid mites that seem especially useful indicators (Lebrun and
van Straalen 1995; Stamou and Argyropoulou 1995; Shimano 2011). Some reasons
for this preference are that adult oribatids are relatively long-lived and populations
tend to be stable in the absence of disturbance. Also, adult oribatids are relatively
easy to sort into morphospecies and species-level identifications are often possible.
This is especially important, because the reaction to some pollutants such as heavy
metals may be species-specific and metal-specific (Skubała and Zaleski 2012). For
example, the oribatid mites Rhysotritia duplicata and Nothrus silvestris have been
shown to quickly accumulate and retain heavy metals (Lugwig et al. 1991). Since
these mites have relatively long life cycles, they may prove useful indicators of
heavy metal pollution events. In contrast, PCP pollution had little effect on large
oribatids in microcosm experiments and small oribatids and astigmatans increased
in abundance at the highest concentrations (500 mg kg−1 PCP).
Many oribatid species have distinct pH preferences (van Straalen and Verhoef
1992; van Straalen 1997) and, therefore, may make good indicators of acidification
caused by acid rain and other air pollutants. For example, Steiner (1995) found that
oribatid mite communities in moss cushions on walls decreased in species richness
and became more uniform in composition along an urban–rural air pollution gradi-
ent in Switzerland. Bdellid predators of oribatid mites also decreased. Air pollution
levels explained the relative abundances of Zygoribatula exilis extremely well, sug-
gesting that this widespread species would be a good indicator. Since lichens are
strongly affected by air pollution, reduction in lichen cover may have been an
important factor in the decline of these mites. However, the nitrogen and sulfur
components of acid rain can act as nutrients in soil systems and mite populations
tend to increase with simulated acid rain pollution in some experiments (Heneghan
and Bolger 1996a, b). Soil mite populations also can respond positively to aboveg-
round perturbations such as herbivory – directly through increased input of drop-
pings and indirectly through effects on root chemistry (Seastedt et al. 1988). High
levels of phosphorus in both organic and chemical fertilizer treatments have been
linked to declines in oribatid mite populations, possibly as an indirect effect of
reduced fungal biomass (Cao et al. 2011).
210 6 Mites in Soil and Litter Systems

Mites and Earthworms

The response of soil mites to invasive plant species is not yet a well-studied area, but
their response to one group of invading animals – earthworms – is relatively well
known (Eisenhauer 2010). Glaciers are rather unforgiving acts of Mother Nature –
life is not only destroyed as a glacier advances, it is obliterated. During the advance
of the icesheets in North America during the Ice Ages, the soil fauna was extermi-
nated throughout most of what are now Canada and the northeastern United States.
As the glaciers began their retreat about 10,000 years ago, mites quickly colonized
the new soils, probably by wind dispersal (Lehmitz et al. 2011) or swept along by
glacial floods, and now are highly diverse even in the far north (e.g. Young et al.
2012). For some unknown reason, but presumably related to low vagility, North
American earthworms have been slow to leave their glacial refugia to move far
north from their southern retreats. European and Asian earthworms, however, have
been more vagile thanks to human commerce. Largely unnoticed until recent
decades, exotic earthworms have spread from cities and farms into more natural
habitats. Perhaps the best known of these is the Dew-worm (Lumbricus terrestris),
a large species (often 20 cm or more) that likes to emerge on laws on wet nights.
Earthworms are often divided into three functional groups (Burke et al. 2011).
The Dew-worm belongs to the Anecic Group: these form deep vertical burrows and
considerable perturbation and mix the soil, but feed at the surface where they can
significantly reduce the amount of litter available and have strong negative effects
on soil microarthropods (Eisenhauer 2010; Burke et al. 2011). Members of the
Epigeic Group like the Red Earthworm (Lumbricus rubellus) also feed on the litter
layer, but make only shallow burrows and cause less soil perturbation. Members of
the Endogeic Group are less well known because they live and feed in deep burrows
in the mineral soil. They include both native and introduced species, e.g. in the
genus Aporrectodea, although unfortunately the introduced species appear to be
replacing the natives (Reynolds 2010).
The effects of earthworms on various components of the ecosystems into
which they have been introduced have been variable and not entirely negative
(Migge-Kleian et al. 2006). Burrows and middens, for example, are enriched in
nutrients and may benefit some of the soil fauna (Eisenhauer 2010). However,
exotic earthworms can have strong negative effects on soil mites, especially
Oribatida. For example, plots of hardwood forest on the Allegheny Plateau in
New York, USA without earthworms averaged almost twice the diversity of orib-
atids as sites with anecic earthworms (but not by worms in the other functional
groups) and abundances of litter dwelling oribatids were drastically reduced in
the presence of any type of exotic earthworm (Burke et al. 2011). Cameron et al.
(2012) tested effects of invasive epigaeic and anecic species of earthworms on
soil microarthropods from an uninvaded site in Alberta, Canada. They found that
abundance was lowest in mesocosms with both worm species present, and that
the oribatid genera Suctobelbella and Tectocepheus were most negatively affected
by the presence of worms.
Summary 211

Palaeoacacrology

Well-sclerotised mites, such as Oribatida and Mesostigmata, are often fossilised


(Erickson 1996; Solhøy and Solhøy 2000) and this tendency has lead to the recent
development of palaeoacarology, i.e. the study of fossil and subfossil mites
(Baker 2009). Although we rarely notice those mites that do not bite us, human
habitations have always had their own acarofauna and many of these unnoticed
companions have tales to tell about our early habits. For example, although fos-
silised human scats can be identified from the eggs of our characteristic parasites,
fossilised Mesostigmata are more useful for the identification of domestic animal
dung (Schelvis 1992). Fossil and subfossil oribatid mites are especially useful
indicators of vegetation, moisture regimes and salinity around ancient human
habitations, at least where the modern fauna is well known (Schelvis 1990; Solhøy
and Solhøy 2000) and can be used to reconstruct aspects of environments thou-
sands of years in the past. For example, oribatid mites have been used to recon-
struct precipitation:evaporation ratio, nutrient sequestration by maturing tundra
vegetation, immigration of taxa and trophic interactions over thousands of years
at a small freshwater lake in west Greenland (Heggen et al. 2010).

Summary

The ‘enigma’ of soil mite biodiversity is that even a handful of forest soil and litter
may contain dozens of species that coexist with seemingly few adaptations to avoid
trophic or habitat overlap. Outside of the soil-litter layers in any forest, numerous
microhabitats with specific assemblages of species can be recognised, but within
these clumps of moss, shelves of fungi, piles of dung and the like, the riddle is end-
lessly repeated. Typically, each microhabitat contains suites of species with exten-
sive within-suite trophic overlap – or at least so it seems. The true enigma is that
although decomposition is the equal of photosynthesis in ecosystem importance and
half or more of all terrestrial biodiversity may be tied to the soil-litter system, the
study of soil biology has been neglected.
In spite of this neglect, certain patterns are clear. Most importantly, because the
diversity and abundance of soil animals is dominated by the Acari, any true under-
standing of the soil system must include understanding the mite fauna. Even in deep
soils or in extreme climates, mites are major constituents that interact with other
components of these systems in a variety of ways. Mites link together many compo-
nents of soil food webs and the study of their behaviour can shed light on food-web
dynamics and stability.
Seven general functional roles of mites in soil are outlined in this chapter: com-
minuting microbivore-detritivores, piercing-sucking microbivores, filter-feeding
microbivores, direct plant parasites, indirect plant parasites, nematophages and
predators of arthropods. This is a general functional classification and any of these
212 6 Mites in Soil and Litter Systems

groups could be subdivided to identify clusters of species for more specific study.
However, it is important to remember that functional groups, guilds and the like are
artificial constructs – user-defined units of study – not pre-existing components of
nature. As such, they should be clearly defined by the user as to which taxa are
included and which excluded.
Predation is the best studied of the soil mite functional roles and it is perhaps one
of the most important. Applied use of soil predators against pests is a recent, but
rapidly developing, branch of biological control. Because of the diversity of preda-
tory soil mites and the ease with which they can be reared and observed, they make
excellent subjects for the study of predatory behaviour, food- web dynamics and the
adaptations of prey to avoid being eaten. Mites are eaten, not just by other mites, but
by many soil invertebrates and by numerous small vertebrates as well. Because
mites tie together so many components of soil food webs, they may make excellent
indicators of human disturbance. Many show strong habitat preference and when
these are known, their subfossils can be used to reconstruct paleoclimates.

References

Abbott, D. T., Seastedt, T. R., & Crossley, D. A., Jr. (1980). Abundance, distribution, and effects
of clearcutting on Cryptostigmata in the Southern Appalachians. Environmental Entomology,
9, 618–623.
Addington, R. N., & Seastedt, T. R. (1999). Activity of soil microarthropods beneath snowpack in
alpine tundra and subalpine forest. Pedobiologia, 43, 47–53.
Aitchison, C. W. (1979). Winter-active subnivean invertebrates in Southern Canada. III. Acari.
Pedobiologia, 19, 153–160.
Alberti, G. (1973). Ernährungsbiologie und Spinnvermögen der Schnabelmilben (Bdellidae,
Trombidiformes). Zeitschrift für Morphologie und Ökologie der Tiere, 76, 283–338.
Alberti, G. (2010). On predation in Epicriidae (Gamasida, Anactinotrichida) and fine-structural
details of their forelegs. Soil Organisms, 82, 179–192.
Alberti, G., & Ehrnsberger, R. (1977). Rasterelektronenmikroskopische untersuchungen zum
Spinnvermögen der Bdelliden und Cunaxiden (Acari, Prostigmata). Acarologia, 19, 55–61.
Ali, O., Dunne, R., & Bennan, R. (1997). Biological control of the sciarid fly, Lycoriella solani by
the predatory mite, Hypoaspis miles (Acari: Lalelapidae) in mushroom crops. Systematic and
Applied Acarology, 2, 71–80.
Andersen, D. C. (1987). Below-ground herbivory in natural communities: A review emphasizing
fossorial animals. The Quarterly Review of Biology, 62, 261–286.
Anderson, J. M. (1975). Succession, diversity and trophic relationships of some soil animals in
decomposing leaf litter. Journal of Animal Ecology, 44, 475–495.
Anderson, J. M. (1978a). Inter- and intra-habitat relationships between woodland Cryptostigmata
species diversity and the diversity of soil and litter microhabitats. Oecologia, 32, 341–348.
Anderson, J. M. (1978b). Competition between two unrelated species of soil Cryptostigmata
(Acarina) in experimental microcosms. Journal of Animal Ecology, 47, 787–803.
Anderson, J. M., & MacFayden, A. (1976). The role of terrestrial and aquatic organisms in decom-
position processes. Oxford: Blackwell.
André, H. M., Noti, M.-I., & Lebrun, P. (1994). The soil fauna: The other last biotic frontier.
Biodiversity and Conservation, 3, 45–56.
Baker, A. S. (2009). Acari in archaeology. Experimental & Applied Acarology, 49, 147–160.
Bal, L. (1982). Zoological ripening of soils. Wageningen: PUDOC.
References 213

Barker, P. S. (1969). The response of a predator, Hypoaspis aculeifer (Canestrini) (Acarina:


Laelapidae), to two species of prey. Canadian Journal of Zoology, 47, 343–345.
Barrett, J. E., Virginia, R. A., Hopkins, D. W., Aislabie, J., Bargagli, R., Bockheim, J. G.,
Campbell, I. B., Lyons, W. B., Moorhead, D., Nkem, J., Sletten, R. S., Steltzer, H., Wall, D. H.,
& Wallenstein, M. (2006). Terrestrial ecosystem processes of Victoria Land, Antarctica. Soil
Biology and Biochemistry, 38, 3019–3034.
Basset, Y., et al. (2012). Arthropod diversity in a tropical forest. Science, 338, 1481–1484.
doi:10.1126/science.1226727.
Beare, M. H., Coleman, D. C., Crossley, D. A., Jr., Hendrix, P. F., & Odum, E. P. (1995). A hierar-
chical approach to evaluating the significance of soil biodiversity to biogeochemical cycling. In
H. P. Collins, G. P. Robertson, & M. J. Klug (Eds.), The significance and regulation of soil
biodiversity. Dordrecht: Kluwer Academic.
Beaulieu, F. (2011). Saproxyly in predatory mites? Mesostigmata in decaying log habitats versus
litter in a wet eucalypt forest, Tasmania, Australia. International Journal of Acarology, 138,
313–323. doi:10.1080/01647954.2011.647072.
Beaulieu, F., & Walter, D. E. (2007). Predation in suspended and forest floor soils: Observations
on Australian mesostigmatic mites. Acarologia, 47, 43–54.
Beaulieu, F., Walter, D. E., Proctor, H. C., & Kitching, R. L. (2010). The canopy starts at 0.5 m:
Predatory mites (Acari: Mesostigmata) differ between rain forest floor soil and suspended soil
at any height. Biotropica, 42, 704–709.
Behan, V. M., & Hill, S. B. (1978). Feeding habits and spore dispersal of oribatid mites in the
North American Arctic. Revue D'Écologie et Biologie du Sol, 15, 497–516.
Behan-Pelletier, V. M., & Hill, S. B. (1983). Feeding habits of sixteen species of Oribatei (Acari)
from an acid peat bog, Glenamoy, Ireland. Revue D'Écologie et Biologie du Sol, 20, 221–267.
Behan-Pelletier, V. M., & Walter, D. E. (2000). Biodiversity of oribatid mites (Acari: Oribatida) in
tree canopies and litter. In D. C. Coleman & P. Hendrix (Eds.), Invertebrates as webmasters in
ecosystems (pp. 187–202). Wallingford: CABI. ISBN 0 85199 394 X.
Behan-Pelletier, V. M., Hill, S. B., Fjellberg, A., Norton, R. A., & Tomlin, A. (1985). Soil inverte-
brates: Major reference texts. Quaestiones Entomologicae, 21, 675–687.
Blackwood, J. S., Croft, B. A., & Schausberger, P. (2001). Jerking in predaceous mites (Acari:
Phytoseiidae) with emphasis on larvae. Experimental & Applied Acarology, 25, 475–492.
Blaszak, C., Ehrnsberger, R., & Schuster, R. (1990). Beiträge zur Kenntnis der Lebensweise der
Litoralmilbe Macrocheles superbus Hull, 1918 (Acarina: Gamasina). Osnabrücker naturwiss.
Mitt., 16, 51–62.
Block, W. (1979). Nanorchestes antarcticus Strandtmann (Prostigmata) from Antarctic ice.
Acarologia, 21, 173–176.
Block, W. (1984). Terrestrial microbiology, invertebrates and ecosystems. In R. M. Laws (Ed.),
Antarctic ecology (Vol. 1, pp. 163–236). Sydney: Academic.
Block, W. (1992). An annotated bibliography of Antarctic invertebrates (terrestrial and freshwater).
Cambridge: British Antarctic Survey.
Block, W., & Convey, P. (1995). The biology, life cycle and ecophysiology of the Antarctic mite
Alaskozetes antarcticus. Journal of Zoology, 236, 431–449.
Block, W., & Stary, J. (1996). Oribatid mites (Acari: Oribatida) of the maritime Antarctic and
Antarctic Peninsula. Journal of Natural History, 30, 1059–1067.
Brodie, E. D., Ducey, P. K., & Baness, E. A. (1991). Antipredator skin secretions of some
tropical salamanders (Bolitoglossa) are toxic to snake predators. Biotropica, 23, 58–62.
doi:10.2307/2388688.
Buitenhuis, R., Shipp, L., & Scott-Dupree, C. (2010). Intra-guild vs extra-guild prey: Effect
on predator fitness and preference of Amblyseius swirskii (Athias-Henriot) and Neoseiulus
cucumeris (Oudemans) (Acari: Phytoseiidae). Bulletin of Entomological Research, 100(2),
167–173.
Burke, J. L., Maerz, J. C., Milanovich, J. R., Fisk, M. C., & Gandhi, K. J. K. (2011). Invasion by
exotic earthworms alters biodiversity and communities of litter- and soil-dwelling microarthro-
pods. Diversity, 3, 155–175.
214 6 Mites in Soil and Litter Systems

Butler, J. F., Camino, M. L., et al. (1979). Boophilus microplus and the fire ant Solenopsis
geminata. In J. G. Rodriguez (Ed.), Recent advances in acarology I (pp. 355–361). New York:
Academic.
Camann, M., Gillette, K., Lamoncha, K. L., & Mori, S. R. (2008). Response of forest soil Acari to
prescribed fire following stand structure manipulation in the southern Cascade Range.
Canadian Journal of Forestry Research, 38, 956–968.
Cameron, E. K., Knysh, K. M., Proctor, H. C., & Bayne, E. M. (2012). Influence of two exotic
earthworm species with different strategies on abundance and composition of boreal microar-
thropods. Soil Biology and Biochemistry, 57, 334–340.
Cancela da Fonseca, J. P. (1975). Notes oribatologiques. Acarologia, 17, 320–330.
Cannon, R. J. C., & Block, W. (1988). Cold tolerance of microarthropods. Biological Reviews, 63,
23–77.
Cao, Z., Han, X., Hu, C., Chen, J., Zhang, D., & Steinberger, Y. (2011). Changes in the abundance
and structure of a soil mite (Acari) community under long-term organic and chemical fertilizer
treatments. Applied Soil Ecology, 49, 131–138.
Castilho, R. C., de Moraes, G. J., Silva, E. S., et al. (2009). The predatory mite Stratiolaelaps
scimitus as a control agent of the fungus gnat Bradysia matogrossensis in commercial pro-
duction of the mushroom Agaricus bisporus. International Journal of Pest Management, 5,
181–185. doi:10.1080/09670870902725783.
Chambers, R. J., Wright, E. M., et al. (1993). Biological control of glasshouse sciarid flies
(Bradysia spp.) with the predatory mite, Hypoaspis miles, on cyclamen and poinsettia.
Biocontrol Science and Technology, 3, 285–293.
Chiba, S., Abe, T., Aoki, J., Imadate, G., Ishikawa, K., Kondoh, M., Shiba, M., & Watanabe, H.
(1975). Studies on the productivity of soil animals in Pasoh Forest Reserve, West Malaysia. I.
Seasonal change in density of soil mesofauna: Acari, Collembola and others. Science Report
Hirosaki University, 22, 87–124.
Choi, R. T., & Beard, K. H. (2012). Coqui frog invasions change invertebrate communities in
Hawaii. Biological Invasions, 14, 939–948. doi:10.1007/s10530-011-0127-3.
Clapperton, M. J., Kanashiro, D. A., & Behan-Pelletier, V. M. (2002). Changes in abundance and
diversity of microarthropods associated with fescue prairie grazing regimes. Pedobiologia, 46,
496–511.
Cohen, J. E. (1978). Food webs and niche space. Princeton: Princeton University Press.
Coineau, Y. (1973). Les Caeculidae (Acariens Prostigmates) quelques aspects de leurs particulari-
tés éco-éthologiques. Bulletin D'Écologie, 4, 329–337.
Coineau, Y., Haupt, J., Delamere Deboutteville, C., & Théron, P. (1978). Un remarquable exemple
di convergenece écologique: l'adaption de Gordialycus tuzetae (Nematalycidae, Acariens) á la
vie dans les interstices des sables fines. Comptes rendus de l'Academie des Sciences, 287,
883–886.
Coleman, D. C. (2008). From peds to paradoxes: Linkages between soil biota and their influ-
ences on ecological processes. Soil Biology and Biochemistry, 40, 271–289. doi:10.1016/j.
soilbio.2007.08.005.
Colloff, M. J., & Cairns, A. (2011). A novel association between oribatid mites and leafy liver-
worts (Marchantiophyta: Jungermanniidae), with a description of a new species of Birobates
Balogh, 1970 (Acari: Oribatida: Oripodidae). Australian Journal of Entomology, 50, 72–77.
Connell, J. H. (1978). Diversity in tropical rainforests and coral reefs. Science, 99, 1302–1310.
Croft, B. A., & McMurtry, J. A. (1971). Comparative studies on four strains of Typhlodromus
occidentalis Nesbitt (Acarina: Phytoseiidae) IV. Life history studies. Acarologia, 13, 460–470.
Cronberg, N., Natcheva, R., & Hedlund, K. (2006). Microarthropods mediate sperm transfer in
mosses. Science, 313, 1255.
Crossley, D. A., Jr. (1977). The roles of terrestrial saprophagous arthropods in forest soils:
Current status of concepts. In W. J. Mattson (Ed.), The role of arthropods in forest ecosystems
(pp. 49–56). New York: Springer-Verlag.
Cummins, K. W. (1974). Structure and function of stream ecosystems. Bioscience, 24, 631–641.
Curl, E. A., & Truelove, B. (1986). The rhizosphere. New York: Springer-Verlag.
References 215

Curry, J. P. (1994). Grassland invertebrates, ecology, influence on soil fertility and effects on plant
growth. London: Chapman & Hall.
Dawes-Gromadzki, T. Z., & Bull, C. M. (1997). Ant predation on different life stages of two
Australian ticks. Experimental & Applied Acarology, 21, 109–115.
De Deyn, G. B., Raaijmakers, C. E., Zoomer, H. R., Berg, M. P., DeRuiter, P. C., Verhoef, H. A., et al.
(2003). Soil invertebrate fauna enhances grassland succession and diversity. Nature, 422, 711–713.
Déchêne, A. D., & Buddle, C. M. (2010). Decomposing logs increase oribatid mite assemblage
diversity in mixedwood boreal forest. Biodiversity and Conservation, 19, 237–256.
Denegri, G. M. (1993). Review of oribatid mites as intermediate hosts of tapeworms of the
Anoplocephalidae. Experimental & Applied Acarology, 17, 567–580.
Donnelly, M. A. (1991). Feeding patterns of the strawberry poison frog Dendrobates pumilio
(Anura: Dendrobatidae). Copeia, 3, 723–730.
Downes, B. J. (1986). Guild structure in water mites (Unionicola spp.) inhabiting freshwater
mussels: Choice, competitive exclusion and sex. Oecologia, 70, 457–465.
Ducarme, X., Andre, H. A., Wauthy, G., & Lebrun, P. (2004a). Comparison of endogeic and cave
communities: Microarthropod density and mite species richness. European Journal of Soil
Biology, 40, 129–138. doi:10.1016/j.ejsobi.2004.10.003.
Ducarme, X., Andre, H. A., Wauthy, G., & Lebrun, P. (2004b). Are there real endogeic species in
temperate forest mites? Pedobiologia, 48, 139–147. doi:10.1016/j.pedobi.2003.10.002.
Eadie, J. M., Broekhoven, L., & Colgan, P. (1987). Size ratios and artifacts: Hutchinson's rule
revisited. American Naturalist, 129, 1–17.
Ebermann, E. (1995). Indication of jumping ability in the mite family Scutacaridae (Acari, Tarso-
nemina). Entomologische Mitteilungen aus dem Zoologischen Museum Hamburg, 11, 205–209.
Edwards, C. A., & Stinner, B. R. (Eds.). (1988). Biological interactions in soil. Amsterdam:
Elsevier.
Eisenbeis, G., & Wichard, W. (1987). Atlas on the biology of soil arthropods. New York:
Springer-Verlag.
Eisenhauer, N. (2010). The action of an animal ecosystem engineer: Identification of the main
mechanisms of earthworm impacts on soil microarthropods. Pedobiologia, 53, 343–352.
Enders, F. (1975). The influence of hunting manner on prey size, particularly in spiders with
long attack distances (Araneidae, Linyphiidae, and Salticidae). American Naturalist, 109,
737–763.
Epsky, N. D., Walter, D. E., & Capinera, J. L. (1988). Potential role of nematophagous arthropods
as biotic mortality factors of entomogenous nematodes (Rhabditida: Steinernematidae,
Heterorhabditidae). Journal of Economic Entomology, 81, 821–825.
Erickson, J. M. (1996). Can paleoacarology contribute to global change research? In R. Mitchell,
D. J. Horn, G. R. Needham, & W. C. Welbourn (Eds.), Acarology IX (Vol. I, pp. 533–537).
Columbus: Ohio Biological Survey.
Faber, J. H. (1991). Functional classification of soil fauna: A new approach. Oikos, 62, 110–117.
Fager, E. W. (1968). The community of invertebrates in decaying oak wood. Journal of Animal
Ecology, 37, 121–142.
Fauth, J. E., Bernardo, J., et al. (1996). Simplifying the Jargon of community ecology: A conceptual
approach. American Naturalist, 147, 282–286.
Fisher, J. R., Skvarla, M. J., Bauchan, G. R., Ochoa, R., & Dowling, A. P. G. (2011). Trachymolgus
purpureus sp. n., an armored snout mite (Acari, Bdellidae) from the Ozark highlands:
Morphology, development, and key to Trachymolgus Berlese. ZooKeys, 125, 1–34. doi:10.3897/
zookeys.125.1875.
Fitzsimons, J. M. (1971). On the food habits of certain Antartic arthropods from coastal Victoria
Land and adjacent islands. Pacific Insects Monograph, 25, 121–125.
Flowers, M. A., & Graves, B. M. (1995). Prey selectivity and size-specific diet changes in Bufo
cognatus and B. woodhousii during early postmetamorphic ontogeny. Journal of Herpetology,
23, 608–612.
Freckman, D. W., & Virginia, R. A. (1997). Low diversity Antarctic soil nematode communities:
Distribution and response to disturbance. Ecology, 78, 363–369.
216 6 Mites in Soil and Litter Systems

Garrett, C. J., Crossley, D. A., Jr., Coleman, D. C., Hendrix, P. F., Kisselle, K. E., & Potter, R. L.
(2001). Impact of the rhizosphere on soil microarthropods in agroecosystems on the Georgia
piedmont. Applied Soil Ecology, 16, 141–148.
Gause, G. F. (1934). The struggle for existence. Baltimore: Williams & Wilkins.
Gerson, U. (1972). Mites of the genus Ledermuelleria (Prostigmata: Stigmaeidae) association with
mosses in Canada. Acarologia, 13, 319–342.
Gill, R. W. (1969). Soil microarthropod abundance following old-field litter manipulation. Ecology,
50, 805–816.
Giller, P. S. (1996). The diversity of soil communities, the ‘poor man’s tropical rainforest’.
Biodiversity and Conservation, 5, 135–168.
Gillespie, D. R., & Quiring, D. M. J. (1990). Biological control of fungus gnats, Bradysia spp.
(Diptera: Sciaridae), and western flower thrips, Frankliniella occidentalis (Pergande)
(Thysanoptera: thripidae), in greenhouses using a soil-dwelling predatory mite, Geolaelaps sp.
nr. aculeifer (Canestrini) (Acari: Laelapidae). Canadian Entomologist, 122, 975–983.
Gless, E. E. (1967). Notes on the biology of Coccorhagia gressitti Womersley and Strandtmann.
In J. L. Gressitt (Ed.), Entomology of Antarctica (Vol. 10, pp. 321–324). Washington, DC:
American Geophysical Union.
Gochenaur, S. E. (1987). Evidence suggests that grazing regulates ascospore density in soil.
Mycologia, 79, 445–450.
Greene, C. H. (1986). Patterns of prey selection: Implications of predator foraging tactics.
American Naturalist, 128, 824–839.
Gressitt, J. L., & Shoup, J. (1967). Ecological notes on free-living mites in North Victoria land.
Antarctic Research Series, 10, 307–320.
Gwiazdowicz, D. J., Kamczyc, J., & Rakowsk, R. (2011). Mesostigmatid mites in four
classes of wood decay. Experimental & Applied Acarology, 55, 155–165. doi:10.1007/
s10493-011-9458-0.
Hågvar, S., & Hågvar, E. B. (2011). Invertebrate activity under snow in a South-Norwegian spruce
forest. Soil Organisms, 83, 187–209.
Hågvar, S., Tolhoy, T., & Mong, C. E. (2009). Primary succession of soil mites (Acari) in a
Norwegian glacier foreland, with emphasis on oribatid species. Arctic, Antarctic, and Alpine
Research, 41, 219–227. doi:10.1657/1938-4246-41.2.
Hairston, N. G., Smith, F. E., & Slobodkin, L. (1960). Community structure, population control,
and competition. American Naturalist, 94, 421–425.
Halffter, G., & Matthews, G. E. (1971). The natural history of dung beetles. A supplement on
associated biota. Revista Latino-Americana de Microbiologia, 13, 147–164.
Halliday, R. B. (2008). Oriflammella n. gen. (Acari: Ologamasidae), a remarkable new genus of
mites from eastern Australia. International Journal of Acarology, 34, 43–53.
Hamlen, R. A., & Mead, F. W. (1979). Fungus gnat larval control in greenhouse plant production.
Journal of Economic Entomology, 72, 269–271.
Hammer, M. (1972). Microhabitats of oribatid mites on a Danish woodland floor. Pedobiologia,
12, 412–423.
Hansen, R. A. (2000). Effect of habitat complexity and composition on a diverse litter microarthro-
pod assemblage. Ecology, 81, 1120–1132.
Hansen, R. A., & Coleman, D. C. (1998). Litter complexity and composition are determinants of
the diversity and species composition of oribatid mites (Acari: Oribatida) in litterbags. Applied
Soil Ecology, 9, 17–23.
Hawkins, C. P., & MacMahon, J. A. (1989). Guilds: The multiple meanings of a concept. Annual
Review of Entomology, 34, 423–451.
Heethoff, M., & Raspotnig, G. (2012). Triggering chemical defense in an oribatid mite using artificial
stimuli. Experimental & Applied Acarology, 56, 287–295. doi:10.1007/s10493-012-9521-5.
Heethoff, M., Koerner, L., Norton, R. A., & Raspotnig, G. (2011). Tasty but Protected – First
evidence of chemical defense in oribatid mites. Journal of Chemical Ecology, 37, 1037–1043.
doi:10.1007/s10886-011-0009-2.
Heggen, M., Birks, H. H., & Anderson, N. J. (2010). Long-term ecosystem dynamics of a small
lake and its catchment in west Greenland. The Holocene, 20, 1207–1222.
References 217

Heidemann, K., Scheu, S., Ruess, L., & Maraun, M. (2011). Molecular detection of nematode
predation and scavenging in oribatid mites: Laboratory and field experiments. Soil Biology &
Biochemistry, 43, 2229–2236.
Heisler, C. (1995). Influence of agricultural traffic and crop management on Collembola and
microbial biomass in arable soil. Biology and Fertility of Soils, 19, 159–165.
Hendricks, P. F., Crossley, D. A., Jr., et al. (1990). Soil Biota as components of sustainable agro-
ecosystems. In C. A. Edwards, R. Lal, P. Madden, R. H. Miller, & G. House (Eds.), Sustainable
agricultural systems (pp. 637–654). Ankeny: Soil, Water and Conservation Society.
Heneghan, L., & Bolger, T. (1996a). Effects of acid rain components on soil microarthropods:
A field manipulation study. Pedobiologia, 40, 413–438.
Heneghan, L., & Bolger, T. (1996b). Effect of components of 'acid rain' on the contribution of soil
microarthropods to ecosystem function. Journal of Applied Ecology, 33, 1329–1344.
Hodkinson, I. D., Coulson, S. J., & Webb, N. R. (2004). Invertebrate community assembly across
proglacial chronosequences in the high Arctic. Journal of Animal Ecology, 73, 556–568.
Hoffmann, D., Vierheilig, H., Pender, S., & Schausberger, S. (2011). Mycorrhiza modulates
aboveground tri-trophic interactions to the fitness benefit of its host plant. Ecological
Entomology, 36, 574–581.
Holland, J. N., Cheng, W., & Crossley, D. A., Jr. (1996). Herbivore-induced changes in plant car-
bon allocation: Assessment of belowground C fluxes using carbon-14. Oecologia, 107, 87–94.
Horn, H. S., & May, R. M. (1977). Limits to similarity among coexisting competitors. Nature, 270,
660–661.
Huber, I. (1978). Prey attraction and immobilization by allomone from nymphs. Acarologia, 20,
112–115.
Huey, R. B., & Pianka, E. R. (1981). Ecological consequences of foraging mode. Ecology, 62,
991–999.
Hunt, H. W., Elliott, E. T., & Walter, D. E. (1989). Inferring trophic transfers from pulse-dynamics
in detrital food webs. Plant and Soil, 115, 247–259.
Hutchinson, G. E. (1959). Homage to Santa Rosalia, or why are there so many kinds of animals?
American Naturalist, 93, 145–159.
Hutchinson, G. E. (1965). The ecological theater and the evolutionary play. New Haven: Yale
University Press.
Hutchinson, G. E., & MacArthur, R. (1959). A theoretical ecological model of size distributions
among species of animals. American Naturalist, 93, 117–126.
Hutu, M. (1991). Reproduction, embryonic and postembryonic development of Trichouropoda
obscurasimilis Hirschmann & Zirngiebl-Nicol 1961 (Anactinotrichida: Uropodina). In R. Schuster
& P. W. Murphy (Eds.), The acari, reproduction, development and life-history strategies
(pp. 279–299). Melbourne: Chapman & Hall.
Inserra, R. N., & Davis, D. W. (1983). Hypoaspis nr. aculeifer: A mite predacious on root-knot and
cyst nematodes. Journal of Nematology, 15, 324–325.
Jaksic, F. M. (1981). Abuse and misuse of the term “guild” in ecological studies. Oikos, 37, 397–400.
Jones, C. G., Lawton, J. H., & Shachak, M. (1997). Positive and negative effects of organisms as
physical ecosystem engineers. Ecology, 78, 1946–1957.
Kaliszewski, M., Athias-Binche, F., & Lindquist, E. E. (1995). Parasitism and parasitoidism in
Tarsonemina (Acari: Heterostigmata) and evolutionary considerations. Advances in
Parasitology, 35, 335–367.
Kaneko, N. (1988). Feeding habits and cheliceral size of oribatid mites in cool temperate forest
soils in Japan. Revue D'Écologie et Biologie du Sol, 25, 353–363.
Kaneko, N. (1995). Composition of feeding types in oribatid mite communities in forest soils. Acta
Zoologica Fennica, 196, 160–161.
Kaneko, N., & Salamanca, N. (1999). Mixed leaf litter effects on decomposition rates and soil
arthropod communities in an oakpine forest stand in Japan. Ecological Research, 14, 131–138.
Karg, W. (1961). Okologische Untersuchungen von edaphischen Gamasiden (Acari: Parasitiformes).
Pedobiologia, 1, 58–98.
Karg, W. (1983). Verbreitung und Bedeutung von raubmilben der Cohors Gamasina als
Antagonisten von Nematoden. Pedobiologia, 25, 419–432.
218 6 Mites in Soil and Litter Systems

Kethley, J. (1990). Acarina: Prostigmata (Actinedida). In D. L. Dindal (Ed.), Soil biology guide
(pp. 667–756). New York: Wiley.
Kinn, D. N., & Witcosky, J. J. (1977). The life cycle and behaviour of Macrocheles boudreauxi
Krantz. Zeitschrift für Angewandte Entomologie, 84, 136–144.
Kitching, R. L. (1987). Spatial and temporal variation in food webs in water-filled treeholes. Oikos,
48, 280–288.
Kitching, R. L., & Pimm, S. L. (1985). The length of food chains: Phytotelmata in Australia and
elsewhere. Proceedings of the Ecological Society of Australia, 14, 123–140.
Klironomos, J. N., & Kendrick, B. (1995). Relationships among microarthropods, fungi and their
environment. In H. P. Collins, G. P. Robertson, & M. J. Klug (Eds.), The significance and regu-
lation of soil biodiversity (pp. 209–233). Dordrecht: Kluwer Academic Publishers.
Knülle, W. (1995). Expression of a dispersal trait in a guild of mites colonizing transient habitats.
Evolutionary Ecology, 9, 341–353.
Koehler, H. H. (1992). The use of soil mesofauna for the judgement of chemical impact on ecosys-
tems. Agriculture, Ecosystems and Environment, 40, 193–205.
Kolesnikov, V. B. (2010). The role of oribatid mites in the process of soil formation. Zashchita i
Karantin Rastenii, 9, 40–41.
Krantz, G. W. (1978). A manual of acarology. Corvallis: Oregon State University Bookstores.
Krantz, G. W. (1983). Mites as biological control agents of dung-breeding flies, with special refer-
ence to the Macrochelidae. In M. A. Hoy, G. L. Cunningham, & L. Knutson (Eds.), Biological
control of pests by mites (Special Publication 3304, pp. 91–98). Berkeley: University of
California, Agriculture Experiment Station.
Krantz, G. W., & Lindquist, E. E. (1979). Evolution of phytophagous mites (Acari). Annual Review
of Entomology, 24, 121–158.
Krantz, G. W., & Walter, D. E. (Eds.). (2009). A manual of acarology (3rd ed.). Lubbock: Texas
Tech University Press. 807 p. 338 b/w illustrations; 60 figures ISBN 978-0-89672-620-8.
Krisper, G. (1990). Das Sprungvermogen der Milbengattung Zetorchestes (Acarida, Oribatida).
Zoologische Jahrbuecher Abteilung fuer Anatomie und Ontogenie der Tiere, 120, 289–312.
Krisper, G. (1991). The saltatory capacity of an oribatid mite. In R. Schuster & P. W. Murphy
(Eds.), The acari. Reproduction, development, and life-history strategies (p. 397). New York:
Chapman & Hall.
Kristin, A. (1993). Diet preferences of the Dunnock in various forest habitats. Vogelwelt, 114, 72–82.
Kuwahara, Y. (1990). Chemical studies on astigmatid mites – Opisthonotal gland secretions and
body surface components with biological functions. Journal of Pesticide Science, 15, 613–619.
Labandeira, C. C., Phillips, T. L., & Norton, R. A. (1997). Oribatid mites and the decomposition
of plant tissues in Paleozoic coal-swamp forests. Palaios, 12, 319–353.
Lasebikan, B. A. (1974). A preliminary communication on microarthropods from a tropical rain-
forest in Nigeria. Pedobiologia, 14, 402–411.
Lavelle, P. (1997). Faunal activities and soil processes: adaptive strategies that determine ecosys-
tem function. Advances in Ecological Research, 27, 93–132.
Lavelle, P., Lattaud, C., Trigo, D., & Barois, I. (1994). Mutualism and biodiversity in soils. Plant
and Soil, 170, 23–33.
Lawton, J. H. (1984). Surface availability and insect community structure: The effects of architec-
ture and fractal dimensions of plants. In B. E. Juniper & T. R. E. Southwood (Eds.), Insects and
the plant surface (pp. 317–331). London: Arnold.
Lebrun, P., & van Straalen, N. M. (1995). Oribatid mites: Prospects for their use in ecotoxicology.
Experimental & Applied Acarology, 19, 361–379.
Leetham, J., & Milchunas, D. G. (1985). The composition and distribution of soil microarthropods
in the shortgrass steppe in relation to the soil water, root biomass and grazing by cattle.
Pedobiologia, 28, 311–325.
Lehmitz, R., Russell, D., Hohberg, K., Christian, A., & Xylander, W. E. R. (2011). Wind dispersal of
oribatid mites as a mode of migration. Pedobiologia, 54, 201–207. doi:10.1016/j.pedobi.2011.01.002.
Leinaas, H. P. (1981). Activity of Arthropoda in snow within a coniferous forest, with special
reference to Collembola. Holarctic Ecology, 4, 127–138.
References 219

Lesna, I., Sabelis, M. W., & Conijin, C. G. M. (1996). Biological control of the bulb mite,
Rhizoglyphus robini, by the predatory mite, Hypoaspis aculeifer, on lillies: Predator–prey inter-
actions at various spatial scales. Journal of Applied Ecology, 33, 369–376.
Levings, S. C., & Windsor, D. M. (1982). Seasonal and annual variation in litter arthropod popula-
tions. In E. G. Leigh, A. S. Rand, & D. M. Windsor (Eds.), The ecology of a tropical forest:
Seasonal rhythms and long-term changes (pp. 355–387). Washington, DC: Smithsonian
Institution Press.
Lilleskov, E. A., & Bruns, T. D. (2005). Spore dispersal of a resupinate ectomycorrhizal fungus,
Tomentella sublilacina, via soil food webs. Mycologia, 97, 762–769.
Lima, A. P., Magnusson, W. E., & Williams, D. G. (2000). Differences in diet among frogs and
lizards coexisting in subtropical forests of Australia. Journal of Herpetology, 34, 40–46.
Lindo, Z., & Gonzalez, A. (2010). The bryosphere: An integral and influential component of the
earth’s biosphere. Ecosystems, 13, 612–627.
Lindo, Z., & Visser, S. (2004). Forest floor microarthropod abundance and oribatid mite
(Acari : Oribatida) composition following partial and clear-cut harvesting in the mixed-
wood boreal forest. Canadian Journal of Forest Research, 34, 998–1006. doi:10.1139/
X03-284.
Lindo, Z., & Winchester, N. N. (2008). Oribatid mite communities and foliar litter decomposi-
tion in canopy suspended soils and forest floor habitats of western redcedar forests,
Vancouver Island, Canada. Soil Biology and Biochemistry, 39, 2957–2966. doi:10.1016/j.
soilbio.2007.06.009.
Lindo, Z., Winchester, N. N., & Didham, R. K. (2008). Nested patterns of community assembly in
the colonisation of artificial canopy habitats by oribatid mites. Oikos, 117, 1856–1864.
doi:10.1111/j.1600-0706.2008.16920.x.
Lindquist, E. E. (1975). Associations between mites and other arthropods in forest floor habitats.
Canadian Entomologist, 107, 425–437.
Lindquist, E. E. (1985). Discovery of sporothecae in adult female Trochometridium Cross, with
notes on analogous structures in Siteroptes Amerling (Acari: Heterostigmata). Experimental
and Applied Acarology, 1, 73–85.
Lindquist, E. E., & Walter, D. E. (1989). Biology and description of Antennoseius janus, new spe-
cies (Mesostigmata: Ascidae), a mesostigmatic mite exhibiting adult female dimorphism.
Canadian Journal of Zoology, 67, 1291–1310.
Lipovsky, L. J. (1954). Studies on the food habits of postlarval chiggers (Acarina, Trombiculidae).
University of Kansas Science Bulletin, 36, 943–958.
Lister, A. (1984). Predation in an Antarctic micro-arthropod community. In D. A. Griffiths
& C. E. Bowman (Eds.), Acarology VI (Vol. 1, pp. 886–892). Chichester: Ellis Horwood.
Lister, A., Usher, M. B., & Block, W. (1987). Description and quantification of field attack rates by
predatory mites: An example using an electrophoresis method with a species of Antarctic mite.
Oecologia, 72, 185–191.
Lister, A., Block, W., & Usher, M. B. (1988). Arthropod predation in an Antarctic terrestrial com-
munity. Journal of Animal Ecology, 57, 957–970.
Lobbes, P., & Schotten, C. (1980). Capacities of increase of the soil mite Hypoaspis aculeifer
Canestrini (Mesostigmata: Laelapidae). Zeitschrift für Angewandte Entomologie, 90, 9–22.
Lugwig, M., Krantzmann, M., & Alberti, G. (1991). Accumulation of heavy metals in two oribatid
mites. In F. Dusábek & V. Bukva (Eds.), Modern acarology (Vol. 1, pp. 431–437). Prague: SPB
Academic Publishing.
Lussenhop, J. (1992). Mechanisms of microarthropod-microbial interactions in soil. Advances in
Ecological Research, 23, 1–33.
Luxton, M. (1966). Laboratory studies on the feeding behaviour of Saltmarsh Acarina. Acarologia,
8, 163–175.
Luxton, M. (1972). Studies on the oribatid mites of a Danish woodland soil. I. Nutritional biology.
Pedobiologia, 12, 434–463.
Luxton, M. (1979). Food and energy processing by oribatid mites. Revue D'Écologie et Biologie
du Sol, 16, 103–111.
220 6 Mites in Soil and Litter Systems

Lynch, J. F. (1985). The feeding ecology of Aneides flavipunctatus and sympatric plethodontid
salamanders in Northwestern California. Journal of Herpetology, 19, 328–352.
Macfadyen, A. (1963). The contribution of the microfauna to total soil metabolism. In J. Doeksen
& J. van der Drift (Eds.), Soil organisms (pp. 3–17). Amsterdam: North Holland Publishing.
Maiorana, V. C. (1978). An explanation of ecological and developmental constants. Nature, 273,
375–377.
Malmstrom, A., Persson, T., & Ahlstrom, K. (2008). Effects of fire intensity on survival and recov-
ery of soil microarthropods after a clearcut burning. Canadian Journal of Forest Research, 38,
2465–2475. doi:10.1139/X08-094.
Manning, M. J., & Halliday, R. B. (1994). Biology and reproduction of some Australian species of
Macrochelidae (Acarina). Australian Entomologist, 21, 89–94.
Maraun, M. A., Erdmann, G., Fischer, B. M., Pollierer, M. M., Norton, R. A., Schneider, K.,
& Scheu, S. (2011). Stable isotopes revisited: Their use and limits for oribatid mite trophic
ecology. Soil Biology and Biochemistry, 43, 877–882.
Marshall, D. J., & Pugh, P. J. A. (1996). Origin of the inland Acari of continental Antarctica, with
particular reference to Dronning Maud Land. Zoological Journal of the Linnean Society, 118,
101–118.
Masuko, K. (1994). Specialized predation on oribatid mites by two species of the ant genus
Myrmecina (Hymenoptera: Formicidae). Psyche, 101, 159–173.
Masuko, K. (2009). Studies on the predatory biology of Oriental dacetine ants (Hymenoptera:
Formicidae). III. Predation on gamasid mites by Pyramica mazu with a supplementary note on
P. hexamerus. Journal of the Kansas Entomological Society, 82, 109–113.
Matsumoto, K., Wada, Y., & Okamoto, M. (1979). The alarm pheromone of grain mites and its
antifungal effect. Recent Advances in Acarology, 1, 243–249.
Matthewman, W. G., & Pielou, D. P. (1971). Arthropods inhabiting the sporophores of Fomes
fomentarius (Polyporaceae) in Gatineau Park, Quebec. Canadian Entomologist, 103, 775–847.
May, R. M. (1978). The dynamics and diversity of insect faunas. In L. A. Mound & N. Waloff
(Eds.), Diversity of insect faunas (pp. 188–204). Oxford: Blackwell Scientific Publications.
Mayr, E. (1969). Principles of systematic zoology. New York: McGraw-Hill.
McAloon, F. M. (2004). Oribatid mites as intermediate hosts of Anoplocephala manubriata,
cestode of the Asian elephant in India. Experimental & Applied Acarology, 32, 181–185.
Mead, L. S., & Boback, S. M. (2006). Diet and microhabitat utilization of two sympatric neo-
tropical Salamanders: Bolitoglossa pesrubra and B. cerroensis. Herpetological Natural
History, 9, 135–140.
Meier, F. A., Scherrer, S., & Honegger, R. (2002). Faecal pellets of lichenivorous mites contain
viable cells of the lichen-forming ascomycete Xanthoria parietina and its green algal photobi-
ont, Trebouxia arboricola. Biological Journal of the Linnaean Society, 76, 259–268.
Messelink, G., & Van Wensveen, W. (2003). Biocontrol of Duponcheria fovealis (Lepidoptera:
Pyralidae) with soil-dwelling predators in potted plants. Communications in Agricultural and
Applied Biological Sciences, 68, 159–165.
Migge-Kleian, S., McLean, M. A., Maerz, J. C., & Heneghan, L. (2006). The influence of invasive
earthworms on indigenous fauna in ecosystems previously uninhabited by earthworms.
Biological Invasions, 8, 1275–1285.
Mizoguchi, A., Murakami, K., Shimizu, N., Mori, N., Nishida, R., & Kuwahara, T. (2005).
S-isorobinal as the female sex pheromone from an alarm pheromone emitting mite,
Rhizoglyphus setosus. Experimental & Applied Acarology, 36, 107–117.
Mollemann, F., & Walter, D. E. (2001). Niche segregation and canopeners: Scydmaenid beetles
as predators of armoured mites in Australia. In R. B. Halliday, D. E. Walter, H. C. Proctor,
R. A. Norton, & M. J. Colloff (Eds.), Acarology: Proceedings of the 10th international congress
(pp. 281–288). Melbourne: CSIRO Publishing.
Moore, J. C., Walter, D. E., & Hunt, H. W. (1988). Arthropod regulation of micro- and mesobiota
in below-ground detrital foodwebs. Annual Review of Entomology, 33, 419–439.
Mortimer, E., van Vuuren, B. J., Lee, J. E., Marshall, D. J., Convey, P., & Chown, S. L. (2011).
Mite dispersal among the Southern Ocean Islands and Antarctica before the last glacial
References 221

maximum. Proceedings of the Royal Society of London Series B Biological Sciences, 278,
1247–1255. doi:10.1098/rspb.2010.1779.
Muraoka, M., & Ishibashi, N. (1976). Nematode-feeding mites and their feeding behaviour.
Applied Entomology and Zoology, 11, 1–7.
Navarro, M.-J., Gea, F.-J., & Escudero-Colomar, L. A. (2010). Abundance and distribution of
Microdispus lambi (Acari: Microdispidae) in Spanish mushroom crops. Experimental &
Applied Acarology, 50, 309–316. doi:10.1007/s10493-009-9326-3.
Neher, D. A., Lewins, S. A., Weicht, T. R., & Darby, B. J. (2009). Microarthropod communities
associated with biological soil crusts in the Colorado Plateau and Chihuahuan deserts. Journal
of Arid Environments, 73, 672–677.
Nicholas, W. L. (1984). The biology of free-living nematodes. Oxford: Clarendon Press.
Nielsen, U. N., Osler, G. H. R., Campbell, C. D., Burslem, D. F. R. P., & van der Wal, R. (2012).
Predictors of fine-scale spatial variation in soil mite and microbe community composition
differ between biotic groups and habitats. Pedobiologia, 55, 83–91.
Norton, R. A. (1985). Aspects of the biology and systematics of soil arachnids, particularly
saprophagous and mycophagous mites. Quaestiones Entomologicae, 21, 523–541.
Norton, R. A. (1994). Evolutionary aspects of oribatid mite life histories and consequences for the
origin of the Astigmata. In M. A. Houck (Ed.), Mites: Ecological and evolutionary studies of
life-history patterns (pp. 99–135). New York: Chapman & Hall.
Norton, R. A., & Behan-Pelletier, V. P. (1991). Calcium carbonate and calcium oxalate as cuticular
hardening agents in oribatid mites (Acari: Oribatida). Canadian Journal of Zoology, 69,
1504–1511.
Norton, R. A., & Behan-Pelletier, V. M. (2009). Suborder Oribatida. In G. E. Krantz & D. E. Walter
(Eds.), A manual of acarology (3rd ed., pp. 430–564). Lubbock: Texas Tech University Press.
Norton, R. A., & MacNamara, M. C. (1976). The common newt (Notophthalmus viridescens)
as a predator of soil mites in New York. The Journal of the Georgia Entomological Society,
11, 89–93.
Norton, R. A., Kethley, J. B., Johnston, D. E., & OConnor, B. M. (1993). Phylogenetic perspec-
tives on genetic systems and reproductive modes of mites. In D. L. Wrensch & M. A. Ebbert
(Eds.), Evolution and diversity of sex ratio in insects and mites (pp. 8–99). New York: Chapman
& Hall.
Norton, R. A., Graham, T. B., & Alberti, G. (1996). A rotifer-eating ameronothroid (Acari:
Ameronothridae) mite from ephemeral pools on the Colorado plateau. In R. Mitchell, D. J. Horn,
G. R. Needham, & W. C. Welbourn (Eds.), Acarology IX (Vol. 1). Columbus: Ohio Biological
Survey.
O'Brien, W. J., Browman, H. I., & Evans, B. I. (1990). Search strategies of foraging animals.
American Scientist, 78, 152–160.
O'Connell, T., & Bolger, T. (1997). Stability, ephemerality and dispersal ability: Microarthropod
assemblages on fungal sporophores. Biological Journal of the Linnean Society, 62, 111–131.
O’Donnell, A. E., & Axtell, R. C. (1965). Predation by Fuscouropoda vegetans (Acarina:
Uropodidae) on the House Fly (Musca domestica). Annals of the Entomological Society of
America, 58, 403–404.
O'Donnell, A. E., & Nelson, E. L. (1967). Predation by Fuscouropoda vegetans (Acarina:
Uropodidae) and Macrocheles muscaedomesticae (Acarina: Macrochelidae) on the eggs of the
little house fly, Fannia canicularia. Journal of the Kansas Entomological Society, 40, 441–443.
Odum, E. P. (1991). Fundamentals of ecology. Philadelphia: Sanders.
Okabe, K., & Amano, H. (1992). Mite species collected from field mushrooms (I): Cryptostigmata.
Journal of the Acarological Society of Japan, 1, 127–135.
Okabe, K., & Amano, H. (1993). Mite species collected from field mushrooms (II): Mesostigmata,
Prostigmata and Astigmata. Journal of the Acarological Society of Japan, 2, 19–28.
Ostle, N., Briones, M. J. I., Ineson, P., Cole, L., Staddon, P., & Sleep, D. (2007). Isotopic detection
of recent photosynthate carbon flow into grassland rhizosphere fauna. Soil Biology and
Biochemistry, 39, 768–777. doi:10.1016/j.soilbio.2006.09.025.
Paine, R. T. (1966). Food web complexity and species diversity. American Naturalist, 100, 65–74.
222 6 Mites in Soil and Litter Systems

Paine, R. T. (1996). Preface. In G. A. Polis & K. O. Winemiller (Eds.), Food webs, integration of
patterns & dynamics (pp. ix–x). New York: Chapman & Hall.
Park, O. (1947). Observations on Batrisodes (Coleoptera: Pselaphidae), with particular reference
to the American species east of the Rocky Mountains. Bulletin of the Chicago Academy of
Sciences, 8, 45–132.
Pengilley, R. K. (1971). The food of some Australian anurans (Amphibia). Journal of Zoology,
163, 93–103. London.
Perdomo, G., Evans, A., Maraun, M., Sunnucks, P., & Thompson, R. (2012). Mouthpart morphol-
ogy and trophic position of microarthropods from soils and mosses are strongly correlated. Soil
Biology and Biochemistry, 53, 56–63.
Petersen, H. (1982). Structure and size of soil animal populations. Oikos, 39, 306–329.
Petersen, H., & Luxton, M. (1982). A comparative analysis of soil faunal populations and their role
in decomposition processes. Oikos, 39, 287–388.
Pielou, D. P., & Verma, A. N. (1968). The arthropod fauna associated with the birch bracket fungus,
Polyporus betulinus, in eastern Canada. Canadian Entomologist, 100, 1179–1199.
Pimm, S. L. (1991). The balance of nature. Chicago: University of Chicago Press.
Polak, M., & Markow, T. A. (1995). Effect of ectoparasitic mites on sexual selection in a Sonoran
fruit fly. Evolution, 49, 660–669.
Polis, G. A. (1991). Complex trophic interactions in deserts: an empirical critique of food web
theory. American Naturalist, 138, 123–155.
Polis, G. A., Myers, C. A., & Holt, R. D. (1989). The ecology and evolution of intraguild preda-
tion: Potential competitors that eat each other. Annual Review of Ecology and Systematics, 20,
297–330.
Polis, G. A., Sissom, W. D., & McCormick, S. J. (1981). Predators of scorpions: Field data and a
review. Journal of Arid Environments, 4, 309–326.
Price, D. W., & Benham, G. S. (1976). Vertical distribution of pomerantziid mites (Acarina:
Pomerantziidae). Proceedings of the Entomological Society of Washington, 78, 309–313.
Price, D. W., & Benham, G. S. (1977). Vertical distribution of soil-inhabiting microarthropods in
an agricultural habitat in California. Environmental Entomology, 6, 575–580.
Pritchard, G., & Scholefield, P. (1978). Observations on the food, feeding behaviour, and associated
sense organs of Grylloblatta campodeiformis (Grylloblattodea). Canadian Entomologist, 110,
205–212.
Pugh, P. J. A. (1993). A synonymic catalogue of the Acari from Antarctica, the sub-Antarctic
Islands and the Southern Ocean. Journal of Natural History, 27, 323–421.
Pugh, P. J. A. (1994). Non-indigenous Acari of Antarctica and the sub-Antarctic islands. Zoological
Journal of the Linnean Society, 110, 207–217.
Pugh, P. J. A., & Convey, P. (2008). Surviving out in the cold: Antarctic endemic invertebrates and
their refugia. Journal of Biogeography, 35, 2176–2186.
Pugh, P. J. A., & King, P. E. (1985). Feeding in intertidal Acari. Journal of Experimental Marine
Biology and Ecology, 94, 269–280.
Qin, T.-K., & Halliday, R. B. (1997). Eriorhynchidae, a new family of Prostigmata (Acarina), with
a cladistic analysis of eupodoid species of Australia and New Zealand. Systematic Entomology,
22, 151–171.
Rabatin, S. C., & Rhodes, L. H. (1982). Acaulospora bireticulata inside oribatid mites. Mycologia,
74, 859–861.
Raspotnig, G. (2006). Chemical alarm and defence in the oribatid mite Collohmannia gigantea
(Acari: Oribatida). Experimental & Applied Acarology, 39, 177–194.
Raspotnig, G., Norton, R. A., & Heethoff, M. (2011). Oribatid mites and skin alkaloids in poison
frogs. Biology Letters, 7, 555–556.
Remen, C., Kruger, M., & Cassel-Lundhagen, A. (2010). Successful analysis of gut contents in
fungal-feeding oribatid mites by combining body-surface washing and PCR. Soil Biology and
Biochemistry, 42, 1952–1957.
Reynolds, J. W. (2010). The earthworms (Oligochaeta: Lumbricidae) of Nova Scotia, Canada,
revisted. Megadrilogica, 14, 77–100.
References 223

Riha, G. (1951). Zur Oekologie der Oribatiden in Kalksteinboeden. Zoologische Jahrbuecher, 80,
407–450.
Rockett, C. L. (1980). Nematode predation by oribatid mites (Acari: Oribatida). International
Journal of Acarology, 6, 219–224.
Rockett, C. L., & Woodring, J. P. (1966). Oribatid mites as predators of soil nematodes. Annals of
the Entomological Society of America, 59, 669–671.
Root, R. B. (1967). The niche exploitation pattern of the blue-gray gnatcatcher. Ecological
Monographs, 37, 317–350.
Ruess, L., Häggblom, M. M., Zapata, E. J. G., & Dighton, J. (2002). Fatty acids of fungi and
nematodes – possible biomarkers in the soil food chain? Soil Biology and Biochemistry, 34,
745–756.
Rusek, J. (1985). Soil microstructures – Contributions on specific soil organisms. Quaestiones
Entomologicae, 21, 497–514.
Saito, Y. (1997). Sociality and kin selection in Acari. In J. C. Choe & B. Crespi (Eds.), Evolution of
social behaviour in insects and arachnids (pp. 443–457). Cambridge: Cambridge University Press.
Salmane, I., & Brumelis, G. (2008). The importance of the moss layer in sustaining biological
diversity of Gamasina mites in coniferous forest soil. Pedobiologia, 52, 69–76.
Salminen, J., & Haimi, J. (1996). Effects of pentachlorophenol in forest soil: A microcosm experi-
ment for testing ecosystem responses to anthropogenic stress. Biology and Fertility of Soils, 23,
182–188.
Samuel, B. (2004). White as a ghost: Winter ticks & moose. Federation of Alberta Naturalists.
New York: Springer
Santos, P. F., & Whitford, W. G. (1981). The effects of microarthropods on litter decomposition in
a Chihuahuan Desert Ecosystem. Ecology, 62, 664–669.
Santos, P. F., Philips, J., & Whitford, W. G. (1981). The role of mites and nematodes in early stages
of buried litter decomposition in a desert. Ecology, 62, 654–663.
Santos, J. C., Coloma, L. A., & Cannatella, D. C. (2003). Multiple, recurring origins of apose-
matism and diet specialization in poison frogs. Proceedings of the National Academy of
Sciences of USA, 100, 12792–12797.
Sanyal, A. K., Basak, S., & Barman, R. P. (2002). Three new species of oribatid mites (Acarina,
Oribatida: Haplochthoniidae) from the Antarctic continent. Acarina, 10, 57–63.
Saporito, R. A., Spande, T. F., Garraffo, M. H., & Donnelly, M. A. (2009). Arthropod alkaloids in
poison frogs: A review of the ‘dietary hypothesis’. Heterocycles, 79, 277–297.
Saporito, R. A., Norton, R. A., Andriamaharavo, N. R., Garraffo, H. M., & Spande, T. F. (2011).
Alkaloids in the mite Scheloribates laevigatus: Further alkaloids common to oribatid mites and
poison frogs. Journal of Chemical Ecology , 37, 213–218.
Savage, A. L., Moorman, C. E., Gerwin, J. A., et al. (2010). Prey selection by Swainson's Warblers
on the breeding grounds. Condor, 112, 605–614. doi:10.1525/cond.2010.
Sayre, R. M., & Walter, D. E. (1991). Factors influencing the efficacy of natural enemies of plant-
parasitic nematodes. Annual Review of Phytopathology, 29, 149–166.
Schausberger, P. (2003). Cannibalism among phytoseiid mites: A review. Experimental & Applied
Acarology, 29, 173–191.
Schausberger, P., Peneder, S., Jürschik, S., & Hoffmann, D. (2012). Mycorrhiza changes plant
volatiles to attract spider mite enemies. Functional Ecology, 26, 441–449.
Schelvis, J. (1990). The reconstruction of local environments on the basis of remains of oribatid
mites (Acari; Oribatida). Journal of Archaeological Science, 17, 559–571.
Schelvis, J. (1992). Mites and archaeozoology, general methods: Applications to Dutch sites.
Ph.D. Thesis, Rijksunversiteit, Groningen, 116 p.
Scheu, S., & Schulz, E. (1996). Secondary succession, soil formation and development of a diverse
community of oribatids and saprophagous soil macro-invertebrates. Biodiversity and
Conservation, 5, 235–250.
Schmelzle, S., Helfen, L., Norton, R. A., & Heethoff, M. (2009). The ptychoid defensive
mechanism in Euphthiracaroidea (Acari: Oribatida): A comparison of muscular elements
with functional considerations. Arthropod Structure & Development, 38, 461–472.
224 6 Mites in Soil and Litter Systems

Schmid, R. (1988). Morphologische Anpassungen in einem Räuber-Beute-System: Ameisenkäfer


(Scydmaenidae, Staphylinoidea) und gepanzerte Milben (Acari). Zoologische Jahrbuecher
Abteilung fuer Systematik Oekologie und Geographie der Tiere, 115, 207–228.
Schneider, K., & Maraun, M. (2005). Feeding preferences among dark pigmented fungi
(“Dematiacea”) indicate trophic niche differentiation of oribatid mites. Pedobiologia, 49,
61–67.
Schneider, K., Migge, S., Norton, R. A., Scheu, S., Langel, R., Reineking, A., & Maraun, M.
(2004). Trophic niche differentiation in oribatid mites (Oribatida, Acari): Evidence from stable
isotope ratios (15N/14N). Soil Biology and Biochemistry, 36, 1769–1774.
Schneider, K., Renker, C., & Maraun, M. (2005). Oribatid mite (Acari, Oribatida) feeding on ecto-
mycorrhizal fungi. Mycorrhiza, 16, 67–72.
Schubart, H. O. R. (1967). Observations préliminaires sur la biologie d’Indotritia acanthophora
Märkel, 1964 (Acari, Oribatei). Revista Brasileira de Biologia, 27, 165–176.
Schubart, H. O. R. (1973). The occurrence of Nematalycidae (Acari, Prostigmata) in Central
Amazonia with a description of a new genus and species. Acta Amazonica, 3, 53–57.
Schuster, R. (1956). Der Anteil der Oribatiden an den Zersetzungsvorgangen in Boden. Zeitschrift
fur Morphologie und Okologie de Tiere, 45, 1–33.
Schuster, R. (1966). Über den Beutefang des Amiesenkäfers Cephennium austiacum Reiter.
Naturwissenschaften, 53, 113.
Schuster, R. K., & Cootzee, L. (2012). Cysticercoids of Anoplocephala magna (Eucestoda:
Anoplocephalidae) experimentally grown in oribatid mites (Acari: Oribatida. Veterinary
Parasitology, 190, 285–288. doi:10.1016/j.vetpar.2012.05.
Seastedt, T. R. (1984). The role of microarthropods in decomposition and mineralization pro-
cesses. Annual Review of Entomology, 29, 25–46.
Seastedt, T. R., Ramundo, R. A., & Hayes, D. C. (1988). Maximization of densities of soil animals by
foliage herbivory: Empirical evidence, graphical and conceptual models. Oikos, 51, 243–248.
Seeman, O. D., & Walter, D. E. (1997). A new species of Triplogyniidae (Mesostigmata:
Celaenopsoidea) from Australian rainforests. International Journal of Acarology, 23, 49–59.
Seyd, E. L., & Seaward, M. R. D. (1984). The association of oribatid mites with lichens. Zoological
Journal of the Linnean Society, 80, 369–420.
Sharma, R. D. (1971). Studies on the plant-parasitic nematode Tylenchorhynchus dubius. Meded.
Landbouwhogesch. Wageningen, 71, 98–104.
Sharma, G. D., Farrier, M. H., & Drooz, A. T. (1983). Food, life-history, and sexual differences of
Callidosoma metzi Sharma, Drooz, and Treat (Acarina: Erythraeidae). International Journal of
Acarology, 9, 149–155.
Shimano, S. (2011). Aoki’s oribatid-based bioindicator systems. Zoosymposia, 6, 200–209.
Shimano, S., Sakata, T., Mizutani, Y., & Kuwahara, Y. (2002). Geranial: The alarm pheromone
in the nymphal stage of the oribatid mite, Nothrus palustris. Journal of Chemical Ecology ,
28, 1831–1837.
Shimizu, N., Noge, K., Mori, N., Nishida, R., & Kuwahara, Y. (2004). Chemical ecology of astig-
matid mites LXXIII. Neral as an alarm pheromone of the acarid mite, Oulenzia sp. (Astigmata:
Winterschmidtiidae). Journal of the Acarological Society of Japan, 13, 57–64.
Siepel, H. (1990). Niche relationships between two panphytophagous soil mites, Nothrus silvestris
Nicolet (Acari, Oribatida, Nothridae) and Platynothrus peltifer (Koch) (Acari, Oribatida,
Camisiidae). Biology and Fertility of Soils, 9, 139–144.
Siepel, H. (1994). Life-history tactics of soil microarthropods. Biology and Fertility of Soils,
18, 263–278.
Siepel, H. (1995). Applications of microarthropod life-history tactics in nature management and
ecotoxicology. Biology and Fertility of Soils, 19, 75–83.
Siepel, H. (1996). The importance of unpredictable and short-term environmental extremes for
biodiversity in oribatid mites. Biodiversity Letters, 3, 26–34.
Siepel, H., & De Ruiter-Dijkman, E. M. (1993). Feeding guilds of oribatid mites based on their
carbohydrase activities. Soil Biology and Biochemistry, 25, 1491–1497.
Siepel, H., & Maaskamp, F. (1994). Mites of different feeding guilds affect decomposition of
organic matter. Soil Biology and Biochemistry, 26, 1389–1394.
References 225

Simberloff, D. S., & Boecklen, W. (1981). Santa Rosalia reconsidered: Size and competition.
Evolution, 35, 1206–1228.
Simberloff, D., & Dayan, T. (1991). The guild concept and the structure of ecological communities.
Annual Review of Ecology and Systematics, 22, 115–143.
Simon, M. P., & Toft, C. A. (1991). Diet specialization in small vertebrates: Mite-eating in frogs.
Oikos, 61, 263–278.
Skubała, P., & Maslak, M. (2010). Succession of oribatid fauna (Acari, Oribatida) in fallen spruce
trees: Deadwood promotes species and functional diversity. In M. W. Sabelis & J. Bruin (Eds.),
Trends in acarology (pp. 123–128). Proceedings of the 12th International Congress. New York:
Springer
Skubała, P., & Zaleski, T. (2012). Heavy metal sensitivity and bioconcentration in oribatid mites
(Acari, Oribatida): Gradient study in meadow ecosystems. Science of the Total Environment,
414, 364–372.
Skubała, P., Marzec, A., & Sokołowska, M. (2006). Accidental acarophagy: Mites found on fruits,
vegetables and mushrooms. Biological Letters, 43, 249–255.
Smit, C. E., Moser, T., & Roebke, J. (2012). A new OECD test guideline for the predatory soil mite
Hypoaspis aculeifer: Results of an international ring test. Ecotoxicology & Environmental
Safety, 82, 56–62. doi:10.1016/j.ecoenv.2012.05.009.
Smith-Meyer, M. K. P., & Ueckermann, E. A. (1997). A review of some species of the families
Allochaetophoridae, Linotetranidae and Tuckerellidae (Acari: Tetranychoidea). International
Journal of Acarology, 23, 67–92.
Solhøy, I. W., & Solhøy, T. (2000). The fossil oribatid mite fauna (Acari, Oribatida) in late glacial
and early holocene sediments in Krakenes Lake, Western Norway. Journal of Paleolimnology,
23, 35–47.
Spaull, V. W. (1973). Distribution of nematode feeding groups at Signy Island, South Orkney
Islands, with an estimate of their biomass and oxygen consumption. British Antarctic Survey
Bulletin, 37, 21–32.
Sprules, W. G., & Bowerman, J. E. (1988). Omnivory and food chain length in zooplankton food
webs. Ecology, 69, 418–426.
Stamou, G. P., & Argyropoulou, M. D. (1995). A preliminary study on the effect of Cu, Pb and Zn
contamination of soils on community structure and certain life-history traits of oribatids from
urban areas. Experimental & Applied Acarology, 19, 381–390.
Stanton, N. L. (1979). Patterns of species diversity in temperate and tropical litter mites. Ecology,
60, 295–304.
Steiner, W. A. (1995). Influence of air pollution on moss-dwelling animals: 3. Terrestrial fauna,
with emphasis on Oribatida and Collembola. Acarologia, 36, 149–173.
Stephen, J. A., & Schweizer, H. (2009). Biological control of Lycoriella ingenua (Diptera:
Sciaridae) in commercial mushroom (Agaricus bisporus) cultivation: A comparison between
Hypoaspis miles and Steinernema feltiae. Pest Management Science, 65, 1195–1200.
Stewart, M. M., & Woolbright, L. L. (1996). Amphibians. In D. P. Reagan & R. B. Waide (Eds.),
The food web of a tropical rain forest (pp. 273–320). Chicago: University of Chicago Press.
Stirling, G. R. (1991). Biological control of plant parasitic nematodes, progress, problems and
prospects. Wallingford: CAB International.
Strandtmann, R. W. (1967). Terrestrial Prostigmata (trombidiform mites). In J. L. Gressitt (Ed.),
Entomology of Antarctica (pp, Vol. 10, pp. 51–80). Washington, DC: American Geophysical Union.
Swift, M. J., Heal, D. W., & Anderson, J. M. (1979). Decomposition in terrestrial ecosystems.
Berkeley: University of California Press.
Szlendak, E., & Lewandowski, M. (2009). Development and reproductive capacity of the preda-
tory mite Parasitus consanguineus (Acari: Parasitidae) reared on the larval stages of Megaselia
halterata and Lycoriella ingenue. Experimental & Applied Acarology, 47, 285–292.
doi:10.1007/s10493-008-9218-y.
Tevis, L. J., & Newell, I. M. (1962). Studies on the biology and seasonal cycle of the giant red
velvet mite, Dinothrombium pandorae (Acari, Trombidiidae). Ecology, 43, 497–505.
Toft, C. A. (1995). Evolution of diet specialization in poison-dart frogs (Dendrobatidae).
Herpetologica, 51, 202–216.
226 6 Mites in Soil and Litter Systems

Usher, M. B. (1975). Some aspects of the aggregations of soil arthropods: Cryptostigmata.


Pedobiologia, 15, 364–374.
Usher, M. B. (1976). Aggregation response of soil arthropods in relation to the soil environment.
In J. M. Anderson & A. Macfadyen (Eds.), The role of terrestrial and aquatic organisms in
decomposition processes (pp. 61–94). Oxford: Blackwell Scientific Publications.
Usher, M. B., & Booth, R. G. (1986). Arthropod communities in Antarctic moss-turf habitat: Life
history strategies of the prostigmatid mites. Pedobiologia, 29, 209–218.
Usher, M. B., & Bowring, M. F. B. (1984). Laboratory studies of predation by the Antarctic mite
Gamasellus racovitzai (Acari: Mesostigmata). Oecologia, 62, 245–249.
Usher, M. B., & Davis, P. R. (1983). The biology of Hypoaspis aculeifer (Canestrini) (Mesostigmata) – Is
there a tendency towards social-behavior? Acarologia, 24, 243–253.
Usher, M. B., Block, W., & Jumeau, P. J. A. M. (1989). Predation by arthropods in an Antarctic ter-
restrial community. In R. B. Heywood (Ed.), University Research in Antarctica. Proceedings of
British Antarctic Survey Antarctic Special Topic Award Scheme Symposium 9–10 November 1988
(pp. 123–129). Cambridge: British Antarctic Survey, Natural Environment Research Council.
Usher, M. B., Booth, R. G., & Sparkes, K. E. (1982). A review of progress in understanding the
organisation of communities of soil arthropods. Pedobiologia, 23, 126–144.
Utzeri, C., Antonelli, D., & Angelini, C. (2004). A note on terrestrial activity and feeding in the
spectacled salamander, Salamandrina terdigitata (Urodela, Salamandridae). Herpetological
Bulletin, 90, 27–31.
Valdecantos, M. S., Arias, F., & Espinoza, R. E. (2012). Herbivory in Liolaemus poecilochromus, a small,
cold-climate lizard from the Andes of Argentina. Copeia, 2, 203–210. doi:10.1643/CE-12-001.
van de Bund, C. F. (1972). Some observations on predatory action of mites on nematodes. Zeszyty
Problemowe Postepow Nauk Rolniczych, 129, 103–110.
van Straalen, N. M. (1997). Community structure of soil arthropods as a bioindicator of soil health.
In C. Pankhurst, B. M. Doube, & V. V. S. R. Gupta (Eds.), Biological indicators of soil health.
Wallingford: CAB International.
van Straalen, M., & Verhoef, H. A. (1992). The development of a bioindicator system for soil acid-
ity based on arthropod pH preferences. Journal of Applied Ecology, 34, 217–232.
Vänninen, I., & Walter, D. E. (2003). Acceptance of thrips pupae as prey by soil-living predatory
mites. 1st international symposium on biological control of arthropods. USDA-Forest Service
FHTET-03-05 (Poster).
Vincent, W. F. (1988). Microbial ecosystems of Antarctica. Melbourne: Cambridge University
Press.
Visser, S. (1985). Role of the soil invertebrates in determining the composition of soil microbial
communities. In A. H. Fitter, D. Atkinson, D. J. Read, & M. B. Usher (Eds.), Ecological inter-
actions in soil (pp. 297–317). Oxford: Blackwell Scientific Publications.
Wallwork, J. A. (1976). The distribution and diversity of the soil fauna. London: Academic.
Wallwork, J. A. (1983). Oribatids in forest ecosystems. Annual Review of Entomology, 28, 109–130.
Walter, D. E. (1985). The effects of litter type and elevation on colonization of mixed coniferous
litterbags by oribatid mites. Pedobiologia, 28, 383–387.
Walter, D. E. (1987a). Life history, trophic behavior and description of Gamasellodes vermivorax
n. sp. (Mesostigmata: Ascidae) a predator of nematodes and arthropods in semiarid grasslands.
Canadian Journal of Zoology, 65, 1689–1695.
Walter, D. E. (1987b). Trophic behavior of "mycophagous" microarthropods. Ecology, 68, 226–229.
Walter, D. E. (1988a). Predation and mycophagy by endeostigmatid mites (Acariformes:
Prostigmata). Experimental & Applied Acarology, 4, 159–166.
Walter, D. E. (1988b). Macrocheles schaeferi (Acari: Mesostigmata: Macrochelidae), a new
species in the subbadius group from grassland soils in the central United States. Annals of the
Entomological Society of America, 81, 386–394.
Walter, D. E. (2000). A jumping mesostigmatan: Saltiseius hunteri, n.g., n. sp. (Acari:
Mesostigmata: Trigynaspida: Saltiseiidae, n. fam.). International Journal of Acarology, 26,
25–31.
References 227

Walter, D. E., & Behan-Pelletier, V. (1999). Mites in forest canopies: Filling the size distribution
shortfall? Annual Review of Entomology, 44, 1–19.
Walter, D. E., & Ikonen, E. K. (1989). Species, guilds and functional groups: taxonomy and behav-
ior in nematophagous arthropods. Journal of Nematology, 21, 315–327.
Walter, D. E., & Kaplan, D. T. (1990a). Feeding observations on two astigmatic mites, Schwiebea
rocketti Woodring (Acaridae) and Histiostoma bakeri Hughes & Jackson, associated with cit-
rus feeder roots. Pedobiologia, 34, 281–286.
Walter, D. E., & Kaplan, D. T. (1990b). A guild of thelytokous mites associated with citrus roots
in Florida. Environmental Entomology, 19, 1338–1343.
Walter, D. E., & Kaplan, D. T. (1991). Observations on Coleoscirus simplex (Acarina: Prostigmata), a
predatory mite that colonizes greenhouse cultures of rootknot nematode (Meloidogyne spp.), and
a review of feeding behavior in the Cunaxidae. Experimental & Applied Acarology, 12, 47–59.
Walter, D. E., & Lindquist, E. E. (1995). The distributions of parthenogenetic ascid mites (Acari:
Parasitiformes) do not support the biotic uncertainty hypothesis. Experimental and Applied
Acarology, 19, 423–442.
Walter, D. E., & Norton, R. A. (1984). Body size distribution in sympatric oribatid mites (Acari:
Sarcoptiformes) from California pine litter. Pedobiologia, 27, 99–106.
Walter, D. E., & Oliver, J. H. (1990). Geolaelaps oreithyiae, n. sp. (Acari: Laelapidae), a thelytok-
ous predator of arthropods and nematodes, and a discussion of clonal reproduction in the
Mesostigmata. Acarologia, 30, 293–303.
Walter, D. E., & Proctor, H. C. (1998a). Feeding behaviour and phylogeny: Observations on early
derivative Acari. Experimental & Applied Acarology, 22, 39–50.
Walter, D. E., & Proctor, H. C. (1998b). Predatory mites in tropical Australia: Local species rich-
ness and complementarity. Biotropica, 30, 72–81.
Walter, D. E., & Proctor, H. C. (1999). Mites: Ecology, evolution and behaviour (p. 322). Sydney/
Wallingford: University of NSW Press/CABI. ISBN 0 86840 529 9.
Walter, D. E., Hudgens, R. A., & Freckman, D. W. (1986). Consumption of nematodes by fungivorous
mites Tyrophagus spp. (Acarina: Astigmata: Acaridae). Oecologia, 70, 357–361.
Walter, D. E., Hunt, H. W., & Elliott, E. T. (1987a). The influence of prey type on the development
and reproduction of some predatory soil mites. Pedobiologia, 30, 419–424.
Walter, D. E., Kethley, J., & Moore, J. C. (1987b). A heptane flotation method for recovering
microarthropods from arid grassland soils, with comparisons to the Merchant–Crossley
high-gradient extraction method and estimates of microarthropod biomass. Pedobiologia,
30, 221–232.
Walter, D. E., Hunt, H. W., & Elliott, E. T. (1988). Guilds or functional groups? An analysis of
predatory arthropods from a shortgrass prairie soil. Pedobiologia, 31, 247–260.
Walter, D. E., Moore, J. C., & Loring, S. J. (1989). Symphylella sp. (Symphyla: Scolopendrellidae)
predators of arthropods and nematodes in grassland soils. Pedobiologia, 33, 113–116.
Walter, D. E., Halliday, R. B., & Lindquist, E. E. (1993). A review of the genus Asca (Acarina:
Ascidae) in Australia, with the description of three new leaf-inhabiting species. Invertebrate
Taxonomy, 7, 1327–1347.
Walter, D. E., Seeman, O., Rodgers, D., & Kitching, R. L. (1998). Mites in the mist: How unique
is a rainforest canopy knockdown fauna? Australian Journal of Ecology, 23, 501–508.
Walton, B. M., & Steckler, S. (2005). Contrasting effects of salamanders on forest-floor macro-
and mesofauna in laboratory microcosms. Pedobiologia, 49, 51–60.
Walton, B. M., Tsatiris, D., & Rivera-Sostre, M. (2006). Salamanders in forest-floor food webs:
Invertebrate species composition influences top–down effects. Pedobiologia, 50, 313–321.
Wardle, D. A. (2006). The influence of biotic interactions on soil biodiversity. Ecology Letters, 9,
870–886. doi:10.1111/j.1461-0248.2006.00931.x.
Wardle, D. A., Bardgett, R. D., Klironomos, J. N., Setala, H., van der Putten, W. H., & Wall, D. H.
(2004). Ecological linkages between aboveground and belowground biota. Science, 304,
1629–1633.
Wauthy, G., Leponce, M., Banai, N., Sylin, G., & Lions, J. C. (1997). Un acarien qui saute et qui
se met en boule. Comptes rendus de l'Academie des Sciences, 320, 315–317.
228 6 Mites in Soil and Litter Systems

Weeks, P. (2000). Red-billed oxpeckers: Vampires or tickbirds? Behavioral Ecology, 11, 154–160.
doi:10.1093/beheco/11.2.154.
Whitford, W. G. (1996). The importance of the biodiversity of soil biota in arid ecosystems.
Biodiversity and Conservation, 5, 185–195.
Wiggins, E. A., & Curl, E. A. (1979). Interactions of collembola and microflora of the cotton
rhizosphere. Phytopathology, 69, 244–249.
Wilkinson, P. P. (1970). Factors affecting the distribution and abundance of the cattle tick in
Australia: observations and hypotheses. Acarologia, 12, 492–508.
Wilson, D. S. (1983). The effect of population structure on the evolution of mutualism: A field test
involving burying beetles and their phoretic mites. American Naturalist, 121, 851–870.
Wilson, E. O. (2005). Oribatid mite predation by small ants of the genus Pheidole. Insectes
Sociaux, 52, 263–265.
Wohltmann, A., Wendt, F.-E., & Waubke, M. (1996). The life cycle and parasitism of the European
grasshopper mite Eutrombidium trigonum (Hermann 1804) (Prostigmata: Parasitengonae:
Microtrombidiidae), a potential agent for biological control of grasshoppers (Saltatoria).
Experimental & Applied Acarology, 20, 545–561.
Wolters, V. (1991). Soil invertebrates – Effects on nutrient turnover and soil structure – A review.
Zeitschrift für Pflanzenernahrung, Dungung und Bodenkunde, 154, 389–402.
Woodring, J. P. (1963). The nutrition and biology of saprophytic Sarcoptiformes. Advances in
Acarology, 1, 89–111.
Woodring, J. P., & Galbraith, C. A. (1976). The anatomy of the adult uropodid Fuscouropoda
agitans (Arachnida; Acari) with comparative observations on other Acari. Journal of
Morphology, 150, 19–58.
Wright, E. M., & Chambers, R. J. (1994). The biology of the predatory mite Hypoaspis miles
(Acari: Laelapidae), a potential biological control agent of Bradysia paupera (Dipt.: Sciardiae).
Entomophaga, 39, 225–235.
Yamanaka, T., White, P. C. L., Spencer, M., & Raffaelli, D. (2012). Patterns and processes in
abundance-body size relationships for marine benthic invertebrates. Journal of Animal Ecology,
81, 463–471.
Yeates, G. W., & Lee, W. G. (1997). Burning in a New Zealand snow-tussock grassland: Effects on
vegetation and soil fauna. New Zealand Journal of Ecology, 21, 73–79.
Yoder, J. A. (1993). An ant-diversionary secretion of ticks – 1st demonstration of an acarine
allomone. Journal of Insect Physiology, 39, 429–435.
Young, O. P., & Welbourn, W. C. (1987). Biology of Lasioerythraeus johnstoni (Acari: Erythraeidae),
ectoparasitic and predacious on the tarnished plant bug, Lygus lineolaris (Hemiptera: Miridae),
and other arthropods. Annuals of the Entomological Society of America, 80, 243–250.
Young, M. R., Behan-Pelletier, V. M., & Hebert, P. D. N. (2012). Revealing the hyperdiverse mite
fauna of subarctic Canada through DNA barcoding. PLoS One, 7, e48755. doi:10.1371/journal.
pone.0048755.
Zacharda, M. (1980). Soil mites of the family Rhagidiidae (Actinedida: Eupodoidea), Morphology,
systematics, ecology. Acta Universitatis Carolinae-Biologica, 1978, 489–785.
Zhang, Z.-Q., & Sanderson, J. P. (1993). Association of Ereynetes tritonymphs (Acari:
Ereynetidae) with the Fungus Gnat, Bradysia impatiens (Diptera: Sciaridae). International
Journal of Acarology, 19, 179–183.
Chapter 7
Acari Underwater, or, Why Did Mites
Take the Plunge?

If one goes back far enough, all terrestrial arthropods are the descendants of marine
ancestors. For some groups, such as isopods, emergence onto dry land was
evolutionarily recent and these animals often appear imperfectly adapted to solid
ground. For most other groups, however, the split between marine and terrestrial
lifestyles happened hundreds of millions of years ago and all traces of aquatic adaptations
have disappeared. So for the many taxa that have since re-invaded water (Fig. 7.1),
ancestral structures that would have been useful (e.g. gills) had long since vanished
and mechanisms for respiration, feeding, and even mating, under water had to
evolve de novo.

Taxonomic Distribution of Secondarily Aquatic Arthropods

Of the terrestrialised arthropods, insects are the most obvious practitioners of


this nouvelle vie aquatique, although only certain families in the Hemiptera (e.g.
Corixidae, Belostomatidae) have all life stages living subaquatically. In other taxa,
pupae or adults remain terrestrial. Most species live in fresh water; only a very
few groups have returned to the ocean and most of these live on the surface film
(e.g. water-striding bugs) or intertidally (beetles, flies). Just one small family of
caddisflies, Chathamiidae, a couple of genera of chironomid midges, Clunio and
Pontomyia, and one genus of beetles, Haemonia (Chrysomelidae), are known to
spend their larval stages completely immersed in the salty brine (Cheng 1976;
Winterbourn and Anderson 1980; Ward 1994).
Life under water has not attracted many myriapods. Centipedes, symphylans and
pauropods are invariably terrestrial. A number of species live in the rocks and wrack
along beaches but they do not willingly enter water (Roth and Brown 1976). Only a
few species of Australian millipedes have invaded fresh water.
Arachnids for the most part seem obstinately landlocked. Even though the
ancestors of scorpions were aquatic (Chap. 1), extant members of this group have

D.E. Walter and H.C. Proctor, Mites: Ecology, Evolution & Behaviour: Life at a Microscale, 229
DOI 10.1007/978-94-007-7164-2_7, © Springer Science+Business Media Dordrecht 2013
230 7 Acari Underwater, or, Why Did Mites Take the Plunge?

Fig. 7.1 Water mites are unusual among freshwater invertebrates in that many taxa are brightly
coloured, such as these orange Hydrodroma (Hydrodromidae), which contrast both with the
pondweed and with the green water mites (Arrenuridae: Arrenurus) to the right of the image
(photo HC Proctor)

gone to the opposite extreme and are now classic desert dwellers. There are no
aquatic amblypygids, uropygids, schizomids or opilionids. One genus of palpi-
grades has been reported to be able to swim in sea water (see Chap. 2) but this is
unlikely to be its normal habitat. However, many of these taxa include species that
make their livings beachcombing at low tide and hiding in wrack or rock crevices
when the tide comes in (Roth and Brown 1976). Some spiders and pseudoscorpions
are exceptional in that they do not hide from the salty water but rather remain in
their silken nests on the undersides of stones or in cracks at high tide (Harvey 1998).
Although many fishing spiders (Pisauridae) and wolf spiders and their relatives
(Lycosidae, Trechaleidae) can run across the surface of fresh water and may sub-
merge briefly to chase prey or avoid enemies, only one of the more than 40,000
named species of spiders spends an appreciable portion of its life under water. The
European water spider Argyroneta aquatica (Cybaeidae, previously Agelenidae)
swims about and catches prey in freshwater ponds but it requires air for respiration.
The spider builds an underwater bell of silk and ferries bubbles to it, thereby creat-
ing a handy store to replenish the layer of air that encases its abdomen. Argyroneta
must also construct diving bells to eat its prey, to moult, to transfer sperm (both
from the male’s genital opening to his palps and from his palps to the female’s geni-
tal opening) and to lay eggs (Braun 1931; Crome 1951; Schütz and Taborsky 2003).
So much for non-acarine arachnids; what about the mites? As in many other
arachnid groups, there are terrestrial mites that live in the intertidal zone (see Luxton
1990 for a review of their biology). According to Roth and Brown (1976), at least
15 families of Mesostigmata, Oribatida, Astigmata and Prostigmata fall into this
ecological category. Intertidal mites often occur in great numbers: Ernst (1996)
observed densities of nearly 200,000 mites/m2 on the rocky shore of the Weser estuary
Taxonomic Distribution of Secondarily Aquatic Arthropods 231

Fig. 7.2 Neomolgus littoralis


(Bdellidae) is a widespread
mite of seashores
in North America (photo,
H. Proctor)

in Germany. On the coasts of North America one can observe the large, red-bodied,
white-booted Neomolgus littoralis (Bdellidae) in great numbers walking around on
the surfaces of tidepools and shoreline rocks (Fig. 7.2). One should include in this
list those mites parasitic on crabs (Laelapidae), seals and sea otters (Halarachnidae),
and marine iguanas and sea snakes (Trombiculidae). Mites parasitising marine
vertebrates dwell in the lungs and nasal passages of their hosts and so are not aquatic
sensu stricto, and it is also possible that the crab mites live in air spaces. There are
many species of mites, especially prostigmatans like Homocaligidae and many
members of the Cohort Parasitengona, that crawl across the surfaces of freshwater
ponds and bogs but do not submerge themselves. Among the oribatids, many species
in the Camisiidae and Mycobatidae inhabit soggy areas in moss and litter next
to water bodies but they are rarely completely immersed (R. Norton, pers. comm.,
Behan-Pelletier and Eamer 2007).
Are there any truly sub-aquatic acarines? By now you should not be surprised to
find that, unlike their less versatile arachnid relatives, many thousands of species
of mites are true permanent residents of underwater habitats (Proctor 2004). In
fact, almost every freshwater and marine habitat imaginable is home to one or
more species of these adaptable arachnids. Mites live in the water that accumulates
in treeholes, pitcher plants, bamboo internodes, bromeliads and leaf axils – structures
collectively termed ‘phytotelmata’ (e.g. Fashing 1994, 1998, 2010; OConnor
1994; Fashing and Chua 2002). They inhabit standing and running fresh water of
all types, from near-freezing glacial meltwater to steaming hot springs, from
the mossy margins of bogs to the depths of Lake Baikal, from temporary ponds
to the stygean permanency of deep groundwater. Marine mites live in tidepools,
interstitially in sandy beaches and at abyssal depths greater than 4 km. Within these
habitats mites act as predators, parasites, herbivores, fungivores, i.e. everything they
do on dry land. Now let us explore the evolution and ecology of the aquatic mites
in more detail.
232 7 Acari Underwater, or, Why Did Mites Take the Plunge?

Repeated Invasions of Water

There have been at least ten independent invasions of water by different lineages of
mites. In some cases, entry appears quite recent in that there are only one or a few
species in a family living underwater while the rest are land-dwellers (e.g.
Blattisociidae: Mesostigmata). In others, whole superfamilies comprising hundreds
or thousands of species are obligatorily aquatic.

Parasitiformes

Very few parasitiformans have ventured below the water’s surface. There are no
aquatic opilioacarids or holothyrans. There are two species of sea-going Amblyomma
ticks (Ixodidae): Amblyomma nitidum parasitises sea snakes of the genus Laticauda,
and A. darwini parasitises the marine iguana Amblyrhynchus cristatus (Hoogstraal
1973; Keirans, J, pers. comm.). Little is known of their biology except that the ticks
initially infest the reptiles on land; iguanas spend a great deal of time hauled out on
rocks and, while most sea snakes spend their entire lives submerged, Laticauda
spends the day outside of the water and only hunts subaquatically at night (Audy
et al. 1960). Ticks of these marine reptiles presumably open their spiracles to
exchange gases only when their hosts are on land or (for those clustered around the
head) when their hosts surface to breathe.
Aquatic species of Mesostigmata have been reported from the Blattisociidae
(Cheiroseius and Platyseius spp.), Dithinozercondae (Caminella peraphora) and
Uropodidae (Phaulodinychus mitis) (Table 7.1). One species in the Phytoseiidae,
Macroseius biscutatus, has been collected from pitcher plants (Chant et al. 1959)
but it is unclear whether it is subaquatic. It may be like Platyseius italicus, which
inhabits wet vegetation where it is likely to be occasionally submerged and has
evolved a plastron to deal with gas exchange under these circumstances (Hinton
1971). Many predatory mesostigmatans inhabit the marine littoral zone but they
typically retreat to air-filled crevices to avoid drowning (Luxton 1990) or they can
survive for only a few days under water (Pugh 1996a).

Oribatida

Oribatids have been more adventurous in their exploration of water (Behan-Pelletier


1996; Behan-Pelletier and Eamer 2007) (Table 7.1). Numerous species feed on
sodden algae in the marine littoral zone (Luxton 1990). Behan-Pelletier and Eamer
(2007) state that members of 15 genera in 9 families of Oribatida are truly freshwa-
ter species (where ‘truly’ requires that the species feed, locomote and mate when
under water). All 20 or so species in the Hydrozetidae are subaquatic in ponds, lakes
(Krantz and Baker 1982) and streams (Proctor, pers. obs.), and numerous freshwater
Table 7.1 Mites that spend the majority of their lives in subaquatic habitats
Suborder Family Habitats References
Ixodida Ixodidae (F) On marine snakes and lizards Hoogstraal (1973)
Mesostigmata Ascidae (F) Treeholes Kitching and Callaghan (1982)
Dithinozerconidae (F) Moss in mountain streams Krantz (1978)
Repeated Invasions of Water

Uropodidae (F) Benthic littoral marine Krantz (1974)


Oribatida Hydrozetidae (A) On/in macrophytes in standing fresh water Krantz and Baker (1982)
Trhypochthoniidae (F) Benthic in running fresh water Norton et al. (1988)
Camisiidae (F) Benthic in springs Schatz and Gerecke (1996)
Ameronothridae (F) Restricted to ephemeral pools Norton et al. (1996)
Astigmata Acaridae (M) Treeholes; bromeliads; freshwater fish tanks; on leeches Fashing (1994), OConnor (1994),
Proctor et al. (1997)
Algophagidae (M) Treeholes; leaf axils; brackish water Fashing (1994), Naeem (1990)
Histiostomatidae (M) Treeholes; bromeliads; pitcher plants; leaf axils; freshwater Fashing (1994), OConnor (1994),
fish tanks (?); on leeches Proctor et al. (1997)
Hyadesiidae (A) Algae in littoral marine, brackish and freshwater; mussel beds Roth and Brown (1976), Krantz (1978)
Prostigmata Halacaroidea (4 families)(A) Benthic littoral and profundal marine; brackish* del* Bartsch (1989), Harvey (1990)
water; standing and running fresh water; interstitial marine
and freshwater; parasitic on/in many marine invertebrates
and in marine/freshwater decapods
Hydracarina (47 families)(A) Benthic and planktonic in standing and running fresh water, Cook (1974), Smith and Cook (1991)
including hot springs and treeholes; interstitial freshwater;
littoral marine; inside freshwater mussels, snails and crayfish
F one or a few aquatic species in that taxon, M many, A all
233
234 7 Acari Underwater, or, Why Did Mites Take the Plunge?

inhabitants are scattered through the families Trhypochthoniidae and Ameronothridae.


A number of other taxa are found in aquatic systems but appear equally at home in
damp moss and other sodden vegetable matter. These ‘amphibians’ include
Limnozetidae, some Trhypochthoniidae and two genera of Malaconothridae.
Zetomimidae are also associated with shallow waters but anecdotal literature sug-
gests they may live on the water’s surface rather than beneath it. There are a number
of other members of the Ameronothridae and related families (Selenoribatidae,
Fortuyniidae) that are intertidal or estuarine. These are active only when their envi-
ronment is wet not when completely submerged (Behan-Pelletier and Eamer 2007).
Pfingstl (2013) experimentally determined that individuals of the Bermudan species
Alismobates inexpectatus, Fortuynia atlantica and Carinozetes bermudensis were
able to survive from 40 to 143 days submerged in seawater, which would easily
allow them to be transported 1,000’s of km in the Gulf Stream to new habitats. The
sub-Antarctic species Trimalaconothrus flagelliformis (Malaconothridae) and
Edwardzetes elongatus (Ceratozetidae) can survive more than 30 days submerged
but they cannot complete their life cycle under water (Pugh 1996a). Because their
exoskeleton is relatively resistant to decomposition, terrestrial oribatids are often
found in aquatic samples after having tumbled into the water and drowned (Schatz
and Gerecke 1996; Proctor, pers. obs.).

Astigmata

Although their thin integument would seem well suited for aquatic gas exchange,
there are few families of astigmatans with subaquatic species (Table 7.1). The algiv-
orous Hyadesiidae inhabit tidal pools and mussel beds (Roth and Brown 1976;
Luxton 1990). Members of the Histiostomatidae, Acaridae and Algophagidae are
common inhabitants of phytotelmata (OConnor 1994). Histiostomatids and acarids
have also been found as associates of freshwater fish and leeches (Proctor et al.
1997). Some Acaridae are amphibious, e.g. Tyrophagus putrescentiae seems equally
content feeding on fungus in mouldy grain or wallowing through cultures of aquatic
algae (Walter, pers. obs.).

Prostigmata

It is among the prostigmatans that one finds the greatest number of subaquatic spe-
cies, concentrated in two major taxa. The Halacaroidea, with two or four families
(depending on the author) has about a thousand mostly marine species (Bartsch
1996), and the ‘true’ water mites, the Hydracarina or Hydrachnidiae, includes about
6,000 mainly freshwater species in 57 families (Di Sabatino et al. 2008). Halacaroids
are thought to be related to the Bdelloidea, a group of free-living terrestrial preda-
tors. Water mites are members of the Cohort Parasitengona, which are characterised
by a complex life cycle with parasitic larvae and predatory post-larval stages.
Number of Invasions into Different Aquatic Habitats 235

Although earlier researchers considered the Hydracarina to be grossly polyphyletic


(e.g. Pennak 1978), studies of larval morphology (I.M. Smith, pers. comm.), struc-
ture of the male genitalia (Barr 1972) and sperm ultrastructure (Alberti 1980)
support the monophyly of seven superfamilies: Hydrachnoidea, Eylaioidea,
Hydryphantoidea, Hydrovolzioidea, Lebertioidea, Hygrobatoidea and Arrenuroidea.
There has been disagreement about whether two superfamilies of aquatic para-
sitengone mites – the Stygothrombidioidea and Hydrovolzioidea, each with one
family – are water mites sensu stricto or if they represent independent invasions of
fresh water (e.g. Cook 1974). Scanning electron microscopic studies show that the
Hydrovolziidae possess structures that were previously thought missing (genital
acetabula) (Alberti and Bader 1990); this observation has removed much of the
rationale for excluding the group from the Hydracarina. Stygothrombidiidae are a
harder case. Although they have the two defining characters of water mites (two
setae on larval palpal genu, combinations of skin glands and trigger hairs called
glandularia) that are lacking in terrestrial Parasitengona, they appear to retain
trichobothria in the adult – a characteristic that has been lost in the water mites
sensu stricto – and they have a well-developed empodial claw in the adult, another
feature not present in the Hydracarina. Molecular data may help to resolve the
placement of this superfamily. The vagueness about which taxa should be included
in the water mites is reflected in the number of names that have been applied to this
group: Hydrachnellae, Hydrachnidia, Hydrachnida and Hydracarina. We prefer the
last name because it literally means ‘water mites’ and we will use it generously to
refer to all eight superfamilies, including Stygothrombidioidea.

Number of Invasions into Different Aquatic Habitats

This is a difficult number to estimate, in part because it is hard to delimit habitats.


Keeping in mind that these habitats can easily be subdivided into smaller units, let
us consider the following list: phytotelmata, temporary fresh water, permanent
standing fresh water, running fresh water, deep interstitial fresh water, brackish
water, marine intertidal, marine deep interstitial, marine subtidal (including abyssal).
Another reason why it is difficult to enumerate invasions is that in many taxa there
have been repeated independent invasions of the same type of aquatic habitat from
land or from another aqueous habitat. So the tallies below should be taken with a
grain of salt.

Phytotelmata

A phytotelm is a water-holding cavity formed by a plant, such as treeholes, pitcher


plants and leaf axils. Phytotelmata are popular subjects of study for community
ecologists because their boundaries are relatively discrete, can be easily replicated
and contained communities tend to have manageably low diversities (e.g. Naeem
236 7 Acari Underwater, or, Why Did Mites Take the Plunge?

1988; Jenkins and Kitching 1990; Kitching 2000; Clarke and Kitching 1993).
Phytotelmata have been independently colonised by the astigmatan families
Acaridae, Histiostomatidae and Algophagidae, by mesostigmatans (Cheiroseius,
Blattiosociidae) and by the water mites Arrenurus kitchingi and an undescribed spe-
cies in the family Anisitsiellidae (Harvey 1998) (Table 7.1). Some pitcher-plant
mites show specificity to particular species of host plants, even when others of the
same genus are present nearby (Fashing and Chua 2002). Phytotelm mites disperse
by walking (e.g. to nearby pitcher plants, OConnor 1994) or by attaching to insects
(e.g., to ants that also inhabit pitchers, Fashing and Chua 2002). The Cheiroseius
species appears to be amphibious, as we have observed it both walking under water
and on the dry sides of containers (Walter, Proctor, pers. obs.). Other species of
Cheiroseius associated with larger bodies of water feed on the floating eggs of mos-
quitoes and disperse via phoresy on large-bodied flies (Smith 1983) so this may also
be true of the phytotelm species.

Temporary Freshwater Bodies

Ephemeral ponds have been invaded by members of several families of Hydracarina


(Hydryphantidae, Eylaidae, Hydrachnidae, Piersigiidae, Arrenuridae, Pionidae)
(Smith and Cook 1991). Water mites escape desiccation in the dry season by bury-
ing themselves in the mud or when their larvae parasitise flying insects that leave the
pools. Smith and Cook (1991) suggest that temporary water bodies are the ancestral
habitats for most of the water mite families listed above and that only the Arrenuridae
and Pionidae secondarily invaded temporary ponds from permanent water.
At least three species of Oribatida in the Ameronothridae inhabit rock pools that
evaporate completely. When this happens, the mites remain in the arid silt for days
or weeks, until the next rains arrive (Norton et al. 1996). Although the three amer-
onorthrid species share an unusual morphological character (a light-sensing lentic-
ulus), they are widely separated geographically with one species in South Africa,
one in Australia and one in North America. It is therefore unclear whether their
morphological similarity indicates convergence or common derivation from a
Pangaean ancestor.

Standing Fresh Water

Permanent ponds, lakes, swamps and bogs have been colonised by many lineages.
In the Oribatida, the family Hydrozetidae is composed mostly of species that live
on macrophytes in standing fresh water. The ancestral hydrozetid may have been a
mite that fed on lake-shore wrack and gradually got its feet wet, evolutionarily
speaking. In the Astigmata, some species in the family Hyadesiidae have been
described from algae in freshwater habitats (Krantz 1978). As most hyadesiids are
Number of Invasions into Different Aquatic Habitats 237

marine or brackish-water inhabitants, the invasion of fresh water may have been
secondary. In the Prostigmata, Smith and Cook (1991) list 33 subfamilies of
Hydracarina with representatives in these habitats but it is unclear how many inva-
sions this represents.
Also within the Prostigmata, the normally sea-going Halacaroidea have repeat-
edly invaded fresh water. Approximately 50 of the >1,000 species of halacaroids
live in non-marine aquatic habitats (Bartsch and Gerecke 2011). In some cases the
move to fresh water may have been passive. Bartsch (1996) hypothesises that
many lake-dwelling halacaroid species originated from brackish-water mites
whose habitats became land-locked and slowly transformed into fresh water. For
example, the ancestor of Astacopsiphagus parasiticus, which lives in the gills of
the Australian freshwater crayfish Euastacus spinifer, probably parasitised a cray-
fish that inhabited saltwater lagoons during the Mesozoic. Geological activity cut
off the lagoons from the sea in the Tertiary. Both crayfish and mite gradually
became adapted to fresh water as their lagoons became less and less brackish. The
same origin is possible for another halacaroid associate of freshwater crayfish in
Australia, Peza daps (Pezidae).
Other lineages of halacaroids may have invaded standing water by gradually
moving up estuaries and adapting to the lower salinity in a leisurely way. Siemer
(1996) examined the distribution of halacarid species along the salt water–fresh
water gradient in Weser estuary in Germany. Although more numerous near the
ocean, Isobactrus uniscutatus occurred at all salinities, while other species (e.g.
Halacarellus basteri) were restricted to high salinities.

Running Fresh Water

Krantz (1978) describes the mesostigmatan mite Caminella peraphora


(Dithinozerconidae) as being ‘virtually aquatic’ in wet moss from mountain streams
in Oregon. However, as he managed to culture this species on wet filter paper, it
seems to qualify more as ‘amphibious’ rather than ‘aquatic’. Platyseius italicus
(Blattisociidae) can be found in great numbers in sewage filter-beds, which perhaps
can be granted status as ‘flowing-water habitat’ (Learner and Chawner 2003).
Among the Oribatida, two species of Mucronothrus in the family Trhypochthoniidae
are benthic in cold running water (Norton et al. 1988). Most other trhypochthoniids
are terrestrial with a few species being amphibious in water and in sodden littoral
vegetation. For these oribatids, it seems reasonable to suggest a mode of invasion
from terrestrial to amphibious to obligatorily aquatic. Mucronothrus nasalis has
been found in streams, springs and the bottoms of very cold lakes throughout the
world. As M. nasalis has no way of actively dispersing, this distribution suggests
that its ancestor invaded water before the break-up of Pangaea 200 million years ago
(Norton et al. 1988). In a survey of springs in the Bavarian and Italian Alps, Schatz
and Gerecke (1996) found that M. nasalis occupied only permanent springs where
it made up about 45 % of all oribatids collected. Many of the oribatids they found
238 7 Acari Underwater, or, Why Did Mites Take the Plunge?

Fig. 7.3 A flexible elongate


body form such as shown
by this Stygothrombium
sp. (Stygothrombidiidae) is
common among aquatic mites
that live in interstitial habitats
(SEM by H. Proctor)

were clearly accidental specimens from nearby terrestrial and arboreal habitats;
however, Platynothrus peltifer (Camisiidae) was also extremely common in their
samples (34 % of individuals). They concluded that it, too, was a true subaquatic
denizen of springs. Unlike M. nasalis, P. peltifer occasionally occurred in intermit-
tent springs.
Halacaroid species from many genera are found in epibenthic substrates of
springs and rivers (Bartsch 1996). There have probably been multiple invasions of
these habitats when mites moved up estuaries into rivers and perhaps via invasion
of springs through hyporheic habitats. Smith and Cook (1991) list 25 subfamilies of
water mites with representatives in running water. It seems likely that many of these
taxa are recent invaders from standing water (e.g. some Unionicola species, pers. obs.
H.P.) while others are ancient taxa that have been in lotic habitats since Pangaean
times (e.g. Protziinae).

Interstitial Fresh Water

Interstitial freshwater areas, including groundwater and the hyporheic regions of


standing and running water, are home to many hydracarines and halacaroids. There
are about 540 species of water mites known from groundwater. Most of these are
from superficial substrates but a fair number, around 80 species, have been collected
from well water that originates several metres underground (Cook 1991). For both
water mites and halacaroids, there have likely been repeated invasions of interstitial
areas from the sediments of springs and rivers and, for some halacaroids, from
sandy marine beaches. Ancient invaders can be recognised by their lack of body and
eye pigment, while more recent invaders and inhabitants of upper interstitial areas
may retain pigmentation. Another morphological characteristic of many lineages of
interstitial water mites is a wormlike body (e.g. Stygothrombidioidea, Wandesiinae)
(Smith and Cook 1991) (Fig. 7.3) or strongly flattened legs (see illustrations in
Cook 1991, 1992), both presumably adaptations for movement through substrates.
Number of Invasions into Different Aquatic Habitats 239

More recent invaders may show no morphological divergence from their non-
interstitial relatives. Cook (1991) mentions that nothing is known of the life history
of deep-interstitial water mites. He hypothesises that, because aerially dispersing
insect hosts would be difficult to find in these habitats, many species may produce
non-parasitic larvae that transform immediately into predatory deutonymphs.

Brackish Water

Water that is ‘brackish’ is made up of a combination of marine and terrestrial


sources. Vacillating ionic gradients in brackish water make adapting to such habitats
difficult for most organisms. Among mites, halacaroids are the only regular inhabit-
ants of brackish water. Perhaps the greatest diversity occurs in the Black Sea, which
has a salinity range of 18–22 %, in which 34 species of halacarid have been recorded
(Bartsch 1989). They also are common inhabitants of brackish water sand flats.
With the exception of the marine family Pontarachnidae, water mites appear to
be particularly averse to salty water. A few species of Eylais (Eylaidae) are able to
deal with brackish marine water but appear to have no special osmoregulatory adap-
tations (Olomski 1991). Timms (1993) recorded six genera of water mites from
Australian lakes ranging in salinity from 0.1 to 19.3 g/L. Species of Arrenurus,
Diplodontus, Eylais and Hydrachna were recorded from mildly saline lakes
(3–20 g/L), while Limnesia and Piona species did not occur at salinities greater than
0.3 g/L. Timms (1993) states that Eylais and Hydrachna were the most salt tolerant
but were abundant only at relatively low salinities (<0.5 g/L).

Marine Intertidal Zone

Surface substrates of the intertidal zone are home to a rich and diverse assemblage
of terrestrial mites (Pugh et al. 1987; Ernst 1996). Some of these species can survive
submersion in sea water for several days. With regard to truly aquatic mites, it seems
likely that the intertidal zone is the ancestral habitat of halacaroids. Many species
spend their lives in sodden shore algae in the splash zone (Bartsch 1989). An even
greater number of halacaroids dwell in more constantly submerged algae as well as
among crevices in barnacle colonies. Although numerous species inhabit the water-
logged crevices of sandy beaches, they are very rare in silty flats (Bartsch 1989).

Marine Subtidal Zone (Including Abyssal)

Many species of halacaroids that occur in the intertidal area also occur in and on
substrates in the permanently submerged subtidal zone. Here, as well as in the inter-
tidal, halacaroids are often associated with sessile invertebrates. Species in the
240 7 Acari Underwater, or, Why Did Mites Take the Plunge?

halacarid genus Bradyagaue live on colonial hydrozoans and have hind legs adapted
for grasping the stolons of their cnidarian hosts (Bartsch 1989). Although halaca-
roids are most common on the continental shelf to about 200 m, nearly 50 species
have been collected at ocean depths greater than 1,000 m (Bartsch 1989). The deep-
est record is for Bathyhalacarus quadricornis at around 7,000 m (Bartsch 1994).
Although living kilometres beneath the ocean surface would seem to require a very
specialised physiology, few of these species are restricted to the deep sea. In some
cases, mites may have been washed into the deep sea from their normal subtidal
habitats. However, there are two monospecific genera (Colobocerasides,
Werthelloides) and a scattering of other species that are restricted to abyssal depths.
All truly abyssal halacarids lack eyes (Bartsch 1994). A gravid female of one rift-
dwelling species, Copidognathus alvinus, contained an egg that enclosed a well-
developed larva. This suggests that C. alvinus may give birth to live young;
larviparity has also been suggested for some species of littoral sub-Antarctic halaca-
rids (Pugh 1996b).
With only a few exceptions, members of the water mite family Pontarachnidae
occur only in the littoral zone of marine waters (Cook 1974) and, rather bizarrely,
in jellyfish (Smit and Alberti 2010). Pontarachnids are unique among water mite
families, both in occupying completely marine waters and in lacking genital
acetabula (see section “Osmoregulation”, below). The hosts of pontarachnid larvae,
if any, are unknown.

(Pre)Adaptations to Subaquatic Life

At the beginning of this chapter we pointed out how few arachnids other than mites
had taken up aquatic lifestyles. One explanation for this lies in basic physiological
and morphological differences between mites and other arachnids. Perhaps most
important among these are differences in modes of gas exchange and feeding, with
sperm transfer also playing a role. In many cases, modifications that evolved in ter-
restrial mites appear to have allowed their descendants to exploit wet habitats. This
pre-adaptation hypothesis, applicable between both mites and non-mites and within
mite lineages, may explain why some acarine groups have been more successful
under water than others.

Gas Exchange

The ancestral mode of gas exchange for chelicerates is the book gill, as exhibited by
extant horseshoe crabs (Xiphosura). The extinct eurypterids and ancient marine
scorpions also possessed these external leaf-like respiratory structures. The invasion
of land by scorpions was accompanied by the invagination of the book gills into the
ventral cuticle and the narrowing of the respiratory opening, protecting the moist
(Pre)Adaptations to Subaquatic Life 241

interior from the desiccating external atmosphere. Liphistiomorph, mygalomorph


and some other primitive spiders exchange gas solely through book lungs but more
derived spiders have tubular tracheal systems in addition to book lungs. In general,
large-bodied arachnids have book lungs (e.g. Amblypygi, Uropygi) while smaller
ones have an insect-like tracheal system. Palpigrades are the odd arachnid out, in
that they appear to have no gas exchange system other than their integument. The
Acari display a range of respiratory structures that characterise the major lineages.
No mites possess book lungs. Rather, they have tracheal systems whose openings
(= stigmata) differ in number and position. The Opilioacarida have four pairs of
dorso-lateral stigmata, the Holothyrida two or three pairs of lateral ones, the
Mesostigmata one pair of lateral stigmata, Prostigmata one pair of prodorsal open-
ings, and the Astigmata and Oribatida (previously called ‘Cryptostigmata’) lack
obvious stigmatal pores (although they are indeed present in the more derived
groups of oribatids, albeit well hidden). Very small mites often have no obvious
structures for gas exchange.
Why do stigmatal/tracheal systems allow mites to exploit water? Alone, they do
not, but combined with their overall smaller size, they do. If an air-breathing arthro-
pod is submerged, its respiratory tubes will fill with water if the openings to these
tubes are large. Small animals, having a smaller mass of respiring tissues, can make
do with very small openings. The smaller the opening, the higher the ratio between
the surface tension of the air/water interface at the spiracular opening and the exter-
nal water pressure, and the less likely it is that water will enter the respiratory tubes.
Spiders that regularly spend time under water avoid drowning by maintaining a
layer of air around their respiratory openings that is supported by special hairs.
This layer serves as a compressible physical gill. Oxygen dissolved in the water
diffuses into the air layer and thence into the spiracular/book lung opening, and
CO2 diffuses out, as does the remaining nitrogen in the layer. Eventually the nitro-
gen is depleted and must be replaced, either by the spider returning to the surface
or to an underwater supply in the case of Argyroneta. Because of the small size and
low oxygen requirements of most terrestrial mites, their respiratory systems often
act as physical gills even without special modifications (Hinton 1971). This allows
terrestrial mites a few minutes grace to escape from flooded areas before their oxy-
gen supply expires. Many marine littoral mites have no clear morphological adap-
tations to their soggy environment (Luxton 1990). For example, littoral
ameronothrid oribatids rely on air held in their extensive tracheal systems when
submerged (Krantz 1978). If a mite wishes to remain submerged for long periods
of time, it would seem a simple matter (evolutionarily speaking) to make the air-
holding cuticular surface wider to provide a greater surface area for gas exchange
by the physical gill. This appears to be just what several aquatic and semi-aquatic
mesostigmatans and oribatids have done.
Hinton (1971) describes the physical gill of Platyseius italicus (Ascidae), a
mesostigmatic inhabitant of marshes and other damp situations. The peritremes of
most mesostigmatans are circular in cross-section, with only a narrow slit running
along the external surface to connect the peritrematal duct to the outside atmosphere
(as if one had taken a long thin strip from a tube). Hinton (1971) suggests that the
242 7 Acari Underwater, or, Why Did Mites Take the Plunge?

narrow aperture prevents large particles from clogging up the stigmatal opening. In
the semi-aquatic P. italicus, a cross-section of the peritreme appears semi-circular;
so the surface area available for gas exchange in this mite is much larger than in its
terrestrial relatives. A lining of microtrichiae presumably helps to keep the layer of
air attached to the peritreme. Krantz (1974) describes a very similar physical gill in
the intertidal uropodid Phaulodinychus miti. Pugh et al. (1987) suggest that the
structure of the leg bases of Phaulodinychus repleta and the coarsely sculptured
integument of Thinozercon michaeli (Thinozerconidae) may capture air bubbles
that act as temporary physical gills when these littoral marine mites are submerged.
The aquatic oribatid Hydrozetes also has an arrangement of micropapillae near the
stigmata that appear to retain a thin layer of air. This layer can be maintained through
a huge range of air pressures (−98 KPa to +550 KPa, Behan-Pelletier and Eamer
2007). Krantz and Baker (1982) note that unsclerotised juvenile Hydrozetes lack
this physical gill, and presumably exchange gases through their thin integument.
Krantz (1974) noted that submerged Phaulodinychus mitis raise the dorsal surface of
their bodies to contact the air/water interface and called this behaviour ‘plastron
capture’. Strictly, however, a plastron is an arrangement of fine hairs that holds an
incompressible physical gill that does not require replenishment (D. Craig, University of
Alberta, pers. comm.). Hinton (1971) states that Platyseius italicus can stay submerged
indefinitely; this suggests that the physical gill borne by P. italicus’ microtrichiae is
incompressible, making the mite more legitimately a plastron-bearer. But many aquatic
mites either have not been observed to make regular trips for air or else live in places
where such journeys would be impossible (interstices or abyssal). These species –
astigmatans, water mites, and halacaroids – undoubtedly respire dermally.
Some of the larger water mites such as Hydrachna and Eylais have complex
systems of closed gas-filled tracheoles that lie under the epidermis and swaddle
metabolically active tissues like the brain and leg muscles (Wiles 1984; Proctor,
pers. obs.). In highly sclerotised water mites such as Arrenurus, the armour is punc-
tuated with thinner areas of integument that allow loops of tracheoles to lie more
proximately to the surface of the mite. Thick-skinned water mites often fan their
hind legs over their dorsal surfaces, presumably to ‘freshen’ the oxygen supply. The
tracheal system of these water mites does not connect functionally with the stig-
mata, which are closed and non-operative (Mitchell 1972; Wiles 1984). Presumably,
the same holds true for the stigmatal openings of halacaroids.
Relative to many co-occurring insects and crustaceans, water mites are highly
resistant to low oxygen levels; however, they may exhibit adaptations to the O2
content of their normal habitat. Young and Rhodes (1974) compared the influence of
different oxygen levels on several vital parameters of one lentic and two lotic species
of water mites. They found that the lentic mite Hydryphantes tenuabilis
(Hydryphantidae) was an O2 regulator (i.e. had a metabolic rate that was constant
over a wide range of O2 concentrations). The lotic mites, Limnesia maculata
(Limnesiidae) and Lebertia quinquemaculosa (Lebertiidae), appeared to alter their O2
consumption when the environmental concentration of the gas changed (i.e. were
O2 conformers). The activity rate of L. quinquemaculosa was greatly slowed by low
O2, L. undulata was affected to a lesser degree, and H. tenuabilis did not alter its
(Pre)Adaptations to Subaquatic Life 243

behaviour at all. The authors felt that these differences may result from H. tenuabilis’
evolution in ponds that experience radical changes in O2 concentrations compared to
the more uniform conditions of the streams the other two species inhabit.

Feeding

With the exception of some harvestmen (Opiliones), all non-acarine arachnids feed on
liquefied prey that they have digested externally. Usually this prey is macerated with
the chelicerae and bathed in salivary products, and the resulting goo is sucked in
through setae that strain out unliquefied matter. As Preston-Mafham and Preston-
Mafham (1984) point out, this mode of external digestion is unsuited to subaquatic
habitats. Digestive enzymes would be diluted to insipidness and the surface tension
of liquefied prey (which allows terrestrial arachnids to ‘sip’ the fluids) would not hold
the dinner together. So the only truly aquatic spider, Agyroneta aquatica, must drag
prey into its air-filled diving bell to eat it. The only way for an arachnid to eat under
water is to swallow whole particles or to somehow avoid dilution of its digestive
enzymes and loss of digested prey. Mites do both.
Particle-engulfing is the normal mode of feeding in the Sarcoptiformes, so it is
not surprising that many sarcoptiformans live subaquatic lives. In contrast, prostig-
matans are fluid feeders that externally digest their food, usually liquefying their
food without macerating it beforehand. Rather, they suck the soup through a punc-
ture hole in the epidermis of their prey or host. The terrestrial relatives of water
mites have this form of feeding and so it is probably ancestral to water mites’ inva-
sion of aquatic habitats. Below we describe in greater detail the feeding behaviour
of aquatic mites, with the exclusion of mesostigmatans, as the diets are not known
for any other than Cheiroseius and Platyseius (mosquito eggs, Smith 1983).

Oribatida and Astigmata: Same Story, Different Substrate

Like their land-locked relatives, aquatic oribatids feed primarily on fungus and
decaying plant matter. Hydrozetes are often found inside stems of aquatic plants and
on damaged and deteriorating leaves but it is not known whether their feeding
causes the damage or if they simply exploit decaying wounds (Krantz 1978; Norton
1994). Norton et al. (1988) examined the gut contents of the widely distributed
spring-dwelling species Mucronothrus nasalis. Adults and immature mites from
Canadian, French, Australian and New Zealand populations contained a mixture of
filamentous algae, diatoms, hyphae, decayed vascular plants and the occasional bit
of arthropod cuticle, suggesting that these mites graze on the periphyton growing on
benthic substrates. An interesting exception to the typical vegetable diet is shown by
a species of ameronothrid that includes bdelloid rotifers in its meals (Norton et al.
1996). These mites have no obvious adaptations for capturing moving prey and it is
likely that the hapless rotifers are engulfed along with detritus as the mites graze.
244 7 Acari Underwater, or, Why Did Mites Take the Plunge?

Most aquatic astigmatans have also maintained diets similar to those of their
terrestrial progenitors, feeding on fungus, decaying plants and bacteria. Fashing
(1994, 1998) compared the feeding methods of three treehole-inhabiting species.
Naiadacarus arboricola (Acaridae) uses its powerful chelicerae to bite off
chunks of decaying leaves and ingests them whole. Algophagus pennsylvanicus
(Algophagidae) uses its more gracile mouthparts to scrape fungus from leaf surfaces.
Hormosianoetus mallotae (Histiostomatidae) has serrate chelicerae that filter
suspended bacteria and bits of detritus that it stirs up using its whip-like palps.
Judging by their gut contents, hyadesiids and other species of freshwater and marine
algophagids are algivores (Fain and Synnot 1981; OConnor 1994).
There are numerous records of aquatic acarids and histiostomatids, and even
oribatids, collected from the bodies of fish, both wild and farmed, and one report of
these mites inhabiting the integument of leeches (Proctor et al. 1997; Olmeda et al.
2011). Although this may appear to represent parasitism, in the case of the leech and
in many of the fish, the ‘host’ animals were diseased and it appears more likely that
the mites were feeding on bacteria and fungus associated with necrotic flesh.

Predation by Water Mites

Let us leave the slime-sucking and vegetarian mites behind and investigate the diets
of predatory aquatic mites, beginning with the Hydracarina. Do their diets differ
much from those of their terrestrial relatives? Water mites belong to the cohort
Parasitengona. Terrestrial parasitengone species are parasitic on arthropods or ver-
tebrates as larvae and predatory on arthropods as nymphs and adults. Insects, par-
ticularly the egg stage, are the main prey, while mites are consumed by some species
of terrestrial parasitengones (Putman 1970; Chen and Zhang 1991; Wohltmann
1996). Often a younger life-stage of the larval host serves as prey for post-larval
stages (A. Wohltmann pers. comm.).
Although there is controversy about which lineage of terrestrial parasitengones
is most closely related to water mites (e.g. Witte 1991 vs. Welbourn 1991), the ear-
lier derivative groups of water mites have retained the egg-eating habit and, like
their terrestrial relatives, often prey on the eggs of those insects they parasitise in
their larval stage (Proctor and Pritchard 1989). Hydryphantids and hydrovolziids
will feed on eggs of other water mite species, sometimes including their own (Martin
2005). This diet is reflected morphologically in the chelate palps of egg-eating mites
in the families Hydryphantidae, Hydrodromidae and Hydrachnidae (Fig. 7.4a–c).
Palps with a claw formed by a dorsal extension of the palp tibia (the ‘palpal thumb-
claw complex’) are also a characteristic of many terrestrial parasitengones and this
feature has presumably been retained in primitive water mites, although it is also
possible that the structure has independently evolved. More derived water mite taxa
have lost this chelate tip, although similar-looking uncate palps (Fig. 7.4d) have
evolved in Arrenurus (Arrenuridae) to grasp the slim appendages of their crustacean
prey. For those species with linear palps, the presence of sharp spines or tubercles
(Pre)Adaptations to Subaquatic Life 245

Fig. 7.4 Water mite palps: (a) Euthyas (Hydryphantidae), (b) Hydrodroma (Hydrodromidae),
(c) Hydrachna (Hydrachnidae), (d) Arrenurus (Arrenuridae), (e) Eylais (Eylaidae), (f) Piona
(Pionidae), (g) Notosperchonopsis (Sperchontidae), (h) Limnesia (Limnesiidae), (i) Uchidastygacarus
(Chappuisididae) (after various illustrations in Cook 1974)

on the ventral side of the second palpal segment (Fig. 7.4g, h) is correlated with a
diet of insect larvae and lack of such tubercles (Fig. 7.4e, f) with a diet of small
crustaceans such as Daphnia.
Even among closely related mites, palp morphology may indicate different diets.
Ullrich (1976) found that Sperchon setiger (Sperchontidae) can prey on vertically
oriented simuliid larvae because it can spread its palps sideways, whereas most
other Sperchon species must turn their bodies sideways to gain purchase. Another
sperchontid, Sperchonopsis verrucosa, is a sit-and-wait predator that stands with its
disproportionately huge palps spread and lifted, waiting for dipteran larvae to crawl
underneath (Ullrich 1976; Martin 2005). This ambush behaviour is similar to that
shown by terrestrial mites in the family Cunaxidae (see Chap. 6). Not all palps can
be readily correlated with diet, however, and direct observations are needed to
explain the function of odd structures. As well as eating the eggs, larvae and pupae
of aquatic insects, water mites prey on small crustaceans, including cladocerans,
ostracods and copepods, on other water mites and sometimes on non-arthropods
such as rotifers, nematodes and oligochaetes.
246 7 Acari Underwater, or, Why Did Mites Take the Plunge?

Fig. 7.5 Hydrodroma


(Hydrodromidae) reaching
into the jelly coat of a
chironomid egg mass
(drawing by HC Proctor)

Fig. 7.6 An Eylais


(Eylaidae) mite feeding
on a daphniid crustacean
(drawing by HC Proctor)

How do water mites catch their prey? Egg-eaters are probably attracted to the
surface chemical of the jelly coat surrounding their food (Fig. 7.5). Wiles (1982)
found that Hydrodroma despiciens (Hydrodromidae) repeatedly probed glass beads
embedded in the jelly of a chironomid egg clutch but showed no interest in chirono-
mid eggs removed from their gelatinous matrix. Hydrachna (Hydrachnidae) feeds
on insect eggs including those that are laid endophytically (e.g. many Odonata) and
avidly investigates small wounds in plant stems by inserting its long, curved ros-
trum. This behaviour led early investigators to assume that Hydrachna was herbivo-
rous (see Böttger 1970). Davids and Belier (1974) found that male Hydrachna
conjecta ate 2–4 corixid eggs a day.
Strongly swimming mites in the families Eylaidae and Pionidae are voracious
predators of small planktonic crustaceans (Fig. 7.6). Ellis-Adam and Davids (1970)
observed Piona alpicola (Pionidae) as it hunted and fed. The mites refused to eat
ostracods or mayfly nymphs. Daphnia, however, were accepted. They occasionally
caught copepods but these prey items often succeeded in escaping from their grasp.
Mites that had not been fed for several days swam about actively and grasped any
(Pre)Adaptations to Subaquatic Life 247

Fig. 7.7 Unionicola


crassipes (Unionicolidae) in
net-stance posture, orienting
to the vibrations of a
swimming Daphnia (drawing
by HC Proctor)

floating particle they came upon. Once a Daphnia was captured, the mite manipulated
it with its legs until its palps achieved a good grip behind the cladoceran’s head.
Immediately after assuming this final position, P. alpicola injected digestive juices
into the prey through a hole torn in its carapace. Mitchell (1972) felt that in compari-
son with terrestrial mites, the great development of the oral digestive glands in water
mites, together with the reduction in musculature in the pharyngeal pump, suggests
that hydrachnids have extremely effective external digestion. Ellis-Adam and Davids
(1970) estimated P. alpicola’s consumption of daphniids to be up to 10 a day.
Riessen (1982) outlined a similar series of events for prey capture by Piona constricta.
Riessen postulated that P. constricta had a low preference for Mesocyclops because
of the copepod’s ability to generate brief spurts of speed (up to 50 cm/s) that allow
the prey to easily outdistance the water mite (whose top speed is 2.5 cm/s). Unlike
Ellis-Adam and David’s observation that copepods often escaped P. alpicola’s grasp,
Riessen found that once caught by P. constricta, Mesocyclops never escaped.
Riessen (1982) determined that two of the favourite prey of Piona constricta
adults (and the only prey of the nymphs) were the cladocerans Bosmina and Chydorus.
He felt that the ‘dead-man’ response of these crustaceans (sinking passively through
the water once disturbed) makes them ideal prey for the mites, especially for the
smaller and slower nymphs. Chydorus was slightly less preferred; this could be due
to its thick carapace which may withstand puncture by Piona’s chelicerae.
Mites in the family Unionicolidae also prey on small crustaceans but, rather than
the above ‘bump-and-grab’ technique, they employ a subtler style of prey capture.
Species of Unionicola and Neumania adopt a special posture termed ‘net-stance’
when hunting, in which the first two pairs of legs are lifted from the substrate and
held outstretched (Fig. 7.7). With the exception of endoparasitic Unionicola (see
section “Parasitism in Post-Larval Water Mites”), the forelegs of these mites are
248 7 Acari Underwater, or, Why Did Mites Take the Plunge?

Fig. 7.8 A Limnochares


(Limnocharidae) waits
patiently for a tube-dwelling
chironomid to move into
striking range (drawing by
H. Proctor)

armed with a battery of spines similar to those of other sit-and-wait acarines (e.g.
rake-legged mites, Caeculidae). The vibrations of passing prey cause the mites to
orient towards and clutch at the stimulus. Proctor and Pritchard (1990) found that
Unionicola crassipes would respond to the vibrating tip of a glass probe as if it were
a prey animal. Male Neumania appear to have taken advantage of the female’s
predatory sensitivity to vibrations by incorporating leg vibrations into their own
courtship (see Chap. 5). Free-living Unionicola species appear to prefer cladocerans
as prey, whereas Neumania prefer copepods (Proctor, pers. obs.).
Crustaceans killed by mites are easily differentiated from those that died another
way. Ellis-Adam and Davids (1970) found that mite-killed daphniids were totally
transparent save for the dark-brown gut. Riessen (1982) also found some dark parts
undigested in Daphnia, while Bosmina was rendered totally transparent. All prey
showed characteristic puncture marks from the chelicerae. Eggs and embryos
in prey animals appear to be undigested, as Riessen observed juvenile Daphnia
escaping from the brood chamber of their mother’s empty husk.
Many water mites are fond of dipteran larvae for supper (Proctor and Pritchard
1989; Martin 2005). Gerecke and Smith (1993) report that Nilotonia longipora
(Anisitsiellidae) feed only on freshly killed larvae and make no attempt to capture
prey themselves; however, the vast majority of mites that feed on dipteran larvae kill
their own prey. The mode of prey capture depends in part on where the larvae live.
Exposed prey (e.g. mosquito larvae and pupae) are encountered and attacked in
the water column apparently by chance. Once they are attacked by one water mite,
long, thin larvae such as chironomids or chaoborids may be swarmed by hungry
hydracarines until there is no space free for another pair of chelicerae (Davids et al.
1981, Proctor pers. obs.).
Tube-dwelling chironomids require a bit more finesse to extract. Limnochares
aquatica (Limnocharidae) is a large, red mite with extrusible mouthparts that has a
unique method of preying on these larvae. The mite slowly approaches a chironomid
tube and walks towards the head end (Fig. 7.8). Presumably, ventilatory currents
created by the encased larva indicate to the mite which end is which. Once at the
(Pre)Adaptations to Subaquatic Life 249

open end of the tube, L. aquatica waits patiently for the larva to extend its head to
graze (Böttger 1969). The mite then shoots out its extendible gnathosoma and grips
the larva’s head capsule with its short, hooked chelicerae. A wild thrashing then
ensues but the mite rarely loses its hold. The larva’s struggles weaken and eventually
cease as its brain is digested, allowing Limnochares to slurp its supper in peace.
Encased chironomid larvae are difficult for less stealthy water mites to extract.
However, such mites sometimes have unwitting vertebrate allies. Ten Winkel and
Davids (1985) observed that the density of Cladotanytarsus mancus, a chironomid
that builds a sturdy sand tube, was significantly lower in areas exposed to bottom-
feeding cyprinid fish than in areas protected by fish exclosures. They ruled out direct
predation by fish because the cyprinids’ ability to sort the small larvae from larger
bits of substrate was very low (about 1 % efficiency). The only other predatory ani-
mals in the man-made lake were two species of water mites, Hygrobates nigro-
maculatus and H. trigonicus (Hygrobatidae). Laboratory studies had shown that
these mites had difficulty preying on the well-protected larvae and captured, on
average, only one larva every 14 days. But when ten Winkel and Davids imitated the
feeding behaviour of cyprinids by sucking up and expelling the substrate with a
pipette, the rate of prey capture by lab-held mites doubled. The authors suggest that
in the field, where bottom-feeding fish sort through 300–400 % of the lake’s sub-
strate annually, water mites rely upon chironomids shaken from their tubes by fish
as their primary food source in the winter. During the summer, daphniids and tube-
less chironomids (Stichtochironomus histrio) are also fed upon by H. nigromaculatus
(ten Winkel et al. 1989).
How do Hygrobates hunt for chironomids when fish are not aiding them? Ten
Winkel et al. (1989) observed an interesting behaviour exhibited by these mites:
burrowing. Hygrobates trigonicus was found to burrow to depths of up to 15 mm in
quest of buried chironomid larvae, while H. nigromaculatus burrowed to a maxi-
mum of 5 mm in the sandy sediments. The chironomid specialist, H. trigonicus,
burrowed irrespective of the presence of prey but H. nigromaculatus only burrowed
under certain circumstances: if alternative prey (Daphnia) were unavailable and if
chemical cues (from real prey or crushed homogenate) indicated that chironomids
were present. The tubeless chironomid Stichtotanytarsus is favoured by these mites.
Similarly, Paterson (1970) found that Piona carnea (Pionidae) had difficulty attack-
ing Glyptotendipes larvae when they were in their silken tubes but readily fed on
those that had been evicted.
Water mites occasionally prey on each other, especially in crowded conditions.
Davids (1973) observed adult Piona coccinea (Pionidae) eating nymphs of Eylais
(Eylaidae) while leaving those of Hydrachna (Hydrachnidae) alone. The author felt
that this discrimination was based on the much softer and more easily pierced skin
of Eylais. Elton (1923) mentioned that Piona longipalpis attacks Limnesia fulgida
(Limnesiidae) and Hydrachna fuscata, and that L. fulgida in turn will prey on Eylais.
Martin (2005) observed a number of lotic water mite taxa feeding on eggs of their
own and other water mite species. They preferred freshly laid eggs to older ones,
possibly because in older clutches the protective clutch-coating (Sokolow 1977) is
better developed.
250 7 Acari Underwater, or, Why Did Mites Take the Plunge?

What effect does water mite feeding have on prey populations? As Proctor and
Pritchard (1989) and Martin (2005) note, there has been little research on the role of
water mites as predators in freshwater communities. However, a few studies of sim-
ple benthic and planktonic food webs have indicated that water mites may some-
times be the top predator. Ten Winkel et al. (1989) estimate that Hygrobates spp. in
the Dutch man-made Lake Maarsseveen reached densities of up to 1,000 mites/m2
of substrate. In winter, when no other prey are available, Hygrobates feeding reduces
the density of the chironomid Cladotanytasus from 28,000 to 12,500/m2. In con-
trast, in caged areas (where dislodgement of chironomids through fish feeding did
not occur) Cladotanytarsus density increased. Mwango et al. (1995, as cited in
Martin 2005) used consumption rates and mite densities in a British stream to esti-
mate that Hygrobates fluviatilis and Lebertia porosa might consume up to 14,000
black fly larvae per generation.
Gliwicz and Biesiadka (1975) discuss the role of Piona limnetica (Pionidae) in
the planktonic food web of a neotropical man-made lake. These mites reached den-
sities of about 150/m3 in the plankton. Laboratory studies showed that feeding rate
on cladocerans was about 3–4 prey/mite/day; at that rate, the mite population would
consume about 50 % of the standing crop of cladocerans per week. However, this
latter study did not have a field component in which the actual decline in prey num-
bers in mite-present and mite-absent enclosures was determined. Other evidence of
potentially high rates of predation come from similar studies of field densities and/
or laboratory observations of prey consumption rates. Riessen (1982) determined
that Piona constricta (Pionidae) reaches densities of about 100/m3 in a Canadian
lake and consumes up to 10 daphniids/day. Laird (1947) describes Limnesia jamu-
rensis (Limnesiidae) as eating 6–8 larvae and up to 18 eggs of mosquitoes per day.
Rajendran and Prasad (1989) calculate the daily predation rate of Encentridophorus
similis (Unionicolidae) to be 40 mosquito larvae/day. Davids et al. (1981) found that
at 24 °C, both Hygrobates longipalpis (Hygrobatidae) and Limnesia undulata
(Limnesiidae) ate one chironomid larva every 2 days. They observed that Unionicola
crassipes (Unionicolidae) would kill on average 18 copepods/day. Other observa-
tions of U. crassipes revealed that mites would kill up to 20 cladocerans equal to
their own body size in 6 h (= 80/day) (Proctor, unpublished data).
Egg-eating mites have also have large appetites. Davids (1991) mentions that an
adult Hydrachna conjecta (Hydrachnidae) consumed about 200 corixid eggs in its
lifespan, which is greater than the total lifetime’s egg production of a female corixid
(140 eggs/female). He suggested that the total loss of corixids to egg predation at his
study site would be at least 30 %. Wiles (1982) observed one female Hydrodroma
despiciens (Hydrodromidae) consume 235 chironomid eggs in 12 h. The high rates
of predation on insect eggs and larvae have led some to consider their use in biologi-
cal control of mosquitoes (Smith 1983; Rajendran and Prasad 1989).
But for several reasons, laboratory studies of predation rates must be taken with
a grain of salt. First, mites may be provided with prey that are normally not encoun-
tered (e.g. planktonic prey presented to a benthic mite), prey that are not allowed
their normal anti-predator behaviour (e.g. chironomids removed from their tubes)
and that are often presented in unrealistic densities. It is likely that in such situations,
(Pre)Adaptations to Subaquatic Life 251

much of the prey mortality is ‘wasteful killing’, in which many prey are killed but
only partially consumed (Johnson et al. 1975). Certainly, if a U. crassipes in the
above example fully consumed the contents of 20 cladocerans in an hour, it would
explode. It is also necessary to maintain the mites in realistic thermal regimes, as
predation rates can vary up to five-fold over a temperature change of 18 °C (Davids
et al. 1981). More field-based studies using mite-present and mite-absent treatments
(such as Ten Winkel and Davids (1985) indirectly achieved) are necessary before
water mites can be considered major controllers of prey populations. However,
mites have one well-studied characteristic that suggests they may share top predator
rank with large piscivorous fish: almost nothing will eat them (see section “Predation:
The Correlation Between Foul Taste and Bright Colour”).

Parasitism in Post-Larval Water Mites

A few groups of water mites have eschewed their ancestral predatory behaviour and
become parasites in their postlarval as well as larval stages. Gravid female
Parasitalbia sumatrensis (Aturidae) attach to the aquatic naiads of ephemerid may-
flies where they become greatly engorged (Cook 1974). Astacocroton molle
(Astacocrotonidae) appears to be parasitic in the gill chambers of an Australian
crayfish (Cook 1974).
Almost all other water mites parasitic during postlarval stages have molluscs as
hosts. The hygrobatid Dockovdia cookarum lives in the mantle cavity of stream-
dwelling snails in Nigeria (Agbolade and Odaibo 2004). In the Pionidae, Najadicola
ingens is parasitic in the gills and pericardial cavity of a wide range of freshwater
mussels (Simmons and Smith 1984). Both adults and deutonymphs inhabit the hosts
and cannot survive when removed. When numerous mites inhabit a single mussel,
the suprabranchial chamber is occupied by a reproductively mature male–female
pair and non-reproductive mites skulk near the gill margin. Simmons and Smith
(1984) suggest that larval N. ingens may transform to deutonymphs without an
intermediate parasitic phase on an insect host; if it does indeed lack an aerially dis-
persing larva, then the geographically distant records from North American and
Thailand indicate that this species may have been present when these land masses
were contiguous (if N. ingens in these two areas are indeed conspecific).
The type genus of the Unionicolidae, Unionicola, is so named because many
species inhabit the mantle cavities of freshwater mussels in the family Unionidae.
More than 50 species of mussel-dwelling unionicolids are known, and six species
from mantle cavities of freshwater snails (Viets 1980; Gledhill 1985; Edwards and
Vidrine 2013). Not all Unionicola species are parasitic; rather, some species merely
use the tissues of other animals (mussels, sponges) as safe, well-oxygenated places
in which to deposit eggs and to pass calyptostatic instars (protonymph, deutonymph)
(Hevers 1980). Three other unionicolid genera parasitise bivalves from the
Unionidae (Vietsatax, Atacella) and Hyriidae (Unionicolopsis) (Cook 1974; Viets
1980). Unlike Najadicola and host-inhabiting halacaroids (see section “Halacaroids
Will Eat Anything”), there is good evidence that Unionicola species act as parasites
252 7 Acari Underwater, or, Why Did Mites Take the Plunge?

(i.e. damage their hosts). Using histological stains, Baker (1976, 1977) showed that
U. intermedia feeds on the blood cells and mucus of its host mussel, Anodonta ana-
tina, and that gill tissue at the site of mite attachment becomes ruptured and swol-
len. Similarly, Fisher et al. (2000) used histological and immunological methods to
determine that U. formosa consumes the mucus and tissues of Pyganodon cata-
racta. The palps of mussel parasites are highly modified relative to free-living
Unionicola and seem to be adapted for tearing rather than for clutching and restrain-
ing small prey (Fig. 7.4).
Inhabiting mussels appears to have drastically altered the sex life of parasitic
unionicolids. All free-living Unionicola for which mating has been observed trans-
fer sperm without copulating; in contrast, direct sperm transfer has evolved repeat-
edly within mussel-dwelling taxa (Proctor 1991). It is possible that the slimy surface
of mussel flesh is a poor substrate for spermatophores; perhaps this has selected for
sperm-transfer through copulation in parasitic species. Another dramatic change in
reproductive behaviour is in male–male interactions. Among non-parasitic union-
icolids, male sexual competition is directed against spermatophores, whereby males
trample or deposit their own spermatophores on those of rival males (Proctor
1992a). But many species of mussel-parasitic Unionicola have highly aggressive
males that attack and attempt to kill rival males entering their mussel (Dimock
1983; Davids et al. 1988; Edwards and Dimock 1991) (see Chap. 5). This behaviour
can result in dramatically female-biased sex ratios; for example, Dimock (1983)
found an average ratio of 30.8 female U. formosa per mussel to 1 male. This aggres-
sive male behaviour not only excludes conspecific males from the harem-master’s
mussel but it may also keep out other more peace-loving species of Unionicola
(Davids et al. 1988).
The most extraordinary form of post-larval parasitism in the water mites is shown
by five species of Hygrobates (Hygrobatidae) that parasitize southeast Asian
newts of the genera Pachytriton and Paramesotriton (Urodela: Salamandridae)
(Goldschmidt and Fu 2011). For example, adult males and females of Hygrobates
aloisii are found on the external skin of their host while the deutonymphs attach to
the inside of the newt’s mouth. The palps of this species have a robust terminal claw
and the cheliceral blade is much longer than in typical invertebrate-eating Hygrobates
species, and probably allow the mites to grasp and breach the skin of their amphib-
ian hosts. Larval H. aloisii do not attach to newts and presumably need insect hosts.

Halacaroids Will Eat Anything

The aquatic mites discussed above, although they may have expanded their diet
relative to that of their terrestrial ancestors, have not made any major switches
between food types. Halacaroids, however, have been more adventurous in their
choice of cuisine. The presumed terrestrial ancestor of the Halacaroidea was a
member of the superfamily Bdelloidea (discussed in Harvey 1990). This taxon is
composed entirely of free-living predators that feed on eggs, larvae and adults of
small arthropods (Krantz and Walter 2009). In striking contrast, halacaroids act as
everything from fungivores to predators to parasites.
(Pre)Adaptations to Subaquatic Life 253

Members of the halacarid subfamily Rhombognathinae are solely algivorous and


live only within the marine photic zone (MacQuitty 1984; Bartsch 1989). They have
long, fine chelicerae appropriate for piercing algal cells and sucking up material that
‘has presumably been subjected to pre-oral digestion’ (Bartsch 1989). The genus
Porohalacarus in the subfamily Halacarinae is also likely to be algivorous, judging by
their often green gut contents (Bartsch 1996), and some species of Halacarellus and
Actacarus also consume algae, although perhaps only incidentally (Bartsch 1989).
Most halacaroid species are thought to be predators, although this is based on a
limited number of direct observations. Bartsch (1996) lists protozoans, oligochaetes
and other small metazoans among their diets. MacQuitty (1984) observed two
marine species of Agauopsis (Halacaridae: Halacarinae) feeding on copepods.
Agauopsis curvata preferentially fed on the earliest instars (nauplii) at a maximum
rate of 10 nauplii/h. Agauopsis filirostris preferred the larger adults at a lower rate
of 2 an hour. The first pair of legs of both species have modified raptorial spines
used to cage and restrain captured prey. The first pair of legs of A. filirostris are
about twice the length of A. curvata’s, which suggests that prey preferences are
based on the ability to handle copepods of different sizes. Acarochelopodia,
Halacarus and Halacarellus also have a spiny first pair of legs, suggesting similar
predatory habits. It is interesting to note that two genera of water mites with similar
modifications to their first pair of legs, Neumania and Unionicola (Unionicolidae),
include copepods in their diets.
A number of halacaroids are parasitic inside larger invertebrates (Bartsch 1987).
Two Copidognathus species are parasites in decapods, one in the gill chambers of
the slipper lobster Parribacus antarcticus and another from the egg packets of the
spider crab Maja squinado. Another species found in the gill chambers of a marine
decapod (Peltarium spinulosum) is Veladeadcarus gasconi. Bartsch (1987) points out
that all three species have well-developed eyes and wonders whether the mites use
light as a cue when searching for hosts. Certainly, water mites of the genus
Unionicola that are parasitic inside freshwater mussels use a combination of host
chemicals and light to sense when they are safe inside a mussel (Welsh 1930; Edwards
and Dimock 1995). In the halacarid subfamily Halixodinae, Halixodes chitonis lives in
the gill and mantle cavities of chitons and gastropods. Parhalixodes travei appears
to be an ectoparasite of the nemertean Cerebratulus hepaticus. Bartsch (1987) feels
that species in the genus Bradyagaue are parasites of hydrozoans. Another probable
parasite is Enterohalacarus minutipalpis (Enterohalacarinae). It lives in the guts of
sea urchins (Plesioderma indicum).
There are a few species of freshwater halacaroids that appear to be parasitic.
Astacopsiphagus parasiticus is a gill parasite of the freshwater crayfish Euastacus
spinifer in Australia (Bartsch 1996). This is a very large halacarid – up to 2 mm long –
that has been placed in its own subfamily, the Astacopsiphaginae. Another halac-
arid, Limnohalacarus wackeri, has been collected from the gill chambers of the
European crayfish Astacus but it has also been found in surface- and groundwater.
Bartsch (1996) suggests it may be an accidental inhabitant of crayfish gills. On
firmer ground is Harvey’s (1990) identification of Peza daps (Pezidae) as a parasite
of the gill chambers of the Australian crayfish Engaeus fultoni. In addition to its
habitat, P. daps lacks eyes, suggesting a cloistered and potentially parasitic life.
254 7 Acari Underwater, or, Why Did Mites Take the Plunge?

On the other hand, Engaeus species are burrowing crayfish that dig tunnels up to a
metre in depth at the margins of swamps and streams (Jones and Morgan 1994); so
it is possible that P. daps is a non-parasitic mite normally inhabiting interstitial areas
that is accidentally ‘breathed in’ by E. fultoni. The exact diet has not been deter-
mined for any presumed parasitic species of halacarid, freshwater or marine, nor has
the effect of halacaroid parasitism on hosts been investigated.

Osmoregulation

Moving from land, where prevention of water loss is a priority, to fresh water, where
loss of soluble ions becomes a problem, has resulted in changes in the osmoregula-
tory structures of aquatic mites. Aquatic astigmatans in the family Algophagidae
have a porous axillary plate between their first and second pair of legs that appears
to function in the same ion-absorbing way as the chloride cells found on aquatic
insect larvae (Fashing and Marcuson 1996). Similarly, the urstigmata and genital
papillae found in Naiadacarus arboricola (Acaridae), Halacaridae and in water
mites (where the papillae are typically called ‘acetabula’) also have the structure of
chloride epithelia (Alberti 1977) (Fig. 7.9). Urstigmata are bulbous or flat porous
plates located between the first and second coxae in many acariform larvae; parasiti-
form mites, in contrast, lack them. The absence of these ‘preadapted’ structures may
be one reason why parasitiformans have so rarely invaded freshwater habitats.
At the opposite extreme of ionic concentration, salty water is anathema to the
vast majority of water mites. The Pontarachnidae is an obvious exception, as most water
mites in this family are restricted to the marine littoral (Cook 1974). In a survey of
the water mite assemblages of Sicily, Gerecke (1991) found that in streams with
conductivities greater than 5 mS/cm, only one of approximately 100 species –
Diplodontus semiperforatus (Hydryphantidae) – was common. In British Columbia,
the water boatman Cenocorixa expleta (Heteroptera: Corixidae) appears to escape
parasitism by water mite larvae because it occupies inland waters too saline for the
mites (Smith 1988).
Why are water mites, otherwise so robust to environmental extremes, so sensitive
to salt? The answer may be in the genital acetabula. If they are designed to
efficiently take up ions from fresh water, then an overabundance of Na+ and other
ions may ‘overload’ a mite’s haemolymph. The primarily marine Pontarachnidae
lack typical genital acetabula, but do have unique cuticular structures that look a
bit like asterisks and may function to pump out excess ions (Smit and Alberti
2010). Many freshwater halacarids have much larger or more numerous acetabula
than their marine cousins (Bartsch and Gerecke 2011). However, Olomski (1991)
found no statistical evidence that Eylais tantilla (Eylaidae) from saline ponds
have any fewer genital acetabula than individuals from less salty waters. Also,
genital papillae of aquatic Oribatida are no larger than those of terrestrial ones
(Behan-Pelletier and Eamer 2007), calling into doubt the generality of these structures
as ionic regulators.
(Pre)Adaptations to Subaquatic Life 255

Fig. 7.9 Diversity in number and arrangement of genital papillae/acetabula in freshwater


Prostigmata. (a–b), freshwater halacarid mite (Halacaridae); (c, d), female of a Mideopsis sp.
water mite (Mideopsidae); (e–f), female of an Arrenurus sp. water mite (Arrenuridae) showing
many dozens of genital acetabula (SEMs by HC Proctor)
256 7 Acari Underwater, or, Why Did Mites Take the Plunge?

Sperm Transfer

Spermatozoa of terrestrial animals do not respond well to being immersed in


hypotonic water. Argyroneta aquatica, the only spider that spends a major amount
of time underwater, transfers sperm while in the air supply of the female’s diving
bell (Bristowe 1958). The relatively high proportion of parthenogenetic versus
sexual taxa of oribatids in aquatic systems may indicate that free-standing sper-
matophores suffer physiological stress when submerged (Behan-Pelletier and
Eamer 2007). In contrast, Witte (1991) suggests that adaptations to keep sperm
moist in terrestrial habitats may have helped the ancestors of water mites to avoid
osmotic damage to their sperm when they invaded fresh water. Terrestrial parasiten-
gones typically transfer sperm in a very indirect way. Males deposit spermatophores
(often in the absence of females) and females may come across the spermatophores
at a later time, at which point they may or may not take up the sperm packet.
Dissociated transfer also occurs in some other prostigmatans and oribatids (see
Chap. 5). The temporal disjunction between deposition and uptake means that
sperm cells are exposed to low humidities prior to being taken up by females. The
osmotic stress placed on sperm of dissociated terrestrial parasitengones has selected
for very clever packaging. The sperm cells are wrapped in coats that are impervious
to water and the fluid surrounding these packets is hygrophilic (Witte 1991).
Spermatophores of these mites ‘osmoregulate’ independently, keeping the sperm
cells surrounded by a liquid.
The impervious wrappings of the sperm not only prevent water from leaving the
cells but also they keep excess water from entering. Witte (1984, 1991) suggests that
it is this characteristic that has allowed the ancestor of water mites to invade water
while retaining external sperm transfer. The spermatophores of terrestrial oribatid
mites also appear to have some osmoregulatory ability (Alberti et al. 1991), sug-
gesting that this may also be a ‘preadaptation’ for invasion of aquatic systems in
oribatids. Astigmatans transfer sperm directly through an aedeagus, thus sperm
cells are not exposed to external conditions. Sperm transfer in mesostigmatans typi-
cally involves a male introducing a sperm packet into the female’s sperm receptacle.
This mode of transfer is unlikely to select for osmoregulatory ability in spermato-
phores and may be one reason why there are no mesostigmatans known to spend
their entire life-history under water. Little is known about sperm transfer in halaca-
rids, although free-standing stalked spermatophores have been described for several
species (see brief review in Pepato and da Rocha 2010).
Although the terrestrial relatives of water mites show no inclination to directly
transfer sperm, copulation has evolved repeatedly within the 5,000 species of water
mites. Using a comparative method, Proctor (1991) tested two hypotheses about
why direct transfer has evolved so often in the hydracarines. There was no evidence
that disruption of pheromonal communication between spermatophores and females
was the selective force, as lineages inhabiting running water showed no greater
tendency to evolve copulation than those in standing water. However, lineages of
swimming mites did show a significantly greater number of independent evolutions
Predation: The Correlation Between Foul Taste and Bright Colour 257

of direct sperm transfer than did lineages of crawling mites. Proctor (1991) suggests
that movement of females into the three dimensional water column decreases the
probability of contact between females and substrate-deposited spermatophores.
In this situation, males that captured females in the water column and transferred
spermatophores directly would be favoured.

Predation: The Correlation Between Foul Taste


and Bright Colour

Water mites are only occasionally found in fish stomachs and it has been assumed
that they are seldom preyed upon. Eriksson et al. (1980) report that of 21,635 food
items identified from the gut contents of perch, only one was a water mite. Keeley
and Grant (1997) found that although lentic water mites made up 14.9 % of all indi-
vidual invertebrates drifting in a stream, they represented only 0.4 % of items in the
stomachs of resident salmon. In contrast, other organisms in the drift appeared to
be fed upon in proportion to their abundance. Modlin and Gannon (1973) examined
the stomach contents of 157 specimens of alewife (Alosa pseudoharengus) from the
Great Lakes in North America and found that fewer than 2 % of the stomachs con-
tained mites. These mites were not noticeably digested, even when found in the
lower intestine. Cupp and Willis (1982) removed 36 living Lebertia mites from the
upper and lower intestines of a single green sunfish, Lepomis cyanellus. The mites
first moved freely about but expired 10 min after removal from the gut. It appears
that water mites not only are rarely preyed upon by fish, they are also extremely
difficult to digest.
Pieczynski (1976) summarises a 3-year study of the effects of increased fish
stocking on the animal communities in a Polish lake. Tench, carp and crucian carp
all contained water mites in their guts, making up respectively 0.5 %, 0.4 % and
2.2 % of the biomass of all gut contents; however, the assemblage of mites in the
guts differed considerably from that in the habitat. Consumed mites were smaller
(500–1,000 μm versus 1,000–2,000 μm) and more dully coloured than the mites that
were common in the habitat but rarely found in the fish guts. Piecsynski suggests
that these small, dull mites are probably “caught by the fish accidentally, together
with other representatives of the invertebrate fauna, constituting the basic compo-
nents of their food”.
This semi-immunity from fish predation has puzzled biologists for many decades.
Especially disconcerting is the fact that even hungry fish will ignore large, brightly
coloured water mites. Elton (1923) was among the first to note a correlation between
the co-occurrence of red-coloured mites and fish in the same body of water. Of the
nine species of hydracarines he collected from a ditch occupied by stickleback,
eight were scarlet. He found that hungry fish would mouth the water mites but
would spit them out immediately; this occurred even if the fish had been starved for
10 days. Elton speculated that the dorsal skin glands (glandularia) common to all
258 7 Acari Underwater, or, Why Did Mites Take the Plunge?

water mites were the source of a foul exudate. The release of this secretion would
be triggered by bending of the large sensory hairs found near the openings of these
glands. The fish, when mouthing a mite, would be punished by spurts of noxious
chemicals and would hastily release the unpleasant mouthful. Elton evoked
Müllerian mimicry to explain the prevalence of red among mites associated with
fish. He also thought that the colour red may have a more striking effect on piscine
retinas than do other colours.
More recent research has had similar results. Riessen (1985) states that Piona
constricta is distasteful to fish, as he observed that after a few tentative attacks on
the mites, fish tend to avoid consuming them. Kerfoot (1982) took an experimental
approach to explaining his observation that red predominates among assemblages
of large-bodied, fish-associated mites. In bodies of water containing fish, he
observed that most zooplankton use transparency to solve the problem of conceal-
ment against a background that continually changes in brightness and colour. Water
mites, however, seem to advertise their presence in open water. Like Elton, Kerfoot
found assemblages of 3–9 species of red water mites in fish-inhabited ponds, while fish-
less ponds contained more yellow-brown species. Using laboratory trials, he found
that if a fish is truly naive, e.g. aquarium-raised individuals, introduction of a water
mite almost always attracts the attention of the fish. Often it will completely swallow
the mite or, more commonly, mouth it and reject it. Rejected mites seldom showed
damage. Subsequent trials indicated a rapid drop in interest as the fish became more
familiar with the mites.
Kerfoot wanted to test Elton’s hypothesis that red colouration enhances learning
in fish, as well as the hypothesis that red mites are especially unpalatable relative to
non-red hydracarines. He found that guppies learned to avoid black balls of food
mixed with powdered Limnochares (a distasteful red mite) just as rapidly as they did
similar red-dyed balls. Thus Elton’s first hypothesis was rejected. But, in general,
powder from red mites was rejected at much lower concentrations than powder from
dully coloured mites. Kerfoot felt that these results confirm ‘that large size and
bright-red coloration are serving an aposematic function in mites and suggest that
the mites generally found in the stomachs of fishes are more poorly protected spe-
cies’. As predicted by Kerfoot, the numerous Lebertia found in the green sunfish
intestines (Cupp and Willis 1982) and those in alewife guts (Modlin and Gannon
1973) were non-red species.
There can be little doubt that red water mites are particularly distasteful to fish.
However, the claim that water mite assemblages are more biased towards red spe-
cies in water bodies containing fish has been called into question. Kerfoot (1982)
collected mites from only eight water bodies (six with fish, two without): a rather
small sample size. To test the generality of Kerfoot’s observations, Proctor and
Garga (2004) collected more than 6,000 water mites from 26 fish-present and 20
fish-absent water bodies in southeastern Ontario, Canada, and classified each mite
according to colour. Contrary to Kerfoot’s findings, they found that assemblages of
mites from sites lacking fish were much more skewed towards redness than those
from sites with fish (Fig. 7.10). Even within particular genera (Limnesia, Piona and
Tiphys), red species were more common in fish-absent than fish-present water bodies.
These results make the hypothesis of fish-driven evolution of redness unlikely.
Predation: The Correlation Between Foul Taste and Bright Colour 259

Fig. 7.10 Collections of Hydracarina in water bodies with fish present (filled circles) and those
with fish absent (open circles) showing the proportion made up of red-coloured individuals. Sites
are ranked from those with the smallest proportion of red mites to those with the largest (Modified
from Garga 1996)

Rather than redness repeatedly arising within the hydracarine clade (as the
Müllerian mimicry hypothesis of Elton would hold), it seems more probable that
redness is a holdover from water mites’ terrestrial relatives. Terrestrial parasiten-
gone species are overwhelmingly orange or red in colour. The related cohort
Anystina also contains primarily large, red, active predators. Those water mites that
are considered ‘primitive’ are also typically red (e.g. Hydryphantidae, Limnocharidae)
and it seems likely that red pigmentation is an ancestral trait in the water mites.
But what of the foul flavour associated with red water mites? Proctor and Garga
(2004) tested the hypothesis that distastefulness as well as redness is a carry-over
from the terrestrial ancestors of water mites. Using Kerfoot’s (1982) methodology,
they presented sunfish with flour balls made with varying proportions of powder
from dried velvet mites, Dinothrombium magnificum (Trombidiidae), a large desert-
dwelling parasitengone. They found that fish rejected the flour balls at concentra-
tions of mite powder as low as Kerfoot had found for red water mites (95 % rejection
at 15 % mite by weight).
These authors hypothesised that red pigmentation had initially evolved as
carotenoid-based photoprotection in the terrestrial ancestor of water mites.
Among the various functions proposed for carotenoid pigments, the best understood
and most researched is protection against photodynamic injury (Goldstrohm and
Lilly 1965). The photoprotective role of carotenoids has been observed in organ-
isms as different as bacteria (Mathews and Sistrom 1959), fungus (Goldstrohm and
Lilly 1965) and copepods (Hairston 1976; Chalker-Scott 1995). Carotenoids have
been found in many species of red water mites and terrestrial parasitengones
(Manunta 1939; Green 1964; Czeczuga and Czerpak 1968a, b, c; Meyer and Kabbe
1991). Proctor and Garga (2004) proposed that redness has been retained in those
hydracarines living in shallow, often temporary water bodies (which are also typically
fishless) but has repeatedly been lost by groups inhabiting deeper water where damage
260 7 Acari Underwater, or, Why Did Mites Take the Plunge?

by light is less likely. Photoprotection by pigments can occur not only at the adult
stage, but in the egg and larval stages. Indirect evidence for carotenoid-based
photoprotection in terrestrial parasitengones is that chiggers (larval Trombiculidae)
dwelling in the nests of their hosts are typically pale yellow, whereas those that
roam about in search of hosts (and hence are exposed to light) are brilliant red
(Krantz 1978). There is also circumstantial evidence from other mite taxa. Norton
et al. (1996) describe oribatids that are active in shallow (2–5 cm) temporary pud-
dles of temporary water on rocky outcrops (see section “Oribatida and Astigmata:
Same Story, Different Substrate”). They consume orange-coloured bdelloid rotifers
and themselves produce eggs that are bright orange, ‘a color not previously reported
for eggs of Oribatida’.
To explain the association between redness and distastefulness, Proctor & Garga
suggest that once redness evolved in terrestrial parasitengones, it put these large
active mites at risk from visual predators such as reptiles, amphibians and birds.
Distasteful exudates evolved to counter this higher visibility and were likewise
maintained in the still visible red water mites, but to rebuff predation by aquatic
rather than by terrestrial vertebrates (Fig. 7.11). Visual vertebrate predators in tem-
porary water bodies would most likely be amphibians. Donnelly (1991) found that
mites composed up to 40 % of the stomach contents of poison arrow frogs and
Flowers and Graves (1995) observed mites in 70 % of examined toad stomachs;
unfortunately, in both of these cases mites were not identified to anything more
specific than ‘Acari’. To our knowledge, there is no published information about
whether velvet mites and their kin are eaten by lizards, frogs, salamanders or birds.
It would be very interesting to test the palatability of red terrestrial and aquatic
parasitengones on potential terrestrial predators. As well, it would be valuable to
test the ‘photoprotection’ hypothesis of Garga (1996) by subjecting differently
coloured water mite eggs (which range from white to red, depending on the
species) to UV light, to determine whether red-pigmented eggs are more likely to
produce healthy larvae.
In contrast to predation on water mites by fish, little work has been done on
invertebrate predators. Elton (1923) stated that an adult Agabus beetle (Dytiscidae),
which had never been exposed to mites before, attacked and ate a red Hydrachna.
Pennak (1978) says that water mites fall prey to numerous invertebrates including
Hydra but does not provide references for this claim. Apparently the foul exudate so
distasteful to fish does not bother those water mites that prey on other hydracarines,
as both Elton (1923) and Davids (1973) mentioned that Eylais, identified by Kerfoot
as a particularly potent red mite, was eaten by other mite species. Böttger (1962a)
found that hydras would grab Piona nodata (Pionidae) but would release them
almost immediately. After a few minutes the stunned mites appeared to recover and
swam off normally. Notonectid and naucorid bugs would not eat them. Larvae of
dytiscid beetles would release the mites after grabbing them (although often the
mites were mortally wounded). Other species of Piona as well as conspecifics
would eat P. nodata, as would zygopteran larvae. Böttger speculated that water
mite glandularia release unpleasant chemicals that deter predation, but he did not
explain why damselfly larvae were unaffected. Red water mite larvae, which lack
Predation: The Correlation Between Foul Taste and Bright Colour 261

Fig. 7.11 Garga’s (1996) hypothesis for the origins of redness and distastefulness in water mites

glandularia, are not avoided by dragonfly predators (Nagel et al. 2011). Proctor
(1992b) found that males of the small blue mite Neumania papillator (Unionicolidae)
were more likely than females to be eaten by odonate larvae (Ischnura, Libellula)
and by the water mite Limnesia sp. (Limnesiidae). She postulated that males were
more susceptible to predation because their active mate-searching behaviour
increased their detectability by or encounter rate with predators.
262 7 Acari Underwater, or, Why Did Mites Take the Plunge?

Like the interaction between Hygrobates spp. and bottom-feeding fish described
above, in which fish increase the number of prey available to the mites, there are
examples of fish presence helping mites by reducing predation on them. Eriksson
et al. (1980) noted a strong positive correlation between lakes that contained fish
and lakes that contained the water mite Piona carnea (Pionidae). In a field experi-
ment they divided a fishless lake in half with a net barrier, and introduced perch to
one part. In the second year of the study, they found that water mites were present
only in the half with fish. Co-occurrence of these two taxa due to a similar prefer-
ence in abiotic surroundings could be ruled out, as conditions in the divided lake
were similar on both sides. Eriksson et al. performed a series of laboratory experi-
ments in which perch were kept together with water mites and a number of potential
insect predators of the mites. Leucorrhinia, Coenagrion (both odonates), Notonecta,
Glaenocorixa, Sigara (all hemipterans) and Chaoborus (Diptera) fed on the pionid
water mites. Acilius larvae (Coleoptera) did not. The perch also ate water mites but
preferred the available insects in all but one case (that of Leucorrhinia). The authors
explain the higher abundance of water mites in lakes containing fish as follows: in
lakes without fish, P. carnea are kept at low densities because of the predation pres-
sure from aquatic insects. In the presence of fish, predatory insects decrease in num-
ber because of increased predation pressure from fish. This implies decreased
pressure from aquatic insects on water mites, which is not compensated for by an
increased predation from fish so long as alternative prey are available.
A similar phenomenon and hypothesis were described by Puncochar and Hrbacek
(1991) to explain the co-occurrence of trout and Piona carnea in a Czechoslovakian
reservoir. When trout were present there were high numbers of P. carnea (and other
mites); when trout-stocking ceased the smaller fish species that the trout normally
consumed increased in number and the water mites (P. carnea in particular)
decreased dramatically. The authors suggest that predation by trout on large insects
reduced predation on mites, and that the trouts’ consumption of small planktivorous
and benthivorous fish kept the mites’ prey and host populations (cladocerans and
chironomids) high (Fig. 7.12).

Locomotion

Aquatic mites have a greater variety of options with regard to locomotion than do
their land-bound relatives. They may crawl, burrow or swim – the last being the
equivalent of flight for terrestrial organisms.

Swimming

Swimming has evolved independently in many lineages of aquatic insects, e.g. in bugs,
beetles, caddisflies and true flies. In contrast, few of the major lineages of aquatic
mites have evolved natatory locomotion. There are no examples of swimming species
Locomotion 263

Fig. 7.12 The effect of stocking a reservoir with trout on the population size of the water mite
Piona carnea (Pionidae) (Modified from Puncochar and Hrbacek 1991)

among the Parasitiformes, Halacaroidea or Oribatida (but see section “Levitation”).


Among the Astigmata, only a few species in the genera Zwickia and Creutzeria
(Histiostomatidae) have exchanged walking for swimming (Fashing 2008). Fashing
et al. (1996). describe the behaviour and morphology of a Creutzeria species from
Australian pitcher plants. Nymphs, males and females have highly modified legs
bearing long whip-like setae at their tips. The male bears a remarkable likeness to a
planktonic copepod nauplius. However, this crustacean-like morphology is the
result of a compromise between sex and swimming. The male uses his strong
forelegs to grasp the female while mating, and the second pair of legs are used
both for swimming and for holding onto a substrate when the first pair are
connubially occupied.
It is among the Hydracarina that we find the greatest number of swimming Acari.
Water mites show a gradation of competence in swimming, from crawling and bur-
rowing forms completely incapable of natatory locomotion to pelagic water mites
that spend much of their life in the plankton. Bottom-dwelling water mites generally
lack the long, slender setae termed ‘swimming hairs’, which are found in more
264 7 Acari Underwater, or, Why Did Mites Take the Plunge?

natatory forms. Swimming ability seems very labile, having arisen and disappeared
repeatedly in different lineages. The invasion of fast-flowing water or interstitial
habitats is often associated with a reduction in size and number, or total loss, of
swimming hairs. Using the presence of swimming hairs as a proxy, Proctor (1991)
examined the distribution of swimming ability among 343 genera of water mites.
She found that 89 genera (26 %) were composed entirely of swimming species, 192
genera (56 %) were entirely non-swimming, and 62 genera (18 %) contained both
swimming and non-swimming species.
Barr and Smith (1979) examined the blade-like swimming hairs of the mite
Limnochares (Cyclothrix) americana. They felt that the swimming setae of L. amer-
icana probably lose some efficiency through their structural characteristics, espe-
cially when compared to the stiffer setae of natatory insects. Barr and Smith
performed depilatory experiments on these mites and found that swimming ability
was eliminated when 80 % of the swimming hairs were removed. Even though nor-
mal swimming movements occurred, the mites could not generate enough lift to
move off the bottom.
In the same study, the setae of L. americana were found to be almost uniform in
width from base to tip with no particular morphological elaborations. The more
pelagic mite, Piona constricta, has more complicated natatory setae (Riessen 1982).
These are widened and flattened at their distal ends to make them spoon-shaped. As
well, the setae themselves are fringed with short hairs, structures that probably add
to swimming efficiency. Cuticular structures other than setae may also be involved
in open-water movement. Halik (1924) proposed that pelagic Unionicola species
(Unionicolidae) could erect thick setae on their legs while drifting to serve as a sort
of parachute. However, as many species of Unionicola and other unionicolids in the
genus Neumania employ these collapsible setae for capturing prey (Fig. 7.7), it is
not clear whether locomotion or predation was the primary selective force behind
their evolution.
Among swimming mites, habitat choice is often correlated with swimming ability.
Böttger (1962a) compared three species of German water mites. Arrenurus globator
(Arrenuridae) was the weakest swimmer and was most often captured in near-shore
vegetation. Eylais infundibulifera (Eylaidae) was the strongest and most perseverant
swimmer of the three and was most commonly found in open water. Piona nodata
(Pionidae) fell between the other two in locomotory efficiency. Predictably, it was
found both among littoral macrophytes and in open areas. Pieczynski (1964) used
trapability in stationary traps vs. dipnets as an estimator of water mite motility. As a
whole, water mites ranked between ceratopogonids and mayflies in motility, with
beetles and corixids topping the list and isopods being the least motile. Within the
water mites, Hygrobates nigromaculatus (Hygrobatidae) was the most motile, fol-
lowed by Hydrodroma despiciens (Hydrodromidae); poorly motile species included
Mideopsis orbicularis (Mideopsidae) and Piona conglobata (Pionidae), which
Pieczynski ranked as sluggish as isopods. Of course, this all depends on whether all
taxa are similarly ‘trap-happy’.
Barr and Smith (1980) found the swimming velocity of Limnochares americana
to be about 1.6 cm/s. Riessen (1982) found that the adult stage of the pelagic water
Levitation 265

mite Piona constricta could swim at speeds of 2.0–2.5 cm/s. These open-water
mites typically swim for short periods of a few seconds on one course and then
switch direction. Riessen states that swimming speeds of various potential prey of
P. nodata are usually less than 2 mm/s. Proctor (1988) determined that Unionicola
crassipes (Unionicolidae), when prodded by an observer, swam at an average rate
of 0.84 cm/s.
Smith and Barr (1977) found that Limnochares primarily uses light direction to
orient itself while swimming. Bottom lighting in an aquarium caused the mite to
swim in loops with the ventral surface inward. Removal of the eye capsule of
Limnochares resulted in normally oriented swimming (dorsum upwards), regardless
of the direction of lighting. This indicates that some cue other than light is also used
for orientation. Böttger (1962b) found that when presented with light from below,
Eylais extendens (Eylaidae) would try to swim upside down but, because of the
shape of its body, it would end up swimming in loops. Hydryphantes ruber
(Hydryphantidae) behaved as Eylais, except that when it contacted the substrate the
tactile sense overrode the visual and the mite would orient so that its leg tips touched
the surface. Böttger found that Piona nodata (Pionidae) initially went into a series
of wild rolls until it crashed into the side of the container but that over time these
mites became acclimated to the inverted light and eventually ‘learned’ to swim in a
normal fashion. Böttger suggested that the trailing fourth pair legs of many species
of water mites act to generate lift when the mite is swimming normally but causes
looping when the mite is upside down.
Positive phototaxis is common in water mites, as evidenced by mites’ greater
activity during the day than at night (Pieczynski 1964). Among lentic taxa, the ten-
dency to be day active is stronger in littoral than in open-water species. Similarly,
lotic mites often have their peak drift during the day (Schmidt 1969). Piona nodata
(Pionidae) and Limnesia undulata (Limnesiidae) are very strongly attracted to light,
which Böttger (1962b) suggests is an adaptation to oxygen-poor deeper waters. He
introduced CO2-laden water to the bottom of an aquarium and the mites responded
by swimming to the surface but when he lit the aquarium from below they swam to
the bottom and suffocated, suggesting that it is Licht über Alles, even when it does
not make sense. However, this attraction to light does not hold for all species. An
illuminated Hydryphantes ruber (Hydryphantidae) will crawl ceaselessly until it
reaches shadow (Böttger 1962b) and most interstitial taxa are also photo-negative.

Levitation

Aquatic oribatids of the family Hydrozetidae do not look much different from ter-
restrial oribatids and move at the same sedate walking pace as they crawl among
water weeds but Hydrozetes makes use of air in a novel way. Newell (1945) noted
that H. lacustris and petrunkevitchi often contain bubbles of gas in their midguts
and that by releasing their grip on the substrate, they would float up to the water’s
surface much more rapidly than they could crawl – rather like underwater
266 7 Acari Underwater, or, Why Did Mites Take the Plunge?

ballooning. He called this activity ‘levitation’, and emphasised that “[t]his most
appropriate term is used in its preferred sense, and is in no way intended to suggest
the unfortunate connotation applied to it in spiritism”. Newell noted that agitating
mites would induce the formation of these bubbles, suggesting that this may be a
form of rapid escape response. Levitation occurred, however, not only after the
mites were agitated but also whenever ambient light conditions fell low enough.
Hydrozetes has a light-sensitive area that would allow it to detect light. Keeping
mites in a thoroughly darkened aquarium did not induce levitation; it appeared that
the mites responded only to a change from high to low intensity. Newell suggested
that levitation would allow animals that had been dislodged from their normal home
on plants in the photic zone to escape starvation and death by the anoxia that would
occur at the dark bottom of a pond. He was uncertain about how Hydrozetes gener-
ates and regulates the gas bubble.

Sensitivity and Diversity: Water Mites


as Environmental Indicators

Organic and inorganic pollution of fresh water is often monitored by subjecting


individuals to water containing a suspected pollutant (bioassays) or by looking at
the species richness and composition of invertebrate assemblages in a water body
(Rosenberg and Resh 1993). Biodiversity is often also used as a measure of the
nebulous quality termed ‘ecosystem health’. Mites have been employed as bioindi-
cators in terrestrial systems (e.g. Lagerlöf and Andrén 1988; Koehler 1992) but
water mites are typically ignored in studies of freshwater health, at least in North
America (Resh and McElravy 1993; Proctor 2007). Promoting one’s favourite taxon
as an indicator of pollution is a popular pastime among freshwater ecologists. It is
therefore disappointing to find that – physiologically – lentic water mites do not
appear particularly susceptible to contaminants. We have observed water mites
swimming energetically through a soup of decomposed zooplankton and insects. As
well, along with beetles, water mites tend to be among that last invertebrates to
expire in collecting jars filled with ethanol. Pieczynski (1976) lists studies that show
water mites to be insensitive to long periods without oxygen, changes in tempera-
ture and dehydration. He suggests that physiological robustness may be particularly
common in lentic littoral species whose habitat naturally presents greatly varying
abiotic conditions.
Lotic water mites may be more sensitive to such extremes. Gerecke and
Schwoerbel (1991) state that almost all studies of lotic mites show that they decline
in polluted vs. unpolluted sections of streams. In Luxembourg streams, water mites
proved to be the taxon of freshwater invertebrates with the greatest proportion of
significant indicator species for streams of different morphological classifications
(Dohet et al. 2008). In Australian streams, water mite and other aquatic mites dis-
play genus-level richnesses similar to that of Trichoptera (Proctor 2007, see also end
Sensitivity and Diversity: Water Mites as Environmental Indicators 267

of this section), a taxon whose richness is often used as an index of water quality
(e.g., Klemm et al. 2003). But regardless of whether mites are ‘better’ than other
taxa at identifying stressed water bodies, it is clear that physical factors play a great
role in determining the distribution, abundance and variety of water mites in a body
of water. The temperature, chemical constitution, depth, substrate and water velocity
in an aqueous environment affect the nature of its hydracarine population.
As well, because water mites depend on a diversity of other organisms for prey,
hosts and even oviposition sites, the presence of certain mites may be used to indicate
the presence of these other organisms. For example, if Unionicola crassipes
(Unionicolidae) is present, then its oviposition and transformation sites – freshwater
sponges – must also be present. Below is an outline of abiotic and biotic factors that
may correlate with the diversity and abundance of water mites.

Temperature

As a taxon, water mites have a broad range of thermal tolerance. Active mites can
often be taken from the frigid waters of ice-bound lakes (Pennak 1978; Proctor,
pers. obs.). Conversely, some species of mites are very tolerant of high temperatures.
Young (1969) found 2 species in thermal springs: Thermacarus nevadensis and
Tyrrellia circularis were taken from water with a temperature of 49.1 °C, near the
upper level of tolerance for most arthropods (Chapman 1982). It is this ability to live
in hot pools that led Philip (1953) to pen this apocalyptic whimsy:
‘…in the futilely protective shadow of a protruding bit of agatized rib of a long-since extinct
whale, … in this shimmering shadow in a now bleak and almost lifeless world steaming in
the writhing heat of a merciless expanding ball of fire in the heavens and beside the ever
boiling cauldron springs of the North Pole, crouches an anxious, solitary water-mite, the
sole survivor of animal life upon this agonized, liquidating planet. There will be no chronicle
of her life cycle, and the validity of her scientific name will be a matter of great unimportance.’

Depth

Water mites tend to be more common in the littoral zone of lakes and seldom venture
into deep water. As samples are taken at progressively greater depths, hydracarine
populations become scanty and diversity decreases (Pennak 1978). Modlin and
Gannon (1973), in a study of Great Lakes water mites, found that only 4 of 32 species
were found in the profundal zone (21–40 m). Rieradevall and Gil (1993) found no
water mites in dredge samples from lake substrates deeper than 13 m, while from 5 to
12 m mites made up at most 0.39 % of the benthic fauna. Mites are also scarce in the
pelagic zone of deep water. The plankton community is usually dominated by com-
paratively few species from even fewer genera, usually in the families Pionidae and
Unionicolidae. Modlin and Gannon (1973) found two species of Piona and one each
268 7 Acari Underwater, or, Why Did Mites Take the Plunge?

of Forelia, Neumania and Unionicola in the plankton of the Great Lakes. Riessen
(1982) found three species of Piona, Huitfeldtia rectipes and Unionicola crassipes in
the open water of Heney Lake, Quebec. In North America Unionicola crassipes often
dominates the pelagic zone; Conroy (1974) found that 90 % of all planktonic mites in
a British Columbian lake were U. crassipes.
Moving from the plankton to the substrate, the intersitial zone may contain lower
numbers of mites but it affords more variety. Cook (1969) says that 33 of the 48 fami-
lies of Hydracarina are known to have at least some representatives in the ground
waters. Only the superfamily Hydrachnoidea goes unrepresented in the phreatic,
possibly because the generally rounded shape of hydrachnoids is ill suited to life in
gravel interstices. I. Smith (pers. comm.) has collected more than 20 species of inter-
stitial mites from a single stream site in Alberta. Much of the subsurface fauna seems
to have evolved from rheophilic ancestors, as the interstitial waters associated with
lakes and ponds are almost completely lacking in water mites (Cook 1969).

Substrate

The majority of water mites remain in close proximity to some form of substrate,
either vegetable or mineral. It has often been remarked that water mites are rare
when there is a lack of macrophytes (Pennak 1978; Young 1969; Modlin and
Gannon 1973). A comprehensive study of biotic and abiotic factors determining
water mite assemblages in Pyrenean streams showed that the presence and type of
moss in a stream was one of the most important features (Angelier et al. 1985). In
his study of Colorado water mites, Young (1969) found that 42 of the 48 lakes that
yielded no mites were either alpine lakes or fluctuating reservoirs. He postulated
that repeated exposure of the lake bed or scouring by ice would exclude both plants
and invertebrate bottom fauna. Hence, he felt that the mutual absence of plants and
mites was due to harsh environmental conditions rather than any vegetable–faunal
interaction. Rooted aquatic plants are not essential to all water mites, although
Hydrachna requires plant stalks in which to lay its eggs. Many species happily exist
in treeholes and temporary bodies of water that lack such flora.
Nevertheless, when rooted macrophytes are available, water mites sometimes
show distinct preferences for certain plants. Cyr and Downing (1988) observed
more mites on the broad-leafed Potamogeton than on Myriophyllum or Vallisneria,
both of which have finely dissected leaves. In Czechoslovakian carp ponds,
Puncochar (1971) found the fewest hydracarines in stands of Elodea, while three
times as many were collected from Potamogeton. Young (1969) found that both
Potamogeton and Myriophyllum were favourite plants for water mites and found
few hydracarines near either Elodea or Chara. Even though large populations of
several species were taken in other parts of a lake, submerged meadows of these two
macrophytes were often devoid of mites, possibly indicating active avoidance.
Young did not investigate the cause of this avoidance, which could be based on a
lack of prey or a surfeit of predators.
Sensitivity and Diversity: Water Mites as Environmental Indicators 269

Mites also seem to show preferences for certain types of mineral substrates.
Sand or gravel shores of lakes are generally poor places to collect water mites (Modlin
and Gannon 1973; Pennak 1978). In streams, sand and gravel are preferred to silty
bottoms or substrates consisting of large boulders (Young 1969; Freundlieb 1979).
Stream velocity seems to influence mite populations through its effect on the substrate.
Young (1969) found that the greatest diversity and abundance of water mites occurred
at stream velocities of 0.3–0.5 m/s, when the stream bed was mainly gravel. Freundlieb
(1979) also noted that the sand and gravel substrate of fast-moving streams had the
highest densities of hydrachnids. Too rapid a flow, especially if it occurs suddenly in
the form of a spate, can reduce the local abundance of water mites and other inverte-
brates by washing them downstream. Boulton et al. (2004) tested whether mites living
in surficial stream substrates avoided the force of a flood by hiding in hyporheic zone,
but found no evidence of active movement into a hyporheic refuge.

Standing Versus Running Water

In his 3-year study of mites in Colorado, Young (1969) compared the lentic and lotic
faunas. He found three major differences between these two groups. The most obvi-
ous was the larger number of lentic vs. lotic species. Of the 99 species he collected,
44 occurred in lotic situations, 66 in lentic, and only 11 species were found in both
habitats. Young felt that three things could account for this phenomenon. Firstly,
environmental conditions in a series of lakes are much more diverse then the condi-
tions occurring throughout a stream system and this provides a greater variety of
niches for those forms with proper adaptations. Secondly, the opportunity for spe-
ciation to occur is greater in lakes and ponds than in lotic situations; the isolation of
populations in bodies of still water results in genetic isolation more frequently than
in continuous stream systems. Finally, it seems likely that lentic situations provide
a greater variety of insect hosts and a better supply of cladocerans prey. This finding
that the lotic fauna is less diverse than the lentic is probably partially an artefact of
Young’s sampling method. He seems to have collected few genera of interstitial
mites, although these can be very diverse in montane streams.
A second major difference between these two faunas was the much higher
between-site frequency of common stream mites. For example, Sperchon glandulo-
sus (Sperchontidae) was found at 90 % of the stream stations, whereas the most
common lake species, Limnesia maculata (Limnesiidae) occurred in only 30 % of
the lentic habitats. This difference probably reflects the greater ease of dispersal in
stream systems as compared to the passive mode of parasitic dispersal that lentic
mites must rely upon.
The third difference cited by Young was the high ratio of ‘euryzonal’ (i.e. wide-
spread) species to species with restricted altitudinal ranges in the stream fauna as
compared to the lake fauna. This likely results from the more uniform conditions
throughout a stream system, permitting even narrowly adapted species to inhabit a
broad altitudinal range.
270 7 Acari Underwater, or, Why Did Mites Take the Plunge?

pH

The pH of a body of water may correlate with the diversity of water mites present
but there is little evidence that acidity affects mites directly. Pennak (1978) says that
acid lakes contain few water mites. Bagge and Meriläinen (1985) found that of 30
species of mites present in the estuary of the River Kyrönjoki in Finland, only
Limnochares aquatica (Limnocharidae) seemed able to survive highly acidic river
waters. Eriksson et al. (1980) found that in acidic lakes where fish were absent,
water mites were absent as well; however, they felt that this was due, not to the acid-
ity, but to the lack of fish which would otherwise eat the insect predators of hydraca-
rines. In a laboratory study, Rousch et al. (1997) found that survivorship of Arrenurus
manubriator (Arrenuridae) was unaffected by acidity until it fell to pH 3 or lower – a
level unlikely to be encountered in most natural water bodies. Mites were entirely
unaffected by high iron levels (added to simulate acidic mine runoff). In Colorado,
several species of water mites seem limited to non-alkaline waters but only a very
few species of Arrenurus show a preference for high alkalinity (Young 1969).
However, it is hard to determine if these preferences are physiologically those of the
mite or those of the mite’s hosts or prey.

Organic Pollution

Organic pollution may virtually eliminate the water mite fauna of a stream. In
Young’s (1969) study of Colorado mites, he found that very few species inhabited
water even 8 miles below the outfall of sewage from a Boulder treatment plant.
Gerecke and Schwoerbel (1991) compare water mite assemblages from the upper
Danube in 1959 with those present in 1984. They found that in zones where pollu-
tion from domestic sewage had increased, species richness had decreased drasti-
cally, whereas in areas where purification plants had improved water quality
compared to 1959, species richness had increased. Hygrobates fluviatilis
(Hygrobatidae) emerged as an indicator of pollution, as its dominance of the water
mite assemblages increased with increasing pollution. At the most polluted sites it
made up more than 90 % of all individuals. In studies of streams in central Italy,
Cicolani and Di Sabatino (1991) observed that diversity of water mites declined
with decreasing water quality. Like Gerecke and Schwoerbel (1991), their study
indicates that H. fluviatilis is extremely tolerant of organic pollutants. It is also inter-
esting to note that Bagge and Meriläinen (1985) found that of 30 species of mites
present in a river estuary, H. fluviatilis and H. longipalpis were the only ones to have
their highest densities in the saltiest reach of the estuary. This suggests that H. flu-
viatilis, and perhaps other congeners, are exceptionally resistant to environmental
extremes and may be good indicators of a stressed water body.
With the exception of these and a handful of other European studies, there is
essentially no work on the effects of man-made disturbances on water mite
assemblages. This is mainly due to the tendency of freshwater ecologists either not
Summary 271

to report the collection of mites at all, to lump them as ‘mites’ without further
taxonomic resolution or to place them with nematodes and oligochaetes in the
extremely uninformative group ‘other’ (Proctor 2007). Do water mites provide
information on water quality that other organisms do not? Do they intensify a signal
given by other macroinvertebrates, or run counter to it? The species richness,
abundance, demonstrated sensitivity to organic pollution and salinity and potential to
act as integrative representatives of benthic communities are strong reasons for the
inclusion of water mites in biological assessments of aquatic systems. So, why this
neglect of mites in the typical freshwater biomonitoring literature? It may be that (1)
mites are relatively small, and are rarely collected by typical sampling methods; (2)
mites show little diversity relative to other, more finely identified, invertebrate taxa.
Proctor (2007) investigated these two possibilities using water mite collections from
Queensland, Australia. First, she identified samples from the Queensland Department
of Natural Resources and Mines river monitoring scheme (663 sites throughout
Queensland). The aquatic taxa included 47 genera of Hydracarina and 4 genera of
other aquatic mites (Mesostigmata, Oribatida). Genus-level richness ranged from 1
to 12 per sample, and > 50 % of samples had 3 or more genera. So, the standardized
collecting techniques did collect a good diversity of aquatic mites. To compare the
diversity of aquatic mites to that of insects, she sampled six stream sites in
southeast Queensland. Mites were identified to genus as were members of the insect
taxa Odonata, Ephemeroptera, Plecoptera, Elmidae (Coleoptera) and Trichoptera.
Overall, aquatic mites had a total generic richness higher than that of any of these
insect groups. Within streams, mite generic richness was greater than or equal to that
of all insect taxa except for caddisflies, which were more diverse than mites in three
of the six streams. Thus there appears to be no good rationale for lumping mites into
the uniformative taxon ‘Acari’, at least in Australian studies of freshwater invertebrate
diversity. Proctor (2007) concluded that the absence of aquatic mites from assessments
of aquatic health in most countries is based on lack of cultural precedent for including
mites rather than an intrinsic property of mites themselves.

Summary

In contrast to other arachnids, mites have invaded aquatic systems at least ten times,
with some lineages giving rise to thousands of water-dwelling species. Every pos-
sible type of aquatic habitat contains mites, from temporary puddles on rocky out-
crops to the stygean permanency of marine rifts. Some acariform lineages,
particularly the Hydracarina and Halacaroidea, have radiated into many different
aquatic habitats. This switch to an aquatic lifestyle was made more feasible by the
respiratory and ionic control systems and ‘waterproof’ feeding methods of these
lineages’ ancestors (and probably explains why parasitiformans have not been suc-
cessful in water). Aquatic oribatids, astigmatans and water mites have not altered
their diet drastically from that of their terrestrial relatives but halacaroids have been
more adventurous in adopting vegetarian and parasitic habits. Some water mites
272 7 Acari Underwater, or, Why Did Mites Take the Plunge?

share top predator rank with vertebrates in freshwater systems because they are
almost immune to predation by fish. Although they are likely to be useful as indicators
of organic pollution, salinity and general ecosystem ‘health’, water mites have very
rarely been included in freshwater assessment studies.

References

Agbolade, O. M., & Odaibo, A. B. (2004). Dockovdia cookarum infection and the prosobranch
gastropod Lanistes libycus host in Omi Stream, Abo-Iwoye, south-western Nigeria. African
Journal of Biotechnology, 3, 202–205.
Alberti, G. (1977). Zur Feinstruktur und Funktion der Genitalnäpfe von Hydrodroma despiciens
(Hydrachnellae, Acari). Zoomorphologie, 87, 155–164.
Alberti, G. (1980). Zur Feinstruktur der Spermien und Spermiocytogenese der Milben (Acari). II.
Actinotrichida. Zoologische Jahrbucher Anatomie, 104, 144–203.
Alberti, G., & Bader, C. (1990). Fine structure of external ‘genital’ papillae in the freshwater mite
Hydrovolzia placophora (Hydrovolziidae, Actinedida, Actinotrichida, Acari). Experimental &
Applied Acarology, 8, 115–124.
Alberti, G., Fernandez, N. A., & Kümmel, G. (1991). Spermatophores and spermatozoa of oribatid
mites (Acari: Oribatida). Part II: Functional and systematical considerations. Acarologia, 32,
435–449.
Angelier, E., Angelier, M. L., & Lauga, J. (1985). Recherches sur l’écologie des Hydracariens
(Hydrachnellae, Acari) dans les eaux courantes. Annales de Limnologie, 21, 25–64.
Audy, J. R., Nadchatram, M., & Liat, L. B. (1960). Malaysian Parasites – XLIX. Host distribution
of Malayan ticks (Ixodoidea). Study, Institute Medical Research, Malaysia, 29, 225–246.
Bagge, P., & Meriläinen, J. J. (1985). The occurrence of water mites (Acari: Hydrachnellae) in the
estuary of the River Kyrönjoki (Bothnian Bay). Annales Zoologici Fennici, 22, 123–127.
Baker, R. A. (1976). Tissue damage and leukocytic infiltration following attachment of the mite
Unionicola intermedia to the gills of the bivalve mollusc Anodonta anatina. Journal of
Invertebrate Pathology, 27, 371–376.
Baker, R. A. (1977). Nutrition of the mite Unionicola intermedia, Koenike and its relationship to
the inflammatory response induced in its molluscan host Anodonta anatina, L. Parasitology,
75, 301–308.
Barr, D. (1972). The ejaculatory complex of water mites (Acari: Parasitengona): morphology and
potential value for systematics. Life sciences contributions/Royal Ontario Museum, 81, 1–87.
Barr, D. W., & Smith, B. P. (1979). The contribution of setal blades to effective swimming in the
aquatic mite Limnochares americana (Acari: Prostigmata: Limnocharidae). Zoological Journal
of the Linnean Society, 65, 55–69.
Barr, D. W., & Smith, B. P. (1980). Stable swimming by diagonal phase synchrony in arthropods.
Canadian Journal of Zoology, 58, 782–795.
Bartsch, I. (1987). Australacarus inexpectatus gen. et spec. nov. (Halacaroidea, Acari), mit einer
Übersicht über parasitisch lebende Halacariden. Zoologischer Anzeiger, 218, 17–24.
Bartsch, I. (1989). Marine mites (Halacaroidea, Acari): A geographical and ecological survey.
Hydrobiologia, 178, 21–42.
Bartsch, I. (1994). Halacarid mites (Acari) from hydrothermal deep-sea sites. New Records.
Cahiers de Biologie Marine, 35, 479–490.
Bartsch, I. (1996). Halacarids (Halacaroidea, Acari) in freshwater. multiple invasions from the
Paleozoic onwards? Journal of Natural History, 30, 67–99.
Bartsch, I., & Gerecke, R. (2011). A new freshwater mite of the marine genus Halacarellus (Acari:
Halacaridae) from the Austrian Alps (Styria, Gesäuse National Park): description and reflection
on its origin. Zoologischer Anzeiger, 250, 151–159.
References 273

Behan-Pelletier, V. M. (1996). Naiazetes reevesi n g, n sp (Acari: Oribatida: Zetomimidae) from


semi-aquatic habitats of eastern North America. Acarologia, 37, 345–355.
Behan-Pelletier, M., & Eamer, B. (2007). Aquatic Oribatida: Adaptations, constraints, distribution
and ecology. In J. B. Morales-Malacara, V. Behan-Pelletier, E. Ueckermann, T. M. Pérez, E.G.
Estrada-Venegas, & M. Badii (Eds.), Acarology XI: Proceedings of the international congress
(pp. 71–82). México: Instituto de Biololgía; Facultad de Ciencias; Universidad Nacional
Autónoma de México; Sociedad Latinoamericana de Acarología. México, 2002.
Böttger, K. (1962a). Zur Biologie und Ethologie der einheimischen Wassermilben Arrenurus
(Megaluracarus) globator (Müll.), 1776 Piona nodata nodata (Müll.), 1776 und Eylais
infundibulifera meridionalis (Thon), 1899 (Hydrachnellae, Acari). Zoologische Jahrbücher
Abteilung für Systematik, 89, 501–584.
Böttger, K. (1962b). Die Bedeutung des Lichtes für die Lage- und Richtungsorientierung einiger
Süsswassermilben (Hydrachnellae, Acari). Zoologische Anzeiger, 169, 476–484.
Böttger, K. (1969). Die Ernährungsweise der Wassermilbe Limnochares aquatica (L.)
(Hydrachnellae, Acari). Zoologischer Anzeiger Supplementband, 33, 85–91.
Böttger, K. (1970). Die Ernährungsweise der Wassermilben (Hydrachnellae, Acari). Internationale
Revue Der Gesamten Hydrobiologie, 55, 895–912.
Boulton, A., Harvey, M., & Proctor, H. (2004). Of spates and species: responses by interstitial
water mites to simulated spates in a subtropical Australian river. Experimental and Applied
Acarology, 34, 149–169.
Braun, F. (1931). Beiträge zur Biologie und Atmungsphysiologie der Argyroneta aquatica Cl.
Zoologische Jahrbücher Abteilung für Systematik, 62, 175–262.
Bristowe, W. (1958). The world of spiders. London: Collins.
Chalker-Scott, L. (1995). Survival and sex ratios of the intertidal copepod, Tigriopus californicus,
following ultraviolet-B (290–320 nm) radiation exposure. Marine Biology, 123, 799–804.
Chant, D. A., Denmark, H. A., & Baker, E. W. (1959). A new subfamily, Macroseiinae nov., of the
family Phytoseiidae (Acarina: Gamasina). Canadian Entomologist, 91, 808–811.
Chapman, R. F. (1982). The insects: Structure and function (5th ed.). Massachusetts: Harvard
University Press.
Chen, P.-R., & Zhang, Z.-Q. (1991). Biology of Allothrombium pulvinum Ewing (Acari,
Trombidiidae) and its impact on twospotted spider mite (Acari, Tetranychidae) in cotton fields.
Journal of Applied Entomology, 112, 31–37.
Cheng, L. (Ed.). (1976). Marine insects. Amsterdam: North Holland Publishing.
Cicolani, B., & Di Sabatino, A. (1991). Sensitivity of water mites to water pollution. In F. Dusbabek
& V. Bukva (Eds.), Modern acarology (Vol. 1, pp. 465–474). Prague: SPB Academic.
Clarke, C. M., & Kitching, R. L. (1993). The metazoan food webs from six Bornean Nepenthes
species. Ecological Entomology, 18, 7–16.
Conroy, J. C. (1974). The taxonomy and ecology of Unionicola crassipes crassipes (Müller), a
water mite parasitic on the freshwater sponge, Spongilla lacustris (Linne). In The 4th
International Congress of Acarology (pp. 959–964). Marion Lake, British Columbia.
Cook, D. R. (1969). The zoogeography of interstitial water mites. In The 2nd International
Congress of Acarology. Budapest: Akademiai Kiado.
Cook, D. R. (1974). Water mite genera and subgenera. Memoirs of the American Entomological
Institute, 21, 1–860.
Cook, D. R. (1991). Water mites from driven wells in New Zealand, the subfamily Notoaturinae
Besch (Acarina, Aturidae). Stygologia, 6, 235–253.
Cook, D. R. (1992). Water mites (Hydracarina), mostly from driven wells in New Zealand: Taxa
other than the Notoaturnae Besch. Stygologia, 7, 43–62.
Crome, W. (1951). Die Wasserspinne. Die Neue Brehm-Bücherei (Vol. 44). Leipzig: Geest &
Portig K.-G.
Cupp, J. K., & Willis, D. W. (1982). Occurrence of the mite Lebertia in a green sunfish (Lepomis
syanellus). Journal of Parasitology, 68, 876.
Cyr, H., & Downing, J. A. (1988). The abundance of phytophilous invertebrates on different
species of submerged macrophytes. Freshwater Biology, 20, 365–374.
274 7 Acari Underwater, or, Why Did Mites Take the Plunge?

Czeczuga, B., & Czerpak, R. (1968a). Pigments occurring in Hydrachna geographica and Piona
nodata (Hydracarina, Arachnoidea). Specialia, 24, 218–219.
Czeczuga, B., & Czerpak, R. (1968b). The presence of carotenoids in Eylais hamata (Koenike,
1897) (Hydracarina, Arachnoidea). Comparative Biochemistry and Physiology, 24, 37–46.
Czeczuga, B., & Czerpak, R. (1968c). Carotenoids in Hydryphantes dispar (Schaub, 1988)
(Hydracarina, Arachnoidea). Comparative Biochemistry and Physiology, 25, 547–552.
Davids, C. (1973). The water mite Hydrachna conjecta, bionomics and relation to species of
Corixidae. Netherlands Journal of Zoology, 23, 363–429.
Davids, C. (1991). Water mites: The impact of larvae and adults on their host and prey populations.
In F. Dusbábek & V. Bukva (Eds.), Modern acarology (Vol. I, pp. 497–501). The Hague:
Academia, Prague and SPB Academic Publishing.
Davids, C., & Belier, R. (1974). The spermatophores of Hydrachna conjecta Koenike and the life
history of the land-living ancestors of this water mite. In Proceedings of the 4th international
congress of acarology (pp. 147–151). Budapest: Akademiai Kiado.
Davids, C., Beintema, E. F., & Weekenstroo, J. E. (1981). Feeding rate and egg production in water
mites in relationship with temperature. Verhandlungen Internationale Vereiningung fuer
Theoretische und Angewandte Limnologie, 21, 1603–1606.
Davids, C., Holtslag, J., & Dimock, R. V., Jr. (1988). Competitive exclusion, harem behaviour and
host specificity of the water mite Unionicola ypsilophora (Hydrachnellae, Acari) inhabiting
Anodonta cygnea (Unionidae). Internationale Revue Der Gesamten Hydrobiologie, 73,
651–657.
Di Sabatino, A., Smit, H., Gerecke, R., Goldschmidt, T., Matsumoto, N., & Cicolani, B. (2008).
Global diversity of water mites (Acari, Hydrachnidia; Arachnida) in freshwater. Hydrobiologia,
595, 303–315.
Dimock, R. V., Jr. (1983). In defense of the harem: intraspecific aggression by male water mites
(Acari: Unionicolidae). Annals of the Entomological Society of America, 76, 463–465.
Dohet, A., Ector, L., Cauchie, H.-M., & Hoffmann, L. (2008). Identification of benthic invertebrate
and diatom indicator taxa that distinguish different stream types as well as degraded from
reference conditions in Luxembourg. Animal Biology, 58, 419–472.
Donnelly, M. A. (1991). Feeding patterns of the strawberry poison frog Dendrobates pumilio
(Anura: Dendrobatidae). Copeia, 1991, 723–730.
Edwards, D. D., & Dimock, R. V., Jr. (1991). Relative importance of size versus territorial
residency in intraspecific aggression by symbiotic male water mites (Acari: Unionicolidae).
Experimental & Applied Acarology, 12, 61–65.
Edwards, D. D., & Dimock, R. V., Jr. (1995). Specificity of the host recognition behaviour of larval
Unionicola (Acari: Unionicolidae): the effects of larval ontogeny and early larval experience.
Animal Behaviour, 50, 343–352.
Edwards, D. D., & Vidrine, M. F. (2013). Patterns of species richness among assemblages of
Unionicola spp. (Acari: Unionicolidae) inhabiting freshwater mussels (Bivalvia: Unionoida) of
North America. Journal of Parasitology, 99, 212–217. http://dx.doi.org/10.1645/GE-3208.1.
Ellis-Adam, A. C., & Davids, C. (1970). Oviposition and post-embryonic development of the
watermite Piona alpicola (Neuman, 1880). Netherlands Journal of Zoology, 20, 122–137.
Elton, C. S. (1923). On the colours of water-mites. Proceedings of the Zoological Society of
London, 82, 1231–1239.
Eriksson, M. O. G., Henrikson, L., & Oscarson, H. G. (1980). Predator–prey relationships among
water-mites and other freshwater organisms. Archiv für Hydrobiologie, 88, 146–154.
Ernst, H. (1996). The distribution and zonation of terrestrial mites inhabiting artificial rocky shores
of estuaries in northern Germany. In R. Mitchell, D. J. Horn, G. R. Needham, & W. C. Welbourn
(Eds.), Acarology IX: Volume 1, proceedings (pp. 529–532). Columbus: Ohio Biological Survey.
Fain, S., & Synnot, R. (1981). Hyadesia australian sp. n. (Astigmata, Hyadesiidae) from South-
Eastern Australia. Bulletin and annales de la Société Royale Belge d’Entomologie, 117,
195–201.
Fashing, N. F. (2008). Mate-guarding in the genus Creutzeria (Astigmata: Histiostomatidae), an
aquatic mite genus inhabiting the fluid-filled pitchers of Nepenthes plants (Nepentheaceae).
Systematic & Applied Acarology, 13, 163–171.
References 275

Fashing, N. J. (1994). Life-history patterns of astigmatid inhabitants of water-filled treeholes. In


M. A. Houck (Ed.), Mites: Ecological and evolutionary studies of life-history patterns (pp.
160–185). New York: Chapman & Hall.
Fashing, N. J. (1998). Functional morphology as an aid in determining trophic behaviour: The
placement of astigmatic mites in food webs of water-filled tree-hole communities. Experimental
& Applied Acarology, 22, 435–453.
Fashing, N. J. (2010). Life history and biology of Hormosianoetus mallotae (Fashing)
(Histiostomatidae: Astigmata), an obligatory inhabitant of water-filled treeholes. International
Journal of Acarology, 36, 189–198.
Fashing, N. J., & Chua, T. H. (2002). Systematics and ecology of Naiadacarus, a new species of
Acaridae (Acari: Astigmata) inhabiting the pitchers of Nepenthes bicalcarata Hook. F. in
Brunei Darussalam. International Journal of Acarology, 28, 157–167.
Fashing, N. J., & Marcuson, K. S. (1996). Fine structure of the axillary organs of Fusohericia
lawrencei Baker and Crossley (Astigmata: Algophagidae). In R. Mitchell, D. J. Horn, G. R.
Needham, & W. C. Welbourn (Eds.), Acarology IX: Volume 1, proceedings (pp. 381–384).
Columbus: Ohio Biological Survey.
Fashing, N. J., OConnor, B. M., & Kitching, R. L. (1996). Adaptations for swimming in the genus
Creutzeria (Astigmata: Histiostomatidae). In R. Mitchell, D. J. Horn, G. R. Needham, & W. C.
Welbourn (Eds.), Acarology IX: Volume 1, proceedings (pp. 385–388). Columbus: Ohio
Biological Survey.
Fisher, G. R., Kuhn, R. E., & Dimock, R. V., Jr. (2000). The symbiotic water mite Unionicola
formosa (Acari: Unionicolidae) ingests mucus and tissue of its molluscan host. The Journal of
Parasitology, 86, 1254–1258.
Flowers, M. A., & Graves, B. M. (1995). Prey selectivity and size-specific diet changes in Bufo
cognatus and B. woodhousii during early postmetamorphic ontogeny. Journal of Herpetology,
23, 608–612.
Freundlieb, U. (1979). Zur Ökologie der Hydrachnellae (Acari) des Schierenseebaches. Archiv für
Hydrobiologie/Supplementband 54:509–538.
Garga, N. (1996). Aposematism in water mites (Acari: Hydracarina): A predator? Defense
mechanism, a phylogenetic hold-over and protection from damaging light. M.Sc. thesis,
Queen’s University, Kingston.
Gerecke, R. (1991). Taxonomische, faunistiche und ökologische Untersuchungen an Wassermilben
(Acari, Actinedida) aus Sizilien unter Berücksichtigung anderer aquatischer Invertebraten.
Lauterbornia., 7, 1–303.
Gerecke, R., & Schwoerbel, J. (1991). Water quality and water mites (Acari, Actinedida) in the
upper Danube region, 1959–1984. In F. Dusbabek & V. Bukva (Eds.), Modern acarology (Vol.
1, pp. 483–491). Prague: SPB Academic.
Gerecke, R., & Smith, I. M. (1993). On the biology of the spring-dwelling water mite Nilotonia
longipora (Walter, 1925) (Acari, Actinedida, Anisitsiellidae). In 2nd symposium (pp. 377–383).
Krynica: EURAAC.
Gledhill, T. (1985). A new species of water-mites, Unionicola (Pentatax) macani (Unionicolidae,
Hydrachnellae, Acari), from the mantle cavity of the prosobranch mollusc Lanistes ovum
Peters in Nigeria, with remarks on some aspects of host/parasite relationships between
unionicolids and molluscs. Archiv für Hydrobiologie, 104, 77–92.
Gliwicz, Z. M., & Biesiadka, E. (1975). Pelagic water mites (Hydracarina) and their effect on the
plankton community in a neotropial man-made lake. Archiv für Hydrobiologie, 76, 65–88.
Goldschmidt, T., & Fu, V. W. K. (2011). Description of Hygrobates aloisii sp. nov. from Hong
Kong, a new species of Hygrobates (Lurchibates) subgen. nov. (Acari, Hydrachnidia,
Hygrobatidae), with data on the parasite–host relationship to the Hong Kong Newt
Paramesotriton hongkongensis (Amphibia, Caudata, Salamandridae). Zoologischer Anzeiger,
250, 19–31.
Goldstrohm, D. D., & Lilly, V. G. (1965). The effect of light on the survival of pigmented and non-
pigmented cells of Dacropinax spathularia. Mycologia, 57, 612–623.
Green, J. (1964). Pigments of the Hydracarine Eylais extendens (Acari: Hydrachnellae).
Comparative Biochemistry and Physiology, 13, 469–472.
276 7 Acari Underwater, or, Why Did Mites Take the Plunge?

Hairston, N. G., Jr. (1976). Photoprotection by carotenoid pigments in the copepod Diaptomus
nevadensis. Proceedings of the National Academy of Science, 73, 971–974.
Halik, R. N. St. L. (1924). Contributions to the physiology of the water mites (Hydracarina) (Vol.
17, pp. 3–9). Prague: Publ. Fac. Sci. Univ. Charles. (in Czechoslovakian with English summary).
Harvey, M. S. (1990). Perzidae, a new freshwater mite family from Australia (Acarina:
Halacaroidea). Invertebrate Taxonomy, 3, 771–781.
Harvey, M. S. (1998). The Australian water mites: A guide to families and genera. In Monographs
on invertebrate taxonomy (Vol. 4). Collingwood: CSIRO Publishing.
Hevers, J. (1980). Biologisch-ökologische Untersuchungen zum Entwicklungszyklus der in
Deutschland auftretenden Unionicola-Arten (Hydrachnellae, Acari). Archiv für Hydrobiologie
Supplementband, 57, 324–373.
Hinton, H. E. (1971). Plastron respiration in the mite, Platyseius italicus. Journal of Insect
Physiology, 17, 1185–1199.
Hoogstraal, H. (1973). Acarina (ticks). In A. J. Gibbs (Ed.), Viruses and invertebrates (pp. 89–103).
Amsterdam: North Holland Publishing.
Jenkins, B., & Kitching, R. L. (1990). The ecology of water-filled treeholes in Australian
rainforests: Food web reassembly as a measure of community recovery after disturbance.
Australian Journal of Ecology, 15, 199–205.
Johnson, D., Akre, B. G., & Crowley, P. H. (1975). Modeling arthropod predation: Wasteful killing
by damselfly naiads. Ecology, 56, 1081–1093.
Jones, D., & Morgan, G. (1994). A field guide to crustaceans of Australian waters. Sydney: Reed
Books.
Keeley, E. R., & Grant, J. W. A. (1997). Allometry of diet selectivity in juvenile Atlantic salmon
(Salmo salar). Canadian Journal of Fisheries and Aquatic Sciences, 54, 1894–1902.
Kerfoot, W. C. (1982). A question of taste: Crypsis and warning coloration in freshwater zooplankton
communities. Ecology, 63, 538–554.
Kitching, R. L. (2000). Food webs and container habitats: The natural history and ecology of
phytotelmata. New York: Cambridge University Press.
Kitching, R. L., & Callaghan, C. (1982). The fauna of water-filled tree holes in box forest in south-
east Queensland. Australian Entomological Magazine, 8, 61–70.
Klemm, D. J., Blocksom, K. A., Fulk, F. A., Herlihy, A. T., Hughes, R. M., Kaufman, P. R., Peck,
D. V., Stoddard, J. L., Thoeny, W. T., Griffith, M. B., & Davis, W. S. (2003). Development and
evaluation of a Macroinvertebrate Biotic Integrity Index (MBII) for regionally assessing Mid-
Atlantic Highlands Streams. Environmental Management, 31, 656–669.
Koehler, H. H. (1992). The use of soil mesofauna for the judgement of chemical impact on
ecosystems. Agriculture, Ecosystems and Environment, 40, 193–205.
Krantz, G. W. (1974). Phaulodinychus mitis (Leonardi 1899) (Acari: Uropodidae) an intertidal
mite exhibiting plastron respiration. Acarologia, 16, 11–20.
Krantz, G. W. (1978). A manual of acarology. Corvallis: Oregon State University Bookstores.
Krantz, G. W., & Baker, G. T. (1982). Observations on the plastron mechanism of Hydrozetes sp.
(Acari: Oribatida: Hydrozetidae). Acarologia, 23, 273–277.
Krantz, G. W., & Walter, D. E. (Eds.). (2009). A manual of acarology (3rd ed.). Lubbock: Texas
Tech University Press. 807 p. 338 b/w illustrations; 60 figures ISBN 978-0-89672-620-8.
Lagerlöf, J., & Andrén, O. (1988). Abundance and activity of soil mites (Acari) in four cropping
systems. Pedobiologia, 32, 129–145.
Laird, M. (1947). Some natural enemies of mosquitoes in the vicinity of Palmalmal, New Britain.
Transactions of the Royal Society of New Zealand, 76, 453–476.
Learner, M. A., & Chawner, H. A. (2003). Macro-invertebrate associations in sewage filter-beds
and their relationship to operational practice. Journal of Applied Ecology, 35, 720–747.
Luxton, M. (1990). The marine littoral mites of the New Zealand Region. Journal of the Royal
Society of New Zealand, 20, 367–418.
MacQuitty, M. (1984). The feeding behaviour of two species of Agauopsis (Halacaroidea) from
California. In D. A. Griffith & C. E. Bowman (Eds.), Acarology VI (pp. 571–580). New York:
John Wiley.
References 277

Manunta, C. (1939). Estraazione e cristallizzazione del pigmento che colora in rosso la pelle di
certi acari del genere Trombidium. Helvetica Chimica Acta, 22, 1154–1155.
Martin, P. (2005). Water mites (Hydrachnidia, Acari) as predators in lotic environments.
Phytophaga, 16, 307–321.
Mathews, M. M., & Sistrom, W. R. (1959). Function of carotenoid pigments in non-photosynthetic
bacteria. Nature, 184, 1892–1893.
Meyer, E., & Kabbe, K. (1991). Pigmentation in water mites of the genera Limnochares
Latr. and Hydrodroma Koch (Hydrachnidia). In R. Schuster & P. W. Murphy (Eds.), The
acari: Reproduction, development and life-history strategies (pp. 379–391). New York:
Chapman & Hall.
Mitchell, R. (1972). The tracheae of watermites. Journal of Morphology, 136, 327–335.
Modlin, R. F., & Gannon, J. E. (1973). A contribution to the ecology and distribution of aquatic
Acari in the St. Lawrence Great Lakes. Transactions of the American Microscopical Society,
92, 217–224.
Mwango, J., Williams, T., & Wiles, R. (1995). A preliminary study of the predator–prey
relationships of watermites (Acari: Hydrachnidia) and blackfly larvae (Diptera: Simuliidae).
The Entomologist, 114, 107–117.
Naeem, S. (1988). Resource heterogeneity fosters coexistence of a mite and a midge in pitcher
plants. Ecological Monograph, 58, 215–227.
Naeem, S. (1990). Resource heterogeneity and community structure: a case study in Heliconia
imbricata phytotelmata. Oecologia, 84, 29–38.
Nagel, L., Zanuttig, M., & Forbes, M. R. (2011). Escape of parasitic water mites from dragonfly
predators attacking their damselfly hosts. Canadian Journal of Zoolgy, 89, 213–298.
Newell, I. M. (1945). Hydrozetes Berlese (Acari, Oribatoidea): the occurrence of the genus in
North America, and the phenomenon of levitation. Transactions of the Connecticut Academy of
Arts and Sciences, 36, 253–275.
Norton, R. A. (1994). Evolutionary aspects of oribatid mite life histories and consequences for the
origin of the Astigmata. In M. A. Houck (Ed.), Mites: Ecological and evolutionary studies of
life-history patterns (pp. 99–135). New York: Chapman & Hall.
Norton, R. A., Williams, D. D., Hogg, I. D., & Palmer, S. C. (1988). Biology of the oribatid mite
Mucronothrus nasalis (Acari: Oribatida: Trhypochthoniidae) from a small coldwater spring-
brook in eastern Canada. Canadian Journal of Zoology, 66, 622–628.
Norton, R. A., Graham, T. B., & Alberti, G. (1996). A rotifer-eating ameronothroid (Acari:
Ameronothridae) mite from ephemeral pools on the Colorado plateau. In R. Mitchell, D. J.
Horn, G. R. Needham, & W. C. Welbourn (Eds.), Acarology IX: Volume 1. Columbus: Ohio
Biological Survey.
OConnor, B. M. (1994). Life-history modifications in astigmatid mites. In M. A. Houck (Ed.),
Mites: Ecological and evolutionary analyses of life-history patterns (pp. 136–159). New York:
Chapman & Hall.
Olmeda, A. S., Blanco, M. M., Pérez-Sánchez, J. L., Luzón, M., Villarroel, M., & Gibello, A. (2011).
Occurrence of the oribatid mite Trhypochthoniellus longisetus longisetus (Acari:
Trhypochthoniidae) on tilapia Oreochromis niloticus. Diseases of Aquatic Organisms, 94, 77–81.
Olomski, R. (1991). Effects of the habitat salinity on the osmotic and ionic regulation of the
freshwater mites Eylais tantilla Koen. and E. planiceps novata Viets (Acari: Hydrachnidia). In F.
Dusbabek & V. Bukva (Eds.), Modern acarology (Vol. 2, pp. 357–364). Prague: SPB Academic.
Paterson, C. G. (1970). Water mites (Hydracarina) as predators of chironomid larvae (Insecta:
Diptera). Canadian Journal of Zoology, 48, 610–614.
Pennak, R. W. (1978). Freshwater invertebrates of the United States (2nd ed.). Toronto: Wiley.
Pepato, A. R., & da Rocha, C. E. R. (2010). On spermiogenesis, sperm cell morphology and
accompanying secretions of Copidognathus (Acari: Halacaridae). Zoologischer Anzeiger, 249,
151–164.
Pfingstl, T. (2013). Resistance to fresh and salt water in intertidal mites (Acari: Oribatida):
Implications for ecology and hydrochorous dispersal. Experimental & Applied Acarology
http://www.ncbi.nlm.nih.gov/pubmed/23456607
278 7 Acari Underwater, or, Why Did Mites Take the Plunge?

Philip, C. B. (1953). Tick talk. Scientific Monthly, 76, 77–84.


Pieczynski, E. (1964). Analysis of numbers, activity, and distribution of water mites (Hydracarina),
and of some other aquatic invertebrates in the lake littoral and sublittoral. Ekologia Polska, 12,
691–733.
Pieczynski, E. (1976). Ecology of water mites (Hydracarina) in lakes. Polish Ecological Studies,
2, 5–54.
Preston-Mafham, R. A., & Preston-Mafham, K. G. (1984). Spiders of the world. London:
Blandford.
Proctor, H. C. (1988). The life history and predatory biology of Unionicola crassipes (Müller)
(Acari: Unionicolidae) in an Albertan foothills pond. M.Sc. thesis, University of Calgary,
Calgary.
Proctor, H. C. (1991). The evolution of copulation in water mites: A comparative test for
nonreversing characters. Evolution, 45, 558–567.
Proctor, H. C. (1992a). Mating and spermatophore morphology of water mites (Acari:
Parasitengona). Zoological Journal of the Linnean Society, 106, 341–384.
Proctor, H. C. (1992b). Discord between field and laboratory sex ratios of the water mite Neumania
papillator Marshall (Acari: Unionicolidae). Canadian Journal of Zoology, 70, 2483–2486.
Proctor, H. C. (Ed.). (2004). Aquatic mites: from genes to communities. Dordrecht: Kluwer
Academic. (book format of Experimental and Applied Acarology 34, issues 1 and 2).
Proctor, H. C. (2007). Aquatic mites in assessments of stream invertebrate diversity. In J. B.
Morales-Malacara, V. Behan-Pelletier, E. Ueckermann, T. M. Pérez, E. G. Estrada-Venegas, &
M. Badii (Eds.) Acarology XI: Proceedings of the International Congress (pp. 105–117).
México: Instituto de Biololgía; Facultad de Ciencias; Universidad Nacional Autónoma de
México; Sociedad Latinoamericana de Acarología. México, 2002.
Proctor, H. C., & Garga, N. (2004). Red, distasteful water mites: Did fish make them that way.
Experimental & Applied Acarology, 34, 127–147.
Proctor, H. C., & Pritchard, G. (1989). Neglected predators: Water mites (Acari: Parasitengona:
Hydrachnellae) in freshwater communities. Journal of the North American Benthological
Society, 8, 100–111.
Proctor, H. C., & Pritchard, G. (1990). Prey detection by the water mite Unionicola crassipes
(Acari: Unionicolidae). Freshwater Biology, 23, 271–279.
Proctor, H. C., Gray, H. M., & OConnor, B. M. (1997). Subaquatic mites (Acari: Astigmata)
associated with adult freshwater leeches (Hirudinea: Erpobdellidae). Journal of Natural
History, 31, 539–544.
Pugh, P. J. A. (1996a). Edaphic oribatid mites (Cryptostigmata: Acarina) associated with an
aquatic moss on sub-Antarctic South Georgia. Pedobiologia, 40, 113–117.
Pugh, P. J. A. (1996b). Using artificial substrata to monitor how cryptofaunal Acari colonize
littoral algae on sub-Antarctic South Georgia. Acarologia, 37, 189–200.
Pugh, P. J. A., King, P. E., & Fordy, M. R. (1987). Structural features associated with respiration in
some intertidal Uropodina (Acarina: Mesostigmata). Journal of Zoology Lond., 211,
107–120.
Puncochar, P. (1971). On the occurrence of water mites (Hydrachnellae) in the submerged
vegetation of a carp pond. In M. Daniel & B. Rosicky (Eds.), Proceedings of the 3rd
international congress of acarology (pp. 173–175) . The Hague: Junk.
Puncochar, P., & Hrbacek, J. (1991). Water mites in the plankton of Hubenov Reservoir and their
relations to fish stock composition. In F. Dusbabek & V. Bukva (Eds.), Modern acarology (Vol.
1, pp. 449–457). Prague: SPB Academic.
Putman, W. L. (1970). Life history and behavior of Balaustium putmani. Annals of the
Entomological Society of America, 63, 76–81.
Rajendran, R., & Prasad, R. S. (1989). Encentridiophorus similis (Acarina: Unionicolidae) an
active predator of mosquito larvae. Current Science, 58, 466–467.
Resh, V. H., & McElravy, E. P. (1993). Contemporary quantitative approaches to biomonitoring
using benthic macroinvertebrates. In D. M. Rosenberg & V. H. Resh (Eds.), Freshwater
biomonitoring and benthic macroinvertebrates (pp. 159–194). New York: Chapman & Hall.
References 279

Rieradevall, M., & Gil, M. J. (1993). Distribution, density and specific composition of water mites
(Acari) in the sublittoral of Lake Banyoles (Spain). Annales de Limnologie, 29, 41–46.
Riessen, H. P. (1982). Pelagic water mites: Their life history and seasonal distribution in the
zooplankton community of a Canadian lake. Archiv für Hydrobiologie Supplementband, 62,
410–439.
Riessen, H. P. (1985). Exploitation of prey seasonality by a planktonic predator. Canadian Journal
of Zoology, 63, 1729–1732.
Rosenberg, D. M., & Resh, V. H. (1993). Freshwater biomonitoring and benthic macroinvertebrates.
New York: Chapman & Hall.
Roth, V. D., & Brown, W. L. (1976). Other intertidal air-breathing arthropod spp. In L. Cheng
(Ed.), Marine insects (pp. 119–150). Amsterdam: North Holland Publishing.
Rousch, J. M., Simmons, T. W., Kerans, B. L., & Smith, B. P. (1997). Relative acute effects of low
pH and high iron on the hatching and survival of the water mite (Arrenurus manubriator) and the
aquatic insect (Chironomus riparius). Environmental Toxicology and Chemistry, 16, 2144–2150.
Schatz, H., & Gerecke, R. (1996). Hornmilben aus Quellen und Quellbächen im Nationalpark
Berchtesgaden (Oberbayern) und in den Südlichen Alpen (Trentino-Alto Adige). Berichte des
Naturwissenschaftlich- Medizinischen Vereins in Innsbruck, 83, 121–134.
Schmidt, H.-W. (1969). Tages- und jahresperiodische Driftaktivität der Wassermilben (Hydrachnellae,
Acari). Oecologia, 3, 240–248.
Schütz, D., & Taborsky, M. (2003). Adaptations to an aquatic life may be responsible for the
reversed sexual size dimorphism in the water spider, Argyroneta aquatica. Evolutionary
Ecology Research, 5, 105–117.
Siemer, F. (1996). Distribution, zonation, phenology, and life cycle of the Halacaridae (Prostigmata)
of artificial rocky shores located at the Weser estuary (Germany). In R. Mitchell, D. J. Horn, G.
R. Needham, & W. C. Welbourn (Eds.), Acarology IX: Volume 1, proceedings (pp. 547–551).
Columbus: Ohio Biological Survey.
Simmons, T. W., & Smith, I. M. (1984). Morphology of larvae, deutonymphs, and adults of the
water mite Najadicola ingens (Prostigmata: Parasitengona: Hygrobatoidea) with remarks on
phylogenetic relationships and revision of taxonomic placement of Najadicolinae. The
Canadian Entomologist, 116, 691–701.
Smit, H. & Alberti, G. (2010). The water mite family Pontarachnidae, with new data on its peculiar
morphological structures (Acari: Hydachnidia). In M. W. Sabelis & J. Bruin (Eds.), Trends in
Acarology: Proceedings of the 12th international congress (pp. 71–79). Amsterdam: Springer.
670 pp.
Smith, B. P. (1983). The potential of mites as biological control agents of mosquitoes. In M. Hoy,
G. Cunningham, & L. Knutson (Eds.), Biological control of pests by mites (pp. 79–85). Agric.
Exp. Sta. Univ. Calif., Special Pub. No. 3304.
Smith, B. P. (1988). Host–parasite interaction and impact of larval water mites on insects. Annual
Review of Entomology, 33, 487–507.
Smith, B. P., & Barr, D. (1977). Swimming by the water mite Limnochares americana Lundblad
(Acari, Parasitengona, Limnocharidae). Canadian Journal of Zoology, 55, 2050–2059.
Smith, I. M., & Cook, D. R. (1991). Water mites. In J. H. Thorp & A. P. Covich (Eds.), Ecology
and classification of North American freshwater invertebrates (pp. 523–592). San Diego:
Academic.
Sokolow, I. I. (1977). The protective envelopes in the eggs of Hydrachnellae. Zoologischer
Anzeiger, 198, 36–46.
ten Winkel, E. H., & Davids, C. (1985). Bioturbation by cyprinid fish affecting the food availability
for predatory water mites. Oecologia, 67, 218–219.
ten Winkel, E. H., Davids, C., & de Nobel, J. G. (1989). Food and feeding strategies of water mites
of the genus Hygrobates and the impact of their predation on the larval population of the
chironomid Cladotanytarsus mancus (Walker) in Lake Maarsseveen. Netherlands Journal of
Zoology, 39, 246–263.
Timms, B. V. (1993). Saline Lakes of the Paroo, Inland New South Wales, Australia. Hydrobiologia,
267, 269–289.
280 7 Acari Underwater, or, Why Did Mites Take the Plunge?

Ullrich, F. (1976). Biologisch-ökologische Studien an rheophilen Wassermilben (Hydrachnellae,


Acari), unter besonderer Berücksichtigung von Sperchon setiger (Thor 1898). Ph.D. thesis,
University of Kiel, Kiel.
Viets, K. O. (1980). New Unionicolidae (Acari, Hydrachnellae) from Australia. Transactions of
the Royal Society of South Australia, 104, 27–40.
Ward, J. (1994). The New Zealand marine caddisflies (Trichoptera). Weta, 17, 18–20.
Welbourn, W. C. (1991). Phylogenetic studies of the terrestrial Parasitengona. In F. Dusbábek & V.
Bukva (Eds.), Modern acarology (2nd ed., pp. 163–170). The Hague: Academia, Prague &
SPB Academic Publishing bv.
Welsh, J. H. (1930). Reversal of phototropism in a parasitic water mite. The Biological Bulletin,
59, 165–169.
Wiles, P. R. (1982). A note on the watermite Hydrodroma despiciens feeding on chironomid egg
masses. Freshwater Biology, 12, 83–87.
Wiles, P. R. (1984). Watermite respiratory systems. Acarologia, 25, 27–31.
Winterbourn, M. J., & Anderson, N. H. (1980). The life history of Philanisus plebius Walker
(Trichoptera: Chathamiidae), a caddisfly whose eggs were found in a starfish. Ecological
Entomology, 5, 293–303.
Witte, H. (1984). The evolution of the mechanisms of reproduction in the Parasitengonae (Acarina:
Prostigmata). In D. A. Griffiths & C. E. Bowman (Eds.), Acarology 6 (Vol. 1, pp. 470–478).
Chichester: Ellis Horwood.
Witte, H. (1991). Indirect sperm transfer in prostigmatic mites from a phylogenetic viewpoint. In
R. Schuster & P. W. Murphy (Eds.), The Acari: Reproduction, development and life-history
strategies (pp. 137–176). London: Chapman & Hall.
Wohltmann, A. (1996). On the life cycle and parasitism of Johnstoniana errans (Johnston) 1852
(Acari: Prostigmata: Parasitengonae). Acarologia, 37, 201–209.
Young, W. C. (1969). Ecological distribution of Hydracarina in north central Colorado. American
Midland Naturalist, 82, 367–401.
Young, W. C., & Rhodes, A. C. (1974). The influence of dissolved oxygen concentrations on three
species of water mites (Hydracarina). American Midland Naturalist, 92, 115–129.
Chapter 8
Mites on Plants

Pondering imponderables may not be scientific, but all hypotheses have to start
somewhere. When Terry Erwin (1982) published his estimate of 30 million living
species of arthropods, up to 20 times that of previous estimates, much brouhaha
ensued. His basic conclusion, however, “that current estimates of Arthropod species
numbers [have been] grossly underestimated” is now common wisdom. Erwin’s
estimate was a towering house of assumptions based on the diversity of beetles
from the canopy of a single tree species in a seasonal forest in Panama. At the time,
little was known of mite diversity in rainforest canopies, so mites played little
role in his estimate, other than being part of the 60 % of described arthropods that
are not beetles. One of Erwin’s assumptions was that the canopy fauna was twice
as diverse as the forest floor fauna. As we learned in Chap. 6, the centres of mite
diversity have long been thought to be in the soil and recent research seems to
support that assumption (e.g. Young et al. 2012). Could mites on plants be twice as
diverse as in the soil? (Fig. 8.1)
Because of their economic importance, mites parasitic on crop plants have long
been known (e.g. Jeppson et al. 1975; Helle and Sabelis 1985a, b; Lindquist et al.
1996). Almost any orchard or field crop is likely to be infested with spider mites,
false spider mites, erinose mites, rust mites, earth mites or any of a host of other
pests – and this is true of rainforest plants as well. Studies in native forests and
plantations throughout the world have demonstrated that trees are aswarm with
free-living mites (e.g. Walter and Proctor 1999; Walter and Behan-Pelletier 1999;
Walter 2004; Lindo and Winchester 2006; Lindo et al. 2008; Affeld et al. 2009;
Beaulieu et al. 2010). In fact, a fine patina of acarines adds its lustre to the outer-
most layer of forests and a cornucopia of Acari thrive between the outer canopy
and the forest floor. Is it reasonable to assume that this arboreal diversity could rival
that of beetles?

D.E. Walter and H.C. Proctor, Mites: Ecology, Evolution & Behaviour: Life at a Microscale, 281
DOI 10.1007/978-94-007-7164-2_8, © Springer Science+Business Media Dordrecht 2013
282 8 Mites on Plants

Fig. 8.1 A red velvet mite (Charletonia sp.) has clambered up a stalk of grass to attack a spider
egg case while horrified spiderlings look on and a mélange of arboreal mites strolls by (Photo by
HC Proctor, SEMs by DE Walter)

Mites on Plants: Where Do They Come From?

The surface and interstices of soil, and the decaying vegetation that accumulates in
this matrix, are the ancestral homes of mites and still contain the most impressive
acarofaunas (see Chap. 6). Except in the driest or most frozen of soils, mite densities
in the tens to hundreds of thousands per square metre are commonly reported, and
a handful of forest humus often contains dozens of species of mites. This detritus-
based diversity is known to extend into the canopy in the suspended soils that collect
in treeholes, crutches and among the leaves of epiphytes (Nadkarni and Longino
1990; Paoletti et al. 1991; Walter 2004; Lindo and Winchester 2006; Beaulieu et al.
2010; Yoshida and Hijii 2011). Although some soil mites wander up tree trunks and
colonise these patches of hanging humus, most of the mite species found on tree
trunks, branches and leaves are truly arboreal and only descend to the forest floor
through mischance (Walter and Behan-Pelletier 1999; Proctor et al. 2002; Beaulieu
et al. 2006; Fischer et al. 2010). Patterns at the family and generic levels indicate
that relatively few lineages have made the transition to living on the surface of
plants but some of these have undergone spectacular radiations.
In the rich Baltic amber fossils of the mid-Tertiary, many of today’s characteristically
arboreal mite taxa can be found. However, these fossil assemblages do not mirror
the common view of arboreal mites derived from agricultural ecosystems. Relatively
few are predators (e.g. Erythraeidae, Anystidae, Cheyletidae, Phytoseiidae, and
Digamasellidae) or herbivores (e.g. Eriophyioidea, Tetranychoidea) (Selden 1993;
Labandera et al. 1997). Instead, most mites in amber have living relatives that feed
Mites on Plants: Where Do They Come From? 283

on fungi, scavenge dead plant matter or graze on algae and lichens growing on
plants: the grazers, browsers and piercing–sucking microbivore functional groups
so common in soil (see Chap. 6), but lineages that have moved up in the world.
Although this is not what we are used to in agricultural systems with high chemical
pesticide usage where most mites are likely to be pests, it is the norm in less con-
tinuously disturbed systems.
For example, the trunks, stems and twigs of trees are home to numerous oribatid
mites that occur in soil samples only through mischance. Most trees have three-
dimensional bark with a vast area of cracks and crevices in which to hide and a rich
flora of algae, fungi, and lichens on which these mites can feed. Yet many genera
and several families of oribatids are characteristically found on bark. Some, such as
members of the Camisiidae, Crotoniidae, Oribatulidae and Plateremaeidae, are
known from Tertiary and Cretaceous fossils, so this association is probably ancient.
The surface of a leaf is a more challenging habitat for any mite, because they are
more exposed to the dehydrating effects of wind and dislodgement by the floods
during rain (Walter 2004). Oribatid mites, since they cannot access water and nutri-
ents in living leaves, must find leaves an especially challenging environment given
their tendency towards slow, ponderous movement and long generation times.
Yet, a number of oribatid mite genera are foliar specialist and exhibit distinctive
morphological attributes associated with leaf-dwelling and probably specialized
behaviours as well (Walter and Behan-Pelletier 1993; Behan-Pelletier and Walter
2007). One of the more characteristic families of arboreal oribatid mites is the
Cymbaeremaeidae and some its genera have undergone impressive radiations on
trees. For example, the genus Scapheremaeus currently has about 115 described
species and can be found throughout much of the world (Colloff 2010, 2011).
Again, this is probably an ancient association because one cymbaeremaeid,
Jujeremus foveolatus, is known from fossils in the late Jurassic.
Armoured mites like the oribatids are the most likely kind of mites to form fos-
sils. However, when coniferous trees with sticky sap appeared, even soft-bodied
mites could be preserved as fossils in amber. For example, tydeid mites are ubiqui-
tous on modern plants and a tydeid has been found in Cretaceous amber (Selden
1993). Many other arboreal mites trapped in amber (Zacharda and Krivolutsky
1985; Krivolutsky and Druk 1986; Selden 1993) are known from the Cretaceous
and Triassic eriophyoid mites trapped in microamber granules have recently been
described (Schmidt et al. 2012).
Australian rainforest canopies often have densities of 1,000–10,000 mites per
metre of leaf area – populations that might terrify an orchardist – but the vast major-
ity of these mites are fungivores, especially oribatids and prostigmatans in the
Tydeidae and Tarsonemidae (Walter and O’Dowd 1995a). Oribatid mites with
pincer-like chelicerae feed on fungal hyphae, algal strands and dead leaf or woody
material. Except for pollen (usually windblown), oribatids mites rarely feed on the
living tissue of green plants, probably because chelate-dentate mouthparts are not
useful for puncturing the epidermis of plants. It is mites with needle-like movable
cheliceral digits that gave rise to mites parasitic on higher plants (Krantz and
Lindquist 1979; Lindquist 1998). Obtaining the leverage to force cheliceral stylets
284 8 Mites on Plants

through the epidermis of a leaf was probably the main problem that these mites had
to overcome. Most of these minute herbivores stab individual plant cells and suck
out their contents, providing a ready source of moisture and allowing them to avoid
localised plant chemical defences (e.g. feeding around vacuoles). Plant parasitism
has evolved repeatedly in different lineages of mites so that today most species of
conifers, monocots and dicots, and many ferns, are attacked by one or more species
of mite (Jeppson et al. 1975; Helle and Sabelis 1985a, b; Lindquist et al. 1996).
In agriculture systems that rely on chemicals to control insect and fungal pests,
secondary outbreaks of plant-parasitic mites are common. The predatory mites that
usually control these pests are more susceptible to pesticides for a number of rea-
sons, but all else being equal, the very nature of a predator does them in: they must
search for prey and have a higher likelihood of encountering a drop of poison than
a sedentary plant-parasite. Fungivores and omnivore-scavengers are also relatively
rare in agricultural systems, direct or indirect victims of fungicides, insecticides,
and acaricides. So in chemical-intensive agricultural systems, most of mites encoun-
tered are likely to be pests.

Plant Parasites

Mites parasitic on the leaves of plants belong to the Acariformes and primarily to
the suborder Prostigmata. As in all Prostigmata, the plant-parasites feed only on
fluids. Some predatory prostigmatans (e.g. Labidostommatidae, Rhagidiidae) have
chelate chelicerae that they use to mash their prey to extract fluids. Most prostig-
matans have stylet-like mouthparts that are used to puncture prey, or in the case of
fungivores and plant-parasites, individual cells and drain their contents (Lindquist
1998). Phytophagy has arisen within at least seven clades of Prostigmata, including
two major radiations of obligate parasites of vascular plants (Eriophyoidea and
Tetranychoidea) and several less evolutionarily successful experiments within
other lineages (Erythraeidae, Eupodoidea, Stigmaeidae, Tydeidae and Tarsonemidae)
that feed on vascular plants, bryophytes, and algae. Balaustium medicagoense
(Erythraeidae) is an especially interesting example because it is a general feeder
considered both a predator of pests and a pest in grain and pasture crops (Arthur
et al. 2011). Tarsonemid mites are an especially interesting example because at least
four independent evolutions of vascular plant-parasitism from mycophagous ances-
tors seem to have occurred (Lindquist 1986, 1998).

Rust, Gall and Erinose Mites: Eriophyoidea

Most of the approximately 6,000 described species of plant-parasitic mites belong


to the three families of the Eriophyoidea: Phytoptidae, Diptilomiopidae and
Eriophyidae. The association between eriophyoid mites and vascular plants appears
Plant Parasites 285

Fig. 8.2 Eriophyoid mites


are strange, worm-like
animals with only two pairs
of legs at the anterior end.
All 4,000 known species are
plant parasites and may can
induce galls. This specimen
from the prairie in Alberta,
Canada, probably feeds on
grasses SEM by HC Proctor

to be an ancient one (Lindquist 1998; Schmidt et al. 2012), which may help explain
the extreme diversity of this group. About 4,000 species of gall and rust mites are
described but most of these are from temperate regions. Amrine and Stasny (1994)
estimate that less than one-fifth of the temperate fauna and that perhaps only one-
twentieth of the tropical fauna has been described. All of these mites are tiny, most
less than 200 μm in length, all post-egg stages (larva, nymph, adult) have worm-like
bodies with only two pairs of legs (Fig. 8.2) and all are parasites of plants. Most
woody dicots, and many herbaceous dicots, monocots, conifers, and ferns, are
attacked by eriophyoids (Amrine and Stasny 1994; Lindquist et al. 1996) and the
vast majority of these mites are host-specific (Oldfield 1996). Individual plant spe-
cies, though, may harbour several species of Eriophyoidea. For example, Sugar
Maple (Acer saccharum) is attacked by eight species of eriophyoids (Patankar et al.
2012). Hosts are located by aerial migration (Zhao and Amrine 1997).
The mouthparts of eriophyoid mites are complex and have not been completely
homologised with those of other prostigmatans. The stylets (7–9 in number) are
encased in a U-shaped channel of the subcapitulum but do not form a feeding
tube (Lindquist 1996). Two methods of attachment to the leaf surface during feeding
are known. Most eriophyoids have the apices of the palp tarsi modified into a
suction pad that allows them to anchor their mouthparts to the leaf epidermis. The
muscles of the palps then contract and force the stylets into the epidermal cell of the
host. In these mites, the palps usually telescope during penetration (Nuzzaci and
Alberti 1996). In contrast, citrus russet mites anchor themselves to the leaf with anal
suckers, arch the body and force the stylets into epidermal cells; palps are reflexed
away from the plant surface. Stylets are short (20 μm or less), penetrate only about
2–7 μm into host tissue and therefore mites feed only on epidermal cells. No feeding
tube is formed by the stylets and they appear to be withdrawn before cell fluids are
sucked out (Westphall and Manson 1996; Oldfield 1996).
Gall, erinose, blister, witches’-broom and big bud mites cause galling and
deformation of leaves, petioles, stems, buds, flowers and fruits (Keifer et al. 1982).
286 8 Mites on Plants

Only plant roots appear to be safe from attack. Perhaps the most unusual of these
cancer-like growths are the felt-like pelts of hair-like processes called erinea (from
the woolly growth of hair-like cells that tend to fill these galls). Mites live among the
hairs and appear to feed on them (Westphall and Manson 1996). Predators can be
seen wandering over the surface of erinea and probing for mites but the erinose
mites appear to be well protected by the growth of hairs. Finding mites in erinea is
no easy task either (especially since erinea persist after the mites have gone), mak-
ing diagnosis difficult for a non-specialist. However, if a leaf with a fresh erineum
is picked and placed under a light, hordes of tiny erinose mites will come scram-
bling to the surface as the leaf dries.
Mites are easier to find in discrete galls (variously called bladder, finger, nail
and pocket galls) and other swollen tissues that can be cut open, often to disclose
hundreds of tiny worm-like mites. The exact mechanism causing the galling
response is not well understood but appears to result from the release of some
component that the mite introduces (or induces) during feeding. The three glands
(an unpaired median gland and a pair of lateral glands) that empty into the feeding
tube are a likely source of this component (Nuzzaci and Alberti 1996). The cell
initially punctured by a gall mite dies but surrounding cells begin to proliferate to
form the protective gall, while differentiation of nutritive cells also occurs
(Westphall and Manson 1996). Considering the economic importance of these
mites and their diversity and commonness, understanding the galling process
should be considered an important area of future research. Since about a dozen
eriophyid mites are known to vector plant viruses or virus-like diseases (Oldfield
and Proesler 1996), the hypothesis that microbial associates are involved in gall
production merits consideration.
The ability to gall plant tissue has evolved several times in acariform mites
(e.g. some Tenuipalpidae and Tarsonemidae cause gall-like growths) but it is
most strongly developed within the Eriophyoidea. However, more than half of all
known eriophyoid species do not live within deformed plant tissues (Amrine and
Stasny 1994). These mites are called rust mites, russet mites or leaf vagrants.
Some species are brightly coloured in orange or purple, some are covered with
waxy plumes, and a few spin small silken webs (Manson and Gerson 1996).
Rust mites wander the leaf surface puncturing cells and causing a silvery to rust-
coloured damage (Keifer et al. 1982). Like all eriophyoid mites that have been
studied, these mites are bisexual and arrhenotokous; the haploid males produce
stalked spermatophores (see Chap. 5). Current evidence suggests that eriophyoid
mites are closely related to tydeid mites (Lindquist et al. 1996) and presumably
plant feeding evolved from scavenging or fungus-feeding leaf vagrants that were
opportunistic leaf cell feeders (as some modern tydeid mites are thought to be).
This switch to obligate plant parasitism probably occurred in the Palaeozoic
since Triassic fossil Eriophyoidea are known (Schmidt et al. 2012). The living
eriophyoid mite with the least derived morphology, Pentasetacus araucariae,
forms galls on the needles of a Chilean species of Araucariaceae, a family that
originated in the Permian period more than 250 million years ago (Boczek and
Shevchenko 1996).
Plant Parasites 287

Fig. 8.3 Earth mites including the economically important Red-legged Earth Mite and Blue Oat
Mite are found on the ground, in mosses and on low-growing vegetation. They are pests of pasture
crops and seedlings. The cuticle is mostly red, easily seen on the legs, but obscured by the dark
body colour (degraded chlorophyll). This mite (Eriorhynchus walteri) lives in the rainforests of
Queensland, Australia (Photo HC Proctor)

Earth Mites: Penthaleidae and Its Kin

In contrast, plant parasitism appears to have arisen more recently within the
Eupodoidea and possibly more than once (Qin and Halliday 1997). As far as is
known, most eupodoids are predatory (e.g. Rhagidiidae) or feed on fungi (e.g. most
Eupodidae) and live in soil, decaying wood, and similar habitats. Penthalodidae are
mysterious moss mites covered in heavy and ornate armour and are thought to be
predators or possibly to feed on algae or lichens (Walter et al. 2009).
In contrast, the earth mites (Penthaleidae, Eriorhynchidae) clearly accumulate
chlorophyll in their bodies and so must be feeding on algae, bryophytes, or higher
plants (or all three). Earth mites are relatively large, long-legged and soft-bodied
mites that are usually found on low-growing herbs, mosses or in the surface litter of
the soil. Generally their integument is orange to red but this colouration is obscured
by the black, blue-black or dark green internal colouration of the body and they
appear to be black-bodied mites with red legs (Fig. 8.3). On the leaves and stems of
Australian rainforest trees, some species that appear to belong to the genus Eupodes
(Eupodidae) have a similar red cuticle (carotenoid pigments that protect from
ultraviolet light – see Chap. 7) and dark green bodies and probably feed on the
bryophytes and algae growing there.
288 8 Mites on Plants

One of the best-known penthaleid mites is a major agricultural pest in the


Southern Hemisphere: the Red-legged Earth Mite (a.k.a. Black Sand Mite)
Halotydeus destructor. Earth mites live up to both their common names: 90 % of a
population is on the soil at any one time and only 10 % on a pasture plant; the mites
are highly susceptible to moisture stress and do best on sandy soils (Ridsdill-Smith
2008). Annual broad-leaved plants and grasses are attacked, usually on the upper
surface of the leaf, and large populations can cause extensive damage to pastures in
the cool seasons in Mediterranean climates (Ridsdill-Smith 1997). When feeding on
the leaves of higher plants, earth mites brace their bodies by gripping the leaf sur-
face with their tarsal claws and then puncture and flatten the convex surface of leaf
epithelial cells during feeding. The needle-like movable digits in these mites are
9–20 μm in length and this appears to determine the depth to which they are able to
feed (Ridsdill-Smith et al. 1996).
Red-legged Earth Mites reproduce sexually and both sexes are diploid (Weeks
et al. 1995). Spermatophores are a simple silken web on which drops of sperm are
deposited by males, even when females are not present (Annells 1995). Eggs are
laid individually on plants during the cooler months but at the beginning of summer
adult females die with their bodies full of diapausing eggs. These cadaverous egg
sacs may be blown on the wind and disperse the mite over long distances. Immature
stages (larva and three nymphs) appear to be less likely to be found on plants and
the life cycle may be completed on mosses, fungi and algae growing on the soil
surface (Ridsdill-Smith 1997).
The Blue Oat Mite, Penthaleus major, is another important pasture pest. Unlike
the Red-legged Earth Mite, the Blue Oat Mite feeds upside-down on the underside
of leaves. Conveniently, this mite has its anal opening on its back (surrounded by an
orange spot in live specimens). Also convenient for lonely females, the Blue Oat
Mite has done away with males and reproduces by thelytokous parthenogenesis
(Weeks et al. 1995). The Blue Oat Mite is active during the autumn and winter and
passes the warmer months as aestivating eggs glued to plant stems (Tashiro 1987).
Although the Blue Oat Mite has a greenish tinge to the adult integument, it is not
blue and, perhaps, winter grain mite is a better common name.
The Penthaleidae as presently defined may be diphyletic and herbivory may have
arisen twice: once in an ancestor of Halotydeus and once in an ancestor of Penthaleus
and its relatives (Qin and Halliday 1997). However, several other related families of
poorly known mites, the Penthalodidae and Eriorhynchidae, complicate interpreting
these transitions. Penthalodid mites are often a bright yellow in colour, strongly
ornamented and well armoured. Two genera are recognised: Penthalodes with a
boreal distribution and Stereotydeus from austral regions (Qin 1994). Species of the
latter are common in mosses and lichens in the Southern Hemisphere, including
Antarctica (see Chap. 6). Eriorhynchus is the only known genus in the recently
described Eriorhynchidae (Qin and Halliday 1997). Five species are currently
known and all are large, hairy, rapidly moving mites with very long legs. Leg span
from the tips of the antenniform front legs to the claws of the hind legs are typically
6–7 mm in adult females. Eriorhynchus walteri can be found in mosses, leaf litter
and in canopy knockdown samples in rainforests in Queensland.
Plant Parasites 289

Fig. 8.4 Spider mites have a stylophore (s) composed of the fused cheliceral bases and the mov-
able digits (md) form a long sucking stylet. The fixed digit (fd) is reduced to a short shealth. The
stylophore can be drawn into the body through a flexible collar-like region that contains the perit-
reme (p). Silk is extruded from the spinneret (sp) on the palp tarsus. (a) Dorsal view of gnatho-
soma; (b) Lateral view (Neonidulus tereotus); (c) Spinnerets (Neonidulus tereotus) (Images by DE
Walter)

Spider Mites and Their Kin

The most familiar and feared of acarine pests of plants are the “spider mites”, so
named for the copious webbing produced by the most damaging species. Although
they have given their common name to their superfamily, the Tetranychoidea, the
ability to produce dense webbing is restricted to species in only a few genera. This
cluster of five families, however, is united by having an eversible stylophore
(Fig. 8.4) formed from the fused cheliceral bases and a long, recurved, hollow stylet
formed from the movable digits (Hislop and Jeppson 1976). Spider mites in the
broad sense are well adapted to feed on plant cells.
All 2,000+ known species of Tetranychoidea are parasites of vascular plants
(Krantz and Walter 2009). Because of the damage caused to economically impor-
tant plants, the Tetranychidae (~1,200 described species) are the best studied.
Although only one species has been demonstrated to vector a plant virus (Robertson
and Carroll 1988), spider mite feeding is often extremely damaging and can result
in death, especially in herbaceous hosts. Spider mites stab individual plant cells
with their mouthparts and suck out the contents. This can lead to necrosis of leaf
tissue and leaf drop. Generation times are usually short (10–14 days) and populations
rapidly reach damaging levels. Spider mites migrate on the wind and readily colonise
290 8 Mites on Plants

new plants. Many respond to chemical miticides by migrating, thereby expanding


the area under attack. Additionally, spider mites rapidly evolve resistance to chemicals,
perhaps because favourable mutations become fixed rapidly in haploid males
(Crozier 1985). Before World War II, spider mites were only rarely pests; however,
since the use of synthetic insecticides and fungicides became common, spider mites
have become major causes of economic loss.
Spider mites get their common name from the silken webbing produced by some
species in the subfamily Tetranychinae; however, many of these mites use silk only
to protect their eggs and those in the subfamily Bryobiinae do not produce silk at all
(Guitierrez and Helle 1985; Gerson 1985). Master-class weavers occur only within
the tribe Tetranychini and many of these mites have developed a subsocial type of
colony. For example, species of Neonidulus, Stigmaeopsis, Schizotetranychus and
some species of Oligonychus (e.g. O. perseae) produce densely woven sheets of
silken cloth under which families of mites feed in relative security. These web nests
(Saito 1985) are as tightly woven as any spider egg case and provide excellent pro-
tection against desiccation and predators. As the nests are constructed, small open-
ings are left and males often guard these entrances and ward off predators and
competitors (Aponte and McMurtry 1997). Immature predators, especially larvae,
that attempt to enter nests of a species of Stigmaeopsis (Schizotetranychus in older
literature) on bamboo are attacked and often killed by thrusts of the chelicerae. Male
and female spider mites appear to cooperate in defence, with females finding and
pursuing the predator and males catching and killing them (Saito 1986a, b, 2010;
Saito et al. 2011).
As in the Hymenoptera, spider mites are arrhenotokous and species like
S. celarius live in colonies, often founded by a single female. The ‘nests’ develop
into complex webs produced by many overlapping generations of mites. Behaviours
in these nests are complicated and involve coordinated sanitation behaviour (feces
are deposited in specific sites). Stigmaeopsis longus is able to use its sticky silk to
remove detritus from the leaf surface – cleaning up the nest and apparently increas-
ing egg survival (Kanazawa et al. 2011). These mites appear to have reached a level
of subsocial organisation unknown in any other mites (Saito 2010). Tetranychus
evansi responds to its predator Phytoseiulus longipes not by producing denser
webbing, as might be expected, but by suspending more eggs in the webbing off the
leaf surface where they are less likely to be encountered by foraging predators
(Lemos et al. 2010).
In general, male spider mites are smaller than the females, often about half the
body mass. Males guard quiescent female nymphs, fend off other males, and at least
in Tetranychus urticae, the first male to copulate with a newly emerged female fer-
tilizes all of the eggs that will develop into daughters (Saito 2010). In contrast the
Australian spider mite Neonidulus tereotus has ‘giant’ males with long, ‘muscular’
front legs (Fig. 8.5). These mites produce a densely woven, tent-like web along the
midvein on the underside of the leaves of the tree Waterhousea floribunda
(Myrtaceae). Males guard quiescent female nymphs and have been observed to pick
up nymphs and move them when disturbed (Beard and Walter 2010).
Plant Parasites 291

Fig. 8.5 Female (left) and male (right) of Neonidulus tereotus. Unlike most spider mites where the
male is smaller than the female, this Australian mite has ‘giant’ males that guard and can carry
quiescent female nymphs (SEMs by DE Walter)

Species of Tetranychus produce the largest and most complex spider mite webs
and they often form dense entanglements of many layers. Some species, such as the
Two-spotted Spider Mite, Tetranychus urticae, are polyphagous pests that can dev-
astate many crops. The Two-spotted Mite holds the record, at over 900 species of
known hosts (Navajas 1998). However, of the approximately 1,200 described spe-
cies in 70 genera (Bolland et al. 1998), only a few dozen spider mites are major
agricultural pests that regularly defoliate and kill plants (Jeppson et al. 1975; Helle
and Sabelis 1985a, b). Unlike many agriculturally important species, most tetrany-
chids appear to have restricted host ranges (Jeppson et al. 1975; Hill and Stone
1985). Most host specialists do not attack economically important crops and are
poorly studied, suggesting that the true diversity of the Tetranychidae is undoubt-
edly considerably higher than 1,200 species (see Chap. 10).
The probable sister group of the Tetranychidae is a clade composed of the
remaining four families of Tetranychoidea (Lindquist 1985). The Tenuipalpidae or
false spider mites are best known, again because of a few economically important
polyphages (Jeppson et al. 1975; Meyer 1979). Most tenuipalpid species, however,
seem to be host specific. More than 600 species have been described in 22 genera
(Meyer 1979; Evans et al. 1993) and, given the diversity of undescribed forms in
Australia, this must be a minor component of the true diversity of the group.
292 8 Mites on Plants

Fig. 8.6 Peacock mites (Tuckerellidae, Tuckerella sp.) get their name from the long setae on the
rear – the setae are flicked back-and-forth over the body to ward off predators. Like their cousins
the spider mites, peacock mites are plant parasites that share the modified mouthparts: stylophore
and long, whip-like stylets (md) (Illustration by DE Walter)

The remaining three families of tetranychoid mites, Tuckerellidae, Linotetranidae


and Allochaetophoridae, are rarely important agriculturally and account for less
than 30 described species in five genera (Meyer and Ueckermann 1997). Linotetranid
mites (Linotetranus, Afrolinotus, and Anoplopalpus) are associated with soil col-
lected among grasses and are unique among the tetranychoids in completely lacking
eyes. Possibly they feed on grass roots or within the crown of grasses. Until recently
allochaetophorid mites, or rather the one known nymph of Allochaetophora, was
one of the great mysteries of acarology. The family was proposed in 1959 to accom-
modate one aberrant tetranychoid nymph from grass in California and for nearly 40
years that was the only known representative of the family. A full developmental
sequence of a second species from Africa (also associated with grasses) has now
been described (Meyer and Ueckermann 1997).
Tuckerellid mites (Tuckerella) are perhaps the least derived of the Tetranychoidea
and they are certainly the most ornate of the spider mites (Fig. 8.6). Sometimes
called peacock mites because of the peacock tail of plumose posterior setae that
trails out behind them their dorsal shields are covered with large, leaf-like setae.
The leaf-like setae are probably protective armour and the peacock tail can be
quickly flicked over the body, almost quicker than the eye can follow, to dislodge or
discourage predators. Described species of Tuckerella fall into two species groups
that segregate ecologically. One group is found in association with grasses (perhaps
feeding on roots) and the other on stems and fruits of woody plants. Some species
of the latter group are pests of citrus (Ochoa et al. 1994) and associated with a
necrosis disease. Peacock mites retain the tritonymphal instar (Quiros-Gonzalez
and Baker 1984), at least in the adult females, although at least two species have
males that develop from deutonymphs skipping the tritonymph (Beard and Ochoa
Plant Parasites 293

2010). As far as is known, the tritonymphs are suppressed in all other Tetranychoidea
and in the closely related Cheyletoidea and most Raphignathoidea except the Xeno-
caligonellidae (Lindquist 1998; Fan 2000; Krantz and Walter 2009). Paedogenesis
has been reported in tritonymphs of Tuckerella knorri (Ochoa 1989).
Unlike other groups of plant-parasitic prostigmatans, the Tetranychoidea appear
to have evolved phytophagy from a predaceous ancestor shared with the Cheyletoidea
or Raphignathoidea. Basal cheyletid mites are predators and derived taxa account
for one of the most impressive radiations of parasites of insects and vertebrates
(Krantz 1978; Evans 1992). No cheyletid mites are known to feed on plants but a
number of predatory Cheyletidae are arboreal and perhaps some ancestor of these
mites first began feeding on plant cells. Most raphignathoid mites are also thought
to be predators, including the Stigmaeidae, a well-known group of biocontrol agents
in agricultural systems. However, at least one genus of stigmaeid mites, Eustigmaeus
(known as Ledermuelleria in the older literature), is plant parasitic, although on
mosses rather than vascular plants (Gerson 1972).
With the exception of the Stigmaeidae, the biologies of species in most families
of Raphignathoidea are very poorly known (Fan and Zhang 2005). The Homocaligidae
is associated with subaquatic to aquatic plants and probably predatory, but an
herbivorous relationship with these plants should at least be considered a possibility.
Also, members of perhaps the most obscure family of the Raphignathoidea, the
Cryptognathidae, have a number of derived characters (e.g. the ability to withdraw
the gnathosoma through a collar-like extension into the idiosoma) that may indicate
a close relationship to the Tetranychoidea. The feeding biology of cryptognathid
mites is not known, but they are often found in moss, on bark, and in other periodically
dry habitats.

Duckweeed and Water Hyacinth Mites

Outside of the Prostigmata, reports of mites feeding on the leaves of living plants
are rare and mostly caused by mistaken assumptions (all mites on leaves are bad)
or by facultative leaf feeding by predators (Porres et al. 1975; Adar et al. 2012) or
by the depredation of omnivores on seedlings (Walter et al. 1986). However, some
species of oribatid mites seem to have made the transition from feeding on fungi
and dead leaves to attacking living plants, especially in aquatic systems. For example,
adult Hydrozetes (Hydrozetidae) are commonly associated with duckweed
(Lemna spp.) and immature stages burrow within the small, floating thallus of the
plant (Fig. 8.7).
Immature stages of another oribatid mite, Orthogalumna terebrantis, burrow in
aquatic plants in the Pontederiaceae, including the noxious weed, water hyacinth,
Eichhornia crassipes. The eggs are laid on pseudolaminae (= leaf-like structures)
and the immature mites excavate tunnels in ‘leaves’. The mite tunnels become
infection portals for a leaf spot fungus, Acremonium zonatum, that is vectored by the
mite. Mite activity stimulates oviposition by the weevil Neochitina eichhorniae,
294 8 Mites on Plants

Fig. 8.7 Right: Eustigmaeus frigida (Prostigmata, Stigmaeidae) a mite that feeds on aquatic and
subaquatic mosses. Left: adult Hydrozetes sp. (~0.5 mm long) an aquatic oribatid mite. Both sus-
pended over duckweed (Lemna sp. ~ 3 mm long) on the surface of a slough. Immature stages of
Hydrozetes burrow in duckweeds and may have caused the damage visible in the picture (Photo by
HC Proctor, SEMs by DE Walter)

causing better control in areas where the mite is present (Cromroy 1983). Even at
relatively low densities, feeding by O. terebrantis reduces water hyacinth photosyn-
thetic rate and the light reaction performance (Marlin et al. 2013). Some Prostigmata
also attack aquatic plants, for example, Eustigmaeus frigida (Fig. 8.7 right) can be
found in numbers feeding on the mosses that line the shores of sloughs and lakes in
Alberta, Canada.

Fruit and Fig Mites

Rotting fruit that has dropped beneath a plant can be colonised by soil-inhabiting
mites and those phoretic on the flies and beetles attracted to the decaying fruit. Fruit
that are damaged but retained on the plant may acquire a similar acarofauna and
fruit with mildews or other fungal growths may support large populations of fungi-
vores and scavengers that have moved out from the leaves and twigs. Intact fruit
may be attacked by species of Tenuipalpidae, Tuckerellidae, Tarsonemidae or
Eriophyidae (e.g. citrus rust mites), each of which may cause characteristic damage
(Jeppson et al. 1975).
One specialised fruit, the fig, has its own specialised group of fig-feeding mites
(Fig. 8.13b). Figs are complex inflorescences composed of many small unisexual
Plant Parasites 295

flowers that are surrounded by a fleshy covering. Within a green fig is a central
cavity carpeted by the stigmas of female flowers, each connected to an ovule by a
style of various lengths. Each of the female flowers can develop into a seed – or a
fig wasp (Hymenoptera: Chalcidoidea: Agaonidae). Entrance to the fig is through
the ‘eye’ or ostiole, an opening that is usually tightly packed with scales that rip the
wings from female fig wasp as she worms her way through the ostiole. Once in the
central cavity, the female wasp spreads pollen across the carpet of stigmas (assuring
development of the fig seeds) and probes styles with her ovipositor, laying eggs in
those ovules she can reach (Janzen 1979). As figs ripen, wingless male wasps
emerge from seeds first and mate with female wasps, often still within their natal
seeds. Female wasps emerge, actively gather pollen that is carried in specialised
folds, pockets or groups of hairs, or in the gut and then exit the fig to search for
another developing fig of the same species. Over 900 species of figs are known;
each is believed to have a host-specific pollinator.
Small tydeid and tarsonemid mites that scavenge or feed as fungivores can make
their way into a fig through the ostiole; however, one tribe of tarsonemid mites
appears to have a much more specific association with figs. Three genera of
Tarsonemellini are currently known with species that feed on fig ovules. The mites
are phoretic on pollinating fig wasps (and possibly on non-pollinators) and a high
diversity of host-specific mites appears to be present (Lindquist 1998; Ho 1994;
Compton 1993; Jauharlina et al. 2012). Although this association has only recently
been discovered, it appears to represent yet another independent evolution of phy-
tophagy among the Tarsonemidae – one potentially represented by many hundreds
of species in tropical forests.

Venereal Diseases of Plants

Although we find many flowers attractive to our sight and nose, only the most
unnatural of commercial varieties produce their flowers solely for our benefit. As
Linnaeus was fond of pointing out (to the discomfort of many of his colleagues),
flowers are the genitalia of plants. As a rude analogy this is true, but a bit more
complicated.
In the vast radiation of plants that we call angiosperms flowers are specialized
reproductive structures that contain the microscopic haploid gametophytes that
eventually produce female gametes inside the ovules and/or the pollen grains that
produce the male gametes. In order for an ovule to develop into a seed, it must be
restored to diploidy and usually this is the result of fertilization by a male gamete
contained in a pollen grain (pollination). Many plants, e.g. most conifers and
grasses, produce copious pollen that depends on wind and gravity to reach an ovule.
Windblown pollen that does not reach an ovule is literally a windfall for many
mites, but feeding on spent pollen is best considered scavenging. In most angio-
sperms and cycads, though, copulation and fertilisation usually occurs as a result of
the transfer of pollen from one flower to another by an animal (the pollinator).
296 8 Mites on Plants

Pollinators are usually rewarded for this servicing with food: pollen, plant secretions
(e.g. nectars, oils or resins) or sometimes plant tissues. Therefore, pollinators also
eat plants and are herbivores. Bumble bees, for example, feed only on pollen and the
secretions of flowers. Like all such interactions that occur over evolutionary time,
pollination by animals is not a simple process, but because both the pollinator and
plant may benefit, pollination is usually considered a mutualism. If an animal feeds
on pollen and/or nectar and does not contribute towards pollination, though, it is
clearly a parasite. If a pollinator transmits an organism that infects plant reproduc-
tive structures, then this is analogous to venereal transmission of diseases in ani-
mals. At least seven, and possibly as many as ten, unrelated groups of Mesostigmata
have evolved a specialised life style that has been appropriately described as a vene-
real disease of plants; they ride pollinators from flower to flower, infest flowers and
consume significant amounts of nectar and pollen (Colwell 1995; Paciorek et al.
1995; Velazquez and Ornelas 2010).
Hummingbird flower mites in the Neotropics are the best studied of these
associations. All species of Rhinoseius and Tropicoseius in the Melicharidae and
lineages of Proctolaelaps (Melicharidae) and Lasioseius (Blattisociidae) live
within the flowers of a variety of shrubs and trees and as adults use hummingbirds
for phoretic transport between plants (Dobkin 1985; Naeem et al. 1985; Colwell
1986; Heyneman et al. 1991; OConnor et al. 1991, 1997; Ohmer et al. 1991;
Colwell and Naeem 1994; Paciorek et al. 1995; Naskrecki and Colwell 1998).
A quick dash down the hummingbird’s bill is necessary for prompt disembarka-
tion and the decision to exit is probably mediated by the olfactory information in
the stream of air generated by the birds as they breathe (100–300 breaths per
minute). The Rhinoseius–Tropicoseius lineage appears to have arisen in the mon-
tane neotropics and to have diverged in association with different families of
plant hosts (Naskrecki and Colwell 1998). At least in flowers that last several
days, hummingbird flower mites can significantly deplete nectar levels (Lara and
Ornelas 2001, 2002).
An Australian system analogous to the neotropical hummingbird–flower mites
has evolved between species of the genus Hattena (Ameroseiidae) and birds that
visit flowers. The mites feed on pollen and nectar in the flowers of mangroves,
mistletoes and forest trees and shrubs, and use honeyeaters, spinebills and probably
lorikeets, to move from plant to plant (Domrow 1979; Seeman 1996; Halliday
1996). For example, Hattena floricola lives in the flowers of Australian shrubs in
the genus Correa that are pollinated by the Eastern Spinebill (Acanthorhynchus
tenuirostris). As has been shown with hummingbird flower mites, these spinebill
mites are able to significantly reduce the amount of nectar present (Scoble and
Clarke 2006). Similarly, Hattena panopla lives in the flowers of the mangrove
Bruguiera gymnorhiza in coastal Queensland. The mite can walk between flowers
that are nearby, but relies on honeyeaters to move longer distances. Unlike
Neotropical hummingbird mites, all post-larval stages are phoretic on birds (Seeman
1996). This association of Hattena with mangrove flowers spans the Pacific Ocean:
Hattena rhizophorae lives in Red Mangrove (Rhizophora mangle) flowers on
the coast of Ecuador (Faraji and Cornejo 2006). Red Mangrove has long been
Hunting on Leaves 297

considered wind pollinated, but recent research indicates hover flies (Syrphidae) are
important pollinators (Sánchez-Núñez and Mancera-Pineda 2012) and insects may
be the phoretic carriers of the Red Mangrove mite.
As well as utilizing hummingbirds and their flowers in the Neotropics, some
African species of Lasioseius are associated with sunbirds (Nectariniidae) and the
flowers they visit (Kaluz et al. 2011). Species of Proctolaelaps also are found in
flowers in South Africa (Ryke 1964) and Western Australia (Domrow 1979), and
utilize honey birds and honey opossums, respectively, for transport among flowers.
Flower-visiting bats are also carriers of flower mites (Tschapka and Cunningham
2004; Lindquist and Moraza 2008).
Flower mites also form phoretic associations with insect pollinators. For exam-
ple, flower-inhabiting species of Proctolaelaps (Melicharidae) are commonly found
on butterflies and moths (Treat 1975; Janzen 1983; Boggs and Gilbert 1987) and
one species that inhabits bromeliad flowers in Brazil is phoretic on the tropical
bumble bee Bombus morio (Guerra et al. 2010). Mites in the genus Xanthippe
(Melicharidae) apparently feed on pollen or nectar in inflorescences of a palm and
may be phoretic on nitidulid beetles that pollinate the palm (Naskrecki and Colwell
1995). Species of Spadiseius (Melicharidae) utilize flower-feeding bats, but also
flower-visiting scarab beetles and bees (Lindquist and Moraza 2008). In Africa,
Australasia and Asia mites in the genera Neocypholaelaps and Afrocypholaelaps
(Ameroseiidae) live in flowers, feed on pollen and nectar, and primarily use bees
and butterflies for phoretic transport (Evans 1963; Baker and Delfinado-Baker 1985;
Eickwort 1994; Seeman and Walter 1995; Halliday 1996).
Flower-mite systems are primarily tropical to subtropical and few species are
known from the temperate zones. In temperate areas, phytoseiid mites (see below)
may sometimes be found in flowers feeding on pollen. For example, Iphiseius
degenerans is used for control of Western Flower Thrips in greenhouse peppers and
adult females are often found in flowers feeding on pollen, nectar or thrips (Faraji
et al. 2002a, b). Other phytoseiids can also be found in flowers, but this seems to be
a relatively rare behaviour. Tarsonemid mites and other prostigmatans also can be
found in flowers, but there is no indication of an obligate relationship. Neither
are any flower-specific oribatid mites are known, although species of Scheloribatidae,
Haplozetidae and Mochlozetidae have been collected from flowers (Aoki 1992;
Seeman and Walter 1995).

Hunting on Leaves

When we walk along a trail under the canopy of a forest on a sunny day, we experi-
ence shade and shelter, but not a lot of arthropods: most of the action is under our
feet (see Chap. 6) or over our heads where sunlight is captured and turned into living
biomass. In the forest canopy the surfaces of leaves (the phylloplane) is regularly
exposed to intense sunlight, driving winds and pouring rain. This is a hostile envi-
ronment to an animal as small as a mite and this must be especially true for
298 8 Mites on Plants

predators that cannot obtain water and nutrition directly from the plant. This probably
explains why relatively few lineages of predatory mites have colonised the phylloplane.
Some of those lineages, however, have undergone spectacular radiations.

Predatory Prostigmata

Few plants in any garden or forest are without free-living prostigmatans and this is
true for one family in particular, the Tydeidae. In Greek mythology, Tydeus was a
murderer on the run who became one of the heroic Seven Against Thebes. Among
arboreal Acari, species of Tydeus and other tydeid mites are one of the spectacular
success stories, although their biology is still mostly a mystery. Algae, fungi and pollen
are known sources of food for tydeid mites but at least some species are facultative
predators (André 1986; Walter and O’Dowd 1995a, b) and others can be important
biocontrol agents of plant-parasitic mites (Hessein and Perring 1976). Similarly,
most foliar Tarsonemidae are fungivores or algivores but a few species are predators
of eriophyoid mites or the eggs of various plant mites (Lindquist 1986).
Tydeid and tarsonemid mites are small, active mites that come in a variety of
pale, beige or greenish shades; however, most predatory prostigmatans on the phyl-
loplane are yellowish orange to bright red in colour, possibly because the carotenoid
pigments that cause these hues protect from ultraviolet light. Commonly encoun-
tered families include the Adamystidae, Anystidae, Bdellidae, Cheyletidae,
Cunaxidae, Erythraeidae and Stigmaeidae. As in soil-litter systems (see Chap. 6),
some are ambush predators and many are cruise predators (Fig. 8.8).
Cheyletid mites are ambush predators and can often be found at the juncture of
the petiole and the leaf blade, lying in wait for small arthropods moving along the
stem-petiole highways that connect one leaf to another. Species of Rubroscirus
(Cunaxidae, Cunaxinae) are also ambush predators but they tend to take up their
positions well out on the leaf blade. As with related genera in soils, Rubroscirus
have snout-like mouthparts and carry their palps reflexed over their bodies when
hunting. The palps are thrown forward to snap-trap prey against the snout. Even
relatively large, active prey such as thrips can be captured using these tactics.
Generally, ambush predators require active prey but Australian species of
Rubroscirus also are able to feed on drops of honeydew and it is possible that they
may feed on inactive prey such as scale insects or eggs as well.
Most predatory prostigmatans actively seek out prey on the phylloplane, although
their pace varies from a constant and deliberate searching of every nook and cranny
to the dizzy, spiralling dash across the leaf characteristic of anystid mites. These
‘whirligig’ or ‘footballer’ mites are general predators that overrun mites and small
insects, stab them with their mouthparts and seemingly paralyse them with a toxin
(Sorensen et al. 1976). Whirligig mites occasionally get carried away with them-
selves and scramble onto people, where they are capable of administering a nasty
bite. Like all prostigmatans, whirligig mites feed only on the fluids of their prey but,
unlike ambush predators, they take a variety of active and sessile prey.
Hunting on Leaves 299

Fig. 8.8 Examples of foliar predatory mites: (a) Cheyletidae – ambush predators that often lie in
wait at the base of a leaf petiole to pounce on passing prey; (b) Stigmaeidae – generally wander the
leaf surface looking for sessile prey such as eggs and moulting individuals; (c) Phytoseiidae – rapidly
searching predators that quickly cover the leaf surface; (d) Cunaxidae (Cunaxinae) – ambush
predators that take up residence on the leaf blade (Illustration by DE Walter)

Many cruise predators are able to search out and feed on the eggs of large insects.
For example, in the Brisbane region, a large (4 mm body, 1 cm leg span) red velvet
mite in the genus Charletonia (Erythraeidae) feeds on the eggs of the Richmond
Birdwing (Ornithoptera or Troides richmondia), one of Australia’s largest butter-
flies. The Birdwing has a restricted distribution in southeast Queensland and north-
east New South Wales and its survival is threatened by habitat fragmentation. The
effects of this predation on birdwing abundance have not been investigated, but
opportunistic predation by another erythraeid mite (Balaustium sp.) on the eggs of
a pierid butterfly can be a major cause of pre-adult mortality (Hayes 1985).
Members of the Stigmaeidae are among the most common predatory prostig-
matans in forests and orchards. Eocene amber fossils of a species of Mediolata
(predators of spider mites and scale insects) are known (Kuznetsov et al. 2010), so
living on trees and feeding on their pests is probably an ancient association. Eggs
seem to be a preferred food for many stigmaeids, including those of plant-parasitic
mites and competing predators, and even cannibalise their own eggs (Clements and
Harmsen 1992). For example, Zetzellia mali wander leaves and feed on the eggs of
a variety of plant-parasitic and predatory mites that they encounter (Santos 1991).
On apple leaves, Z. mali spend about three-quarters of their foraging time on and
around the midrib, resulting in preferential consumption of eggs of species that
oviposit in this area (MacRae and Croft 1996). As with most plant-inhabiting stig-
maeids, Z. mali oviposit on leaf hairs or strands of silk, presumably reducing the
chance of cannibalism. Agistemus exsertus, for example, will feed on its own eggs
300 8 Mites on Plants

when those of spider mites are not available (Rasmy and Saber 2012). In contrast,
Agistemus olivi is a specific predator of eriophyoid mites and can be useful in
biological control (Momen 2012).

Foliar Mesostigmata

In contrast to the diversity of predatory prostigmatans, only a single family of meso-


stigmatans, the Phytoseiidae, is characteristically found on leaves in temperate for-
ests. This lineage, however, has undergone a spectacular radiation on vegetation and
accounts for about 20 % of the known diversity of Mesostigmata. Most of the 2,300
species have been described since World War II as a direct result of their importance
as biological control agents (Chant and McMurtry 1994; Kostiainen and Hoy 1996).
They are small (200–600 μm in length) and suppress many of the body and leg setae
found in their soil-inhabiting relatives. Temperate phytoseiid faunas are relatively
well known but tropical faunas are extremely diverse and have scarcely been touched
(Walter and Beard 1997; Walter and Proctor 1998). Both the local and regional spe-
cies diversity of phytoseiid mites increases from temperate to tropical regions and
additional lineages of Mesostigmata (e.g. Blattisociidae, Ologamasidae, Uropodidae)
begin appearing in the phylloplane fauna (Walter et al. 1994; Walter and Proctor
1998). For example, in tropical Australia species in the Lasioseius porulosus group
and the Asca foliata group (both Ascidae) are often more abundant on leaves than
phytoseiid mites (Walter et al. 1993, 1994; Walter and Lindquist 1997).
Producing a balance between predatory and plant-parasitic mites in agricultural
systems is one of the goals of Integrated Pest Management (IPM). However, preda-
tory mites actively search for prey and are much more likely to encounter a pesticide
droplet than are more sedentary plant parasites. For this and other reasons, pesticide
use often disrupts populations of predatory mites. Before World War II, phytoseiid
mites were a curiosity and few species had been described. In the post-World War II
DDT era, however, accounts of damaging outbreaks of spider mites after the appli-
cation of pesticides such as DDT and parathion to control pests became all too com-
mon. This is when the hunt for biological control agents of spider mites began and
in unsprayed fields and orchards phytoseiids was the primary groups of predators
encountered.

Development and Reproduction of Phytoseiid Mites

The total life cycle of phytoseiid mites, from egg to egg, usually takes 8–10 days at
22–24 °C (Helle and Sabelis 1985b; Sabelis and Janssen 1994) but varies with sub-
family (Amblyseiinae have faster average rates than Phytoseiinae–Typhlodrominae)
and with genera within subfamilies (Zhang 1995). Species in some amblyseiine
genera, e.g. Phytoseiulus, have generation times as short as 5 days at summer
Hunting on Leaves 301

temperatures and, given enough food, females can lay 3–5 eggs a day for several
weeks. The host plant on which the prey feeds can influence development and repro-
duction. For example, Phytoseiulus persimilis reared on Two-spotted Spider Mites
grown on carnation had less than half the fecundity (33 eggs per female) of those
reared on two-spotted mites grown on soybean (69 eggs per female) (Popov and
Khudyakova 1989).
Except for a few known thelytokous species (Hoy 1985), phytoseiids must mate in
order to lay eggs. Female deutonymphs near maturity produce a pheromone that
produces an arrestment response in males (Rock et al. 1976; Hoy and Smilanick
1979). A single mating apparently is sufficient to fertilise a female’s entire comple-
ment of eggs in some species, whereas in others, females apparently run out of sperm
and oviposition stops until they mate again. Males develop from fertilised eggs, but
have haploid chromosome contents as adults. The paternal set of chromosomes is lost
or heterochromatised in the early development of the eggs destined to become males.
This has been termed both parahaploidy and pseudoarrhenotoky (see Chap. 5).
All species that have been studied have several generations a year when suffi-
cient prey is available a year. In warm, humid tropical or subtropical climates all
stages can be found essentially anytime of year. In dry tropical climates, reproduc-
tion appears to be more seasonal. In temperate climates, the adult female is the
overwintering stage and a reproductive diapause has been demonstrated for a num-
ber of species. Photoperiod is the prime factor inducing diapause but it often inter-
acts with temperature, relative humidity and nutrition. For example, in some species
carotenoids are essential for photoperiodic induction of diapause and can be sup-
plied in part by the pollen of some plants (Van et al. 1981; Overmeer et al. 1989).
For a comprehensive review of diapause in the Phytoseiidae, see Veermann (1992).
Diapause is a complicating factor when phytoseiids are used for pest control in
greenhouses and non-diapausing strains are much sought after and selected for
(Hoy 2009).

Feeding Biology of Phytoseiid Mites

Much of what we know about the feeding biology of phytoseiid mites is based on
the few species that are useful in biological control (McMurtry and Croft 1997).
Preeminent among this small group are the predators of spider mites in the genus
Tetranychus that produce dense webbing (CW-u type of Saito 1983). One of the
most useful of all of these biological control agents is the Chilean Predatory Mite
Phytoseiulus persimilis (Fig. 8.9a, b), perhaps the most intensively studied of all
acarine biological control agents. A bibliography of the Phytoseiidae (Kostiainen
and Hoy 1996) lists nearly a thousand citations indexed to P. persimilis (21 % of the
4,634 items). Like its spider mite prey, the Chilean Predatory Mite forages for new
resources both by foot and on the wind (Charles and White 1988), and the predator
responds to chemical signals both from its prey and the prey’s host plant (Bruin
et al. 1995). Although host plant signals may most effective at close range or in
302 8 Mites on Plants

Fig. 8.9 Representative Phytoseiidae: (a-b) Chilean Predatory Mite Phytoseiulus persimilis: long
legs, long dorsal setae and chelicerae are used to penetrate dense spider mite webbing. (c)
Phytoseius oreillyi: an inhabitant of hairy leaf surfaces. (d) Amblyseius sp.: a generalist predator/
pollenivore. (e-f) Neoseiulus sp. the chelate-dentate chelicerae are used to both bite through the
cuticle of prey and pick-up pollen grains (clasped on the inner face shown). Structures such as the
tectum (e arrow) and deutosternal denticles that control fluid movement during feeding are reduced
in these arboreal predators (SEMs by DE Walter)

enclosed areas like greenhouses, prey cues can be followed in open environments
(Yano and Osakabe 2009).
Although Phytoseiulus mites do not hunt together, numbers of them often pene-
trate the same webs and they are able to recognise familiar conspecifics and this
increases their hunting efficiency (Strodl and Schausberger 2012). Chilean Predatory
Mites have elongate dorsal shield setae that help them to penetrate the dense silken
webs of Tetranychus (Sabelis and Bakker 1992) and use their chelicerae and palps
to cut prey webbing (Shimoda et al. 2009). Their long legs undoubtedly help them
Hunting on Leaves 303

step over sticky strands. As a result, the predators can take up residence in the midst
of their prey. Eggs are laid on the spider mite silk and the nymphs and adults hunt
through the web. When other predators are present, females appear to be able to
alter their oviposition behaviour to reduce the risk of intraguild predation on their
young (Walzer and Schausberger 2010).
The Chilean Predatory Mite and other species of Phytoseiulus appear to require
spider mites that develop in dense colonies with heavy webbing, especially species
of Tetranychus. This degree of trophic specialisation is unusual in the Phytoseiidae,
although species of Phytoscutus are associated with foliar acarid mites in the genus
Neotropacarus. A few other phytoseiid mites, e.g. Galendromus occidentalis,
Neoseiulus fallacis and Typhlodromus pyri, seem to prefer spider mites but they also
take other prey and feed to a limited extent on plant products such as nectar and
pollen (digested externally). Most phytoseiid mites, however, are polyphagous. For
example, Amblyseius swirskii is used as a biocontrol agent against a host of green-
house pests including Chilli Thrips, Western Flower Thrips, Tobacco Whitefly,
Greenhouse Whitefly, Broad Mite and Tomato Russet Mite (Meng et al. 2012).
Unlike some phytoseiids (see below), A. swirskii does not feed on leaf cells (Adar
et al. 2012).
Polyphagous phytoseiids often do best when they have a high proportion of pol-
len in their diets (McMurtry and Croft 1997). Pollen is handled quite differently
from prey, where the chelicerae slice through the cuticle, the salivary stylets (lying
atop the closely inserted corniculi at the tip of the hypostome) are inserted into the
wound, and digestive enzymes are secreted that digest host tissues. Instead, the
gnathosoma is flexed towards the leaf surface, individual pollen grains are snatched
so that they are held on the inner side on one chelicera, brought up to the levelled
subcapitulum, then manipulated by both chelicerae to rupture the exine. A slurry of
digestive enzymes and pollen material is then sucked in by the pharynx and the
pollen grain collapses (Flechtmann and McMurtry 1992).
The diet of some Euseius species can be very complex and include prey, pollen,
honeydew and plant exudates (Bruce-Oliver et al. 1996). Some phytoseiid mites
also feed on leaf cells. Adar et al. (2012) have studied the feeding biology of Euseius
scutalis on leaves and artificial membranes. When engaged in leaf-feeding the teeth
at the tip of the moveable digit allow the chelicerae to grip the surface as the move-
able digit penetrates in a ‘crimping’ motion; the distal tip of the subcapitulum
(where the flattened corniculi are closely inserted) is then inserted into the wound
and fluid is drawn up by the pharynx. This is very similar to how phytoseiid mites
feed on spider mites (Flechtmann and McMurtry 1992). Unlike most phytoseiids,
species of Euseius are common on smooth, hairless leaves where moisture stress
may be significant and plant feeding an important component of survival. The rela-
tively short, broad chelicerae of E. scutalis are characteristic of the genus and simi-
lar chelicerae are found in other genera (e.g. Iphieus, Kampimodromus,
Typhlodromus) and may be a morphological clue to the phytoseiids that leaf-feed
(Adar et al. 2012). However, some of these species (e.g. Kampimodromus aberrans,
Typhlodromus pyri) prefer hairy leaves (Croft et al. 2004), so leaf feeding is not
always associated with smooth leaves.
304 8 Mites on Plants

The Chilean Predatory Mite is an archetypal phytoseiid spider mite specialist


(Type I of McMurtry and Croft 1997) and exhibits characteristics such as preferring
host eggs to larvae, not feeding on pollen, and rarely taking prey other than spider
mites (Croft et al. 2004). These characteristics allow the Chilean Predatory Mite to
be used like a bioinsecticide: predators are applied at a recommended prey density
and disappear after the prey have been eaten below a threshold. However some gen-
eralist phytoseiids are considered effective biocontrol agents of spider mites and
their success may depend on their ability to obtain nutrients from other sources. For
example, Typhlodromus pyri and Kampimodromus aberrans can obtain enough
nutrients from the conidia of mildews to support development and low rates of ovi-
position (Zemek and Prenerova 1997; Lorenzon et al. 2012). In natural systems,
these generalist phytoseiid mites are abundant and relatively uniform in their distri-
bution compared to most plant-parasitic mites. In the subtropical rainforests in
southeast Queensland, for example, plant-parasitic mites are rare, but one doesn’t
have to look at many leaves before finding a phytoseiid. This is especially true if the
leaves have hairs on their underside or special structures in the vein axils called
domatia (Walter 1996, 2004).

Mites and Leaf Domatia

From the perspective of an animal as minute as a mite, the surface of a leaf will
seem a complex and variable habitat. Since mites are small, they can seek out small
crevices in which to shelter. But being small also means that even a single leaf may
be an immense distance to cross. The large smooth, rain-shedding leaves of many
tropical trees must be especially challenging places to live. In contrast, leaves cov-
ered by forests of hairs (tomenta) or with assorted nooks and crannies offer refugia
for anything small enough to fit in. The differences in foliar mite populations
between adjacent trees with differences in pubescence can be striking. Hairy leaves
in tropical rainforests in Australia typically average three times as many species and
five times as many individual mites as smooth leaves at the same site. Factors such
as the density of hairs on leaf veins and in vein axils can be especially important in
the distribution of predatory mites (Karban et al. 1995; Loughner et al. 2009). Only
recently have the responses of mites to leaf architecture been considered but it is
clear that leaf surface structures alter mite abundances, influence predator–prey
interactions and are central to an understanding of the relationship between phyl-
loplane mites and plants (Walter 1992; Walter and O’Dowd 1995b). For example,
leaf hairs provide refugia from predators of phytoseiids and their eggs (Roda et al.
2001; Romero and Benson 2005; Seelmann et al. 2007; Ferreira et al. 2008).
More than a century ago a Swedish worker, Axel Lundström (1887), noticed that
tufts of hairs found in the vein axils on the underside of the leaves of lime trees often
sheltered mites. He called these tufts of hairs (similar to Fig. 8.11a), and other tiny
structures that protected a small pocket of leaf tissue, acarodomatia (mite houses)
and compared them to the much larger cavities that housed ants (formicaria or ant
Mites and Leaf Domatia 305

domatia) on the leaves of some tropical plants. Unfortunately for Lundström, botanists
did not take his observations seriously and his hypothesis that mites lived in leaf
domatia and that plants benefited from the association was considered crackpot at
best (O’Dowd and Willson 1989). It did not help that Lundström mistakenly believed
that mites induced acarodomatia. Leaf domatia are under the genetic control of
the plant, unlike the hair-like erinose patches, pocket galls and other deformities
induced by eriophyoid mites. The rolled leaf margins (orlettes) of some plants
appear to be analogous to domatia.

Structure and Distribution of Leaf Domatia

In temperate forests of the Northern Hemisphere, trees and shrubs with leaf doma-
tia are very common but the domatia themselves are mostly unremarkable struc-
tures: simple tufts of hairs in shallow pits in vein axils on the underside of leaves
(Willson 1991; O’Dowd and Pemberton 1994). These clumps of hairs may be use-
ful for telling one oak or one blackberry species from another but attributing an
adaptive function to them seems to be straying into the kind of reductionist story-
telling that Stephen Jay Gould and Richard Lewontin (1978) criticised so effec-
tively with their example of the spandrels in the dome of the great San Marco
Cathedral in Venice. The spandrels, referring to the tapering triangular surfaces
between pairs of intersecting arches that support the dome, each contain an elabo-
rate mosaic mural that perfectly fits the space and smoothly integrates the pictorial
story inlaid across the arches and ceiling of the cathedral. Gould and Lewontin
argued that it would be easy to be overwhelmed by the perfection of the iconogra-
phy, invert cause and effect, and assume that the triangular mosaics required the
use of spandrels for their display.
In the case of leaf domatia, spandrel may be a particularly apt term because a
domatium fills the triangular space formed by intersecting leaf veins. Veins often
bear hairs, so one might expect the juncture of two hairy veins to have a high density
of hairs – a plant architectural inevitability with no need for an adaptationist expla-
nation. Most botanists seem to have adopted such an iconoclastic view of leaf
domatia: they are useful taxonomic characters that needed no further explanation
(O’Dowd and Willson 1989).
However, many woody plants in subtropical to tropical forests have leaf domatia
in the form of pits sunken deep within the leaf lamina, igloo-like domes that rise
above the surface or pocket-like sheets of tissue that connect the veins (Figs. 8.10
and 8.11). These domatia are very house-like, at least at a microscale, and are not as
easily dismissed as hair tufts. Even the domatia that Lundström studied on lime tree
leaves are striking structures: large cottony tufts that rise well above the leaf surface.
As with our own armpit hair, these tangles seem poorly explained by traditional
concepts of the function of hair.
Given the diversity of structures that have been called leaf domatia, one might
wonder if the definition is too diffuse? Leaf domatia usually occur in vein axils on
Fig. 8.10 Leaf domatia or
acarodomatia. (a) Leaves
of a Blue Quandong
Elaeocarpus angustifolius,
an Australian rainforest tree
with deep pit domatia;
(b) Leaf of Kangaroo Vine
Cissus antarcticus,
a rainforest liana with
large, raised domatia
(Photos by DE Walter)

Fig. 8.11 Leaf domatia vary in structure, sometimes even within a single tree, from relatively
simple pits with tufts of hairs (a) to elaborate cave-like retreats in the leaf lamina to raised, pocket-
like pouches (b) (SEMs by DE Walter)
Mites and Leaf Domatia 307

the underside of leaves. This is true even when veins are weakly expressed.
Similarity of position is a basic criterion of homology and finding similar structures
in the same location argues for continuity through descent, at least among related
taxa. In some genera of rainforest trees (e.g. Cryptocarya, Lauraceae), certain
species have simple tuft domatia, while others have more elaborate domes or pock-
ets in the same location. An even greater diversity of domatial forms can be found
in shrubs in the genus Viburnum (Adoxaceae) (Weber et al. 2012) anda diversity of
forms can be wound even within a single leaf (Nishida et al. 2005). This pattern of
differently formed structures found in the same position in related species is moder-
ately strong evidence for functional homology with a genetic basis.
Another supporting pattern is that the structures called leaf domatia are found in
more than 30 families of angiosperms and it seems likely that they have evolved
convergently numerous times (O’Dowd and Willson 1989). Vein axils must be
predisposed to the production of leaf domatia, much as heads are predisposed sites
for the evolution of horns in various animals. Exceptions tend to prove this rule.
For example, the tropical tree Timonius timon has pit domatia that occur on the
upper surface of the leaves, but again, in vein axils (O’Dowd and Willson 1989;
Walter and O’Dowd 1995a). Short-lived plants (annuals, rosette-forming biennials)
appear to lack domatia: domatia are found primarily on the leaves of woody
perennials. For example, the Chilli Pepper (Capsicum annuum, Solanaceae) has hair-
tuft domatia in its vein axils. Although the species name implies it is an annual, and
it is often grown as an annual crop, in frost-free areas it is a perennial woody shrub.
As well as positional similarity, leaf domatia share a basic structural similarity
that seems more important than their differences in construction. Most leaf domatia
enclose small spaces, typically a millimetre or less in diameter. Access to the inte-
rior of a domatium is via a minute pore, slit or gap. For example, tuft domatia often
resemble the tangles of brush that are pulled together for temporary shelter by hunt-
ers in many tribal societies (e.g. the wickiup of the Apache and the humpy or gun-
yah of Australian Aborigines). Access to the sparsely filled centre of such a tangle
is limited by the space between the shafts of the hairs that form the outer margin of
the domatium – typically less than 100 μm. Pit domatia open through similarly
small pores. Pocket domatia consist of sheets of tissue that roof over the vein junc-
tures. These often have relatively broad openings (1–3 mm) that could be accessible
to small, flattened insects but many pocket domatia sport growths of hairs that
reduce the effective size of these openings. Except for a few very large domatia
found on some tropical plants, the size of the space enclosed and the size of the
opening to that space ensure that only very small organisms, such as mites and min-
ute insects, can enter.

What Lives in Leaf Domatia?

Although Lundström first reported mites in domatia, many botanists subsequently


claimed they were unable to find any arthropods consistently associated with leaf
domatia (O’Dowd and Willson 1989). It is certainly possible to find leaf domatia
308 8 Mites on Plants

that are uninhabited, especially on newly flushed leaves, on trees in city parks or in
vineyards with heavy pesticide use (many grape varieties have well- developed leaf
domatia). However, every serious examination of acarodomatia throughout the
world has found that small arthropods, their cast skins and their faeces are consis-
tently present.
With few exceptions, mites have been found to be the predominant inhabit-
ants of both tuft domatia and the more elaborate pits, pockets, pouches and
domes (Walter 1996). An exception is an extremely small species of thrips
(Thysanoptera) described from leaf domatia in Costa Rica (Mound 1993). Large
black and red thrips can be found in some of the larger leaf domatia in Australian
rainforests, but rarely: these thrips appear to be predators. The nymphs of small
bugs (Heteroptera) that may be predatory also are occasionally found in leaf
domatia. However, even these small insects are a bit too large to fit into most
leaf domatia comfortably. In contrast, relatively large scale insects (Homoptera,
Coccoidea) can be found completely filling domatia. Adult female scales are
bloated, sessile, plant parasites but they develop from a minute crawling stage.
The older leaves of some Australian rainforest trees sometimes have all of their
leaf domatia clogged with scale insects or their parasitised bodies. Garden trees
and shrubs, such as bay leaf and elms, may also have their leaf domatia clogged
with scale insects. Overall, however, insects are rare in leaf domatia: mites are
commonplace.
Plant-parasitic mites, however, usually are not common in domatia except during
outbreaks and, if predatory mites are present, quickly decline in relative abundance
compared to predators and fungivores (Walter 1996). An exception is discussed
below, but some reports in the literature of ‘herbivorous’ mites inhabiting domatia
are of tydeid or tarsonemid mites with obscure feeding habits. For example, Nishida
et al. (2005) found “herbivorous or fungivorous Tarsonemidae mites” the second
most common inhabitants of the domatia of Camphor Laurel (Cinnamomum cam-
phora, Lauraceae). The (Fig. 8.13c) or Yellow Mite Polyphagotarsonemus latus
(Tarsonemidae) is a pest of Camphor Laurel, but at least in Australia, fungivore
tarsonemids are common residents on Camphor Laurel leaves (Walter, unpub-
lished). Romero et al. (2011) reported Lorryia formosa (Tydeidae) as a phytopha-
gous mite on Coffea arabica (Rubiaceae). Although L. formosa is an obligate foliar
mite, there is no indication that it is a plant parasite: it occurs on a large number of
different plant species and is often the most abundant mite reported, but it has not
been shown to damage. It has been shown to be a fungivore and pollenivore (Wiggers
et al. 2005).
There is an exception to the ‘no plant-parasitic mites in domatia’ rule, though:
Gall and Rust Mites, Eriophyoidea. On saplings of the South American rainforest
tree Cupania vernalis (Sapindaceae), Romero and Benson (2004) found that erio-
phyoids were the most common mites in domatia and far more common on leaves
with open as opposed to blocked domatia, although without obvious damage to the
plants. Nishida et al. (2005) also found eriophyoid mites the most common group in
the domatia of Camphor Laurel, although when predatory phytoseiid mites were
present eriophyoids were found only in domatia with openings too small for the
Mites and Leaf Domatia 309

predators. During outbreaks of eriophyoids in Queensland rainforests, the leaf


undersides including any domatia present can harbour large numbers of eriophyoid
mites, but this is usually a transient phenomenon: eriophyoid numbers quickly drop
(presumably due to predation) and domatia are rarely galled. However, at one site
near Mt Warning in New South Wales, the pit domatia of a rainforest tree
(Elaeocarpus angustifolius, Elaeocarpaceae) appeared to be preferentially galled by
an eriophyoid that induced pocket galls. Within many of the galls, however, erio-
phyoids were rare or absent and elongate tarsonemid mites, probably a predatory
species, were in residence (Walter unpublished).
Predatory mites and those that scavenge and feed on fungi and other microbes
usually represent over 90 % of the arthropods in domatia (Walter 1996). None of
these mites are known to harm plants. In fact, plants with significant populations of
predators and scavengers are likely to receive considerable benefits from reduced
parasite load and from the mobilisation of nutrients sequestered by leaf fungi, algae
and other epiphylls such as bacteria, yeasts, liverworts and lichens as mites graze
and defecate.
Fungivores are especially common inhabitants of domatia in humid climates. For
example, on a wild grape in Florida, populations of the tydeid Orthotydeus nr.
lindquisti and the astigmatan Czenspinskia transversostriata increased dramatically
during the onset of the rainy season. Individual leaves often had several hundred
fungivores and domatia were packed with eggs and moulting mites (Walter and
Denmark 1991). Grape Powdery Mildew (Uncinula necator) is a significant prob-
lem on domestic grapes and one of the main factors limiting commercial production
in Florida; however, this mite-infested wild grape species thrives there. Farther
north, leaf domatia on grape have been shown to increase the number of fungivores
per leaf and reduced damage by mildews (Norton et al. 2000). Orthotydeus lambi
(Tydeidae) appears to be the most important of these fungivores and when in present
in sufficient density (which is positively influenced by domatia presence and size) is
capable of suppressing Grape Powdery Mildew on grape varieties that are not
unusually susceptible (English-Loeb et al. 2007).
In another study, leaves of a dozen species of rainforest plants had leaf domatia
blocked with a bitumen paint. Five months later, fungivores were six times more
common on leaves with domatia left open than on adjacent leaves with the domatia
painted shut (Walter and O’Dowd 1995b). Across a rainfall gradient in Panama, the
domatia of Psychotria horizontalis (Rubiaceae) have smaller openings and more
hairs on drier sites and during the dry season at all sites. More fungivorous mites
were present in humid sites and more scavenger-omnivores in drier sites (Richards
and Coley 2012). Therefore, leaf domatia appear to elevate populations of fungivo-
res but whether the plant gains from having herds of these mites roaming its leaves
likely various with the fungal species, plant species, and weather. Fungi vary in their
effect on plants from beneficial to deleterious and the importance of elevated num-
bers of fungivores on leaves with domatia is difficult to evaluate in natural systems.
An easier test is to look for effects of domatia on predatory mites that are known to
act as ‘bodyguards’ against plant parasitic mites in agricultural systems (Dicke and
Sabelis 1988).
310 8 Mites on Plants

Domatia as a Constitutive Plant Defence

Predatory mites are often active animals that move rapidly and frequently in their
search for food. Therefore, chance alone is a sufficient explanation for an occasional
predatory mite found loafing within a leaf domatium; however, a more consistent
association between predatory mites and leaf domatia would be more difficult to
explain. Mite use of leaf domatia has been evaluated in North America, Costa Rica,
Hawaii, Australia, New Guinea, New Zealand, Bangladesh and Korea (see review
in Walter 1996). In all of these studies, predatory mites are common inhabitants of
leaf domatia and leaves with domatia have more predators than leaves without
domatia. This is true even within a species (see Fig. 8.4 in Walter and Proctor 1999).
Even having a single domatium significantly increases the chance that a leaf also
will have a resident predatory mite. Spider mites are known to avoid plants with
resident predators (Pallini et al. 1999), so having predators in place is likely an
advantage to a plant.
That predatory mites seek out appropriately sized spaces on the plants they
inhabit is not in itself surprising but that plants produce structures that have no
known function except to house mites calls for an explanation. This is especially
true because plant-parasitic mites are at best rare and ephemeral inhabitants of
domatia. Even during outbreaks, when dozens to hundreds of pests occupy each
leaf, few plant-parasitic mites seek refuge within domatia. This could be do to the
‘cave bear effect’ – wandering into a domatium that a predator may use is likely to
be unfortunate for the wanderer. For example, Nishida et al. (2005) have shown that
on Camphor Laurel eriophyoid mites will occupy leaf domatia, but when predatory
phytoseiid mites are present eriophyoids survive only in the domatia with the
smallest openings.
Plant-parasitic mites affect plant fitness through reduction in yield (growth and
reproduction) and death of plants. The evidence for this in agricultural systems is
overwhelming, especially since the advent of synthetic insecticides elevated spider mites
to major secondary pests. Plant-parasitic mites are small, use stylet-like mouthparts
to puncture individual plant cells (and thus can avoid many plant defences) and have
short generation times and high reproductive outputs, allowing them to overexploit
their hosts (Dicke and Sabelis 1988). However, plant-parasitic mites must run a
gauntlet of predators when they attempt to colonise leaves with domatia.
Structurally complex leaves tend to accumulate predatory mites compared to
structurally simple leaves. For example, even in the absence of prey, the orchard
biocontrol agent Typhlodromus pyri prefers leaves with hairs over those without
them (Loughner et al. 2009). By the simple expedient of producing tufts of hair in
vein axils, a plant can acquire ‘bodyguards’ before attack by plant parasites.
Although herbivores don’t necessarily flee a plant with predatory mites (Meng et al.
2012), they are less likely to colonise one where predators are present. Therefore, by
retaining predatory mites on leaves, domatia can be considered a type of constitutive
defence against herbivory. One ready prediction of this hypothesis is that the removal
of domatia should result in fewer bodyguards. Walter and O’Dowd (1992a, b, 1995b)
Mites and Leaf Domatia 311

used two techniques to ‘remove’ domatia from leaves. The simplest technique was
to shave off tuft domatia with a small scalpel. The leaf surface below each doma-
tium was scrapped with the scalpel on another treatment to control for any effects
caused by shaving. After 2 weeks, predatory mites (mostly the stigmaeid Agistemus
longisetus and the phytoseiid Euseius elinae) were 4–7 times more abundant on
leaves where the domatia had been left intact (Walter and O’Dowd 1992a).
Tuft domatia in shallow pits can be removed by shaving but most leaf domatia,
even most tuft domatia, are complex structures and cannot be cut away from plants
without causing significant damage to leaves. Walter and O’Dowd’s solution to this
problem was to paint the entrances to domatia closed with bitumen paint (thickened
with ground chalk) on newly flushed leaves. As with the shaving experiment, an
appropriate control treatment was needed and in this experiment the tar paint was
dabbed below the domatia on a second set of leaves. To date, this technique has been
applied to 13 species of trees, shrubs and lianas in Australia and in every case clos-
ing domatia with paint resulted in drastic reductions in the numbers of predatory
mites in comparison to leaves with paint dabbed below the domatia (Walter and
O’Dowd 1992b, 1995b). Similar blocking studies have been conducted in other
countries since with similar results (e.g. Romero and Benson 2004, Monks et al.
2009), although domatia openings have to be large enough to allow entrance of
mites for an effect to be observed (Romero et al. 2011). When domatia of various
shapes and openings with a range of diameters are present, as on the leaves of
Camphor Laurel, mites have been shown to sort themselves into appropriately sized
domatia (Nishida et al. 2005). Similarly, average mite and domatium size have been
shown to match with wet forests having large domatia and mites and dry forests
smaller domatia and mites (Richards and Coley 2005).
Laboratory experiments have taken these results one step further to demonstrate
a protective effect of leaf domatia (Walter and O’Dowd 1992a, b). Shoots of an
ornamental shrub with domatia shaved off had higher numbers of spider mites and
lower numbers of predators than similarly treated shoots with domatia left intact.
Similarly, Grostal and O’Dowd (1994) demonstrated that predators ate more spider
mite eggs when domatia were present.
The reverse holds true as well. In an innovative series of experiments, Shelley
Rozario (1994) added tufts of polyester fibres to the vein axils of grape varieties
with poorly expressed domatia and observed the effect on the phytoseiid predator
Galendromus occidentalis. Leaves with artificial tuft domatia retained more of the
predator and these mites had higher reproductive outputs than on the same grape
variety without polyester tufts. In the field, results were more variable but in some
experiments tydeid and phytoseiid mite populations were significantly higher when
artificial domatia were present. Similar results have been found in other systems
(Agwaral 1997) and the domatia concept has been applied to cotton plants using
relatively large tufts of cotton (4 mm diameter) glued to leaves at their juncture with
the petiole. These tufts of cotton harboured predatory thrips and bugs and, as a
result, plants with artificial domatia experienced less damage by spider mites and a
30 % increase in fruit production over cotton plants without simulated domatia
(Agwaral and Karban 1997).
312 8 Mites on Plants

What’s in It for the Mites?

Given the potential of plant-parasitic mites to reduce plant fitness and the ability of
predatory mites to protect plants from attack, the independent derivation of domatia
in many lineages of angiosperms is easily understood. Under attack by plant-
parasitic mites, individual plants with leaves having resident predatory mites would
experience enhanced survival, growth and reproduction. Plant parasites would
avoid leaves with predators or be picked off as they attempted to colonise leaves
and before they could produce large, damaging populations or have much chance
to vector a plant disease. But how do domatia benefit predatory mites?
Predators with narrow prey ranges or those that require large populations of
prey for reproduction are unlikely to find leaves with domatia especially favour-
able habitats. If mites gain no particular benefit from entering domatia or if preda-
tors are lured away from better hunting grounds by hideaways on barren leaves,
then the mites could be exploited by the plants. In the field, specialised predators
such as species of Phytoseiulus are rarely found in leaf domatia; they live in the
webbing of their spider mite prey and enter domatia only at low prey densities.
Most of the predatory mites associated with leaf domatia are members of the the
Phytoseiidae, Cheyletidae, Stigmaeidae and Cunaxidae that can be found on many
different plant hosts in a region and have broad prey preferences. Many of these
mites are also able to persist on wind-blown pollen, honeydew and fungi; there-
fore, generalist predator-omnivores may find leaves having convenient shelters
especially attractive.
The most obvious benefit of leaf domatia to predators is protection and shelter,
especially for eggs and moulting mites. Several studies have demonstrated that phy-
toseiid mites preferentially use leaf domatia for oviposition and moulting (Walter
1996). In the field, more than three-quarters of all eggs laid by phytoseiid mites
were found within domatia. As discussed above, predation on phytoseiid mites is
reduced on hairy leaves. This is also true of domatia: predation against phytoseiid
eggs by Western Flower Thrips (Faraji et al. 2002a, b) and intraguild predation of
larvae by phytoseiids (Ferreira et al. 2008, 2011) is reduced in the presence of
domatia. Although Iphiseius degenerans hunts for food (thrips, pollen, nectar) in
flowers, it lays its eggs in leaf domatia far from the source of food for its offspring
(Ferreira et al. 2011).
Abiotic factors also are buffered by leaf domatia. In a series of laboratory experi-
ments, Grostal and O’Dowd (1994) demonstrated that leaf domatia buffered the
effect of low humidity and when domatia were present the spider-mite specialist
predator Galendromus occidentalis laid more than twice as many eggs as when
domatia were absent. However, although the generalist predator Typhlodromus
doreenae did best when domatia were present, an effect of humidity was apparently
only at the highest relative humidities tested (70, 98 %) (Rowles and O’Dowd
2009). In the field, though, the domatia of Psychotria horizontalis have been shown
to vary in a manner that is consistent with providing better protection in dry environ-
ments (Richards and Coley 2012).
Arboreal Scavengers and Fungivores 313

Arboreal Scavengers and Fungivores

Butterworts (Pinguicula spp.) are low-growing carnivorous plants found in moist


soils and they have leaves coated with glandular hairs that secrete drops of adhesive.
Small insects that land on a leaf are ensnared by the adhesive traps, die and are
digested by the plant. The undigested remains accumulate on the surface and can
support the growth of fungi that also can attack the leaves. However, a small oribatid
mite, Oribatula tibialis, is able to wander among the glandular hairs without becom-
ing trapped. The mites feed on fungi and on the undigested remains of prey, thereby
reducing the resources that support the growth of pathogenic fungi (Antor and
Garcia 1995).
Intuitively, leaves do not seem a suitable habitat for oribatid mites, archetypal
denizens of the soil, but species in 17 families have been recorded from leaf surfaces
from low-growing plants like butterwort to the tops of the highest rainforest trees
(Walter and Behan-Pelletier 1999). Other mite scavengers also have colonised the
phylloplane, probably in the same manner as oribatid mites – via a long evolution-
ary progression from the soil, up the trunk and stems, to the leaves. Leaves are a
good place to start in our consideration of arboreal scavengers because the phyllo-
plane is as different a habitat from the soil as one can find in forest canopies.
Temperature fluctuations, light conditions, wind currents, humidity and water flow
all tend to be buffered and predictable in soil but vary rapidly and extremely on the
phylloplane. Safe nooks and crannies are at a premium and on many smooth-
surfaced leaves a small boundary zone alone the midrib is the only refugium.

Scavenging on Leaves

Oribatid mites living on leaves are readily identifiable by their morphology, espe-
cially by the short, clubbed to globose prodorsal sensilla that are characteristically
present (Norton and Palacios-Vargas 1982; O’Dowd et al. 1991; Hunt 1996). As an
organism moves, it displaces a volume of air and creates local air currents. This is
especially true in closed spaces such as exist in the soil. Most soil-inhabiting oriba-
tids are blind but they have elongate, and often plumose, prodorsal trichobothria
(sensillus) that may provide them with a radar-like image of animals moving in
their vicinity. Mites living on leaves, however, must be exposed to an excess of wind
current information and hence the short, thick, usually globose, bothridial seta typi-
cally found on arboreal oribatid mites.
Cerotegument, a waxy cuticular secretion, can be especially thick on arboreal
oribatids and probably aids in water conservation (Alberti et al. 1997). Other modi-
fications to sense and secretory organs may reflect arboreal adaptations (Walter and
Behan-Pelletier 1999) but perhaps the clearest adaptation to life on the phylloplane
is the convergent evolution of adhesive empodia on the tarsi of many leaf-inhabiting
oribatid mites. For example, Adhaesozetes polyphyllos has a single large claw on
314 8 Mites on Plants

Fig. 8.12 Some adaptations common in foliar oribatid mites: (a) Phylleremus leei (arrow points
to clubbed sensillus); (b) Tarsus of same with suction pad (arrow); (c) Adhaesozetes polyphyllos
tarsus with adhesive pad (arrow); (d) Scapheremaeus sp. With lenticulus (arrow). All mites from
leaves in Queensland, Australia (SEMs by DE Walter)

each tarsus beneath which is a large rectangular pad that adheres to the leaf surface
(Walter and Behan-Pelletier 1993) (Fig. 8.12c). The claw probably serves both to
grip rough surfaces and to lever the pad free from smooth surfaces. Similarly, spe-
cies of Phylleremeus (Fig. 8.12a, b) have a similar suction-cup like pad (Behan-
Pelletier and Walter 2007).
An interesting, but largely unexplored, characteristic of arboreal oribatid mites is
that sexual dimorphism tends to be well developed. In soils, male oribatid mites
tend to be half the mass of females but otherwise tend to be unremarkable. Males of
arboreal taxa, however, are often covered with bumps, processes or horn-like setae
not present on females (Karasawa and Behan-Pelletier 2007). For example, in the
Oripodoidea, male Sellnickia caudata (Sellnickiidae) have a nose-like extension of
the rostrum that fits over a tubercle on the rear of the female; males of Nasozetes
(Scheloribatidae) have a similar ‘nose’. Symbioribates papuensis (Symbioribtidae)
was described from a mite living in lichens growing on the backs of weevils in New
Guinea (Aoki 1966) and S. aokii from forest canopy in Japan. Another species of
Symbioribates is common on leaves in rainforests in Queensland. The males of
these mites have elongate, inflated prodorsal setae that resemble the horns of bulls
or, given their heavy armour, ceratopsian dinosaurs. In other arboreal oribatids, the
sexual dimorphism is less noticeable, usually having to do with differences in the
expression of the porose areas in the cuticle (Behan-Pelletier et al. 2001).
Arboreal Scavengers and Fungivores 315

Soil-inhabiting oribatid mites appear to have a primarily indirect and dissociated


mating system where spermatophores are deposited by males with few or no cues
from females (see Chap. 5). Presumably males choose favoured microhabitats
where females are likely to visit and concentrate on producing lots of attractive
spermatophores. Arboreal habitats appear to require a more intimate association
between the sexes and a more morphologically expressive kind of sexual selection.
Similar sexual dimorphism is known for oribatids that have colonized other extreme
habitats, e.g. the rocky intertidal zone (Krisper and Schuster 2008) and semi-aquatic
habitats (Behan-Pelletier 1996). At least in their latter instars, crotoniine oribatids
are largely associated with the epiphytes growing on bark and tree limbs, and they
are unusual in being the only sexual lineage (with moderate sexual dimporphism) in
a cluster of parthenogenetic lineages (Colloff and Cameron 2009). Some partheno-
genetic lineages of Crotonioidea also are arboreal, but it is interesting that sexual
dimorphism and extreme habitats are again associated.
In Australian rainforests, the oribatid fauna of leaves has is composed of several
lineages of unrelated mites that have colonized the phylloplane. For example, only
Oripodoidea (members of several families) and the genus Scapheremaeus
(Cymbaeremaeoidea) are diverse on leaves. Other lineages are represented by few
genera and species. This pattern holds true for the most impressive of all sarcopti-
form radiations, the Astigmata. Only two genera of astigmatans, each from a differ-
ent family, are commonly found on leaves in Australian forests. Species of
Neotropacarus (Acaridae) are relatively large mites that occur in moderately dense
aggregations that suggest a degree of subsociality may be present. Leaves with
smooth surfaces seem to be preferred. In contrast, species of Czenspinskia
(Winterschmidtiidae [formerly Saproglyphidae]), including the cosmopolitan par-
thenogen C. transversostriata, are found on hairy leaves, and there is no indication
of aggregation behaviour. Species of Czenspinskia have been reported as common
leaf inhabitants from many parts of the world, and sometimes are mistaken for
plant-parasites, but gut contents contain detritus (e.g. senescent leaf hairs) and fungi
(Walter and Denmark 1991). They thrive on sooty moulds growing on aphid honey-
dew and may be mildly beneficial to plants (Atcheson 1963).
In mesic to humid forests, densities of oribatid mites can exceed several hun-
dred per leaf; however, over all leaf and forest types the most characteristic aca-
rine inhabitants of the phylloplane are tydeid mites. Some of these have been
shown to be important consumers of foliar fungal pathogens (English-Loeb et al.
2007), but many also are suspected of being opportunistic predators, especially of
eriophyoid mites. These tydeids tend to be pale to apricot in colour. Species in
other genera, however, often take on deep greenish hues, presumably from feed-
ing on algae and lichens. Herbivory is suspected in some foliar tydeids but we
have been unable to find convincing evidence that any tydeid mites have negative
effects on higher plants.
When leaves or fruit are damaged or attacked by parasitic fungi, rampant fungal
growth often occurs on the damaged and dying tissues. Mites that feed on fungi, e.g.
oribatids, astigmatans, tydeids and tarsonemid mites, are attracted to such areas and
can produce large populations. Even healthy leaves often have a fine covering of
316 8 Mites on Plants

Fig. 8.13 As in many other families of mites, trophic behaviour in the Tarsonemidae varies con-
siderably: (a) Fungitarsonemus sp. lives on leaves and feeds on fungi and probably other microbes;
(b) Paratarsonemella giblindavisi is inhabits figs, is phoretic on fig wasps and probably feeds on
galled ovules; (c) Two Broad Mites, Polyphagotarsonemus latus, clinging to the leg of a Whitefly
(Bemisia sp.). Broad Mites are plant parasites and important economic pests (SEMs by DE Walter)

opportunistic fungi and honeydew from feeding by homopterans often produces dense
growths of mildews that can support populations of scavengers and fungivores.
Another group of fungivores that has done well on the phylloplane is the
Tarsonemidae. One particularly successful group is Fungitarsonemus (Fig. 8.13a).
Many of these mites have the unusual habit of collecting tiny bits of detritus – pol-
len, algal balls, bits of leaf hairs – and gluing them onto their backs. Presumably this
camouflages them from predators and it certainly can astonish one to see a minute
ball of fluff dashing across a leaf. Some species also glue extensive fields of detritus
to the leaf surface in which eggs are laid and larvae secrete themselves before the
moulting to adults. The latter is an especially interesting behaviour because in the
Tarsonemidae quiescent larvae (with the developing female inside) are usually car-
ried on the sucker-like genital disc.
Most species of Tarsonemus are thought to be fungivores, including species in
the waitei-group that have sometimes been suspected of phytophagy (Suski 1967,
1972a, b; Lindquist 1986; Nucifora and Vacante 1986). They are commonly found
in protected areas on leaves and fruit (e.g. in leaf domatia and under the calyx) and
are usually associated with growths of fungi. However, some species of Tarsonemus
and members of the genera Acaronemus and Dendroptus are predators of eriophy-
oid mites or the eggs of Tetranychoidea. For example, Tarsonemus acerbilis larvae
were found in 40 % of Maple Spindle Galls induced by Vasates aceriscrumena and
most of these galls lacked eggs of the gall mite (Patankar et al. 2012). Many other
Arboreal Scavengers and Fungivores 317

genera of Tarsonemidae can be found on leaves, especially old rainforest leaves that
are encrusted with growths of lichens and liverworts. However, these mites, e.g.
species of Daidalotarsonemus and Ceratotarsonemus, are usually more abundant
on stems and trunks where growths of epiphytes are more extensive.

Moss and Lichen Mites

Oribatid mites often are called moss mites (moosmilben in German) because they
dominate the acarofauna living in the bryophytes. Species in the families Achipteriidae,
Ameronothridae, Camisiidae, Carabodidae, Ceratozetidae, Crotoniidae, Damaeidae,
Eremaeidae, Galumnidae, Hammeriellidae, Liacaridae, Liodidae, Megeremaeidae,
Micreremidae, Mochlozetidae, Mycobatidae, Nanhermanniidae, Oppiidae, Oribatulidae,
Oripodidae, Pedrocortesellidae, Peloppiidae, Phenopelopidae, Phthiracaridae, Tecto-
cepheidae and Xenillidae all are commonly found associated with moss and other
‘cryptogams’ on tree trunks and branches (Walter and Behan-Pelletier 1999). In spite
of this association and strong evidence that oribatid mites and other microarthropods
have a ‘pollinator-like’ relationship with mosses (Cronberg et al. 2006; Rosenstiel
et al. 2012), there is little indication that oribatid mites feed on moss leaves. However,
in the eucalyptus forests of north Queensland, Birobates hepaticolus (Oripodidae),
both lives in and feeds on a bark-inhabiting leafy liverwort (Colloff and Cairns 2011).
Liverworts that live on the surface of leaves (epiphylls) in Queensland’s rainforest
often have mites hiding in them, but no similar relationship has yet been
demonstrated.
Epiphytes (literally, ‘on plants’) often grow in patches and mites living in
these clumps have highly aggregated distributions. Some specificity to different
types of epiphytes is also known. For example, species of Dometorina are found
in crustose lichens, those of Eueremaeus and Trichoribates are sheltered by foli-
ose lichens (André 1984, 1985). More general patterns of host-specificity may be
present as well. In a series of studies on different tree species in North America,
Europe and Africa, Nicolai (1986, 1989, 1993) demonstrated that both tree spe-
cies and bark type affected the development of epiphytes and the resulting oriba-
tid diversity. Of 19 species collected from the bark of five species of coniferous
and five species of deciduous trees in North America only five mites were com-
mon to both. Among deciduous species, those with structured bark had the highest
numbers of species and individuals, whereas the reverse was true for coniferous
tree species (Nicolai 1993).
Epicorticolous oribatids graze lichens, algae and fungi, and also may have more
complex ecological roles. For example, species of Humerobatidae are known to
disperse lichen soridia (Stubbs 1995), and other oribatid mites may spread fungal
pathogens. Nanelli and Turchetti (1993) demonstrated that Scheloribates latipes
from bark habitats on chestnut can be cultured on virulent strains of the fungus that
causes chestnut canker, that virulence is not reduced in passage through the mite
guts and that this species also can transmit the disease through carrying fungal
318 8 Mites on Plants

spores on the body. Species of the ceratozetid genera Humerobates and Diapterobates
are particularly common on twigs, branches and buds (Murphy and Balla 1971;
Momen 1987), and sometimes alarm orchardists. However, these mites scavenge
on dead plant and animal material and eat living fungi and algae: they pose no
threat to the trees.
In Japan, a stem-inhabiting oribatid, Diapterobates humeralis, has an unex-
pected relationship with the hemlock woolly adelgid (Adelges tsugae), a destruc-
tive pest of hemlock trees in North America. The female adelgid feeds on young
shoots and before dying lays about 200 eggs in a sac covered with waxy filaments.
The mite feeds primarily on the hyphae and spores of fungi; however, it has also
developed a taste for the woolly filaments covering the eggs and while grazing on
them dislodges the egg sacs, which then fall to the ground. Displacement of egg
sacs is extremely high in sites with large populations of mites and survival of dis-
placed eggs is near zero, so this oribatid mite acts as an inadvertent predator
(McClure 1995).

Fungal Sporocarps

Large, fleshy to woody sporocarps are produced by many fungi growing on standing
dead wood and others sprout from the heartwood of infected living trees. These
include traditional mushroom-like structures that persist only for short periods of
time and more elaborate, longer-lived conks, shelf fungi and bracket fungi. All of
these sporocarps soon become inhabited by mites, primarily through phoretic trans-
port by insects, especially the beetles and flies that breed in the fungi. All except the
woodiest sporocarps are ephemeral resources, so colonisation, development and
emigration must occur within a restricted span of time. This is especially true during
the warmer months in temperate to subtropical climates and at any time in the trop-
ics. Therefore, the mite fauna of sporocarps in these climates tends to be dominated
by species with short generation times, such as those found in many Astigmata and
Mesostigmata.
In temperate forests, the Mesostigmata assemblages on bracket fungi are com-
posed of species in about a dozen genera (Pielou and Verma 1968; Matthewman
and Pielou 1971; Lindquist 1975, 1995; OConnor 1984; O’Connell and Bolger
1997; Gwiazdowicz and Lakomy 2002). In more tropical realms, a diversity of
unusual forms can be found (Moraza and Lindquist 2011). For example, in sub-
tropical to tropical Australia, a single sporocarp may have more than a dozen spe-
cies of mesostigmatans (Walter and Proctor 1998; Walter et al. 1998). These mites
primarily belong to the families Ameroseiidae, Blattisociidae, Cercomegistidae,
Digamasellidae, Laelapidae, Macrochelidae, Sejidae and Uropodidae (Lindquist
1975; OConnor 1984; Walter and Proctor 1998). Most mesostigmatans found in
bracket fungi are mycetophiles that feed as predators on the insects, mites and
nematodes that inhabit decaying sporocarps but Ameroseiidae (Ameroseius) and
some other Mesostigmata are fungivores (Fig. 8.14).
Arboreal Scavengers and Fungivores 319

Fig. 8.14 Some


representatives of the
Ameroseiidae, a family of
mycophagous Mesostigmata:
(a) Epicroseius walteri;
(b) Ameroseius sp.;
(c) Asperolaelaps rotundus
(SEMs by DE Walter)

Mycophagous mesostigmatans appear to have arisen several times from preda-


tory mycetophilic ancestors (Lindquist 1975, 1995; OConnor 1984). For example,
species of Hoploseius (Blattisociidae) are found only on living bracket fungi, where
they feed on fungal tissue (Lindquist 1963, 1995; Walter 1998). Most species live
on the surface of the bracket fungi where they can be observed probing the spore
tubes with their mouthparts; however, a Mexican species, H. tenuis, has an extraor-
dinarily elongate body that allows it to live in the spore tubes on the underside of the
fungal shelf (Lindquist 1965). An unrelated mite from bracket fungi in western
Canada and the United States, Mycolaelaps maxinae (Melicharidae), has conver-
gently evolved a similar elongate body and tube-inhabiting life style (Lindquist
1995).
Histiostomanatidae is one of the most common groups of astigmatans in fleshy
fungi. These mites crawl through rotting sludge, filter-feeding on bacteria, yeasts
and very fine particulate organic matter (Walter and Kaplan 1990). In boreal to
cool temperate forests, the acarofauna in long-lived bracket fungi is often
dominated by oribatid mites (Pielou and Verma 1968; Matthewman and Pielou
1971; O’Connell and Bolger 1997). Annual and perennial bracket fungi have a
much more diverse fauna of oribatid mites than that of the more ephemeral agaric
and bolete fungi. Oribatids in fungal sporocarps show no evidence for feeding
specificity on a particular fungus or avoidance of the toxic metabolites in certain
fungi (O’Connell and Bolger 1997) and these mites may originate from the fauna
associated with dead wood.
320 8 Mites on Plants

Under Bark

Dead wood on living trees forms a significant component of aboveground biomass


and decomposition may proceed for some time before the stem or trunk crashes to
the forest floor. This standing dead wood is home to a variety of oribatid mites and
other detritivores that invade from the bark or in association with wood- boring
insects. Beetle galleries within living wood are also likely to harbour a distinctive
mite fauna; however, few studies of this fauna exist except for those associated with
moribund conifers in the Northern Hemisphere.
Pine, spruce, fir and hemlock all are attacked by a number of scolytid, ceramby-
cid and buprestid beetles that cause significant economic damage. The galleries
constructed under bark and the burrows that enter the sapwood of the trees are home
to numerous Astigmata, Mesostigmata and Prostigmata (Kinn 1971; Moser and
Roton 1971; Cardoza et al. 2008). Tarsonemid mites are often prominent among the
latter and include parasitoids of beetle eggs (Lindquist 1986) and fungivores that
often have intricate relationships with the fungi, beetles and trees (Bridges and
Moser 1986; Moser 1985; Moser et al. 1989). Some of these mites are implicated in
the transmission of fungal disease such as Dutch Elm (Moser et al. 2010) and others
as vectors of fungi that compete with the fungal resources used by the beetles and
negatively affect their fitness (Hofstetter et al. 2006a, b; 2007). Mesostigmata,
including many of the genera found in decaying fungi, are especially prominent as
predators of nematodes and mites in galleries, and as phoretics on beetles (Lindquist
1965, 1975; Lindquist and Hunter 1965; McGraw and Farrier 1969; Kinn 1971;
Moser and Roton 1971; Kinn 1987; Lindquist and Wu 1991).

Mites and Biological Control

The first mite reported in the literature as a biological control agent of a pest was not
a phytoseiid mite – without a doubt one of the most useful of all families of biocon-
trol agents – nor did it belong to any of the other families of mites usually reported
as predators on economically important plants. Instead, a seemingly inoffensive
astigmatan, Hemisarcoptes malus (Hemisarcoptidae), was found feeding on the
oyster scale Lepidosaphes ulmi on apple in 1868 (Gerson and Smiley 1990). Rather
than being microbivores or parasites of insects or vertebrates, as are most Astigmata,
all species of Hemisarcoptes are predators of armoured scales (Homoptera:
Diaspididae). Mite eggs are laid under scales and larvae, non-dispersing nymphs
and adults feed on scales, especially ovipositing female scales. Deutonymphs are
phoretic under the elytra of coccinellid beetles (Chilocorus spp.) (Gerson and
Izraylevich 1997).
Phytoseiid mites (De Moraes et al. 2004) are still the mainstay of biological
control of mite and small insect pests, especially in greenhouse production and orchards,
but other mesostigmatans, especially in the family Laelapidae are becoming better
Mites and Biological Control 321

known. It seems strange that given the diversity of predatory mites in the world, so
few are exploited as biocontrol agents, but we really know very little about most of
their critical behaviours and microhabitat requirements (Beaulieu and Weeks 2007;
Sabelis et al. 2008). What we do know has been strongly influenced by those mites
that are easily cultured, such as the Phytoseiulus persimilis-Tetranychus urticae system.
Recent research involving interactions belowground interactions is reviewed in
Chap. 6. Here we briefly touch on the aboveground aspects.

Infochemicals

Recent advances in understanding mites as biological control agents have been just
as surprising as the original discovery of a predatory astigmatans, although in a
quite different manner. Any one who has watched a phytoseiid foraging on a leaf
might think that any encounter between predator and prey was a matter of luck,
made probable only by the energy and persistence of the predator. Phytoseiid mites,
as in all Mesostigmata, are blind and lack the distinctive chemosensory solenidia
characteristic of acariform mites. Therefore, phytoseiid mites were thought to sense
prey only from short distances and primarily by touch and contact chemoreception.
However, it has now been convincingly shown that not only do phytoseiid mites
such as Phytoseiulus persimilis sense their prey from a distance but they also
respond to plant volatiles that can call in the predators (Dicke and Sabelis 1988;
Dicke et al. 1990, 1998; Vet and Dicke 1992; Bruin et al. 1995; Takabayashi and
Dicke 1996; De Boer and Dicke 2006, and see section “Feeding Biology of
Phytoseiid Mites” above).
The realisation that plants are actively involved in their own defence should not
have been so unexpected; it is a direct prediction of tritrophic-level theory (Price et al.
1980). Predatory mites inhabit a complex chemical landscape composed of cues from
plants, pests, competitors and conspecifics, and they must be able to distinguish
among these signals properly. While responding, mites must negotiate a complex
physical landscape that includes structures like leaf hairs, domatia and nectaries that
are also, in part, produced by the plant for its own defence. The effectiveness of phy-
toseiid mites as biocontrol agents is a function not only of long-distance attraction to
prey kairomones and arrestment within patches of prey but also of the ability to rec-
ognise the information contained in volatiles released by plants in response to damage
by herbivores, to utilise plant produced structures such as domatia and to navigate
between other plant components such as glandular hairs (Walter 1996).

Induced Resistance

Spider mites (Tetranychidae) are major pests of grapes in most of the world’s vine-
yards but outbreaks are often prevented or controlled by phytoseiid mites (Helle and
Sabelis 1985a, b; Duso 1992, 1993). In California, however, Richard Karban and his
322 8 Mites on Plants

coworkers at the University of California, Davis, have found an unexpected interaction


between two species of spider mites. Zinfandel grapes that are attacked by Willamette
Spider Mite (Eotetranychus willamettei) early in the season are resistant to attack by
Pacific Spider Mite (Tetranychus pacificus) later in the season.
Willamette Mite feed primarily on grape, overwinter under bark and form small
colonies with little webbing on the upper surface of young leaves in early spring.
Damage to plants can be serious but often populations fail to reach damaging levels
and in the Central Valley of California Willamette Mite is not considered a serious
pest. In contrast, Pacific Spider Mite attacks a broad range of plants, including many
weeds, forms large colonies and produces dense webbing that protects them both
from chemical miticides and from all but specialised predators. Pacific Mite attacks
grapes in the Central Valley later in the season and thrives during hot, dry conditions
and on mature leaves. Damage to grapes causes the leaves to appear burned, leaf
drop is extensive and yield (sugar level and fruit volume) is drastically reduced
(Karban et al. 1997).
Treating vineyards with broad-spectrum miticides has been generally ineffective
and has caused resurgence of Pacific and Willamette Mite populations. Pacific Mite
problems have become chronic and acute. The use of chemicals less disruptive to
predator populations provides some help; however, Karban and his coworkers found
that the best method for controlling Pacific Mite was to avoid miticide applications
and release Willamette Mite in vineyards early in the season. When chemicals are
not used, Willamette Mite caused little damage to grapes but these vines showed
increased resistance to Pacific Mite resulting in little or no Pacific Mite damage.
Although elevated predator populations may explain some of this effect, resistance
to spider mite feeding induced by early season herbivory appears to be more impor-
tant (Karban et al. 1997).
Not all scenarios are so rosy for plant production: spider mites can also inhibit
plant defences. For example, Tetranychus evansi has been shown to both suppress
the induction of tomato defensive compounds and to suppress the release of vola-
tiles that attract predators of the spider mite (Sarmento et al. 2011).

Transgenic Mites

In the first edition of this book, we speculated that mites might be the first arthropod
biocontrol agents engineered using recombinant DNA techniques to be released
into the field and in a limited trial release of a transgenic phytoseiid mite Galendromus
(= Metaseiulus) occidentalis was carried out (Li and Hoy 1996; Presnail et al. 1998;
Hoy 2000). Although the lack of appropriate risk assessment guidelines has stifled
further efforts (Hoy 2009), phytoseiid mites are still excellent candidates for such a
release. An extensive body of work exists that shows that wild-type strains of many
species of phytoseiid mites have been released over and over again without causing
environmental harm (Hoy 1992). There is some concern that transgenic predators
will be used primarily for short-term solutions, such as producing pesticide-resistant
Mites and Biological Control 323

predators to augment chemical control programs (Hoy 1992). However, the potential
environmental benefits of using transgenic mites are considerable and one suspects
that eventually they will be used.

Biocontrol of Weeds

Biological control of mite pests by other mites is a common tool in many IPM sys-
tems and the use of mites against weeds appears to be a rapidly developing field.
However, the first demonstration of the potential of mites for biocontrol of weeds
was entirely by accident. Prickly pear and other cacti are noxious weeds in many
parts of Australia, especially Queensland, where in the early part of the twentieth
century 25 million hectares were infested. In 1924, however, the cactus spider mite
Tetranychus desertorum was discovered to be established and spreading at a rate of
80 km per year. The spider mite soon became a significant control agent of the
prickly pear Opuntia inermis. Heavily infested plants produced fewer fruit and
prickly pear densities dropped 75 % in parts of Queensland. Although the desert
spider mite looked to be the first example of a good mite biocontrol agent of a weed,
mite populations declined dramatically when cactus moth (Cactoblastis spp.) popu-
lations devastated prickly pear (Hill and Stone 1985).
Research into spider mites as biological control agents of weeds has been limited
by the commonness of polyphagous crop pests and by the difficulty of distinguish-
ing species in the most promising genus: Tetranychus (Andres 1983; Cromroy
1983). However, Tetranychus lintearius is being tested against gorse (Ulex euro-
paeus) in New Zealand, Australia, Chile, Hawaii, and the west coast of mainland
USA. Early results have been promising, but naturalised populations of Chilean
Predatory Mite (Phytoseiulus persimilis), endemic phytoseiids, and coccinellid bee-
tles (Stethorus spp.) appear to be limiting the mite’s effectiveness (Davies et al.
2007; Palevsky et al. 2013).
In contrast to spider mites, eriophyoid mites tend to be highly host specific or
even tissue specific. Collection data (as opposed to host specificity tests) indicate
that 80 % of eriophyoids have been recorded from only one host species, 99 % from
a single plant family and that gall-inducing species are more host specific than leaf
vagrants (Skoracka et al. 2010). Some of the apparent generalists, e.g. Abacarus
hystrix (known from 60 species of grasses) show geographical preferences in host
choice and overall show high degrees of genetic divergence (Skoracka and Dabert
2009), suggesting a complex of sibling species may be present. Although the impact
of these tiny parasites on their hosts may be hard to evaluate, those species that
attack shoots and flowers, and so reduce seed set, or that kill their hosts, are poten-
tially excellent biological control agents (Briese and Cullen 2001).
Rush Skeleton Weed, Chondrilla juncea (Asteraeceae) is a native of Europe,
but an introduced weed in temperate to semi-arid open scrublands in Argentina,
the western United States and south eastern Australia (Rosenthal 1996) and a
major weed in dryland wheat farming. Rush Skeleton Weed is an apomictic
324 8 Mites on Plants

perennial that overwinters as a rosette. In the spring the rosette bolts to produce a
skeletal flowering stalk which can yield over 20,000 wind-dispersed seeds. Even
root fragments can produce new plants. The rosette reforms after autumn rains
(Parsons and Cuthbertson 1992).
The skeleton weed gall mite, Aceria chondrillae, is highly host specific. The mite
attacks only skeleton weed and mites from different areas of Europe prefer (or are
limited to) different strains of the plant (Carèsche and Wapshere 1974; Cullen and
Moore 1983). The mites overwinter in the rosette and ascend the stem to the flower
buds during the bolt. Summer generation times are around 10 days and females
form galls in vegetative and flower buds that can reach 5 cm in diameter and harbour
over 500 mites (Rosenthal 1996). At high levels of infestation, seed production and
plant growth are suppressed, apical development is stunted and plant longevity
reduced. In Australia and the United States, skeleton weed gall mites have been
released as part of a complex of natural enemies that includes a gall midge and a rust
(Sobhian and Andres 1978; Cullen et al. 1982), and there seems to be general agree-
ment that the mite is a useful part of the complex, although its importance varies
from place to place (Smith et al. 2010). Other species of Aceria, for example the
lantana gall mite Aceria lantanae, and a variety of other eriophyoid mites are being
considered for use against a variety of weeds (Cromroy 1983; Amrine 1996;
Rosenthal 1996). In general, though, no magic bullets have emerged from the use of
eriophyoid mites in classical biological control of weeds (Smith et al. 2010;
Boughton and Pemberton 2011).
In contrast to a primary need of classical biological control – host specificity –
the greatest potential for eriophyoid mites to control invasive weeds may require the
ability to switch hosts and to act as a disease vector. For example, Phyllocoptes
fructiphilus appears to be a natural parasite of roses endemic to North America, but
it has colonized the invasive Asian Rosa multiflora. The mite vectors a virus that
causes Rose Rosette Disease, apparently also native to North America, and can
cause high mortality in dense stands of Rosa multiflora (Smith et al. 2010).

Summary

During their long evolutionary association, mites and terrestrial plants have devel-
oped a variety of interactions. Some seeds with sticky, mucilaginous coats (myxo-
spermy) are reported to trap unwary soil mites; but in general, it is the mites that
have direct or indirect effects on plants, from beneficial to severely detrimental.
Mite parasitism of vascular plants has evolved at least a dozen times, including a
number of independent colonisations of plant reproductive organs by pollen- and
nectar-feeding mesostigmatans and by gall mites. Since these mites directly affect
plant fitness, they must be important selective agents. The Eriophyoidea, which
includes most gall-forming mites, is the most successful radiation of acarine plant
parasites (about 4,000 described species). Many ferns and most conifers and woody
angiosperms are hosts to eriophyoids. Although the infections often appear to be
relatively benign, eriophyoids are capable of severely damaging or killing plants
References 325

through their own feeding and by transmission of malign plant viruses. In agricultural
systems, however, it is usually the damage caused by spider mites that results in the
most economic loss. Many spider mites produce extensive webs and appear to have
reached a subsocial level of colony organisation. Gall mites and web-producing
spider mites are well protected from environmental extremes and also from preda-
tory mites by their homes, but predatory mites are at the mercy of wind, rain and
larger predators on the phylloplane.
Predatory mites naturally seek out protected areas on leaves and many species
of woody plants have taken advantage of this behaviour by producing tufts of
hairs and more elaborate pits, pouches, pockets and domes in their vein axils.
These mite-sized ‘little houses’ are called leaf domatia and extensive evidence
now exists to show that the mite–leaf domatia association is widespread and
generally mutually beneficial. Plants benefit because many plant-parasitic mites
are capable of over-exploiting their hosts, strongly reducing reproductive output
and even killing large trees. However, a variety of voracious predators, especially
phytoseiid mites, are capable of suppressing populations of plant parasites.
The mite–leaf domatia mutualism and other recently discovered (or rediscovered
in the case of leaf domatia) tritrophic interactions between mites and plants hold
promise for increasing the effectiveness of natural enemies in agricultural systems
and a better understanding of why outbreaks of pest mites appear rare in
undisturbed systems.
Agricultural fields and horticultural plantings, however, have numerous pest out-
breaks and atypical, low diversity arboreal acarofaunas. In contrast, forests through-
out the world are covered and completely infiltrated by a fine dusting of acarines. In
undisturbed forests, fungivores and scavengers dominate this fauna, often at incred-
ible densities. Given the antiquity of this association, there has been time enough for
interactions as interesting as those between plants and herbivores and predators to
develop and recent research supports the hypothesis that plants obtain substantial
protection from foliar pathogens from the feeding of fungivorous mites. The discov-
ery of the diversity and abundance of mites in rainforest canopies is recent, however,
and our state of knowledge about these animals and their associations can only be
described as rudimentary.

References

Adar, E., Inbar, M., Shira, G., Doron, N., Zhang, Z.-Q., & Palevsky, E. (2012). Plant-feeding and
non-plant feeding phytoseiids: Differences in behavior and cheliceral morphology. Experimental
& Applied Acarology, 58, 341–357.
Affeld, K., Worner, S., Didham, R. K., Sullivan, J., Henderson, R., Olarte, J. M., Thorpe, S.,
Clunie, L., Early, J., Emberson, R., Johns, P., Dugdale, J., Mound, L., Smithers, C., Pollard, S.,
& Ward, J. (2009). The invertebrate fauna of epiphyte mats in the canopy of northern rata
(Myrtaceae: Metrosideros robusta A. Cunn.) on the West Coast of the South Island, New
Zealand. New Zealand Journal of Zoology, 36, 177–202.
Agwaral, A. A. (1997). Do leaf domatia mediate a plant–mite mutualism? An experimental test of
the effects on predators and herbivores. Ecological Entomology, 22, 371–376.
326 8 Mites on Plants

Agwaral, A. A., & Karban, R. (1997). Domatia mediate plant–arthropod mutualism. Nature, 387,
562–563.
Alberti, G., Norton, R. A., Adis, J., Fernandez, N. A., Franklin, E., Kratzmann, M., Moreno, M. I.,
Weigmann, G., & St, W. (1997). Porose integumental organs of oribatid mites (Acari,
Oribatida). 2. Fine structure. Zoologica, 146, 33–114.
Amrine, J. W., Jr. (1996). Phyllocoptes fructiphilus and biological control of multiflora rose. In E.
E. Lindquist, M. W. Sabelis, & J. Bruin (Eds.), Eriophyid mites: Their biology, natural enemies
and control (6th ed., pp. 741–749). Amsterdam: Elsevier.
Amrine, J. W., Jr., & Stasny, T. A. (1994). Catalog of the Eriophyoidea (Acarina: Prostigmata) of
the world. West Boomfield: Indira Publishing House.
André, H. M. (1984). Notes on the ecology of corticolous epiphyte dwellers. 3. Oribatida.
Acarologia, 25, 385–396.
André, H. M. (1985). Associations between corticolous microarthropod communities and epiphytic
cover on bark. Holarctic Ecology, 8, 113–119.
André, H. M. (1986). Notes on the ecology of corticolous epiphyte dwellers. 4. Actinedida
(especially Tydeidae) and Gamasida (especially Phytoseiidae). Acarologia, 26, 107–115.
Andres, L. A. (1983). Considerations in the use of phytophagous mites for the biological control
of weeds. In M. A. Hoy, G. L. Cunningham, & L. Knutson (Eds.), Biological control of pests
by mites (Vol. 3304, pp. 53–56). Berkeley: University of California Agriculture Experiment
Station Special Publication.
Annells, A. J. (1995). The reproductive biology and mating behaviour of redlegged earth mite: An
overview. Proceedings of the Second National Workshop on Redlegged Earth Mite, Lucerne
Flea and Blue Oat Mite, (pp. 29–31). Victoria: Victoria Printing.
Antor, R. J., & Garcia, M. B. (1995). A new mite–plant association: Mites living amongst the
adhesive traps of a carnivorous plant. Oecologia, 101, 51–54.
Aoki, J.-i. (1966). Epizoic symbiosis: An oribatid mite, Symbioribates papuensis, representing a
new family, from cryptogamic plants growing on the backs of Papuan weevils (Acari:
Cryptostigmata). Pacific Insects, 8, 281–289.
Aoki, J.-i. (1992). Oribatid mites inhabiting orchid plants in greenhouse. Journal of the Acarological
Society of Japan, 1(1), 7–13.
Aponte, O., & McMurtry, J. A. (1997). Damage on ‘Hass’ avocado leaves, webbing and nesting
behaviour of Oligonychus perseae (Acari: Tetranychidae). Experimental & Applied Acarology,
21, 265–272.
Arthur, A. L., Weeks, A. R., Hill, M. P., et al. (2011). The distribution, abundance and life cycle of
the pest mites Balaustium medicagoense (Prostigmata: Erythraeidae) and Bryobia spp.
(Prostigmata: Tetranychidae) in Australia. Australian Journal of Entomology, 50, 22–36.
doi:10.1111/j.1440-6055.2010.00778.
Atcheson, W. C. (1963). An ecological study of three species of mites on American Linden.
Journal of Economic Entomology, 46, 705.
Baker, E. W., & Delfinado-Baker, M. (1985). An unusual new species of Neocypholaelaps from the
nests of stingless bees (Apidae: Meliponnae). International Journal of Acarology, 11, 227–232.
Beard, J., & Ochoa, R. (2010). Ontogenetic modification in the Tuckerellidae (Acari: Tetranychoidea).
International Journal of Acarology, 36, 169–173. doi:10.1080/01647950903555459.
Beard, J., & Walter, D. E. (2010). New spider mite genus (Prostigmata: Tetranychidae) from
Australia & New Zealand, with a discussion of Yezonychus Ehara. Zootaxa, 2578, 1–24.
Beaulieu, F., & Weeks, A. R. (2007). Free-living mesostigmatic mites in Australia: Their roles in
biological control and bioindication. Australian Journal of Experimental Agriculture, 47, 460–
478. doi:10.1071/EA05341.
Beaulieu, F., Walter, D. E., Proctor, H. C., Kitching, R. L., & Menzel, F. (2006). Mesostigmatid
mites (Acari: Mesostigmata) on rainforest tree trunks: Arboreal specialists, but substrate
generalists? Experimental & Applied Acarology, 39, 25–40.
Beaulieu, F., Walter, D. E., Proctor, H. C., & Kitching, R. L. (2010). The canopy starts at 0.5 m:
Predatory mites (Acari: Mesostigmata) differ between rain forest floor soil and suspended soil
at any height. Biotropica, 42, 704–709.
References 327

Behan-Pelletier, V. M. (1996). Naiazetes reevesi n g, n sp. (Acari: Oribatida: Zetomimidae) from


semi-aquatic habitats of eastern North America. Acarologia, 37, 345–355.
Behan-Pelletier, V. M., Clayton, M., & Humble, L. (2001). Parapirnodus (Acari: Oribatida:
Scheloribatidae) of canopy habitats in western Canada. Acarologia (Paris), 42, 75–88.
Behan-Pelletier, V. M., & Walter, D. E. (2007). Phylleremus n. gen., from leaves of deciduous trees
in eastern Australia (Oribatida: Licneremaeoidea). Zootaxa, 1387, 1–17.
Boczek, J., & Shevchenko, V. G. (1996). Ancient associations: Eriophyoid mites on gymnosperms.
In E. E. Lindquist, M. W. Sabelis, & J. Bruin (Eds.), Eriophyoid mites: Their biology, natural
enemies and control (pp. 217–226). Amsterdam: Elsevier.
Boggs, C. L., & Gilbert, L. E. (1987). Spatial and temporal distribution of Lantana mites phoretic
on butterflies. Biotropica, 19, 301–305.
Bolland, H. R., Gutierrez, J., & Flechtmann, C. H. W. (1998). World catalogue of the spider mite
family (Acari: Tetranychidae). Brill: Leiden.
Boughton, A. J., & Pemberton, R. W. (2011). Limited field establishment of a weed biocontrol
agent, Floracarus perrepae (Acariformes: Eriophyidae), against old world climbing fern in
Florida – A possible role of mite resistant plant genotypes. Environmental Entomology, 40,
1448–1457. doi:10.1603/EN11030.
Bridges, J. R., & Moser, J. C. (1986). Relationship of phoretic mites (Acari: Tarsonemidae) to the
blue-staining fungus Ceratocystis minor in trees infested by southern pine beetle (Coleoptera:
Scolytidae). Environmental Entomology, 15, 951–953.
Briese, D. T., & Cullen, J. M. (2001). The use and usefulness of mites in biological control of
weeds. In R. B. Halliday, D. E. Walter, H. C. Proctor, R. A. Norton, & M. J. Colloff (Eds.),
Acarology: Proceedings of the 10th International Congress of Acarology (pp. 453–463).
Melbourne: CSIRO Publishing.
Bruce-Oliver, S. J., Hoy, M. A., & Yaninek, J. S. (1996). Effect of some food sources associated
with cassava in Africa on the development, fecundity and longevity of Euseius fustis (Pritchard
and Baker) (Acari: Phytoseiidae). Experimental & Applied Acarology, 20, 73–85.
Bruin, J., Sabelis, M. W., & Dicke, M. (1995). Do plants tap SOS signals from their infested
neighbours? Trends In Ecology and Evolution, 10, 167–170.
Cardoza, Y. J., Moser, J. C., Klepzig, K. D., & Raffa, K. F. (2008). Multipartite symbioses among
fungi, mites, nematodes, and the spruce beetle, Dendroctonus rufipennis. Environmental
Entomology, 37, 956–963. doi:10.1603/0046-225X (2008)37[956:MSAFMN]2.0.CO;2.
Carèsche, L. A., & Wapshere, A. J. (1974). Biology and host specificity of the Chondrilla gall mite
Aceria chondrillae (G. Can.) (Acarina, Eriophyidae). Bulletin of Entomological Research, 64,
183–192.
Chant, D. A., & McMurtry, J. A. (1994). A review of the subfamilies Phytoseiinae and
Typhlodrominae (Acari: Phytoseiidae). International Journal of Acarology, 20, 223–310.
Charles, J. G., & White, V. (1988). Airborne dispersal of Phytoseiulus persimilis (Acarina:
Phytoseiidae) from a raspberry garden in New Zealand. Experimental & Applied Acarology, 5,
47–54.
Clements, D. R., & Harmsen, R. (1992). Stigmaeid–phytoseiid interactions and the impact of
natural enemy complexes on plant-inhabiting mites. Experimental & Applied Acarology, 14,
327–341.
Colloff, M. J. (2010). The hyperdiverse oribatid mite genus Scapheremaeus (Acari: Oribatida:
Cymbaeremaeidae) in Australia, with descriptions of new species and consideration of
biogeographical affinities. Zootaxa, 2475, 1–38. 112 species.
Colloff, M. J. (2011). A new genus of oribatid mite, Spineremaeus gen. nov. and three new species
of Scapheremaeus (Acari: Oribatida: Cymbaeremaeidae) from Norfolk Island, South-west
Pacific, and their biogeographical affinities. Zootaxa, 2828, 19–37.
Colloff, M. J., & Cairns, A. (2011). A novel association between oribatid mites and leafy liverworts
(Marchantiophyta: Jungermanniidae), with a description of a new species of Birobates Balogh,
1970 (Acari: Oribatida: Oripodidae). Australian Journal of Entomology, 50, 72–77.
Colloff, M. J., & Cameron, S. L. (2009). Revision of the oribatid mite genus Austronothrus
Hammer (Acari: Oribatida): sexual dimorphism and a re-evaluation of the phylogenetic
328 8 Mites on Plants

relationships of the family Crotoniidae. Invertebrate Systematics, 23, 87–110. doi:10.1071/


IS08032.
Colwell, R. K. (1986). Community biology and sexual selection: Lessons from hummingbird
flower mites. In J. Diamond & T. J. Case (Eds.), Community ecology (pp. 406–424). New York:
Harper and Row.
Colwell, R. K. (1995). Effects of nectar consumption by the hummingbird flower mite Proctolaelaps
kirmsei on nectar availability in Hamelia patens. Biotropica, 27, 206–217.
Colwell, R. K., & Naeem, S. (1994). Life-history patterns of hummingbird flower mites in relation
to host phenology and morphology. In M. A. Houck (Ed.), Mites: Ecological and evolutionary
analyses of life-history patterns (pp. 23–44). New York: Chapman & Hall.
Compton, S. G. (1993). One way to be a fig. African Entomology, 1, 151–158.
Croft, B. A., Blackwood, J. S., & McMurtry, J. A. (2004). Classifying life-style types of phytoseiid
mites: Diagnostic traits. Experimental & Applied Acarology, 33, 247–260.
Cromroy, H. L. (1983). Potential use of mites in biological control of terrestrial & aquatic weeds.
In M. A. Hoy, G. L. Cunningham, & L. Knutson (Eds.), Biological control of pests by mites
(3304th ed., pp. 61–66). Berkeley: University of California Agriculture Experiment Station
Special Publication.
Cronberg, N., Natcheva, R., & Hedlund, K. (2006). Microarthropods mediate sperm transfer in
mosses. Science, 313, 1255.
Crozier, R. H. (1985). Adaptive consequences of male-haploidy. In W. Helle & M. W. Sabelis
(Eds.), Spider mites, their biology, natural enemies and control (1Ath ed., pp. 201–222).
Amsterdam: Elsevier.
Cullen, J. M., & Moore, A. D. (1983). The influence of three populations of Aceria chondrillae on
three forms of Chondrilla juncea. Journal of Applied Ecology, 20, 235–243.
Cullen, J. M., Groves, R. H., et al. (1982). The influence of Aceria chondrillae on the growth and
reproduction capacity of Chondrilla juncea. Journal of Applied Ecology, 19, 529–537.
Davies, J. T., Ireson, J. E., & Allen, G. R. (2007). The impact of the gorse spider mite, Tetranychus
lintearius, on the growth and development of gorse, Ulex europaeus. Biological Control, 41, 86–93.
De Boer, J. G., & Dicke, M. (2006). Olfactory learning by predatory arthropods. Animal Biology,
56, 143–155. doi:10.1163/157075606777304221.
De Moraes, G. J., McMurtry, J. A., Denmark, H. A., & Campos, C. B. (2004). A revised catalog of
the mite family Phytoseiidae. Zootaxa, 434, 1–494.
Dicke, M., & Sabelis, M. (1988). How plants obtain predatory mites as bodyguards. Netherlands
Journal of Zoology, 38, 148–165.
Dicke, M., Sabelis, M. W., Takabayashi, J., Bruin, J., & Posthumus, M. A. (1990). Plant strategies
of manipulating predator–prey interactions through allelochemicals: Prospects for application
in pest control. Journal of Chemical Ecology, 16, 3091–3118.
Dicke, M., Takabayashi, J., Posthumus, M. A., Schutte, C., & Krips, O. E. (1998). Plant-phytoseiid
interactions mediated by herbivore-induced plant volatiles: Variation in production of cues in
responses of predatory mites. Experimental & Applied Acarology, 22, 311–333.
Dobkin, D. S. (1985). Heterogeneity of tropical floral microclimates and the response of hum-
mingbird flower mites. Ecology, 66, 536–543.
Domrow, R. (1979). Ascid and ameroseiid mites phoretic on Australian mammals and birds.
Records of the Western Australian Museum, 8, 97–116.
Duso, C. (1992). Role of Amblyseius aberrans (Oud.), Typhlodromus pyri Scheuten and Amblyseius
andersoni (Chant) (Acari, Phytoseiidae) in vineyards. Journal of Applied Entomology, 114,
455–462.
Duso, C. (1993). Factors affecting the potential of phytoseiid mites (Acari: Phytoseiidae) as
biocontrol agents in North-Italian vineyards. Experimental & Applied Acarology, 17, 241–258.
Eickwort, G. C. (1994). Evolution and life-history patterns of mites associated with bees. In M. A.
Houck (Ed.), Mites: Ecological and evolutionary analyses of life-history patterns (pp. 218–
251). New York: Chapman & Hall.
English-Loeb, G., Norton, A. P., Gadoury, D., Seem, R., & Wilcox, W. (2007). Biological control
of grape powdery mildew using mycophagous mites. Plant Disease, 91, 421–429.
References 329

Erwin, T. (1982). Tropical forests: Their richness in Coleoptera and other arthropod species. The
Coleopterists Bulletin, 36, 74–75.
Evans, G. O. (1963). The genus Neocypholaelaps Vitzthum (Acari: Mesostigmata). Annals and
Magazine of Natural History, 6, 209–230.
Evans, G. O. (1992). Principles of acarology. Wallingford: CAB International.
Evans, G. A., Cromroy, H. L., & Ochoa, R. (1993). The Tenuipalpidae of Honduras (Tenuipalpidae:
Acari). Florida Entomologist, 76, 126–155.
Fan, Q.-H. (2000). The morphology of Xenocaligonellidus smileyi (Acari: Xenocaligonellidae).
In Y.-L. Zhang (Ed.), Systematic and faunistic research on Chinese insects (pp. 290–297).
Beijing: China Agriculture Press.
Fan, Q.-H., & Zhang, Z.-Q. (2005). Raphignathoidea (Acari: Prostigmata). Fauna of New Zealand,
52, 1–400.
Faraji, F., & Cornejo, X. (2006). A new Hattena Domrow (Acari: Ameroseiidae) from Ecuadorian
mangroves and a new generic record for South America. International Journal of Acarology,
32, 287–291.
Faraji, F., Janssen, A., & Sabelis, M. W. (2002a). The benefits of clustering eggs: The role of egg
predation and larval cannibalism in a predatory mite. Oecologia, 131, 20–26.
Faraji, F., Janssen, A., & Sabelis, M. W. (2002b). Oviposition patterns in a predatory mite reduce
the risk of egg predation caused by prey. Ecological Entomology, 27, 660–664.
Ferreira, J. A. M., Eshuis, B., Janssen, A., & Sabelis, M. W. (2008). Domatia reduce larval
cannibalism in predatory mites. Ecological Entomology, 33, 374–379.
Ferreira, J. A. M., Dalyson, F. S. C., Pallini, A., Sabelis, M. W., & Janssen, A. (2011). Leaf
domatia reduce intraguild predation among predatory mites. Ecological Entomology, 36,
435–441.
Flechtmann, C. H. W., & McMurtry, J. A. (1992). Studies on how phytoseiid mites feed on spider
mites and pollen. International Journal of Acarology, 18, 157–162.
Gerson, U. (1972). Mites of the genus Ledermuelleria (Prostigmata: Stigmaeidae) association with
mosses in Canada. Acarologia, 13, 319–342.
Gerson, U. (1985). Webbing. In A. W. Helle & M. W. Sabelis (Eds.), Spider mites, their biology,
natural enemies and control (1st ed., pp. 223–232). Amsterdam: Elsevier.
Gerson, U., & Izraylevich, S. (1997). A review of host utilization by Hemisarcoptes (Acari:
Hemisarcoptidae) parasitic on scale insects. Systematic and Applied Acarology, 2, 33–42.
Gerson, U., & Smiley, R. L. (1990). Acarine biocontrol agents: An illustrated key and manual.
Melbourne: Chapman & Hall.
Grostal, P., & O’Dowd, D. J. (1994). Plants, mites and mutualism: Leaf domatia and the abundance
and reproduction of mites on Viburnum tinus (Caprifoliaceae). Oecologia, 97, 308–315.
Guerra, T. J., Romero, G. Q., & Benson, W. W. (2010). Flower mites decrease nectar availability
in the rain-forest bromeliad Neoregelia johannis. Journal of Tropical Ecology, 26, 373–379.
Guitierrez, J., & Helle, W. (1985). Evolutionary changes in the Tetranychidae. In W. Helle &
M. W. Sabelis (Eds.), Spider mites: Their biology, natural enemies and control (1ath ed.,
pp. 91–116). New York: Elsevier.
Gwiazdowicz, D. J., & Lakomy, P. (2002). Mites (Acari: Gamasida) occurring in fruiting bodies of
Aphyllophorales. Fragmenta Faunistica (Warsaw), 45, 81–89.
Halliday, R. B. (1996). Revision of the Australian Ameroseiidae (Acarina: Mesostigmata).
Invertebrate Taxonomy, 10, 179–201.
Hayes, J. L. (1985). The predator–prey interaction of the mite Balaustium sp. and the pierid
butterfly Colais alexandra. Ecology, 66, 300–303.
Helle, W., & Sabelis, M. W. (Eds.). (1985a). Spider mites: Their biology, natural enemies, and
control (1Ath ed.). New York: Elsevier.
Helle, W., & Sabelis, M. W. (Eds.). (1985b). Spider mites: Their biology, natural enemies, and
control (1Bth ed.). New York: Elsevier.
Hessein, N. A., & Perring, T. M. (1976). Feeding habits of the Tydeidae with evidence of
Homeopronematus anconai (Acari: Tydeidae) predation on Aculops lycopersici (Acari:
Eriophyidae). International Journal of Acarology, 12, 215–221.
330 8 Mites on Plants

Heyneman, A. J., Colwell, R. K., Naeem, S., Dobkin, D., & Hallet, B. (1991). Host plant
discrimination: Experiments with hummingbird flower mites. In P. W. Price, T. M. Lewisohn,
G. W. Fernandes, & W. W. Benson (Eds.), Plant animal interactions: Evolutionary ecology in
tropical and temperate regions (pp. 455–485). New York: Wiley.
Hill, R. L., & Stone, C. (1985). Spider mites as control agents for weeds. In W. Helle &
M. W. Sabelis (Eds.), Spider mites: Their biology, natural enemies and control (1Bth ed.,
pp. 443–448). Amsterdam: Elsevier.
Hislop, R. G., & Jeppson, L. R. (1976). Morphology of the mouthparts of several species of
phytophagous mites. Annals of the Entomological Society of America, 69, 1125–1135.
Ho, C.-C. (1994). A new genus and two new species of Tarsonemidae from Ficus spp. (Acari:
Heterostigmata). International Journal of Acarology, 20, 189–197.
Hofstetter, R. W., Cronin, J. T., Klepzig, K. D., Moser, J. C., & Ayres, M. P. (2006a). Antagonisms,
mutualisms and commensalisms affect outbreak dynamics of the southern pine beetle.
Oecologia, 147, 679–691. doi:10.1007/s00442-005-0312-0.
Hofstetter, R. W., Klepzig, K. D., Moser, J. C., & Ayres, M. P. (2006b). Seasonal dynamics of
mites and fungi and their interaction with southern pine beetle. Environmental Entomology, 35,
22–30. doi:10.1603/0046-225X-35.1.22.
Hofstetter, R. W., Dempsey, T. D., Klepzig, K. D., & Ayres, M. P. (2007). Temperature-dependent
effects on mutualistic, antagonistic, and commensalistic interactions among insects, fungi and
mites. Community Ecology, 8, 47–56. doi:10.1556/ComEc.8. 2007.1.7.
Hoy, M. A. (1985). Recent advances in genetics and genetic improvement of the Phytoseiidae.
Annual Review of Entomology, 30, 345–370.
Hoy, M. A. (1992). Criteria for release of genetically-improved phytoseiids: An examination of the
risks associated with release of biocontrol agents. Experimental & Applied Acarology, 14,
393–416.
Hoy, M. A. (2000). Transgenic arthropods for pest management programs: Risks and realities.
Experimental & Applied Acarology, 24, 463–495.
Hoy, M. A. (2009). The predatory mite Metaseiulus occidentalis: Mitey small and mitey large
genomes. Bioessays, 31, 581–590. doi:10.1002/bies.200800175.
Hoy, M. A., & Smilanick, J. M. (1979). A sex pheromone produced by immature and adult females
of the predatory mite Metaseiulus occidentalis (Acarina: Phytoseiidae). Entomologia
Experimentalis et Applicata, 26, 291–300.
Hunt, G. S. (1996). Description of predominantly arboreal plateremaeoid mites from Eastern
Australia (Acarina: Cryptostigmata: Plateremaeoidea). Records of the Australian Museum,
48, 303–324.
Janzen, D. (1979). How to be a fig. Annual Review of Ecology and Systematics, 10, 13–51.
Janzen, D. H. (1983). Costa Rican natural history (p. 816). Chicago: The University of Chicago
Press.
Jauharlina, J., Lindquist, E. E., Quinnell, R. J., Robertson, H. G., & Compton, S. G. (2012).
Fig wasps as vectors of mites and nematodes. African Entomology, 20, 101–110.
Jeppson, L. R., Keifer, H. H., & Baker, E. W. (1975). Mites injurious to economic plants. Berkeley:
University of California Press.
Kaluz, S., Literak, I., Capek, M., Konecny, A., & Koubek, P. (2011). A new mite species of the
genus Lasioseius (Acarina: Gamasina, Blattisociidae) associated with the flowers of Englerina
lecardii and Chalcomitra senegalensis (Aves: Nectariniidae) in Senegal. International Journal
of Acarology, 37, 511–524.
Kanazawa, M., Sahara, K., & Saito, Y. (2011). Silk threads function as an ‘adhesive cleaner’ for
nest space in a social spider mite. Proceedings of the Royal Society B, 278, 1653–1660.
doi:10.1098/rspb.2010.1761.
Karasawa, S., & Behan-Pelletier, V. (2007). Description of a sexually dimorphic oribatid mite
(Arachnida: Acari: Oribatida) from canopy habitats of the Ryukyu Archipelago, southwestern
Japan. Zoological Science, 24, 1051–1058.
Karban, R., English-Loeb, G., Walker, M. A., & Thaler, J. (1995). Abundance of phytoseiid mites
on Vitis species: Effects of leaf hairs, domatia, prey abundance and plant phylogeny.
Experimental & Applied Acarology, 19, 189–197.
References 331

Karban, R., English-Loeb, G., & Hougen-Eitzman, D. (1997). Mite vaccinations for sustainable
management of spider mites in vineyards. Ecological Applications, 7, 183–193.
Keifer, H. H., Baker, E. W., Kono, T., Delfinado, M., & Styer, W. E. (1982). An illustrated guide to
plant abnormalities caused by Eriophyid mites in North America (USDA Agricultural
Handbook, Vol. 573). Washington, DC: United State Department of Agriculture.
Kinn, D. N. (1971). The life cycle and behaviour of Cercoleipus coelonotus (Acarina:
Mesostigmata): Including a survey of phoretic mite associates of California Scolytidae. UC
Publications in Entomology, 65, 1–66.
Kinn, D. N. (1987). Incidence of pinewood nematode dauer larvae and phoretic mites associated
with long-horned beetles in central Louisiana. Canadian Journal of Forest Research, 17,
187–190.
Kostiainen, T. S., & Hoy, M. A. (1996). The Phytoseiidae as biological control agents of pest mites
and insects (Agricultural Experimentation Monograph, Vol. 17). Gainesville: University of
Florida.
Krantz, G. W. (1978). A manual of acarology. Corvallis: Oregon State University Bookstores.
Krantz, G. W., & Lindquist, E. E. (1979). Evolution of phytophagous mites (Acari). Annual Review
of Entomology, 24, 121–158.
Krantz, G. W., & Walter, D. E. (Eds.), (2009). A manual of acarology (3rd ed). Lubbock: Texas Tech
University Press. 807 pages; 338 b/w illustrations; 60 figures, ISBN 978-0-89672-620-8.
Krisper, G., & Schuster, R. (2008). Fortuynia atlantica sp. nov., a thalassobiontic oribatid mite
from the rocky coast of the Bermuda Islands (Acari: Oribatida: Fortuyniidae). Annales
Zoologici, 58, 419–432. doi:10.3161/000345408X326753.
Krivolutsky, D. A., & Druk, A. (1986). Fossil oribatid mites. Annual Review of Entomology, 31,
533–545.
Kuznetsov, N. N., Khaustov, A. A., & Perkovsky, E. E. (2010). First record of mites of the family
Stigmaeidae (Acari, Raphignathoidea) from Rovno amber with description of a new species of
the genus Mediolata. Vestnik Zoologii, 44, 549–551.
Labandera, C. C., Phillips, T. L., & Norton, R. A. (1997). Oribatid mites and the decomposition of
plant tissues in Paleozoic coal-swamp forests. Palaios, 12, 319–353.
Lara, C., & Ornelas, J. F. (2001). Nectar ‘theft’ by hummingbird flower mites and its consequences
for seed set in Moussonia deppeana. Functional Ecology, 15, 78–84.
Lara, C., & Ornelas, J. F. (2002). Flower mites and nectar production in six hummingbird-polli-
nated plants with contrasting flower longevities. Canadian Journal of Botany, 80, 1216–1229.
doi:10.1139/B02-109.
Lemos, F., Sarmento, R. A., Pallini, A., Dias, C. R., Sabelis, M. W., & Janssen, A. (2010). Spider
mite web mediates anti-predator behaviour. Experimental & Applied Acarology, 52, 1–10.
Lewontin, R. C. (1978). Adaptation. Scientific American, 239(3), 156–169.
Li, J., & Hoy, M. A. (1996). Adaptibility and efficacy of transgenic and wild-type Metaseiulus
occidentalis (Acari: Phytoseiidae) compared as part of a risk assessment. Experimental &
Applied Acarology, 20, 563–573.
Lindo, Z., & Winchester, N. N. (2006). A comparison of microarthropod assemblages with
emphasis on oribatid mites in canopy suspended soils and forest floors associated with ancient
western redcedar trees. Pedobiologia, 50, 31–41.
Lindo, Z., Winchester, N. N., & Didham, R. K. (2008). Nested patterns of community assembly in
the colonisation of artificial canopy habitats by oribatid mites. Oikos, 117, 1856–1864.
doi:10.1111/j.1600-0706.2008.16920.x.
Lindquist, E. E. (1963). A taxonomic review of the genus Hoploseius Berlese (Acarina:
Blattisocidae). Canadian Entomologist, 95, 1175–1185.
Lindquist, E. E. (1965). An unusual new species of the genus Hoploseius Berlese (Acarina:
Blattisociidae) from Mexico. Canadian Entomologist, 97, 1121–1131.
Lindquist, E. E. (1975). Associations between mites and other arthropods in forest floor habitats.
Canadian Entomolologist, 107, 425–437.
Lindquist, E. E. (1985). Discovery of sporothecae in adult female Trochometridium cross, with
notes on analogous structures in Siteroptes Amerling (Acari: Heterostigmata). Experimental &
Applied Acarology, 1, 73–85.
332 8 Mites on Plants

Lindquist, E. E. (1986). The world genera of Tarsonemidae (Acari: Heterostigmata): A


morphological, phylogenetic, and systematic revision, with a reclassification of the family-
group taxa in the Heterostigmata. Memoirs of the Entomological Society of Canada, 136, 1–517.
Lindquist, E. E. (1995). Remarkable convergence between two taxa of ascid mites (Acari:
Mesostigmata) adapted to living in pore tubes of bracket fungi in North America, with
description of Mycolaelaps new genus. Canadian Journal of Zoology, 73, 104–128.
Lindquist, E. E. (1996). 1.1.1 External Anatomy and notation of structures. In E. E. Lindquist,
M. W. Sabelis, & J. Bruin (Eds.), Eriophyid mites. Their biology, natural enemies and control
(Vol. 6, pp. 3–31). Amsterdam: Elsevier.
Lindquist, E. E. (1998). Evolution of phytophagy in trombidiform mites. Experimental & Applied
Acarology, 22, 81–100.
Lindquist, E. E., & Hunter, P. E. (1965). Some mites of the genus Proctolaelaps Berlese (Acarina:
Blattisociidae) associated with forest insect pests. Canadian Entomologist, 97, 16–32.
Lindquist, E. E., & Moraza, M. L. (2008). A new genus of flower-dwelling melicharid mites
(Acari: Mesostigmata: Ascoidea) phoretic on bats and insects in Costa Rica and Brazil.
Zootaxa, 1685, 1–37.
Lindquist, E. E., Sabelis, M. W., & Bruin, J. (1996). Eriophyoid mites, their biology, natural ene-
mies and control. Amsterdam: Elsevier.
Lindquist, E. E., & Wu, K. W. (1991). Review of mites of the genus Mucroseius (Acari:
Mesostigmata: Ascidae) associated with sawyer beetles (Cerambycidae: Monochamus and
Mecynippus) and pine wood nematodes [Aphelenchoididae: Bursaphelenchus xylophilus
(Steiner and Buhrer) Nickle], with descriptions of six new species from Japan and North
America, and notes on their previous misidentification. Canadian Entomologist, 123, 875–927.
Lorenzon, M., Pozzebon, A., & Duso, C. (2012). Effects of potential food sources on biological
and demographic parameters of the predatory mites Kampimodromus aberrans, Typhlodromus
pyri and Amblyseius andersoni. Experimental & Applied Acarology, 58, 259–278. doi:10.1007/
s10493-012-9580-7.
Loughner, R., Wentworth, K., Loeb, G., & Nyrop, J. (2009). Leaf trichomes influence predatory
mite densities through dispersal behaviour. Entomologia Experimentalis et Applicata, 134,
78–88. doi:10.1111/j.1570-7458.2009.00939.x.
Lundström, A. N. (1887). Planzenbiologische Studien. II. Die Anpassungen der Planzen an Thiere.
Nova acta Regiae Societatis scientarum upsaliensis, 13(3), 1–87.
MacRae, I. V., & Croft, B. A. (1996). Differential impact of egg predation by Zetzellia mali (Acari:
Stigmaeidae) on Metaseiulus occidentalis and Typhlodromus pyri (Acari: Phytoseiidae).
Experimental & Applied Acarology, 20, 143–154.
Manson, D. C. M., & Gerson, U. (1996). Web spinning, wax secretion and liquid secretion by
eriophyoid mites. In E. E. Lindquist, M. W. Sabelis, & J. Bruin (Eds.), Eriophyoid mites: Their
biology, natural enemies and control (6th ed., pp. 199–216). Amsterdam: Elsevier.
Marlin, D., Hill, M. P., Ripley, B. S., Strauss, A. J., & Byrne, M. J. (2013). The effect of herbivory
by the mite Orthogalumna terebrantis on the growth and photosynthetic performance of water
hyacinth (Eichhornia crassipes). Aquatic Botany, 104, 60–69. doi:10.1016/j.aquabot.2012.09.005.
Matthewman, W. G., & Pielou, D. P. (1971). Arthropods inhabiting the sporophores of Fomes
fomentarius (Polyporaceae) in Gatineau Park, Quebec. Canadian Entomologist, 103, 775–847.
McClure, M. S. (1995). Diapterobates humeralis (Oribatida: Ceratozetidae): An effective control
agent of hemlock woolly adelgid (Homoptera: Adelgidae) in Japan. Environmental Entomology,
24, 1207–1215.
McGraw, J. R., & Farrier, M. H. (1969). Mites of the superfamily Parasitoidea (Acarina:
Mesostigmata) associated with Dendroctonus and Ips (Coleoptera: Scolytidae). Technical
Bulletin, North Carolina Agricultural Experiment Station, 192, 1–162.
McMurtry, J. A., & Croft, B. A. (1997). Life-styles of phytoseiid mites and their roles in biological
control. Annual Review of Entomology, 42, 291–321.
Meng, R.-X., Sabelis, M. W., & Janssen, A. (2012). Limited predator-induced dispersal in
whiteflies. PLoS One, 7(9), e45487. doi:10.1371/journal.pone.0045487.
References 333

Meyer, M. K. P. S. (1979). The Tenuipalpidae (Acari) of Africa with keys to the world fauna.
Entomology Memoir Department of Agriculture Technical Services Republic of South Africa,
50, 1–135.
Meyer, M. K. P. S., & Ueckermann, E. A. (1997). A review of some species of the families
Allochaetophoridae, Linotetranidae and Tuckerellidae (Acari: Tetranychoidea). International
Journal of Acarology, 23, 67–92.
Momen, F. M. (1987). The mite fauna of an unsprayed apple orchard in Ireland. Zeitschrift fuer
Angewandte Zoologie, 74, 417–434.
Momen, F. M. (2012). Influence of life diet on the biology and demographic parameters of
Agistemus olivi Romeih, a specific predator of eriophyid pest mites (Acari: Stigmaeidae and
Eriophyidae). Tropical Life Sciences, 23, 25–34.
Monks, A., O’Connell, D. M., Lee, W. G., Bannister, J. M., & Dickson, K. J. M. (2009). Benefits
associated with the domatia mediated tritrophic mutualism in the shrub Coprosma lucida.
Oikos, 116, 873–881.
Moraza, M. L., & Lindquist, E. E. (2011). A new genus of fungus-inhabiting blattisociid mites
(Acari: Mesostigmata: Phytoseioidea) from Middle America, with a key to genera and
subgenera of the subfamily Blattisociinae. Zootaxa, 2759, 1–25.
Moser, J. C. (1985). Use of sporothecae by phoretic Tarsonemus mites to transport ascospores of
coniferous bluestain fungi. Transactions of the Brisish Mycological Society, 84, 750–753.
Moser, J. C., & Roton, L. M. (1971). Mites associated with southern pine beetle in Allen Parish,
Louisiana. Canadian Entomologist, 103, 1775–1798.
Moser, J. C., Perry, T. J., & Solheim, H. (1989). Ascospores hyperphoretic on mites associated
with Ips typographus. Mycological Research, 93, 513–517.
Moser, J. C., Konrad, H., Blomquist, S. R., & Kirisits, T. (2010). Do mites phoretic on elm bark
beetles contribute to the transmission of Dutch elm disease? Naturwissenschaften, 97, 219–
227. doi:10.1007/s00114-009-0630-x.
Mound, L. A. (1993). The first thrips species (Insecta) inhabiting leaf domatia: Domatiathrips
cunninghamii gen. et sp. nov. (Thysanoptera: Phlaeothripidae). Journal of the New York
Entomological Society, 101, 424–430.
Murphy, P. W., & Balla, A. N. (1971). The bionomics of Humerobates rostrolamellatus
Grandjean (Cryptostigmata: Ceratozetidae) on fruit trees. In M. Daniel & B. Rosick (Eds.),
Proceedings of the 3rd International Congress of Acarology (pp. 97–104). The Hague: Junk
B.V. Publishers.
Nadkarni, N. M., & Longino, J. T. (1990). Invertebrates in canopy and ground organic matter in a
neotropical montane forest, Costa Rica. Biotropica, 22, 286–289.
Naeem, S., Dobkin, D., & OConnor, B. M. (1985). Lasioseius mites (Acari: Gamasida: Ascidae)
associated with hummingbird-pollinated flowers in Trinidad, West Indies. International
Journal of Entomology, 27, 338–353.
Nanelli, R., & Turchetti, T. (1993). Interazioni acari corticoli – Cryphonectria parasitica: Probabile
ruolo delgi acari nella diffusione degli isolati del parassita (pp. 157–163) in Covassi, M.
(Coord.) M.A. F. Convegno Piante Forestali, Firenze 1992, 1st. Sper. Pat. Veg., Roma.
Naskrecki, P., & Colwell, R. K. (1995). New genus and two new species of Melicharini from
Venezuela (Acari: Mesostigmata, Ascidae). Annals of the Entomological Society of America,
88, 284–293.
Naskrecki, P., & Colwell, R. K. (1998). Systematics and host plant affiliations of hummingbird
flower mites of the genera Tropicoseius Baker and Yunker and Rhinoseius Baker and Yunker
(Acari: Mesostigmata: Ascidae) (Thomas Say Foundation Monographs). Lanham: Entomological
Society of America. ISBN 0-938522-67-1.
Navajas, M. (1998). Host plant associations in the spider mite Tetranychus urticae (Acari:
Tetranychidae): Insights from molecular phylogeography. Experimental and Applied Acarology,
22, 201–204.
Nicolai, V. (1986). The bark of trees: Thermal properties, microclimate and fauna. Oecologia, 69,
148–160.
334 8 Mites on Plants

Nicolai, V. (1989). Thermal properties and fauna on the bark of trees in two different African
ecosystems. Oecologia, 80, 421–430.
Nicolai, V. (1993). The arthropod fauna on the bark of deciduous and coniferous trees in a mixed
forest of the Itasca State Park, MN, USA. Spixiana, 16, 61–69.
Nishida, S., Naiki, A., & Nishida, T. (2005). Morphological variation in leaf domatia enables
coexistence of antagonistic mites in Cinnamomum camphora. Canadian Journal of Botany,
83(1), 93–101.
Norton, R. A., & Palacios-Vargas, J. G. (1982). Nueva Belba (Oribatei: Damaeidae) de musgos
epifitos de Mexico. Folia Entomologica Mexicana, 52, 61–73.
Norton, A. P., English-Loeb, G., Gadoury, D., & Seem, R. C. (2000). Mycophagous mites and foliar
pathogens: Leaf domatia mediate tritrophic interactions in grapes. Ecology, 81, 490–499.
Nucifora, A., & Vacante, V. (1986). Laboratory observations on the biology of Tarsonemus waitei
banks. In R. Cavalloro & E. Di Martin (Eds.), Integrated pest control in citrus-groves, Acireale
26–29 Mar 1985 (pp. 202–207). Rotterdam: A.A. Balkema.
Nuzzaci, G., & Alberti, G. (1996). Internal anatomy and physiology. In E. E. Lindquist, M. W.
Sabelis, & J. Bruin (Eds.), Eriophyoid mites: Their biology, natural enemies and control (6th
ed., pp. 101–167). Amsterdam: Elsevier.
Ochoa, R. (1989). A note on paedogenesis in Tetranychoidea. International Journal of Acarology,
2, 117–118.
Ochoa, R., Aguilar, H., & Vargas, C. (1994). Phytophagous mites of Central America: An
illustrated guide. Turrialba: CATIE.
O’Connell, T., & Bolger, T. (1997). Stability, ephemerality and dispersal ability: Microarthropod
assemblages on fungal sporophores. Biological Journal of the Linnean Society, 62, 111–131.
OConnor, B. M. (1984). Acarine–fungal relationships: The evolution of symbiotic associations. In
Q. Wheeler & M. Blackwell (Eds.), Fungus–insect relationships: Perspectives in ecology and
evolution (pp. 354–381). New York: Columbia University Press.
OConnor, B. M., Colwell, R. K., & Naeem, S. (1991). Flower mites of Trinidad II. The genus
Proctolaelaps (Acari: Ascidae). Great Basin Naturalist, 51, 348–376.
OConnor, B. M., Colwell, R. K., & Naeem, S. (1997). The flower mites of Trinidad III. The genus
Rhinoseius (Acari: Ascidae). Miscellaneous Publications Museum of Zoology University of
Michigan, 184, 1–32.
O’Dowd, D. J., & Pemberton, R. W. (1994). Leaf domatia in Korean plants: Floristics, frequency,
and biogeography. Vegetatio, 114, 137–149.
O’Dowd, D. J., & Willson, M. F. (1989). Leaf domatia and mites on Australian plants: Ecological
and evolutionary implications. Biological Journal of the Linnean Society, 37, 191–236.
O’Dowd, D. J., Brew, C. R., Christophel, D. C., & Norton, R. A. (1991). Mite–plant associations
from the Eocene of southern Australia. Science, 252, 99–101.
Ohmer, C., Fain, A., & Schuchmann, K.-L. (1991). New ascid mites of the genera Rhinoseius
Baker and Yunker, 1964, and Lasioseius Berlese, 1923 (Acari: Gamasida: Ascidae) associated
with hummingbirds or hummingbird-pollinated flowers in southwestern Columbia. Journal of
Natural History, 25, 481–497.
Oldfield, G. N. (1996). Diversity and host plant specificity. In E. E. Lindquist, M. W. Sabelis, &
J. Bruin (Eds.), Eriophyoid mites: Their biology, natural enemies and control (6th ed., pp.
199–216). Amsterdam: Elsevier.
Oldfield, G. N., & Proesler, G. (1996). Eriophyoid mites as vectors of plant pathogens. In E. E.
Lindquist, M. W. Sabelis, & J. Bruin (Eds.), Eriophyoid mites: Their biology, natural enemies
and control (6th ed., pp. 259–275). Amsterdam: Elsevier.
Overmeer, W. P. J., Nelis, H. J. C. F., De Leenheer, A. P., Calis, J. N. M., & Veerman, A. (1989).
Effect of diet on the photoperiodic induction of diapause in three species of predatory mite
Amblyseius potentillae, Amblyseius cucumeris and Typhlodromus pyri. Experimental &
Applied Acarology, 7, 281–288.
Paciorek, C. J., Moyer, B. R., Levin, R. A., & Halpern, S. L. (1995). Pollen consumption by the
hummingbird flower mite Proctolaelaps kirmsei and possible fitness effects on Hamelia patens.
Biotropica, 27, 258–262.
References 335

Palevsky, E., Gerson, U., & Zhang, Z.-Q. (2013). Can exotic phytoseiids be considered ‘benevolent
invaders’ in perennial cropping systems? Experimental & Applied Acarology, 59, 11–26.
Pallini, A., Janssen, A., & Sabelis, M. W. (1999). Spider mites avoid plants with predators.
Experimental & Applied Acarology, 23, 803–815.
Paoletti, M. G., Stinner, B. R., Stinner, D., Benzig, D., & Taylor, R. (1991). Diversity of soil fauna in the
canopy and forest floor of a Venezuelan cloud forest. Journal of Tropical Ecology, 7, 373–383.
Parsons, W. T., & Cuthbertson, E. G. (1992). Noxious weeds of Australia. Melbourne: Inkata Press.
Patankar, R., Beaulieu, F., Smith, S. M., & Thomas, S. C. (2012). The life history of a gall-inducing
mite: Summer phenology, predation and influence of gall morphology in a sugar maple canopy.
Agricultural and Forest Entomology, 14, 251–259.
Pielou, D. P., & Verma, A. N. (1968). The arthropod fauna associated with the birch bracket
fungus, Polyporus betulinus, in eastern Canada. Canadian Entomologist, 100, 1179–1199.
Popov, N. A., & Khudyakova, O. A. (1989). Development of Phytoseiulus persimilis (Acarina:
Phytoseiidae) fed on Tetranychus urticae (Acarina: Tetranychidae) on various food plants. Acta
Entomologica Fennica, 53, 43–46.
Porres, M. A., McMurtry, J. A., & March, R. B. (1975). Investigations of leaf sap feeding by three
species of phytoseiid mites by labelling with radioactive phosphoric acid. Annals of the
Entomological Society of America, 68, 871–872.
Presnail, J. K., Hoy, M. A., & Jeyaprakash, A. (1998). Developing transgenic phytoseiids for
biological control programs, in Acarology IX: Volume 2, Symposia. Columbus: Ohio Biological
Survey.
Price, P. W., Bouton, C. E., Gross, P., McPheron, B. A., Thompson, J. N., & Weis, A. E. (1980).
Interactions among three trophic levels: Influence of plants on the interactions between insect
herbivores and natural enemies. Annual Review of Ecology and Systematics, 11, 41–65.
Proctor, H. C., Montgomery, K. M., Rosen, K. E., et al. (2002). Are tree trunks habitats or
highways? A comparison of oribatid mite assemblages from hoop-pine bark and litter.
Australian Journal of Entomology, 4, 294–299. doi:10.1046/j.1440-6055.2002.00309.x.
Qin, T.-K. (1994). Revision of Australian Penthalodidae (Acarina). Invertebrate Taxonomy, 8,
1305–1323.
Qin, T.-K., & Halliday, R. B. (1997). Eriorhynchidae, a new family of Prostigmata (Acarina), with
a cladistic analysis of eupodoid species of Australia and New Zealand. Systematic Entomology,
22, 151–171.
Quiros-Gonzalez, M. J., & Baker, E. W. (1984). Idiosomal and leg chaetotaxy in the Tuckerellidae
Baker & Pritchard: Ontogeny and nomenclature. In D. A. Griffiths & C. E. Bowman (Eds.),
Acarology VI (1st ed., pp. 166–173). Chichester: Ellis Horwood.
Rasmy, A. H., & Saber, S. A. (2012). Effect of cannibalism on predation, oviposition and
longevity of the predacious mite, Agistemus exsertus Gonzalez (Acari: Stigmaeidae).
Archives of Phytopathology and Plant Protection, 45, 977–985. doi:10.1080/03235408.201
2.655147.
Richards, L. A., & Coley, P. D. (2012). Domatia morphology and mite occupancy of Psychotria
horizontalis (Rubiaceae) across the Isthmus of Panama. Arthropod-Plant Interactions, 6, 129–
136. doi:10.1007/s11829-011-9161-4.
Ridsdill-Smith, T. J. (1997). Biology and control of Halotydeus destructor (Tucker) (Acarina:
Penthaleidae): A review. Experimental & Applied Acarology, 21, 195–224.
Ridsdill-Smith, T. J. (2008). Mass rearing an earth mite to screen plants for resistance: A review.
Entomological Reviews, 38, S22–S27.
Ridsdill-Smith, T. J., Craig, S., & Beaton, C. D. (1996). Microsopic examination of feeding
damage to subterranean clover cotyledons caused by Halotydeus destructor (Tucker)
(Penthaleidae). In R. Mitchell, D. J. Horn, G. R. Needham, & W. C. Welbourn (Eds.), Acarology
IX: Volume 1, proceedings (pp. 55–57). Columbus: Ohio Biological Survey.
Robertson, N. L., & Carroll, T. W. (1988). Virus-like particles and spider mite intimately associated
with a new disease of barley. Science, 240, 1188–1190.
Rock, G. C., Monroe, R. J., & Yergan, D. R. (1976). Demonstration of a sex pheromone in the
predaceous mite Neoseiulus fallacis. Environmental Entomology, 5, 264–266.
336 8 Mites on Plants

Roda, A., Nyrop, J., Dicke, M., & English-Loeb, G. (2001). Trichomes and spider-mite webbing
protect predatory mite eggs from intraguild predation. Oecologia, 125, 428–435.
Romero, G. Q., & Benson, W. W. (2004). Leaf domatia mediate mutualism between mites and a
tropical tree. Oecologia, 140, 609–616. doi:10.1007/s00442-004-1626-z.
Romero, G. Q., & Benson, W. W. (2005). Biotic interactions of mites, plants and leaf domatia.
Current Opinion in Plant Biology, 8, 436–440.
Romero, G. Q., Daud, R. D., Salomão, A. T., Martins, L. F., Feres, R. J. F., & Benson, W. W.
(2011). Mites and leaf domatia: No evidence of mutualism in Coffea arabica plants. Biota
Neotropica, 11(1), 27–34. http://dx.doi.org/10.1590/S1676-06032011000100002
Rosenthal, S. S. (1996). Aceria, Epitrimerus and Aculus species and biological control of weeds.
In E. E. Lindquist, M. W. Sabelis, & J. Bruin (Eds.), Eriophyid mites: Their biology, natural
enemies and control (6th ed., pp. 729–739). Amsterdam: Elsevier.
Rosenstiel, T. N., Shortlidge, E. E., Melnychenko, A. N., Pankow, J. F., & Eppley, S. M. (2012).
Sex-specific volatile compounds influence microarthropod-mediated fertilization of moss.
Nature, 489, 431–433. doi:10.1038/nature11330.
Rowles, A. D., & O’Dowd, D. J. (2009). Leaf domatia and protection of a predatory mite
Typhlodromus doreenae Schicha (Acari: Phytoseiidae) from drying humidity. Australian
Journal of Entomology, 48, 276–281. doi:10.1111/j.1440-6055.2009.00716.
Rozario, S. (1994). Domatia and mites: Effects of leaf morphology on beneficial mites in semi-
natural and managed systems. Ph.D thesis, Monash University, Melbourne, p. 179.
Ryke, P. A. J. (1964). Acarina associated with Protea flowers in the cape Province. Journal of the
Entomological Society of South Africa, 26, 337–354.
Sabelis, M. W., & Bakker, F. M. (1992). How predatory mites cope with the web of their tetranychid
prey: A functional view on dorsal chaetotaxy in the Phytoseiidae. Experimental & Applied
Acarology, 16, 203–225.
Sabelis, M. W., & Janssen, A. (1994). Evolution of life history patterns in the Phytoseiidae. In
M. A. Houck (Ed.), Mites, ecological and evolutionary analyses of life-history patterns
(pp. 70–98). New York: Chapman & Hall.
Sabelis, M. W., Janssen, A., Lesna, I., Aratchige, N. S., Nomikou, M., & van Rijn, P. C. J. (2008).
Developments in the use of predatory mites for biological pest control. In Enkegaard A (Ed)
Bulletin OILB/SROP (Vol. 32, pp. 187–199). International Organization for Biological and
Integrated Control of Noxious Animals and Plants, West Palearctic Regional Section (IOBC/
WPRS): Daarmstadt.
Saito, Y. (1983). The concept of ‘life types’ in Tetranychidae – An attempt to classify the spinning
behaviour of Tetranychidae. Acarologia, 24, 377–391.
Saito, Y. (1985). Life types of spider mites. In W. Helle & M. W. Sabelis (Eds.), Spider mites:
Their biology, natural enemies and control (1ath ed., pp. 253–264). New York: Elsevier.
Saito, Y. (1986a). Prey kills predator: Counter-attack success of a spider mite against its specific
phytoseiid predator. Experimental & Applied Acarology, 2, 47–62.
Saito, Y. (1986b). Biparental defence in a spider mite (Acari: Tetranychidae) infesting Sasa
bamboo. Behavioural Ecology and Sociobiology, 18, 377–386.
Saito, Y. (2010). Plant mite and sociality – Diversity and evolution (p. 187). Tokyo: Springer.
Saito, Y., Chittenden, A. R., & Kanazawa, M. (2011). Counterattack success of a social spider mite
against two predominant phytoseiid predator species. Experimental & Applied Acarology, 55,
249–258.
Sánchez-Núñez, D. A., & Mancera-Pineda, J. E. (2012). Pollination and fruit set in the main
neotropical mangrove species from the Southwestern Caribbean. Aquatic Botany, 103, 60–65.
Santos, M. A. (1991). Searching behaviour and associational response by Zetzellia mali (Acarina:
Stigmaeidae). Experimental & Applied Acarology, 11, 81–87.
Sarmento, R. A., Lemos, F., Bleeker, P. M., Schuurink, R. C., Pallini, A., Oliveira, M. G. A., Lima,
E. R., Merjin, K., Sabelis, M. W., & Janssen, A. (2011). A herbivore that manipulates plant
defence. Ecology Letters, 14, 229–236. doi:10.1111/j.1461-0248.2010.01575.x.
Schmidt, A. R., Jancke, S., Lindquist, E. E., Ragazzi, E., Roghi, G., Nascimbene, P. C., Schmidt,
K., Wappler, T., & Grimaldi, D. A. (2012). Arthropods in amber from the Triassic period. PNAS
109: 14796–14801. www.pnas.org/cgi/doi/10.1073/pnas.1208464109
References 337

Scoble, J., & Clarke, M. F. (2006). Nectar availability and flower choice by eastern spinebills
foraging on mountain correa. Animal Behaviour, 72, 1387–1394. doi:10.1016/j.anbehav.
2006.03.024.
Seelmann, L., Auer, A., Hoffmann, D., & Schausberger, P. (2007). Leaf pubescence mediates
intraguild predation between predatory mites. Oikos, 116, 807–817.
Seeman, O. D. (1996). Flower mites and phoresy: The biology of Hattena panopla Domrow and
Hattena cometis Domrow. Australian Journal of Zoology, 44, 193–203.
Seeman, O. D., & Walter, D. E. (1995). Life history of Afrocypholaelaps africana (Evans)
(Parasitiformes: Ameroseiidae), a mite inhabiting mangrove flowers and phoretic on honeybees.
Journal of the Australian Entomological Society, 34, 45–50.
Selden, P. A. (1993). Arthropoda (Aglaspidida, Pycnogonida and Chelicerata). In M. J. Benton
(Ed.), The fossil record (2nd ed., pp. 297–320). New York: Chapman & Hall.
Shimoda, T., Kishimoto, H., Takabayashi, J., Amano, H., & Dicke, M. (2009). Comparison of
thread-cutting behavior in three specialist predatory mites to cope with complex webs of
Tetranychus spider mites. Experimental & Applied Acarology, 47, 111–120. doi:10.1007/
s10493-008-9205-3.
Skoracka, A., & Dabert, M. (2009). The cereal rust mite Abacarus hystrix (Acari: Eriophyoidea) is
a complex of species: Evidence from mitochondrial and nuclear DNA sequences. Bulletin of
Entomological Research, 100(3), 263–272. doi:10.1017/S0007485309990216.
Skoracka, A., Oldfield, G., Smith, L., Cristofaro, M., Ochoa, R., & Amrine, J. W., Jr. (2010). Host
plant specificity and specialization in eriophyoid mites. Experimental & Applied Acarology, 51,
93–113.
Smith, L., de Lillo, E., & Amrine, J. W., Jr. (2010). Effectiveness of eriophyid mites for biological
control of weedy plants and challenges for future research. Experimental & Applied Acarology,
51, 115–149.
Sobhian, R., & Andres, L. A. (1978). The response of the skeletonweed gall midge, Cystiphora
schmidti (Diptera: Cecidomyiidae), and gall mite, Aceria chondrillae (Eriophyidae) to North
American strains of rush skeletonweed (Chondrilla juncea). Environmental Entomology, 7,
506–508.
Sorensen, J. T., Kinn, D. N., Doutt, R. L., & Cates, J. R. (1976). Biology of the mite Anystis agilis
(Acari: Anystidae), a California vineyard predator. Annals of the Entomological Society of
America, 65, 905–910.
Strodl, M. A., & Schausberger, P. (2012). Social familiarity modulates group living and foraging
behaviour of juvenile predatory mites. Naturwissenschaften, 99, 303–311. doi:10.1007/
s00114-012-0903-7.
Stubbs, C. S. (1995). Dispersal of soredia by the oribatid mite, Humerobates arborea. Mycologia,
87, 454–458.
Suski, Z. W. (1967). Tarsonemid mites on apple trees in Poland. IX. Tarsonemus pauperoseatus
n.sp. (Acarina: Heterostigmata). Bulletin de l’Académie polonaise des sciences. Série des
sciences biologiques, 15, 267–272.
Suski, Z. W. (1972a). Tarsonemid mites on apple trees in Poland. X. Laboratory studies on the
biology of certain mite species in the family Tarsonemidae (Acarina, Heterostigmata). Zeszyty
Problemowe Postępów Nauk Rolniczych, 129, 111–137.
Suski, Z. W. (1972b). Tarsonemid mites on apple trees in Poland. XI. Field observations on the
distribution and significance of Tarsonemidae (Acarina, Heterostigmata) in apple orchards.
Zeszyty Problemowe Postępów Nauk Rolniczych, 129, 139–157.
Takabayashi, J., & Dicke, M. (1996). Plant–carnivore mutualism through herbivore-induced
carnivore attractants. Trends in Plant Science, 1, 109–113.
Tashiro, H. (1987). Turfgrass insects of the United States and Canada. Ithaca: Cornell University
Press.
Treat, A. E. (1975). Mites of moths and butterflies. Ithaca: Cornell University Press.
Tschapka, M., & Cunningham, S. A. (2004). Flower mites of Calyptrogyne ghiesbreghtiana
(Arecaceae): Evidence for dispersal using pollinating bats. Biotropica, 36, 377–381.
Van, W. P., Zon, A. Q., Overmeer, J., & Veerman, A. (1981). Carotenoids function in photoperiodic
induction of diapause in a predacious mite. Science, 213, 1131–1133.
338 8 Mites on Plants

Veerman, A. (1992). Diapause in phytoseiid mites: A review. Experimental & Applied Acarology,
14, 1–60.
Velazquez, T., & Ornelas, J. F. (2010). Pollen consumption by flower mites in three hummingbird-
pollinated plant species. Experimental & Applied Acarology, 50, 97–105. doi:10.1007/
s10493-009-9309-4.
Vet, L. E. M., & Dicke, M. (1992). Ecology of infochemical use by natural enemies in a tritrophic
context. Annual Review of Entomology, 37, 141–172.
Walter, D. E. (1992). Leaf surface structure and the distribution of Phytoseius mites (Acarina:
Phytoseiidae) in south-east Australian forests. Australian Journal of Zoology, 40, 593–603.
Walter, D. E. (1996). Living on leaves: Mites, tomenta, and leaf domatia. Annual Review of
Entomology, 41, 101–114.
Walter, D. E. (1998). Hoploseius australianus, sp. nov. (Acari: Mesostigmata: Ascidae), a unique
element in the Australian acarofauna. The Australian Entomologist, 25, 69–74.
Walter, D. E. (2004). Hidden in plain sight: Mites in the canopy. In M. D. Lowman & H. B. Rinker
(Eds.), Forest canopies (2nd ed., pp. 224–241). Amsterdam: Elsevier Academic Press.
Walter, D. E., & Beard, J. J. (1997). A review of the Australian Phytoseiinae (Acari: Mesostigmata:
Phytoseiidae). Invertebrate Taxonomy, 11, 823–860.
Walter, D. E., & Behan-Pelletier, V. (1993). Systematics and ecology of Adhaesozetes polyphyllos
sp. nov. (Acari: Oribatida: Licneremaeoidea), a leaf-inhabiting mite from Australian rainforests.
Canadian Journal of Zoology, 71, 1024–1040.
Walter, D. E., & Behan-Pelletier, V. (1999). Mites in forest canopies: Filling the size distribution
shortfall? Annual Review of Entomology, 44, 1–19.
Walter, D. E., & Denmark, H. A. (1991). Use of leaf domatia on wild grape (Vitis munsoniana) by
arthropods in central Florida. Florida Entomologist, 74, 440–446.
Walter, D. E., & Kaplan, D. T. (1990). Feeding observations on two astigmatic mites, Schwiebea
rocketti Woodring (Acaridae) and Histiostoma bakeri Hughes & Jackson, associated with citrus
feeder roots. Pedobiologia, 34, 281–286.
Walter, D. E., & Lindquist, E. E. (1997). Australian species of Lasioseius (Acari: Mesostigmata:
Ascidae): The porulosus group and other species from rainforest canopies. Invertebrate
Taxonomy, 11, 525–547.
Walter, D. E., & O’Dowd, D. J. (1992a). Leaves with domatia have more mites. Ecology, 73,
1514–1518.
Walter, D. E., & O’Dowd, D. J. (1992b). Leaf morphology and predators: Effect of domatia on the
distribution of phytoseiid mites (Acari: Phytoseiidae). Environmental Entomology, 21, 478–484.
Walter, D. E., & O’Dowd, D. J. (1995a). Beneath biodiversity: Factors influencing the diversity
and abundance of canopy mites. Selbyana, 16, 12–20.
Walter, D. E., & O’Dowd, D. J. (1995b). Life on the forest phylloplane: Hairs, little houses, and
myriad mites. In M. D. Lowman & N. M. Nadkarni (Eds.), Forest canopies (pp. 325–351).
Sydney: Academic Press.
Walter, D. E., & Proctor, H. C. (1998). Predatory mites in tropical Australia: Local species richness
and complementarity. Biotropica, 30, 72–81.
Walter, D. E., & Proctor, H. C. (1999). Mites: Ecology, evolution and behaviour (p. 322).
Wallingford: University of NSW Press, Sydney and CABI. ISBN 0 86840 529 9.
Walter, D. E., Hudgens, R. A., & Freckman, D. W. (1986). Consumption of nematodes by
fungivorous mites Tyrophagus spp. (Acarina: Astigmata: Acaridae). Oecologia, 70, 357–361.
Walter, D. E., Halliday, R. B., & Lindquist, E. E. (1993). A review of the genus Asca (Acarina:
Ascidae) in Australia, with the description of three new leaf-inhabiting species. Invertebrate
Taxonomy, 7, 1327–1347.
Walter, D. E., O’Dowd, D. J., & Barnes, V. (1994). The forgotten arthropods: Foliar mites in the
forest canopy. Memoirs of the Queensland Museum, 36, 221–226.
Walter, D. E., Seeman, O., Rodgers, D., & Kitching, R. L. (1998). Mites in the mist: How unique
is a rainforest canopy knockdown fauna? Australian Journal of Ecology, 23, 501–508.
Walter, D. E., Lindquist, E. E., Smith, I. M., Cook, D. R., & Krantz, G. M. (2009). Order
Trombidiformes. In G. W. Krantz & D. E. Walter (Eds.), A manual of acarology (3rd ed., pp.
233–420). Lubbock: Texas Tech University Press.
References 339

Walzer, A., & Schausberger, P. (2010). Threat-sensitive anti-intraguild predation behaviour:


Maternal strategies to reduce offspring predation risk in mites. Animal Behaviour, 81,
177–184.
Weber, M. G., Clement, W. L., Donoghue, M. J., & Agrawal, A. A. (2012). Phylogenetic and
experimental tests of interactions among mutualistic plant defense traits in Viburnum
(Adoxaceae). American Naturalist, 180, 450–463.
Weeks, A. R., Fripp, Y. J., & Hoffmann, A. A. (1995). Genetic structure of Halotydeus destructor
and Penthaleus major populations in Victoria (Acari: Penthaleidae). Experimental & Applied
Acarology, 19, 633–646.
Westphall, E., & Manson, D. C. M. (1996). Feeding effects on host plants: Gall formation and
other distortions. In E. E. Lindquist, M. W. Sabelis, & J. Bruin (Eds.), Eriophyoid mites: Their
biology, natural enemies and control (pp. 251–258). Amsterdam: Elsevier.
Wiggers, M. S., Pratt, P. D., Tipping, P. W., Welbourn, C., & Cuda, J. P. (2005). Within-plant
distribution and diversity of mites associated with the invasive plant Schinus terebinthifolius
(Sapindales: Anacardiaceae) in Florida. Environmental Entomology, 34, 953–962.
Willson, M. F. (1991). Foliar mites in the Eastern Deciduous forest. American Midland Naturalist,
126, 111–117.
Yano, S., & Osakabe, M. (2009). Do spider mite-infested plants and spider mite trails attract
predatory mites? Ecological Research, 24, 1173–1178. doi:10.1007/s11284-009-0598-1.
Yoshida, T., & Hijii, N. (2011). Microarthropod colonization of litter in arboreal and soil
environments of a Japanese cedar (Cryptomeria japonica) plantation. Journal of Forest
Research, 16, 46–54. doi:10.1007/s10310-010-0205-x.
Young, M. R., Behan-Pelletier, V. M., & Hebert, P. D. N. (2012). Revealing the hyperdiverse mite
fauna of Subarctic Canada through DNA barcoding. PLoS One, 7, e48755. doi:10.1371/journal.
pone.0048755.
Zacharda, M., & Krivolutsky, D. A. (1985). Prostigmatic mites (Acarina: Prostigmata) from the
Upper Cretaceous and Paleocene amber of the USSR. Vestnik Ceskoslovenske Spolecnosti
Zoologicke, 49, 147–152.
Zemek, R., & Prenerova, E. (1997). Powdery mildew (Ascomycotina: Erysiphales) – An alternative
food for the predatory mite Typhlodromus pyri Scheuten (Acari: Phytoseiidae). Experimental
& Applied Acarology, 21, 405–414.
Zhang, Z.-Q. (1995). Variance and covariance of ovipositional rates and developmental rates in the
Phytoseiidae (Acari: Mesostigmata): A phylogenetic consideration. Experimental & Applied
Acarology, 19, 139–146.
Zhao, S., & Amrine, J. W., Jr. (1997). Investigation of snowborne mites (Acari) and relevancy to
dispersal. International Journal of Acarology, 23, 209–213.
Chapter 9
Animals as Habitats

To most non-biologists, a ‘mite’ is an almost invisible creature that lives in the


carpet and gives you asthma or that burrows in your skin to produce socially
unacceptable scabies. Cat owners may curse the ear mites that infest their pets
and those who know that ticks are mites may mention Lyme Disease. Such
medically important associations between mites and mammals are discussed in
most parasitology texts, and mites associated with human diseases are dis-
cussed in our Chap. 10; therefore, in this chapter we cover ticks and other
human- and livestock-associated mites superficially, concentrating instead on
lesser known relationships between mites and the animals they use for room and
board. These associations are not always negative; in fact, many seemingly par-
asitic mites have no impact on their hosts or may even be beneficial (see section
“Mutualism”, below).
One of the recurring themes of this book is that mites live in the most astonishing
places. This is particularly apparent among animal-associated acarines. A complete
list of associations between mites and other animals would occupy an entire book
but to whet the reader’s appetite we have provided Table 9.1, where we list the
associations that acarologists use to shock jaded students or to start conversations at
dull parties. After perusing this list many will be compelled to exclaim, ‘What’s a
mite doing there?’ or ‘How did a mite get there?’. The first question is a proximate
one – what is the ecology of the symbiotic mite? – and the second an ultimate
one – what evolutionary history and selective pressures have resulted in this asso-
ciation? In the first section of this chapter we discuss the ecological terminology
dealing with symbiosis and then outline some major hypotheses about how different
symbioses may evolve. The next two sections discuss the life histories and behaviour
of mites associated with invertebrates and vertebrates respectively and the fourth
section surveys the diverse effects that parasitic mites have on their hosts. Finally,
the last section addresses the topic of coevolution between hosts and symbionts
from an acarine point of view.

D.E. Walter and H.C. Proctor, Mites: Ecology, Evolution & Behaviour: Life at a Microscale, 341
DOI 10.1007/978-94-007-7164-2_9, © Springer Science+Business Media Dordrecht 2013
342 9 Animals as Habitats

Types of Ecological Interactions

One way to categorise the diversity of interactions between pairs of individuals is by


the effect each has on the other’s fitness. Table 9.2 depicts the categorisation we will
follow. Neutralism is a situation in which two interactors have no effect on each
other, positive or negative. Competition negatively affects both individuals, while
mutualism increases the fitness of both interactors. Amensalism, although seldom
discussed, is a common phenomenon in which one individual is disadvantaged by
another that is unaffected by the interaction (e.g. a mite is stepped on by an elephant).
Commensalism occurs when one individual benefits by associating with another, the
host, and the host neither gains nor loses. Another ecological category, facilitation,
involves one species benefitting from the act of another, but need not involve a
physically proximate relationship between individuals (e.g., bumblebees may build
their nests in abandoned mouse holes).
The final three categories, parasitism, parasitoidism and predation, cannot be
distinguished by differences in the effects that one individual has on the other.
In each case, one individual gains from the interaction (the parasite, parasitoid
or predator) and the other loses (the host or prey). The difference between the
interactions is one of degree. A parasite (Fig. 9.1) does not have to kill its host in
the process of achieving its own fitness. Parasitoids and predators invariably kill to
achieve their ends but a parasitoid kills only one host in its lifetime (usually slowly)
and a predator kills more than one prey. However, the organisms we place in
these pigeonholes do not always follow the rules. A water mite larva may act as a
parasite in one situation but in another, for example on a weaker or smaller host, as
a parasitoid. Likewise, Poecilochirus mites have no effect on their burying beetle

Fig. 9.1 Top: a Lestes sp. damselfly bearing a cluster of parasitic water mites on its thorax; bottom:
a gecko (Hemidactylus frenatus) with parasitic pterygosomatid mites between its toes (arrows)
(Images by HC Proctor and DE Walter)
Table 9.1 The 22 most astonishing mite–animal associations according to D.A. Crossley, M.L. Goff, B. Halliday, K. Kamerik, A. Kinnear, L. Lundqvist,
P. Martinez, R. Ochoa, B. OConnor, P. Naskrecki, G. Needham, H. Proctor, D. Sammataro, H. Schatz, H. Schwarz, and D. Walter during a survey on the
Acarology Listserver
Mite Site and host Nature of association References
Acarapis woodi (Tarsonemidae) In the tracheae of honey bees Parasitic on host haemolymph Sammataro and
Needham (1996)
Antennequesoma On the terminal antennal segments Steals food? Elzinga (1982)
(Trachyuropodidae) of army ants
Various genera of Cheyletidae In the feather shafts of birds Mutualistic – prey on parasitic mites Atyeo et al. (1984)
inside feather
Types of Ecological Interactions

Cloacaridae In the cloacal mucosa of turtles Parasitic on host blood and mucosal Camin et al. (1967)
tissue?
Coreitarsonemus (Tarsonemidae) In the ‘stink’ glands of bugs (Hemiptera) Parasitic? Krantz (1978)
Demodex folliculorum, D. brevis In facial pores and eyelash and eyebrow Commensal (only occasionally causing Krantz (1978)
(Demodicidae) follicles in humans pimples or dermatitis)
Dicrocheles (Dermanyssidae) In the ears of noctuid moths Feed on haemolymph; avoid deafening Treat (1975)
host by occupying only one ear,
thereby allowing moth to hear bats
Enterohalacarus minutipalpis In the digestive system of sea urchins Parasitic? Bartsch (1987)
(Halacaridae)
Entonyssinae (Laelapidae) In the lungs of snakes Parasitic Domrow (1987)
Gastronyssus bakeri In the stomach mucosa of fruit-eating bats Parasitic? Fain and Hyland (1985)
(Gastronyssidae)
Hypodectes propus (Hypoderidae) In the subcutaneous fat of pigeons Parasitically absorb nutrients from Fain (1969)
host fat through their integument
(hypopi have no mouths)
Kennethiella trisetosa In the acarinaria (‘mite pockets’) of wasps Venerally transmitted parasites Cowan (1985)
(Winterschmidtiidae) (and mutualists?) of wasp larvae
Larvamima (Larvamimidae) In the brood chambers of army ants Mimics ant larvae; may prey on larvae Elzinga (1993)
when ants are not looking
(continued)
343
Table 9.1 (continued)
344

Mite Site and host Nature of association References


Macrocheles lukoschusi Near the anus of sloths Hypothesized to feed on parasitic Krantz (1983)
(Macrochelidae) nematodes associated with the host’s
dung or rectum
Macrocheles rettenmeyeri Attached to the ventral side of army ant feet Parasitic on host blood? minimizes Gotwald (1996)
(Macrochelidae) ambulatory impairment by functioning
as the ant’s foot
Opsonyssus (Gastronyssidae) On the eyeballs (corneas and orbits) Parasitic? Fain (1969)
of fruit bats
Orthohalarachne (Halarachnidae) In the nasal passages and lungs of seals Parasitic on blood and mucosal tissue Kim (1985a)
of hosts
Paraspinturnix globosus Inside the anus of hibernating Myotis bats Parasitic, or just overwintering? Radovsky (1985)
(Spinturnicidae)
Rhinoseius (Ascidae) In the nostrils of hummingbirds Use the birds as transport between Colwell and Naeem
flowers, their true sources of food (1994)
Ricardoella limacum (Erynetidae) In the pulmonary chambers of terrestrial Parasitic Baker (1970)
slugs
Trochometridium In the brood chambers of alkali bees Kill the bee eggs and consume the fungi Kaliszewski et al. (1995)
(Trochometridiidae) that grow on them and the provisions
in their brood cells
Vatacarus ipoides (Trombiculidae) In the tracheal passages of sea snakes Parasitic Nadchatram (2006)
Note: A question mark indicates that the exact nature of the association is unknown
9 Animals as Habitats
Types of Ecological Interactions 345

Table 9.2 Types of interactions between two individuals, categorised by the


effects on fitness when they are interacting compared to when they are not
Effect on fitness of A and B
When not interacting When interacting
A B A B
Neutralism 0 0 0 0
Competition 0 0 − −
Proto-cooperation 0 0 + +
Mutualism − − + +
Amensalism 0 0 − 0
Commensalism − 0a + 0
Parasitism − 0a + −
Parasitoidism − 0a + −
Predation − 0b + −
Note: Table modified from Cheng (1991)
a
B is host
b
B is prey

hosts in some circumstances ( = commensalism) but increase their host’s fitness in


others ( = mutualism). In this chapter we will discuss commensalism, mutualism,
parasitism and parasitoidism between mites and other animals.
Another way to categorise interactions is by the location of the two individuals.
Symbiosis occurs when one individual lives in or on another. It may be a highly
integrated relationship, such as that between tree roots and mycorrhizal fungi, or
loose, as in clownfish living among the tentacles of sea anemones. Although some-
times used as if synonymous with mutualism, symbiosis has no necessary connotation
of being beneficial to both parties. The prefix ‘ecto-’ is used for a symbiont that lives
on the outside of its host (e.g. ectocommensal) and ‘endo-’ for one that lives inside.
Phoresy is a particular type of symbiosis that is defined by the function of the
interaction – to allow a smaller individual (the phoretic) to move from one place to
another on a larger (the host or carrier). Despite the apparent simplicity of this
relationship there is controversy about what constitutes phoresy (see section
“Phoresy and Dispersal”). A nidicole is an animal that lives in the nest or den of
another (e.g. dust mites in human habitations) and an inquiline is a particular type
of nidicole that dwells in the nests of social insects.

Evolutionary Pathways Between Interactions

The nature of ecological relationships between taxa is fluid over time. For example,
within the lifespan of our own species the parasitic relationship between humans and
the honey bee Apis mellifera has become primarily mutualistic. We benefit from the
bees’ production of honey, wax, and royal and from their services as pollinators; the
bees benefit from our protection, custom-built housing and worldwide distribution.
346 9 Animals as Habitats

When lineages start down the road of symbiosis, specialized behaviours may
evolve in the more derived taxa, but the older, more general types of interactions do
not necessarily disappear. For example, some derived species in the Pyroglyphidae
(Astigmata) are parasitic on the skin of birds, while others retain the ancestral
relationship of being detritivorous in birds’ nests (Bochkov and Mironov 2011, but
see Klimov and OConnor 2013). In many symbiotic taxa, however, free-living
relatives are either extinct or unknown (e.g. for lice, fleas, cestodes). In contrast,
non-parasitic relatives of parasitic mites are common (e.g. Treat 1975; Radovsky
1985; Houck and OConnor 1991). Closely related taxa that exhibit a variety of eco-
logical relationships allow the testing of hypotheses about the direction of evolution
(e.g. ‘phoresy leads to parasitism’) and about forces behind such changes.
Water mites, in which related parasitic and non-parasitic species often co-occur
(Smith 1998), are a particularly good group for testing such hypotheses. Some other
taxa also show evidence of repeated gains or losses of parasitism (e.g. leeches,
Siddall and Burreson 1995; copepods, Poulin 1995), but mites provide the most
numerous and clearest examples of switches in lifestyle. It is therefore strange that
acarologists have produced few phylogenetically based comparative tests of the
forces selecting for and maintaining parasitism. Evolutionary narratives are common
(e.g. Houck and OConnor 1991) but phylogenetic trees demonstrating that purported
causative factors did indeed precede hypothesised results are rare. So some of
the following scenarios are what some scoffers might refer to as adaptationist
story-telling (Gould and Lewontin 1978) but they are actually hypotheses waiting to
be tested. It is also important to note that these categories are not mutually exclusive
but simply different ways of subdividing the literature into themes.

Cohabitation Predisposes Mites to Parasitism

Waage (1979) suggested two types of evolutionary routes to parasitism. In Type A,


adaptations for association between parasite and host preceded adaptations for
feeding. In Type F routes, adaptations that would eventually allow feeding on the host
preceded adaptations for association. Parasites that evolved through the A pathway
typically have free-living nidicolous relatives and clearly have secondarily evolved
mouthparts adapted to feeding on host products (e.g. blood). Followers of the Type F
pathway had the right mouthparts in the first place but only secondarily developed
an association with the host.
Many authors describe lineages of parasitic mites evolving from nidicolous
ancestors (Type A route). Moss (1979) presumes that the ancestral harpyrhynchid
evolved from cheyletid-like nest-dwelling predators. Kim (1985a) suggests that the
ancestors of mesostigmatans parasitic on mammals and of argasid ticks were origi-
nally nidicoles that fed on wastes and remnants of the host’s meals. Fain and Hyland
(1985) suggest that parasitic psoroptoids (in mammals) and analgoids (in birds)
probably arose from nidicolous Pyroglyphidae. Bochkov and Mironov (2011) pro-
vide one of the few phylogeny-based explorations of the origin of host-associations
Types of Ecological Interactions 347

of the Psoroptoidea, and argue that obligatory symbioses arose repeatedly from
nidicolous ancestors. However, their scenario maps out on their tree in a way that
also allows for the alternative hypothesis of a single origin of symbiosis followed by
one or more reversals to nidicoly in the Pyroglyphidae. Indeed, in an impressive
multigene phylogenetic analysis of several 100 species of Psoroptidia and related
outgroups, Klimov and OConnor (2013) found that the nidicolous habits of dust
mites were likely derived once or twice from an ancestral lifestyle of living on a
vertebrate host.
Unionicolid water mites may have followed a different Type A route. Water
mites, like all members of the Parasitengonina, have two calyptostatic instars
(protonymph and tritonymph) that act as immobile ‘pupal’ stages between the active
larva, deutonymph and adult. In the genus Unionicola, the calyptostatic instars are
spent embedded in the well-aerated tissues of other animals, primarily sponges
and mussels. For species associated with sponges, the relationship goes no further
than housing during these transformation stages. Some mussel-associated species
(termed ‘temporary mussel mites’ by Hevers 1980) also have this relatively benign
interaction, and deutonymphs and adults are free-living predators in the water
column. But other Unionicola species are parasitic on mussel tissues in their
post-larval stages (‘permanent mussel mites’ sensu Hevers 1980). Perhaps further
study on the ecology of these mites will demonstrate an additional, transitional
type of Unionicola that dwells in mussels but feeds on zooplankton drawn in by
the host.

Dietary Adaptations Precede Parasitism

What of Waage’s (1979) Type F routes? Piercing-sucking feeding and physogastric


development of young may have set the stage for parasitoidism in the Heterostigmatina
(Kaliszewski et al. 1995; see section “Parasitism and Parasitoidism” below).
Radovsky (1985) felt that mesostigmatic mites are morphologically preadapted for
parasitism and that the switch from a free-living to a parasitic lifestyle is often
accompanied by only slight physical changes. The typical chelicerae in free-living
Mesostigmata have musculature that can be used for penetration, are very close to
the shape of a piercing stylet and require little modification to perform like the
mouthparts of blood-sucking insects. However, unlike classical Type F examples,
such as the evolution of blood sucking by nectar-sucking mosquitoes, these meso-
stigmatans likely went through a period of close physical association with their
eventual hosts.
The Parasitengonina (formerly Parasitengona) parasitise both vertebrates and
invertebrates and likely evolved from ancestors that had free-living predatory larvae
(Wiggins et al. 1980). Mite larvae feeding on large prey that remained motile and
were not necessarily killed by the feeding activity could have made the transition
from a predatory to a parasitic relationship. Wiggins et al. (1980) suggest that the
advantages of transport conferred upon such larvae would select for parasitism.
348 9 Animals as Habitats

Phoresy Precedes Parasitism

Phoresy appears to have favoured evolution towards parasitism in several lineages


of mites (Athias-Binche 1991, 1995). Treat (1975) considers that many instances
of parasitism in mesostigmatans associated with Lepidoptera arose from phoretic
associations. Phoresy in these moth mites may have originated as accidental
stumbling upon by the mites, which benefited from hitching a ride on a larger and
more vagile animal. Some mites combine phoresy with occasional parasitism.
Treat quips (p. 89) “a mite, normally predatory, attaches itself to an insect as a
phoretic…. During its ride it taps its host for a light snack. Is it on its way to a career
of crime? Perhaps so. Even among moth mites it is not always easy to tell a hitchhiker
from a hijacker.”
Treat felt that the transition from phoresy to parasitism is particularly apparent
in Blattisocius (Mesostigmata: Blattisociidae). The hypothetical ancestral type
can be represented by Blattisocius dentriticus and B. keegani, which have been
collected from nests of pack rats, various plants and animals, and from the tympanic
recesses of moths. Closer to parasitism is Blattisocius tarsalis, which eats the eggs
of several species of moths. Female mites are phoretic on the adult moths. Treat
found that in most cases, there was no evidence that B. tarsalis fed upon its host;
however, he sometimes observed scars on the moths and mites apparently in the act
of feeding. Finally there is Blattisocius patagiorum, in which nymphs and adults
of both sexes occur on moths, feed on haemolymph and can reproduce without
any other food.
When present, the heteromorphic deutonymph of the Astigmata is typically
a non-feeding phoretic stage but a few have become parasitic (OConnor 1982).
Athias-Binche (1991) suggests that endofollicular hypopi (e.g. in the Hypod-
eroidea), which appear to gain weight while in the follicle, evolved from purely
phoretic hypopi.
Houck and OConnor (1991) and Houck (1994) describe how deutonymphs of
Hemisarcoptes cooremani (Hemisarcoptidae) appear to be in the midst of a transition
from phoresy to parasitism. In all members of this genus, free-living stages are
predators of diaspidid scale insects and deutonymphs are transported by lady beetles
in the genus Chilocorus (Coccinellidae). Deutonymphs of H. cooremani swell up
while on their host Chilocorus cacti and cannot develop to tritonymphs if they are
not allowed this association. Deutonymphs stay on the hosts for 5–21 days, which the
authors felt could not be explained by an obligatory need to disperse because this
multi-day association also occurs when there is plenty of food for H. cooremani’s
active stages. Haemolymph from reflexive bleeding of the host frequently bathes the
subelytral zone where the deutonymphs reside but the caustic fluid does not harm
them. Radio-labelling showed that substances move from beetle to deutonymphs,
apparently through the attachment sucker. The authors suggest that the interaction
has changed from the serendipitous use of reflexed haemolymph controlled by the
host to the deutonymphs extracting fluids as required. Somewhat teleologically
they state “the ultimate evolutionary significance of the deutonymphal morph is to
serve as a transition from a free-living existence in temporary environments to a
Types of Ecological Interactions 349

more stable parasitic existence.” Athias-Binche (1995) considers that the honey-bee
parasite Varroa jacobsoni also evolved parasitic tendencies via a phoretic route,
with paraphagy (feeding on host exudates or secretions) as an intermediate step.

Loss of Parasitism

Radovsky et al. (1997) report that an undescribed species from the Macronyssidae,
a family otherwise composed of parasites, is a voracious predator of other mites that
live in the guano that coats bat caves. Based on its morphology, the mite is clearly
derived from parasitic ancestors and the authors state that it represents “a rare
and hitherto undocumented evolutionary shift from parasitism to predation…”
It is not the only case, however, as A. Wohltmann notes (pers. comm.) that larvae
of Johnstoniana rapax (Johnstonianidae) and a few species of Abrolophus
(Erythraeidae) are predatory. Both are from the Parasitengonina, a group characterised
by parasitism in the larval stage. Another example of a dietary shift is that of larval
Balaustium (Erythraeidae) that have switched from the normal parasitic behaviour
of erythraeid larvae to feeding on plant fluids and pollen (Newell 1963). Perhaps the
most extreme case of diet-switching is in the cave-dwelling tick Antricola delacruzi
(Argasidae). Larvae of A. delacruzi are parasitic on bats, but nymphs and adults
have reduced mouthparts and don’t feed on blood. Instead, they appear to consume
the nutrient-rich bat guano on which they spend most of their time, and have salivary
glands very different from those of normal blood-feeding ticks (Ribeiro et al. 2012).
Klimov and OConnor (2013) provide strong evidence that free-living dust mites
in the family Pyroglyphidae arose from one or two switches from an obligatory
on-host lifestyle, but it is not clear that the diet of the mites changed, as the ancestors
of dust mites may also have fed on keratin (albeit on the body of the host rather than
in the host’s nest).
Although there are few examples of lineages in which a particular life-history
stage has switched from parasitism to predation, herbivory or coprophagy, there
are many instances in the Hydracarina in which feeding in the larval stage has been
completely suppressed. Smith (1998) lists 29 species of water mites from eight
families in which the larva is not parasitic. Larvae either hatch, swim about briefly
and then transform to the calyptostatic protonymph without feeding, or transform to
the protonymph without emerging from the egg clutch at all. Smith feels that loss of
larval parasitism has evolved at least 21 times within the water mites. Non-parasitic
species can often be matched with closely related species that produce parasitic
larvae (e.g., Bohonak et al. 2004), suggesting that while loss of parasitism is an
ongoing process, it is unlikely to lead to distinct lineages of non-parasitic species.
Such species pairs may be so morphologically similar as to be confused for a single
species (e.g. Davids 1997a). In the terrestrial Parasitengonina, only one example of
suppression of larval feeding is known, in a species of Allothrombium (Trombidiidae)
(Smith 1998). It is not clear whether this propensity to dispense with parasitism is a
true difference between terrestrial and aquatic parasitengones or whether it reflects
a paucity of studies on land-dwelling species.
350 9 Animals as Habitats

Why have water mites repeatedly given up parasitism? Or more generally, what
would lead to the suppression of a feeding stage, especially one that also serves as
a transport stage when on a flying host?
Suppression of a feeding larva has evolved repeatedly in marine invertebrates
(Strathmann 1978). Species producing smaller eggs tend to have larvae that feed for
a relatively long time while drifting in the plankton. Species that produce large,
yolk-rich eggs tend to have non-feeding, non-planktonic larvae that rely on yolk
reserves for development. Lampreys (Agnatha) are another example of a taxon in
which a parasitic stage is sometimes skipped (Youson and Beamish 1991). Smith
(1998) points out that certain characteristics of lampreys and marine invertebrates
that have given up juvenile feeding – relatively larger eggs and smaller size at
adulthood – are shared by water mites with non-feeding larvae. So as well as giving
up transport, non-parasitic water mites tend to produce fewer, larger larvae per
clutch (Cook et al. 1989; Davids 1997b). The water mite Arrenurus rufopyriformis,
which lacks larval parasitism and hence has reduced dispersal capacity, has a more
scattered distribution and less genetic connectivity among populations than its
putative sister species, A. angustilimbatus, which does have parasitic larvae
(Bohonak et al. 2004).
What benefit could possibly outweigh loss of dispersal ability and reduced
fecundity? The riskiest stage in the life cycle of a water mite is likely to be the
parasitic one, as the probability of finding a host and then successfully detaching
over water appears to be quite low. Collins (1975) estimates that most mortality of
Wandesia thermalis occurs in the transition from a searching to a parasitic larva:
only 10 % of larvae successfully encounter a brine fly host, avoid its defences and
leap onto the host. A larva may encounter only one potential host in its lifespan.
Meyer (1985) calculated the host-finding success of Hydrodroma despiciens larvae
(Hydrodromidae) in a lake in Germany. Only 2060 of 181,500 mites attached to a host;
an astonishing mortality of 98.9 %. So within a given body of water, production of
larvae with a ‘non-parasitic’ strategy would result in greater fitness for a female
water mite than would production of parasitic larvae. Collins (1975) suggests that
Wandesia thermalis has not abandoned parasitism, even though it is so risky, because
the hot-spring pools they inhabit are so ephemeral that it needs the flies more for
dispersal than for nutrition.
How much energy does a larva gain through parasitism? For water mites, a larva
may increase its volume from 3 to 600-fold, depending on the species (Smith 1988;
Davids 1991). For some species of Eylais (Eylaidae), the larval nutrition provides
up to 45 % of the adult body mass. In terrestrial parasitengones, the range of increase
during parasitism is 10–576-fold in fresh weight (A. Wohltmann, pers. comm.).
Thus to an Eylais larva, foregoing parasitic feeding would have a greater effect on
fitness than it would for many Unionicola or Limnesia larvae, whose volume
increases less than fivefold during engorgement.
The above leads to the prediction that loss of parasitism should be more likely in
water mite species that (a) live in stable habitats and (b) have larvae that engorge
little during their parasitic phase. However, Smith (1998) found no evidence that
water mites that have forgone parasitism inhabit unusually stable habitats; in fact,
some even inhabit temporary pools. There is indirect evidence for (b). None of the
Types of Ecological Interactions 351

water mites in Smith’s list are from the families Hydrachnidae or Eylaidae, the taxa
in which larvae show the greatest engorgement. Most are concentrated in the
Hygrobatidae, Limnesiidae, Pionidae and certain subgenera of Arrenuridae. These
taxa are primarily parasites of Diptera, on which the larvae spend only a few days
and engorge little.
The lack of species-rich non-parasitic lineages in water mites and the common-
ness of related pairs of parasitic and non-parasitic species suggests that while
non-parasitism is beneficial in the short term, it is a dead end in the long term.
Lack of dispersal may doom non-parasitic lineages to extinction through loss of
habitat or deleterious inbreeding. To test the latter idea, Bohonak (1998) collected
over 1,100 Arrenurus from five parasitic and three non-parasitic species from
southern Ontario and New York. He divided the parasites into three categories of
potential dispersal ability: mites on odonates were ranked as having greatest dispersal
potential, those on mosquitoes second and those on midges third. Non-parasitic
ones were last. He predicted that heterozygosity would be higher in more widely
dispersing species. In general, these species did tend to possess higher average
heterozygosities than poorly dispersing species; however, the rank correlation with
dispersal ability was not significant. Bohonak also found little evidence of population
differentiation among non-parasitic species, even across hundreds of kilometres.
He suggested that in these recently glaciated areas, time since end of glaciation had
been insufficient to allow genetic differentiation of the populations.
Is there evidence of life-history flexibility within a single species? Two species
of water mites have been reported to produce different sizes of eggs. Crowell (1960)
found that one ‘form’ of Thyas stolli (Hydryphantidae) produced eggs with an
average diameter of 0.142 mm and another form produced fewer, larger eggs with
diameters of 0.24 mm (volumes of 1.5 and 7.2 × 10−3 mm3, respectively). Böttger (1962)
observed that females of Piona nodata (Pionidae) lay two different sizes of eggs.
Well-fed females held in warm water laid eggs about 0.27 mm in diameter. He held
the same females for 6 weeks without food at 5–6 °C and when they were brought
up to room temperature they laid small, yolk-poor eggs of about 0.19 mm in diameter.
From the big eggs hatched big larvae that transformed into protonymphs without
having to undergo a parasitic stage. The small eggs produced small larvae that died
after 10 days without a host. Böttger suggested that after winter, females produce
small, parasitic larvae and that during the summer they produce large, non-parasitic
ones. It may be that such a phenotypically plastic ‘mother’ species gives rise, perhaps
repeatedly, to non-parasitic daughter ‘species’ that repeatedly go extinct.

Loss of Phoresy

OConnor (1982) mentions that the astigmatans Glycyphagus domesticus and


G. destructor are stored-product pests with ‘inert’ deutonymphs that are dispersed by
wind rather than on a host. Likewise, Acarus species that live in nests normally have
deutonymphs phoretic on fleas but this stage has been lost in stored-product species.
Some astigmatan associates of beetles that remain on their hosts for their entire
life no longer express the deutonymph (Eickwort 1983). The 30 families in the
352 9 Animals as Habitats

group Psoroptida, permanent symbionts of vertebrates, also have lost the deutonymphal
stage (Houck and OConnor 1991).
OConnor (1982) considers that the repeated loss of the phoretic deutonymph in
the Astigmata is correlated with permanently parasitic lifestyles and with the transi-
tion from living in ephemeral resources to life in stable, widespread habitats, such
as litter and soil. In such situations, transport via phoresy would be less useful or
even disadvantageous. A supporting example is that of the pitcher-plant dwelling
astigmatans Sarraceniopus (Histiostomatidae). The Sarracenia pitcher plants typi-
cally occur in colonies in peat bogs. Deutonymphs of Sarraceniopus move from
aged pitchers to new ones by walking (OConnor 1994). Although deutonymphs
retain the physical modifications for phoresy and may occasionally use these for
long distance dispersal, most migrate to other pitchers of the same plant or nearby
plants. OConnor (1994) suggests that an ambulatory dispersal mode may also be
common in astigmatan associates of Old World pitcher plants (Nepenthes; however,
this is not true for all Nepenthes inhabitants, see Fashing and Chua 2002). Species
in the marine Hyadesiidae have completely lost the ability to form deutonymphs,
presumably a reflection of the continuous nature of their algal food source.
In summary, ecological interactions between mites and other animals are fluid
over evolutionary and ecological timescales. Certain behaviours or structures appear
to predispose lineages to switch from one type of association to another. However,
there is no single path from free-living existence to symbiosis. For example, parasitism
may arise from nidicolous, phoretic or predatory ancestors, and can be lost through
changes in diet or through suppression of parasitic stages. Now let us examine some
of these relationships in greater detail.

Life with Invertebrates

As hinted at in Table 9.1, mites are associated with a wide range of invertebrates.
Very few of the major terrestrial or freshwater taxa have escaped and several marine
groups also have their mite associates. But it is arthropods, including other acarines,
that provide the bulk of invertebrate associates of mites.

Taxonomic Survey of Associates

Below, we review the major mite taxa associated with other invertebrates before
analysing the functional classifications of these relationships.

Parasitiformes

There are no known symbiotic associations between the Opilioacarida and other
invertebrates or vertebrates. The Holothyrida are also free of such relationships.
Of the other parasitiform taxa, the Ixodida (ticks) is composed entirely of vertebrate
Life with Invertebrates 353

associates. In contrast, relationships with invertebrates abound in the Mesostigmata.


All known invertebrate associations are with arthropods. Hunter and Rosario (1988)
note that 25 of the 44 monogynaspid (56 %) and 20 of the 24 trigynaspid families
(83 %) have species associated with arthropods, 95 % of them with insects. Insects
that dwell in confined quarters, such as ground-nesting Hymenoptera and gallery-
dwelling beetles, are particularly popular hosts for mesostigmatans. Sixteen families
are associated with Apidae, 34 species with honey bees alone. Twenty families are
associated with Formicidae. Myrmecophilous (ant-loving) mesostigmatans display
some of the most extravagant mimicry of their hosts' bodies, a phenomenon termed
Wasmannian mimicry (Gotwald 1996). For example, Macrocheles rettenmeyeri
(Macrochelidae) is associated with the army ant Eciton dulcius on which it attaches
by inserting its chelicerae into the pulvillus (toe pad) of the hind leg. The curved
hind legs of the mite act as the ant’s ‘claws’ while the host walks and interacts with
other ants. Another example of such Wasmannian mimicry in ant-associated
Mesostigmata is shown in mites of the genus Planodiscus (Uropodina) that attach
to the host’s leg. The sculpturing on the mite matches that of the ant’s leg, down to
the arrangement of setae. A final example is the Larvamimidae, which mimic ant
larvae and are carried about by workers when the colony moves (Elzinga 1993).
The dung- and wood-dwelling beetle families Scarabaeidae, Scolytidae and
Passalidae are used as hosts by many species of mesostigmatans, as are the
predatory Carabidae and carrion-feeding Silphidae (e.g., Knee et al. 2012a, b).
The mutualistic relationship between silphids and their mites is discussed in section
“Mutualism”. Other major insect hosts are found in the Diptera and Lepidoptera.
The astounding behaviour of Dicrocheles moth-ear mites is discussed in section
“Parasitism and Parasitoidism”. Non-insect hosts for mesostigmatans include:
beach-hoppers (Crustacea: Amphipoda) for some Eviphididae; terrestrial hermit
crabs some Cercomegistidae; millipedes for some Blattisociidae, Paramegistidae,
Parantennulidae and all species in the Heterozerconidae and Neotenogyniidae; and
centipedes for the Discozerconidae and some Paramegistidae and Parantennulidae
(Krantz 1978; Farfan and Klompen 2012).

Acariformes

The major groups in the Acariformes vary in the proportion of taxa associated with
other invertebrates. With one exception, no species in the almost 140 families of
Oribatida has a clearly obligatory relationship with another taxon. The exception is
Aribates javensis (Aribatidae), which appears to be an obligatory inquiline associ-
ated with ants (Aoki et al. 1994). Oribatids have been found clasping setae and hairs
of a variety of invertebrates and vertebrates (Norton 1980; Miko and Stanko 1991)
but it is not clear whether such phoresy is a necessary part of these species’ life
histories.
Members of the Prostigmata show a greater range of associations, with many
taxa involved in phoretic, parasitic and possibly mutualistic relationships. A few
groups have no known symbiotic relationships with other invertebrates (Bdelloidea,
Tetranychoidea). Other taxa have phoretic or commensal relationships with insects,
354 9 Animals as Habitats

Fig. 9.2 Parasitengone larva


(Trombidiidae) parasitizing
a parasitengone adult
(Smaridiidae) (Illustration
by HC Proctor)

especially social insects (e.g. Scutacaridae, Cheyletidae). Parasitoidism is frequent


in the cohort Heterostigmatina, the only mite taxon in which this particular form of
symbiosis is known (Kalizewski et al. 1995). Other members of the Heterostigmatina
are purely phoretic on their arthropod hosts. Several very species-rich prostigmatid
taxa are entirely parasitic. The cohort Parasitengonina comprises ~ 11,000 named
species placed in more than 50 families. Its members are termed protelean parasites,
meaning that the larvae are parasitic but the post-larval stages are free-living predators.
Larvae of two of the terrestrial families, Leeuwenhoekiidae and Trombiculidae, and
of one family of water mites, Thermacaridae, are parasites on vertebrates as larvae
but those of the remaining groups parasitise arthropods. All major orders of insects
and arachnids have been described as hosts (Welbourn 1983). Hosts include other
parasitengones (Wendt et al. 1994) (Fig. 9.2), though based on a general failure to
of larvae to engorge when feeding on other mites, most such parasitism is accidental
(Makol et al. 2012). In a few families of water mites, the post-larval stages are parasitic
on freshwater mussels, snails, crayfish, larval mayflies and salamanders (see Chap. 7).
A few species in the mainly marine group Halacaroidea are thought to be parasitic
on crayfish, sea urchins, nemerteans and hydroids. The Pterygosomatidae parasitise
scorpions, cockroaches and reduviid bugs, as well as reptiles. Some Ereynetidae are
parasitic on terrestrial slugs (Baker 1970).
When the Astigmata arose within the Oribatida (see Chap. 3), one of the
changes in life history was a specially modified deutonymphal stage, the hypopus,
adapted for phoretic dispersal on insects (OConnor 1982). Although numerous
lineages of Astigmata have become permanently associated with mammal and
bird hosts, there are many families that still show the ancestral association with
arthropods. Beetles, bees and wasps are favoured hosts, and carry deutonymphs
of Acaridae, Glycyphagidae, Carpoglyphidae, Chaetodactylidae, Gaudiellidae,
Histiostomatidae, Platyglyphidae, Canestriniidae and Winterschmidtiidae. Other
phoretic hosts include dipterans and lepidopterans.
Many astigmatans associated with bees and wasps live in the nests of their
hymenopteran hosts, where they may act as commensals, mutualists or parasites
Life with Invertebrates 355

(e.g. Winterschmidtiidae). The Ewingiidae live in the gills and vasa deferentia of
pagurid land crabs but their feeding ecology is unknown (OConnor 1994). Some
species in the Histiostomatidae are found in the egg cocoons of freshwater leeches
and terrestrial earthworms; likewise, a histiostomatid and an acarid species have
been found grazing on the epidermis of a freshwater leech (Proctor et al. 1997).
For these annelid-associated astigmatans, it is unclear whether the relationships are
predatory or parasitic, or whether the mites feed on associated microbes. Most of
the invertebrate-associated astigmatans appear to have purely phoretic relationships
with their hosts but there are exceptions. Deutonymphs of Hemisarcoptes cooremani
(Hemisarcoptidae) are reported to be parasitic on their host coccinellid beetles
(Houck and OConnor 1991). The epidermoptid Myialges macdonaldi is hyperparasitic
on hippoboscid flies, themselves parasites of birds (Krantz and Walter 2009).

Phoresy and Dispersal

Acarologists have been long fascinated by phoretic associations between mites


and other invertebrates and there are numerous reviews of this phenomenon in
mites in general (e.g. Binns 1982; Athias-Binche 1991, 1995) and for particular taxa
(e.g. Hunter and Rosario 1988; Houck and OConnor 1991; OConnor 1994).
The earliest fossil record of phoresy is for a small astigmatan, possibly a member of
the Histiostomatidae, associated with a spider preserved in 44–49 MY old amber
(Dunlop et al. 2012); however, the ecological phenomenon is undoubtedly much
older than that. The term ‘phoresy’ has a history of controversy. It was coined in 1896
in response to the observation of small insects riding on larger ones (see Wheeler
1919 for a history). Many definitions have been proposed and none has yet been
accepted by all acarologists. In 1971, Farish and Axtell restricted phoresy to refer
to attachment of one animal to another only when attachment was temporary and
actively sought, with neither feeding nor ontogenetic changes taking place during
attachment. Athias-Binche (1991) uses a similar definition but states that while on
the host, the phoretic is ‘quiescent’. Houck and OConnor (1991) dislike the stipulation
of ‘actively seeking hosts’ because deutonymphal astigmatans may be stationary
while waiting for a host to pass their way. Their more complex definition is that
phoresy occurs when “one organism (the phoretic) receives an ecological or
evolutionary advantage by migrating from the natal habitat while superficially
attached to a selected interspecific host for some portion of the individual phoretic’s
lifetime”. An additional requirement of Houck and OConnor is that the phoretic
cannot benefit nutritionally or developmentally while on the host.
To continue the tradition of tailoring phoresy to one’s own preferences, in this
book we use a verbally simpler definition: phoresy is a type of temporary symbiosis
whose function is to allow a smaller individual (the phoretic) to move from one
place to another on a larger individual (the host or carrier). This definition allows
phoretics to make more than one use of a host. For example, the pseudoscorpion
Cordylochernes scorpioides (Chernetidae) disperses from its normal habitat of
356 9 Animals as Habitats

decaying trees by riding giant harlequin beetles ( Acrocinus longimanus ,


Cerambycidae) (Zeh and Zeh 1992). While on the host, they feed on mites also
riding on the beetle, and male pseudoscorpions defend territories and mate with
females. Feeding and mating activities on the host would eliminate C. scorpioides
from the category of phoretic in the other definitions but none of the above authors
would deny that the primary benefit received by the pseudoscorpions is transport to
new habitats. Similarly, water mites that are transported between water bodies as
parasites benefit from the transport, as they have no other means of moving overland
between aquatic habitats. Even deutonymphs of Poecilochirus carabi (Parasitidae),
famous examples of phoretics, sometimes feed on nematodes or oral secretions of
their beetle hosts (Brown and Wilson 1994).
An example of something that is sometimes called phoresy (e.g. in Athias-Binche
1991) but that does not meet our criteria is the pre-parasitic attendance shown by
water mites, in which larvae attach non-parasitically to a juvenile insect and wait
until the adult insect emerges. It is not phoresy because there is no loss of fitness to
the mite larva if the juvenile insect does not move during this period. Water mites
that parasitise adult black flies (Diptera: Simuliidae) associate with the sedentary
pupae beforehand. Likewise, larvae of terrestrial parasitengones in the families
Johnstonianidae and Microtrombidiidae attend immobile pupae of their dipteran
hosts for up to 23 days (Wohltmann 1996; Wohltmann and Wendt 1996). Clearly, no
transport benefit of association is accrued in these cases. In this section we will
concentrate on mite taxa that gain no apparent benefit from phoresy other than
transport, and will discuss phoretic parasites and mutualists in subsequent sections.
What is the value of phoresy? The short-term benefits probably include escape
from physically deteriorating habitats, emigration from overcrowded natal sites or
transport to new sources of food. Phoretic species are especially common in ephem-
eral and patchy habitats such as rotting wood or fruit, dung, fungi, carrion, and
temporary water bodies (OConnor 1984; Hunter and Rosario 1988). In contrast,
mites in continuous and relatively constant habitats, such as leaf litter and foliage,
rarely show adaptations for phoresy. It is interesting to note that species of Siteroptes
(Pymephoroidea) that feed on grass-associated fungus – a continuous resource – have
no phoretic phase, whereas those that feed on patchily distributed fungi disperse
phoretically (OConnor 1984).
Being small, wingless and having sense organs with a limited range, mites that
live in ephemeral habitats benefit from associating with larger animals that
can move farther, faster and can detect appropriate sites from a greater distance.
For example, under its own power the astigmatan Sancassania ( = Caloglyphus)
rodionovi can walk only 5.2 mm in a minute (Binns 1982), while as a deutonymph
phoretic on a winged insect it can cover 1,000 times the distance in the same time.
Potential hosts vary in their suitability as dispersal agents. Deutonymphal Uropoda
orbicularis (Uropodidae) rarely attach to very small hydrophilid beetles, which may
not be able to disperse from dung pats when laden with phoretic mites (Bajerlein
and Przewoźny 2012). Long-term benefits of dispersal and migration by phoresy
include prevention of inbreeding and access to new areas for colonisation (e.g. oceanic
islands, Athias-Binche 1995).
Life with Invertebrates 357

Table 9.3 Mite taxa that have stages modified for phoresy (excluding parasitic taxa)
Taxon Stage modified Modifications
Uropodidae Deutonymph Anal pedicel in some spp.
Ascidae (Antennoseius ♀ phoretomorph Body flatter and broader, integument and
janus) setae smoother than non-phoretomorphs
Astigmata ♂ & ♀ deutonymphs Reduced legs and mouthparts, dorsoventral
flattening, increased sclerotization,
attachment structures (suckers, claspers)
Pygmephoridae ♀ phoretomorph Smaller and more compact body, leg I and
leg I tarsal claw much thicker than
non-phoretomorphs
Pyemotidae (scolyti group) ♂ & ♀ phoretomorphs ♀ body shorter and broader than
non-phoretomorphic ♀s, legs I of both
sexes thicker and with strong claw
Scutacaridae (Archidispus, ♀ phoretomorph ♀ body broader, with thicker setae than
Lamnacarus, non-phoretic ♀s; many differences
Scutacarus spp.) in legs, tarsal claws of legs I larger
(often absent in non-phoretomorphs)

Examples of phoretic relationships with invertebrates can be found in all


acariform suborders and in the Mesostigmata. Typically, only one life-history stage
is phoretic in a given species (Table 9.3). In the Astigmata the phoretic stage is
the deutonymph, while in the Prostigmata it may be the fertilised adult female
(e.g. Pygmephoridae), the larva (e.g. Parasitengonina) and, in one case, the tritonymph
(Ereynetes sp., Ereynetidae) (Zhang and Sanderson 1993). In Mesostigmata,
deutonymphs are phoretic in Sejida, Microgyniina, Uropodina, Parasitiae and early
derivative Dermanyssiae (e.g. Digamasellidae). In the Trigynaspida and derived
Dermanyssiae, adults (typically females) are the phoretics. Mitchell (1970) points
out that when dispersal occurs prior to mating, in order to be fairly certain (87.5 %)
that at least one individual of both sexes are present, at least four mites must meet
after phoretic migration (assuming equal sex ratios and probabilities of dispersal).
Hunter and Rosario (1988) note that for Parasitus spp. (Parasitidae) on bumble
bees, mites that disperse prior to mating tend to do so as groups rather than as
individuals. Zhang (1992) suggests that Allothrombium pulvinum (Trombidiidae)
larvae parasitise already parasitised aphids (superparasitism) in order to increase
the chance of finding a mate in the new habitat. This may be a general explanation
for the clumped distribution of larvae on hosts that is typically observed for
parasitengone mites (e.g. Smith 1988).
The phoretic stage may be specifically modified for transport on the host
(Table 9.3). The best known of these is the astigmatan deutonymph, often termed
the hypopus (plural hypopi). Hypopus was originally a genus name applied to
strange, flattened mites with reduced legs that were found attached to insects. Later
it was realised that these mites represented a stage common to many astigmatans.
The phoretic deutonymphs of astigmatans have a variety of modifications, in part
358 9 Animals as Habitats

dependent on the host taxon. Entomophilous forms attach to arthropods using a


caudoventral sucker. Pilicolous forms have reduced suckers but use a different set
of modified setae to attach to insect setae or to mammal hair. In endofollicular
types, living in the skin of birds and mammals, attachment organs are highly reduced
or absent (Houck and OConnor 1991). Hypopi lack functional mouthparts and
resemble neither the preceding protonymphal stage nor the subsequent tritonymphs
and adults; they are thus termed a heteromorphic stage. Larvae of Parasitengonina
are both parasitic and phoretic, and are also extremely heteromorphic relative to
the subsequent stages. The most highly modified of mesostigmatan phoretics are
the heteromorphic deutonymphs of families in the Uropodina. As the name implies
(uro = tail, pod = foot), most uropodines attach to their arthropod hosts by a stalk
secreted from the anus (Bajerlein and Witaliński 2012), although others use caudal
suckers, claws or chelicerae.
Expression of the heteromorphic stage may be facultative. In the Astigmata, the
phoretic deutonymph is often skipped, with protonymphs transforming directly
to the tritonymph. A few species of water mites may have facultatively non-
parasitic (and hence non-phoretic) larvae. Many species in the Pygmephoridae and
Scutacaridae (Prostigmata) have two types of females, one that is not phoretic and
another that is the phoretomorph (Moser and Cross 1975; Ebermann 1991a, b).
The scolyti group of Pyemotes (Pyemotidae) has phoretomorphic males and females
(Krantz 1978). In these taxa, phoretomorphs are distinguished by enlarged tarsal
claws on the first pair of legs, used to cling to the host, as well as other characteristics
(Table 9.3). Binns (1982) notes that phoretic morphs of many different taxa share
similar morphologies: dorsoventral flattening, oval or circular bodies and flanges
covering all or some appendages. This convergent morphology may serve to reduce
loss of moisture when on the host and to present a smooth dorsal surface making it
difficult for the host to remove mites by grooming.
The striking differences between phoretic and non-phoretic females has often
resulted in a single species being described as two separate species. The most confusing
situation, termed by Lindquist (pers. comm.) “a taxonomic and nomenclatural can
of worms”, is the Pediculaster-Siteroptes dyad. Members of the genus Pediculaster
(typically placed in the Pygmephoridae) appear to be phoretic morphs of Siteroptes
species. As the genus Siteroptes is often placed in its own family, Siteroptidae, the
confusion is at both the generic and familial level. Cross and Kaliszewski (1988)
sidestepped the problem by simply describing Pediculaster flechtmanni as a member
of the Pygmephoroidea, a superfamily that includes both contentious families.
Similarly, there is one known example of phoretomorphy in the Mesostigmata, the
ascid genus Antennoseius (Lindquist and Walter 1989; Beaulieu et al. 2008), in
which the phoretomorphs and homeomorphs were originally thought to represent
different species.
Phoretic phases may lie in wait for passing hosts or they may actively seek them
out (Binns 1982). Deutonymphs of Histiostoma laboratorium (Histiostomatidae)
adopt a characteristic ‘questing’ posture in which they stand on their third and
fourth pair of legs with their bodies propped against their posterior margins,
gently waving their first and second pair of legs. Should a host insect walk or fly
Life with Invertebrates 359

past at low altitude, the H. laboratorium deutonymph is capable of springing up to


5 cm to gain attachment.
Active phoretics may wander about until they contact a host of the correct taxon
and stage. For example, female Blattisocius tarsalis (Blattisociidae) ignore pupae of
their moth hosts until the insect is within a few hours of eclosion (Binns 1982).
Presumably some chemical associated with the adult moult acts as an attractant to
the mites, which congregate around the pupa and rush onto the moth as it emerges.
Sometimes a combination of active and passive searching is involved, as many
astigmatan hypopi crawl upwards towards the highest point in their local environ-
ment, where they wait in questing postures for larger organisms (Binns 1982).
The trigger for dismounting from a host varies with the phoretic. The only reliable
way to induce Poecilochirus carabi (Parasitidae) to dismount and transform into
adults is to place them on a mated pair of Nicrophorus beetles burying a carcass
(Brown and Wilson 1994). Many species dismount when hosts oviposit but to our
knowledge there have been no demonstrations that oviposition per se rather than
changes in the physical environment at the oviposition site induce detachment from
the host. Farish and Axtell (1971) observed that Macrocheles muscaedomesticae
(Macrochelidae) were much more likely to dismount from their housefly hosts if the
fly had landed on fresh, moist manure than on a dry patch. The pygmephorid
Brenddandania lambi drops from its phorid fly host only in the presence of the fungal
mycelia on which it feeds (Clift and Larsson 1987). Binns (1982) suggests that high
humidity is a cue for detachment for many astigmatan phoretics. Some mites will
detach from moribund hosts. Larvae of Arrenurus water mites (Arrenuridae) leave
hosts killed by having their heads crushed more rapidly than they do hosts killed in
other ways (Smith 1988), suggesting that a factor released from the brain may be a
normal trigger for dismounting. Although leaving a dead host would seem a sensible
thing to do, one sometimes finds dead insects still covered by phoretic mites. This
suggests that such taxa require a specific dismounting cue and will not terminate the
phoretic phase until this signal has been received. Some phoretic mites may benefit
from host death. Histiostoma polypori (Astigmata: Histiostomatidae) rides on
earwigs, and feeds on the corpses of juvenile earwigs that die accidentally in the
nests of their mothers; this lifestyle is termed necromenic and is also exhibited by
many nematode species (Wirth 2009).
For species with facultative phoretic stages, indications that the environment is
growing less conducive to survival and reproduction appear to trigger the change to
the phoretic form (OConnor 1994). The proportion of dispersing females in popula-
tions of the fungivorous mushroom pest Pediculaster flechtmanni (Pygmephoroidea)
increases as the colony ages (Cross and Kaliszewski 1988). Overcrowding induces
production of the smooth-skinned dispersal morph of Antennoseius janus (Lindquist
and Walter 1989). Athias-Binche (1995) describes several different cues that affect
phoresy in mites. She notes that whether female Scutacarus acarorum (Scutacaridae)
develop into phoretic or non-phoretic morphs during the larval stage depends on the
quality of the food offered to their mothers. A 6-year study of the log-dwelling uro-
podid Allodinychus flagelliger (Uropodina) showed that the ‘emigration rate’ of
phoretic deutonymphs was correlated with the degree of decomposition of the wood.
360 9 Animals as Habitats

In taxa without obvious phoretomorphs, the behavioural switch to phoresy also


seems cued by deteriorating conditions. For example, female macrochelids may
become phoretic after food is eliminated, if they become overcrowded, or if O2 gets
low (Athias-Binche 1991). Treat (1975) suggests that the ‘jostling’ behaviour
exhibited by moth-ear mites, Dicrocheles (Laelapidae), under crowded conditions may
induce the exodus of young females from their natal ears. The propensity to switch
from a non-phoretic to a phoretic ‘program’ appears to be heritable. Athias-Binche
(1995) describes studies by Knülle (1987, 1991) in which hypopus formation of
Lepidoglyphus destructor was induced by feeding with inferior food (oats rather
than yeast) but ranged from 3 % in old laboratory strains to 96 % in field populations.
When laboratory strains were crossed with field populations, the proportion of
protonymphs able to generate hypopi was intermediate in the progeny. This suggests
that in stable laboratory habitats, the fitness advantage of phoresy is too low to be
maintained by selection (see also section “Loss of Phoresy”, above).

Commensalism

When one species benefits by permanently associating with a second species, with
the second neither gaining nor losing by this association, the first species is termed
a commensal. It is often difficult to know whether a mite is truly commensal or if its
beneficial or harmful effects on the host simply have not been determined. Thus the
mites listed in this section may later prove to be more appropriately classed as
parasitic or mutualistic. Many of them may be kleptoparasites, which steal food from
hosts. If food is superabundant, kleptoparasites may not necessarily cause harm to
their victims and hence are not always ‘parasites’ in the strict sense. Commensal
mites may live on the host or they may live in the host’s nest. In the first case,
they may feed on exudates and secretions of the host (paraphagy). In the latter,
they may also feed on host products or they may consume other nidicoles. Some
species of Otopheidomenidae live under the wings of hemipterans, where they
appear to feed by scraping the host’s integument (Eickwort 1983). Yoder (1996)
found that the Androlaelaps ( = Gromphadorholaelaps) schaeferi (Dermanyssidae)
that ride about on the Madagascar hissing cockroach, feed on cockroach saliva
expressed by the host during feeding, grooming and defence. Yoder termed this
behaviour ptyalophagy (Gr. ptyalon spittle). Kaliszewski et al. (1995) note that
Tarsonemus dispar (Tarsonemidae), all of whose instars live under the elytra of the
passalid beetle Popilius disjunctus, may be paraphagous. Treat (1975) suggests that
the moth-riding Pronematulus pyrrohippeus (Tydeidae) may prey on other passengers.
The consumption of other symbiotic mites may benefit the moth (in which case it
would be mutualism) but irrespective of this, it certainly benefits the mite.
Eickwort (1990) reviews the relationships between mites and those great
nest-builders, social insects. One of the most interesting is Antennophorus
(Antennophoridae) in the nests of Lasius ants. The mite rides clasped beneath its
host’s head where it appears to intercept trophallaxis (social feeding) by mimicking
Life with Invertebrates 361

the ants’ food-exchange cues. Army ants (Eciton spp.) have especially diverse arrays
of mite associates, with members of 55 mite families having been identified with the
nests and bodies of only a single army ant species, Eciton burchellii (Rettenmeyer
et al. 2011) (but note that even in temperate zones, an ant species may host > 30 spp.
of mites, Campbell et al. 2012). Acarines associated with army ants are reviewed by
Gotwald (1996). Mites are not the only arthropods that trespass in the nests of social
insects. Beetles, silverfish and caterpillars are also common inquilines, sometimes
causing no harm to their hosts, sometimes stealing food and sometimes being
clearly detrimental. The toleration shown by social insects to these mites and insects
is puzzling, particularly because these insects recognise and attack conspecifics
belonging to different nests. Perhaps nidicoles coat themselves in the surface chem-
icals of their hosts and, together with morphological mimicry, this allows them to
pass unnoticed among the workers. But it is also possible that they produce some
chemical that their hosts find attractive. Wheeler (1923) pointed out that almost all
ant inquilines produce a tasty exudate that seems irresistible to the hosts, who pro-
vide the inquilines with food in exchange for the attractive secretions. What the
housing of these foreign species indicated was not stupidity on the ants’ part, argued
Wheeler, but simply a result of the habit of trophallaxis. He provided a litany of
social evils in human life that had parallels in ant societies (p. 197): ‘We not only
tolerate but even foster in out midst whole parasitic trades, institutions, castes and
nations, hordes of bureaucrats, grafting politicians, middle men, profiteers and usu-
rers, a vast and varied assortment of criminals, hoboes, defectives, prostitutes,
white-slavers and many other purveyors of antisocial proclivities…’.
It is possible that the myrmecophilous oribatid Aribates javensis (Aribatidae) is
one such intoxicating inquiline. Itu and Takaku (1994) feel that these rotund mites
are completely dependent on Myrmecina ants for transport, food provision and care
of their eggs. The mites do not walk on their own and are carried from nest to nest
by the ants along with the ants’ own brood. The ants regularly lick the mites, each
mite being groomed up to eight times per hour. Presumably, the mites produce a
tasty exudate that Myrmecina enjoys. Mite eggs are removed from the oribatids’
ovipositors by the ants, licked and placed together with the ants’ eggs. Itu and Takaku
did not observe trophallaxis between Aribates and Myrmecina, and assumed that
mites feed on fungus growing in the nests. Mites maintained without ants died
within 6 days but those with ants lived more than four times longer. The author’s
statement that the mites are unable to walk on their own is somewhat puzzling given
that they found both males and females in ant nests. With one known exception
(see Chap. 5), oribatids transfer sperm by depositing spermatophores on a substrate;
it is difficult to understand how immobile males and females would be able to mate.

Parasitism and Parasitoidism

As we have stressed above, the difference between being phoretic or commensal


and being parasitic is not necessarily large. Phoretic mites may not interfere with a
362 9 Animals as Habitats

Fig. 9.3 A chigger or scrub


itch mite (Trombiculidae) – a
larval parasitengone that
attaches to the skin of
vertebrates (SEM by
DE Walter)

host’s locomotion when in small numbers but may seriously impede the host when
abundant. Commensals may be innocuous when food supplies are in excess but
become debilitating kleptoparasites in lean times. Parasites may cause no more than
a momentary itch or they may greatly weaken the host. However, because parasitoids
invariably kill their hosts, they can be nothing but extremely bad. In this section we
describe the ecology and behaviour of selected groups of mites that have clearly
negative effects on their invertebrate hosts.

Parasitengonina

The Parasitengonina is the largest group of mites that have parasitic relationships
with invertebrates. Although we may be more familiar with the chiggers and scrub
itch mites that bite vertebrates (Fig. 9.3), the vast majority of parasitengones have
larvae that attack arthropods. Welbourn (1983) details the host–parasite relation-
ships of terrestrial members of this cohort, and Smith and Oliver (1986), and Smith
and Cook (1991) those of the aquatic ones (Hydracarina). Other than host associa-
tions, little is known of the parasitic behaviour of the former (but see Wendt et al.
1994; Wohltmann 1996; Wohltmann and Wendt 1996). In contrast, there is a great
deal known about the behaviour and ecology of water mite larvae, at least those in
Europe and North America.
Water mite larvae primarily parasitise the adults of freshwater insects. They reach
the adults by attending the pupae or final instar larvae of their hosts and rapidly trans-
ferring to the adult when it emerges. Some taxa parasitise insects that have extensive
contact with the water as adults (Coleoptera) or as both juveniles and adults
(Hemiptera), and do not show prolonged pre-parasitic attendance. Few families and
fewer orders of aquatic insects escape water mite parasitism. The exceptions are
the Ephemeroptera (mayflies) and Megaloptera (dobsonflies, alderflies). These
exceptions may be explained by their exceptional life histories. Mayflies are the only
Life with Invertebrates 363

extant order of insects to display a winged moult, from the subimago with its heavy
opaque wings, to the imago, with its lighter, more functional wings. These secondary
moults would likely dislodge any mite larvae that had attached to the subimago.
Megalopterans and species of Neuroptera that inhabit water (e.g., spongilla flies,
Sisyridae) are holometabolous insects that undergo their pupal stage in cocoons or
chambers on dry land. The adults do not return to water except to oviposit and
there is little opportunity for a mite larva to contact a host.
Insects are not the only hosts of parasitic water mites. Postlarval stages of some
taxa are parasites of mussels and snails (Chap. 7). The Astacocrotonidae have
been described as parasitic on the gills of a freshwater crayfish. Several species of
Hygrobates parasitize salamanders. There are also isolated reports of postlarval
water mites parasitising fish, either in their digestive tracts (Cupp and Willis 1982;
Yankovskaya and Fernando 1982) or on their gills (Tedla and Fernando 1970).
Rather than these instances representing parasitism, it seems more likely that the
mites were swallowed by the fish and remained undigested in the first case or tried
to crawl out through the gills in the second.
Smith and Cook (1991) use a two-tiered system to classify water mite larvae
based on where they search for their hosts. ‘Terrestrial’ larvae search for their hosts
on the upper side of the surface film, while ‘aquatic’ larvae search for hosts within
the water column or on the benthic substrate. The terrestrial search mode appears to
be ancestral, judging by the more primitive morphology of larvae and adults of such
taxa (Hydrovolzioidea, Hydryphantoidea, Eylaioidea). The aquatic search mode
evolved at least three times, in the Stygothrombidiidae, Hydrachnidae and in
the basal lineage leading to the Lebertioidea, Arrenuroidea and Hygrobatoidea.
As discussed above (section “Loss of Parasitism”), parasitic behaviour in water
mite larvae has been lost in at least 20 different lineages.
The process of attachment, from searching for the host to the initiation of para-
sitic feeding, has attracted the attention of many workers. Mitchell (1960) observed
the host-searching behaviour of Arrenurus fissicornis, which parasitises dragonflies.
He considered the movements of the larvae to be random and concluded that there
was no distance perception involved in finding a host. Davies (1959) had a different
opinion and felt it was possible that some chemical factor, released as the pupal skin
splits, stimulates larval Sperchon (Sperchontidae) to congregate near emerging
adult black flies. Gledhill et al. (1982) studied Sperchon setiger larvae in England
and saw that the mites would enter the pupal cases of blackflies long before the
insects were ready for ecdysis. This early colonisation appears to refute Davies’
concept of a chemical cue leading to the congregation of mites at times of mass
emergence. Smith and McIver (1984a), in their study of Arrenurus parasitism on
Aedes mosquitoes, observed that mite larvae appeared to be attracted to mosquito
pupae over short distances. The attraction was not a response to vibrations because
mite larvae were also attracted to anaesthetised pupae. They felt a water-borne
chemical cue was most likely, although other cues could not be disregarded.
Prior to attachment for feeding, water mite larvae can be easily removed from the
pre-adult host without any apparent harm. Upon beginning parasitism, the larva
introduces its chelicerae and produces a stylostome (discussed below). In this case
364 9 Animals as Habitats

the mouthparts become firmly embedded in the body of the host and prevent the
easy removal of the parasites. The window of opportunity for transfer may be very
brief and some larvae may be left behind on the host’s exuviae (Proctor, pers. obs.).
Perhaps this danger has selected for an interesting mode of pre-parasitic attachment
shown by Unionicola (Unionicolidae), Hygrobates and Atractides (Hygrobatidae).
Larvae of these genera bite through the pupal skin of their dipteran hosts to embed
their chelicerae in the integument of the adult developing within (Ullrich 1976).
When the adult ecloses, the mite larvae are dragged through the pupal exuviae rather
than having to actively transfer. These larvae are very long and narrow compared to
mite larvae that transfer in a more conventional manner, most likely probably to
allow them to be pulled more easily through the old pupal skin.
Many authors have described tube-like structures that penetrate into the body
of the host at the site of cheliceral penetration. These structures, which may be
flask-shaped or ramified, are termed stylostomes. The larvae of some terrestrial
parasitengones also produce stylostomes, although these are of a slightly different
form (Åbro 1988). Smith (1988) says it is not clear how many species of water mites
use stylostomes and that Thyas barbigera (Hydryphantidae) apparently does not.
Within a species of mite, stylostomes tend to have the same structure regardless of
host species but structure can vary between congeners (Lanciani and Smith 1989).
Åbro (1984, 1991) has examined the process of stylostome formation microscopi-
cally. After transferring to its adult damselfy host, an Arrenurus larva (Arrenuridae)
bites through the host integument and injects a bleb of fluid. Into the interior of
this droplet another fluid is injected that condenses along its outer surface to form a
tube. Åbro feels that the role of the vesicle is to provide a buffer for this developing
stylostome against immune reactions from host tissues. As more of this fluid is
injected, the tube elongates until it reaches an open haemolymph space in the host’s
body. The original bleb darkens (presumably because of the host’s immune response)
and solidifies and cements the larva’s chelicerae to the host’s integument. In water
mites the stylostome ends blindly but in the terrestrial chiggers it appears to be an
open-ended tube. It is easy to understand how chiggers might suck up the liquefied
tissues of their hosts through such a drinking straw but it is less apparent how
water mites get nutrients through their closed stylostomes. Åbro (1984) feels that
Arrenurus larvae feed by pumping out digestive enzymes into the stylostome.
The stylostome’s wall is permeable to this fluid, which is forced out to liquefy
host tissues. Subsequent sucking by the larval pharynx draws the surrounding fluids
back into the stylostome and to the mite’s mouth. The long stylostome tube may
allow a greater area of the host’s body to be probed than would be available to mites
feeding on tissues only at the site of penetration. Åbro (1984) suggests that the
blind-ended tube of water mites is adaptive in preventing fragments of cuticle from
being sucked into the narrow oesophagus of the larval mite; chiggers can get by
with open-ended tubes because they feed on non-chitinised hosts. In Arrenurus spp.
a stylostome begins to form within 10 min of attachment to a host and reaches its
final shape after 1 day; however, in larvae of Hydrachna spp., which remain on their
hosts for a much longer time, the stylostome grows considerably over a period of
months (Smith 1988).
Life with Invertebrates 365

Detachment and re-entry mechanisms are not well understood. It is clear that
water mite larvae on aerial hosts have a more uncertain fate than those parasitic on
aquatic hosts like beetles or bugs. Ellis-Adam and Davids (1970) found that larvae of
Piona alpicola dropped from their chironomid host apparently at random, regardless
of the midge’s proximity to water. Ullrich (1976) also states that Sperchon parasites
of simuliids leave their host randomly, whether over open water or not.
It has often been stated that mite larvae should prefer the host sex that is more
likely to be in close proximity to water. Smith and McIver (1984b) state that
Arrenurus larvae detach from their mosquito hosts when the mosquitoes return to
water to oviposit. However they found no consistent preference of the mites for
female mosquitoes. Davies (1959) and Booth (1978) found significantly more mite
larvae on female dipterans but Booth attributed this greater load to the larger size of
the female flies. Gledhill et al. (1982), on the other hand, found little evidence of
preference for female black flies over males as hosts for Sperchon larvae. Mitchell
(1967) and Robinson (1983) both found that male odonates were more commonly
parasitised than were females. Mitchell suggested that the larvae might be choosing
males because this sex tends to remain near territories around water.
Oviposition behaviour – irrespective of host sex – may affect detachment rate.
Rolff (1997) found that Arrenurus cuspidator larvae were more likely to detach from
mating pairs of the damselfly Coenagrion haustulatum than from pairs of C. puella.
He attributed this to C. haustulatum ovipositing under water, thereby wetting the
integument of the pair, while C. puella oviposits in floating aquatic vegetation.

Dicrocheles (Laelapidae)

Asher Treat’s (1975) book Mites of Moths and Butterflies is a lyrical and entertaining
exploration of the mites associated with lepidopterans. Although Treat describes
many fascinating acarines, the most astonishing are the moth-ear mites. Many
moths have paired tympanic chambers on the abdomen or thorax. Each chamber is
divided into an interior and exterior ‘room’ by a thin tympanic membrane that
acts as an eardrum. The main function of a moth’s ears is to listen for the ultrasonic
cries of hunting bats. About a dozen species of mites have been taken from the outer
ears of noctuid moths but only those of the genus Dicrocheles are known to invade
the inner chamber.
Species of Dicrocheles are tympanicolous and are found on noctuids on all
continents (except, of course, Antarctica). Occupancy of moth ears by the best-known
species, D. phalaenodectes, is unilateral. The occupied ear may be right or left but,
regardless of which, the other ear is almost invariably left undisturbed. Infaunation
appears to occur when moths feed at flowers, as it does for mesostigmatans phoretic
on hummingbirds (Colwell 1995). When a female D. phalaenodectes is placed on a
restrained moth, she first explores both ‘shoulders’ of the moth repeatedly and even-
tually rushes for one ear. She revisits the junction between thorax and abdomen
several times but always returns to the same ear. Treat feels that this repeated travel
reflects the laying of a pheromone trail by the first mite because D. phalaenodectes
366 9 Animals as Habitats

individuals subsequently introduced to the same moth almost always take the same
route as the first. He found no evidence that mites exploring ‘virgin’ moths had a
preference for the left or right side.
Soon after arriving at the chosen ear, the female removes both the tympanic and
countertympanic membranes and engorges on host haemolymph. She lays eggs in
the tympanic air sac and on the conjunctival membrane, up to 80 eggs in her lifetime.
Treat noted that ‘public sanitation’ appeared well regulated in Dicrocheles colonies.
Mites excrete only at the outer margin of the tympanic recess or at the rear of the
countertympanic cavity. When the young reach adulthood they mate. At this point
the ear may be too crowded to accommodate more offspring. Treat observed that the
departure of young mated females appeared to be encouraged by jostling movements
of more established individuals. Disembarkment tends to occur at nightfall, when
the moth actively searches for flowers. Mites move from the ears to the ventral side
of the head. If a flower is offered to the moth, several mites may quickly transfer on to
it via the moth’s proboscis. Although D. phalaenodectes often remained in flowers for
several hours, Treat did not see them feed on nectar or pollen. These young females
readily mount other moths that come to feed, in quest of uncrowded ears to colonise.
It would be an unwise mite that would damage the auditory ability of its host.
What effect does being deafened in one ear have on the escape response of moths
parasitised by Dicrocheles phalaenodectes? Treat occasionally collected moths
with ‘ghost colonies’ in which evidence of previous occupancy of an ear was obvi-
ous but no mites remained, suggesting that the moths had no problem surviving bat
attacks both during and after infestation. Treat also observed infested moths show-
ing the characteristic jinking and diving indicative of bat avoidance. It is possible
that bilateral colonies do occur but are not observed in moths collected at light traps
because such completely deafened hosts tend to be caught and eaten by bats.
In the summer of 1975, when his book had already gone to press, Treat was rather
shocked to find moths at his collecting light that had both ears heavily colonised by
what looked like Dicrocheles phalaenodectes (Treat 1975). Since he had seen only
ten bilateral infestations in 23 years of collecting, Treat felt that something was
amiss. Upon closer inspection, they proved to be Dicrocheles scedastes, a species
previously reported to produce bilateral infestations in New Zealand moths. Of 26
experimental infaunations of various moth hosts, Treat found that 20 were bilateral.
But although D. scedastes typically invaded both ears, in most instances the
tympanic membranes and acoustic sensilla were not destroyed and moths retained
acoustic sensitivity. Another difference between Dicrocheles phalaenodectes and
D. scedastes was that the latter took a subalar (beneath the wing) route to the moth’s
ear rather than a dorsal midline approach. This could be one of the reasons why
D. scedastes produces bilateral colonies: pheromone paths laid by pioneering mites
may not be noted by mites that approach the moth from the other side.
Another group of ear-dwelling mites also avoids the danger of deafening the host
by not puncturing their eardrums. ‘Otopheidomenidae’ means ‘ear-sparing’ and
refers to the fact that the mites of this family do not damage tympanic areas. Eggs
are usually laid in both ears. Many otopheidomenids live on sphingid moths, which
do not have ears, while others parasitise hemipterans (Krantz and Walter 2009).
Life with Invertebrates 367

Acarine Parasitoids

A parasitoid slowly but inevitably kills its victim, only one of which is required
for the parasitoid to complete its development. This places its strategy between
parasitism and predation. The best-known parasitoids are the tiny wasps often
used for biocontrol of insect pests. Parasitoidism is also common in the Diptera
(e.g. Tachinidae). Among parasitoid insects, the larval stage feeds on the host/prey
and the adult females are free-living and search for victims for their offspring.
The situation is somewhat different in mites, as adult females both search and ‘para-
sitoidise’. Parasitoidism is common in the prostigmatan cohort Heterostigmatina
but has not yet been recorded among other groups of mites (Kaliszewski et al. 1995).
Members of the Pyemotidae and Acarophenacidae are parasitoids of a wide
range of insect taxa and attack eggs, larvae, pupae and soft-bodied adults. Female
mites search out and attach to a victim by their mouthparts. Pyemotids are known to
inject a toxin that immobilises the insect. Species of Pyemotes, hay- or straw-itch
mites, often bite non-target organisms including livestock and humans (Chap. 10).
The injected toxin causes intensely itchy and painful blisters. During feeding on insect
hosts, the body of the female parasitoid swells up to contain many eggs. Swelling of
the posterior body is termed physogastry and is also seen in fungus-feeding
members of the Heterostigmatina (e.g. Pygmephoridae). Kaliszewski et al. (1995)
suggest that physogastry in fungivorous ancestors was a preliminary step towards
parasitoidism, rather than parasitoids evolving from predatory ancestors. Eggs typi-
cally remain in the female’s body until they reach an advanced stage of development,
either hatching out as adults upon oviposition or even within the body of the mother
(see Chap. 4). So the parental female acts as a conduit for all nutrition needed
for complete development of her offspring. Wrensch and Bruce (1991) found that
Pyemotes tritici females produce on average 254 offspring, of which 92 % are
females, resulting in a population doubling time of only 1.1 days. The rapid rate of
population increase achieved by such truncated life histories has encouraged some
to consider use of pyemotids as biocontrol agents of stored-product pests. However,
Eickwort (1983) notes that the limited powers of dispersal of these mites, plus the
nasty side effects on humans, would seem to make them poor choices for biocontrol
of insects near human habitations.
Some parasitoid mites may have solved the problem of moving to new sites.
Adult female Adactylidium (Acarophenacidae) are parasitoids on eggs of thrips.
Unengorged female Adactylidium have been found attached to adult thrips, which
implies a phoretic relationship that allows the mites a higher probability of finding
food (Eickwort 1983). This may be similar to the pattern in water mites of larvae
parasitising adult insects whose eggs, larvae or pupae are consumed by postlarval
stages (Proctor and Pritchard 1989).
Sometimes the results of mite behaviour on host fitness are identical to parasit-
oidism, even though a host is not eaten. In the fungivorous Trochometridium
tribulatum (Pygmephoridae), adult females are phoretic on female alkali bees,
Nomia melanderi (Halictidae) (Lindquist 1985; Kaliszewski et al. 1995). As the
female bee provisions each brood cell with pollen sufficient to feed a single
368 9 Animals as Habitats

offspring, a female T. tribulatum disembarks. The bee oviposits and seals the cell.
The mite then empties her spore-bearing pockets (sporothecae) onto the pollen.
The spores germinate and develop into a fungal mycelium growing on the pollen
store. Lindquist (1985) hypothesises that the female T. tribulatum kills the bee egg
or larva almost immediately, as living bee larvae have not been found in cells that
contain mites. The developing bee would have been a competitor with the mite’s
fungus for the brood provisions. The mite feeds on the mycelium, undergoes physo-
gastry and reproduces. After mating, her adult daughters escape from the brood
cell and seek female bees to carry them. Kaliszewski et al. (1995) feel that because
T. tribulatum reproduces only in cells that do not contain a living larva, the effect is
the same as that of a parasitoid.
Likewise, females of Iponemus species (Tarsonemidae) act like parasitoids when
feeding on the eggs of bark beetles from the scolytid subfamily Ipinae. Lindquist
(1969) provides a detailed description of the taxonomy and biology of this group.
Iponemus females use adult bark beetles as phoretic hosts and dismount to feed
when eggs are laid in galleries in pine trees. The mites invariably kill the eggs and
each mite typically does not require more than one egg to complete its development.
Feeding females become physogastric and lay from 40 to 80 eggs. The eggs develop
directly into adults without feeding. Males mate with newly eclosed females, which
then seek out young beetles on which to hitch a ride to the next tree. Lindquist noted
a positive correlation between the size of the unengorged female and that of the host.
He felt this correlation was determined not by the phoretic interaction but by the
parasitoidic one, as larger beetles lay larger eggs suitable for feeding larger mites.

Mutualism

There have been few unequivocal demonstrations of mutualism among mites and
hosts (Eickwort 1990). The most obvious benefit that a mite could provide is reduction
of harmful parasites or competitors. Eickwort (1994) notes that queen bumble bees
bearing Parasitellus (Mesostigmata: Parasitidae) have low levels of parasitic
nematodes. Hunter and Rosario (1988) suggest that some phoretic species of
mesostigmatans benefit their beetle hosts by preferentially feeding on harmful
nematodes (e.g. Dendrolaelaps neodisetus on the southern pine beetle and
Cercoleipus coelonotus on the bark beetle Ips confusus) or on competing species
(e.g. Poecilochirus necrophori pierces fly eggs deposited on mouse carcasses
prepared by its burying beetle host). Eickwort (1990) suggests that fungus-feeding
astigmatans, tydeids and pygmephoroids, and necrophagous mesostigmatans and
acarids may serve a sanitary role, as would histiostomatids that remove bacteria
from host bodies and provisions. For example, Parapygmephorus costaricanus
(Pygmephoridae) feeds on the mecomium of the larvae of its phoretic host, the bee
Agapostemon nasutus, possibly reducing disease in the larvae (Athias-Binche
1991). Likewise, Houck and OConnor (1991) mention that Anoetus (Histiosto-
matidae) occurs with soil-dwelling bees where it filter feeds on microorganisms
Life with Invertebrates 369

Fig. 9.4 Burying beetles such as this Nicrophorus sp. (Silphidae) frequently act as phoretic hosts
for nymphal mites of the family Parasitidae (Mesostigmata) (Photos by HC Proctor)

by skimming the surfaces of provisions and larvae. Scissuralaelaps (Laelapidae),


found on burrowing cockroaches (Macropanesthia, Geoscapheus and Neogeos-
capheus), may contributes to the hygiene of the cockroach burrows by feeding on
potentially damaging astigmatans (Halliday 1993).
The relationships between carrion-feeding burying beetles (Silphidae) and their
ubiquitous mites, species of Poecilochirus (Parasitidae), have been studied by many
authors (e.g. Schwarz and Müller 1992; Brown and Wilson 1994; Schwarz 1996;
Grossman and Smith 2008). Burying beetles (Silphidae) of the genus Nicrophorus
(a.k.a. Necrophorus) are attracted to the odour of rotting vertebrate flesh. If the
carcass is small enough, a male/female pair will bury it and form it into a fleshy
‘nest’ for their larvae. The parents remain with their offspring until the larvae are in
an advanced stage of development and then leave to search for other carcasses.
Deutonymphal Poecilochirus are phoretic on the beetles (Fig. 9.4) and dismount to
feed on fly larvae and nematodes that compete with the beetle larvae for the carcass,
complete their development, mate and reproduce in the nest chamber. The next
generation of mites may depart with the parents or with the newly emerged young
beetles. For at least one species, large male beetles are preferentially selected as
carriers, possibly on the basis of chemical cues (Grossman and Smith 2008).
Poecilochirus deutonymphs are also able to switch from their original host to a new
one, should two beetles meet and interact (Brown and Wilson 1994). Wilson and
Knollenberg (1987) performed 18 experiments on several species of Nicrophorus
and their Poecilochirus. In most experiments, the mites had neither positive nor
negative effects on host fitness but in four experiments the mites had positive effects.
Negative effects were observed in a few replicates in which mites existed at densi-
ties rarely observed in nature. Previous studies of the interactions between mites and
370 9 Animals as Habitats

the beetle N. tomentosus suggested that deeply buried carcasses were protected
from fly oviposition, while shallowly buried ones received many fly eggs. If fly
larvae hatched in the absence of the mite P. necrophori they outcompeted the beetle
larvae (Wilson 1983); so mites appeared most beneficial when carcasses were
shallowly buried. However, in Wilson and Knollenberg’s experiments they found
that N. defodiens, which puts carcasses in shallow depressions covered by leaves,
did not suffer when its mites were removed. They modified (but did not test)
Wilson’s (1983) hypothesis by suggesting that flies may be leery of walking under-
neath leaves to oviposit for fear of spiders. One benefit for the beetle N. orbicollis
was that mites reduced the number of nematodes carried to future carcasses.

Acarinaria

The most puzzling aspect of mite/invertebrate symbioses is that hosts often appear
to encourage their own colonisation by having structures seemingly designed to
carry mites. The presence of these pits or pockets, termed acarinaria, has led
many to assume that the mites benefit their hosts in some way. Carpenter bees
(Megachilidae) often have prominent acarinaria. Skaife (1952) studied the relationship
between one species of African carpenter bee, Mesotrichia caffra, and its laelapid
pocket mite, Dinogamasus braunsi.
Most female carpenter bees have acarinaria (on abdomens), but all males and
some females lack them. Up to 17 young female D. braunsi can occupy each pocket.
At the time of Skaife’s paper, 36 species of Dinogamasus had been found, all asso-
ciated with bees. During the autumn and winter the mites remain in the acarinaria
and die if removed from the host, even if honey and pollen are provided. When the
female bee nests in the spring, one or two mites disembark with each oviposition
event and are sealed in a chamber with the bee egg and food mass. Because the bee
starts by laying female eggs and ends with male eggs, the first mites to disembark
are imprisoned with a female larva and later mites get males. Mites do not do much
until the larva has fed and enters the pre-pupal stage. At this point the mites creep
over the resting larvae and become very fat, “as thick from above downwards as
they are wide”. The D. braunsi die at this point if removed from the pre-pupa but, if
placed in a glass tube with nothing but the larva, they flourish. Thus the mites do not
feed on provisions or on other species of invertebrates typically present in the cham-
bers. Skaife concluded that they must feed on exudate from the bee’s skin. When the
bee larva pupates, the mites lay their eggs on the pupa. Skaife never observed males
and so assumed the mites were totally parthenogenetic. Larvae and nymphs feed on
pupal exudate. After 4–5 weeks the adult bees emerge and demolish the divisions
between cells. Mites and bees freely intermix and the mites crawl into the acarinaria
of female bees (so those locked in with male eggs are not doomed).
Skaife observed that the mites do not appear to harm their hosts in any way.
Pupae heavily infested with D. braunsi produced adults as vigorous as those from
pupae without any mites at all. Conversely, they did not seem to benefit their hosts.
Skaife removed all the mites from some nests and the bees reared in them behaved
Life with Invertebrates 371

Fig. 9.5 This Ancistrocerus


antilope wasp (Vespidae) is
bearing a heavy load (arrow)
of phoretic Kennethiella
trisetosa mites (Astigmata,
Winterschmidtiidae) that
hope to share brood chambers
with the offspring of the wasp
(Photo by HC Proctor)

normally. Furthermore, a small percentage of female M. caffra lack the abdominal


pouch and do not harbour mites. The presence of mites does not cause acarinaria, as
pupae without mites did not grow up lacking these structures. It is possible that the
benefit of D. braunsi for their hosts was too subtle for Skaife to observe or that the
benefits would manifest only in certain conditions (e.g. if the nests were infested
with dangerous fungus or bacteria, or with parasitoids like Trochometridium,
see section “Acarine Parasitoids”).
Another astonishing study of hymenopterans with acarinaria is that of Cowan (1984),
on the solitary wasp Ancistrocerus antilope (Vespidae: Eumeninae) and its symbiotic
astigmatan Kennethiella trisetosa (Winterschmidtiidae) (Fig. 9.5). Female wasps
nest in tubular cavities in wood, where they leave paralysed caterpillars as food
for their offspring. Only one wasp larva develops in each cavity. Deutonymphs of
K. trisetosa are phoretic on adult wasps; when the female wasp oviposits, several
deutonymphs crawl off into the cell. The mites rest on the wasp egg and, although
they do not appear to feed, they do get fatter. It is possible that they absorb water or
egg fluids through their integument. After moulting to tritonymphs, the mites share
the haemolymph of the paralysed caterpillar with the wasp larva.
By the time the wasp pupates, the mites are adult. They then feed on haemo-
lymph of the quiescent pre-pupal wasp. The mites endeavour to stay at the posterior
dorsal region of the host when not feeding, possibly because this is the hardest
place for the host to reach with its mouthparts. Nevertheless, Cowan found that
female wasp larvae were able to decimate their mites (only 16 survived of 123 on
female larvae versus 132 of 137 on male larvae). Cowan found no evidence that mite
deutonymphs distinguished between eggs destined to become males or females.
372 9 Animals as Habitats

A virgin female K. setosa develops a single large egg that hatches internally into
a larva, which immediately moults into a male protonymph. Females then give birth
to the protonymphs, which are very small (probably because the larva does not
feed). The male protonymphs transform to tritonymphs without an intervening
deutonymphal stage. After giving birth to these puny males, females may mate with
large males that developed from male deutonymphs that entered the wasp’s cell
before it was sealed. However, Cowan found only 9 % of all deutonymphs entering
cells became males. So, most females (114 of 132) had to mate with small males.
Arrhenotokous production of small males appears to be a bet-hedging strategy by
females in case they are trapped without any mates. Large males are not aggressive
to small ones, possibly because they do not represent serious rivals; however, small
males attack and kill each other.
Mated female mites deposit their eggs (up to 125) on the wasp pupa and mite
larvae and protonymphs feed on the pupa apparently without causing it harm. By
the time the wasps emerge from their pupal cases, the mites are deutonymphs and
attach only to the acarinaria on the wasps’ propodeum. The maximum number that
can squeeze into an acarinarium is 400, and any extra mites end up falling or being
scraped off. As mentioned above, female wasp larvae kill all the mites in their cells
prior to pupation but male wasp larvae do not. So adult female wasps emerge with-
out mites but the acarinaria of males are packed with them.
How then does K. setosa get into the egg chambers of the next generation of
wasps? Females are infested venereally when they copulate (male wasps may
remain attached to females for several minutes), at which point the deutonymphs
run off the male and into the female’s genital chamber, there to await her oviposition.
Deutonymphs that do not transfer to female wasps are doomed because males do not
visit nests. Female wasps often mate with different partners, allowing the possibility
of different mite genotypes boarding a given female. The occasional mite that
avoids being killed by female wasp larvae also increases the potential for outbreeding.
The aggression shown by small male mites suggests that mixing of genotypes in a
single cell is common.
Why do female wasp larvae kill mites but male larvae do not? If the mites are
being beneficial (e.g. destroying parasites), why would this be good for males but
not females? And why do the females have acarinaria if they almost never carry
mites? Cowan wondered whether the presence of mites in the female’s venereal
chamber may inhibit intromission by subsequent males and this acarine chastity belt
may be advantageous to males in sexual competition. Cowan also noted that the
mites a male wasp transmits do not persist on his descendants but only on his mate’s;
because of the haplodiploid genetic system, male hymenopterans have no sons (haploid)
and their diploid daughters usually destroy the mites transmitted by their fathers.
Athias-Binche (1991) expresses great puzzlement over the relationship between
Kennethiella and Ancistrocerus. Why would the host develop an acarinarium for
this not obviously beneficial mite? Are acarinaria to protect mites and ensure their
presence or to reduce disturbance to the host caused by having irritating little mites
all over one’s body (the ‘mite pocket’ hypothesis; see section “Mites Associated
with Reptiles”)? Clearly, a great deal is unknown about the evolutionary processes
Life with Vertebrates 373

resulting in the apparent adaptation for mite carrying in many bees and wasps,
as well as for other potential mutualisms. Manipulative experiments such as
those of Wilson and Knollenberg (1987), rather than observational studies, seem the
only way to solve these mysteries. And this is exactly what Okabe and Makino
(2008a, b, 2010) did to test the pros and cons for another species of eumenine wasp
of carrying mites.
The wasp Allodynerus delphinalis (Eumeninae) has acarinaria that house
Ensliniella parasitica mites (Winterschmidtiidae). Juvenile wasps are often attacked
by parasitoid hymenopterans (Melittobia acasta, Eulophidae) in the nest. Okabe &
Makino manipulated the densities of mites in a host cell and showed that the mite
protected wasp juveniles from M. acasta by continuously harassing the parasitoid
and stabbing it with their chelicerae. A density of ten mites per cell resulted in
100 % mortality of the parasitoid. But they also observed that the benefits to the
wasp of inoculating their brood cells with mites depended upon the presence of
M. acasta. When the parasitoid was absent, the mites could potentially do more
harm than good by feeding on haemolymph of the juvenile wasps. Increases in wasp
mortality were only observed when the mite densities were experimentally increased
to more than the maximum load observed in the field, however, and so it may be that
E. parasitica typically acts as a harmless commensal when M. acasta is absent.

Life with Vertebrates

As mentioned in the introduction to this chapter, we will describe the associations


between parasitic mites and vertebrates rather superficially, as acarines of medical
and veterinary importance are well covered in other texts and in Chap. 10. This,
however, does not make them any less fascinating.

Mammals and Their Homes

As clothes make the man, so hair makes the mammal. Some mammals are hairier
than others: heavily furred artiodactyls have 233–6,916 hairs per square centimetre,
while otters have 35,000–51,000 (Kim 1985a). Hair provides a wonderful climate-
controlled environment for the mites that swarm over and in the skin surfaces of
most species of mammals. Although sweat glands are lacking in rodents and
elephants, most mammals have numerous skin glands that have been invaded by
mites. According to Kim (1985a), bats have perhaps the most diverse array of both
sudoriferous and sebaceous glands found within the Mammalia. This rich diversity
of glands harbours a similar diversity of mites, although apparently no insects have
invaded the glands of mammals. The ecological and evolutionary relationships
between this class of vertebrates and their acarine associates are thoroughly
discussed in the book edited by Kim (1985b) and much of what follows comes
from this source.
374 9 Animals as Habitats

Mites in Mammalian Nests and Larders

Although we consider birds to be the quintessential nest builders among vertebrates,


most mammals also live in nests of some sort. Such abodes range from long-term
colonies of thousands of bats to the mats of boughs built nightly by chimpanzees.
Nests and dens are places to rest, to give birth and to nurse young, and to eat. So
nests accumulate such potential foodstuffs as arthropods, food bits, fungus, faeces,
skin flakes and blood. As well, nests are warm and humid when the host is home.
Among mammal-associated mites, everything from free-living nidicoles to obligate
haematophages often co-exist in the same nest (Radovsky 1985). For example, species
in several genera of Laelapidae occur as free-living predators as well as nidicoles in
vertebrate and invertebrate nests. Species of Androlaelaps and Haemogamasus
(Laelapidae) vary in their requirement for blood for successful development and
this requirement appears to vary with the species’ ability to draw blood. Evolutionarily
speaking, nest mites may acquire a ‘taste’ for blood by initially preying on blood-
feeding parasites (see review by Durden 1987) and then moving onto fresh
sources of blood.
For humans and some rodents (e.g. squirrels) and lagomorphs (e.g. pikas),
construction activities are not confined to nest making but include the manufacture
of larders for storage of food. To our knowledge there have been no detailed studies
of mites of rodent larders but there is a large body of literature on human stored-
product mites (e.g. Hughes 1976; OConnor 1979, 1982). Species in 34 genera in ten
families have independently invaded stored products and/or house dust (OConnor
1979; Colloff 2009). Most of these mites are astigmatans that feed on fungus also
infesting the stores.
OConnor (1979) feels that because man has been storing food in quantity for
only about 10,000 years, this is insufficient time to allow the evolution of such a
large number of taxa purely through association with man. Rather, stored-product
Astigmata have repeatedly invaded human habitations from similar natural habitats.
OConnor divides these mites into four groups. Some are ancestrally associated
with a specific product that humans bring inside (e.g. Carpoglyphus occurs on
rotting fruit). A second group is associated with widespread ‘field’ resources,
e.g. Tyrophagus spp. are widely distributed in grassland and litter habitats. The largest
group is that associated with mammal nests. An example of this is Acarus, which is
found in a wide range of mammal nests including bat roosts. Finally, there are those
associated with bird nests such as Dermatophagoides, the infamous house-dust
mites. Solid food is not the only refreshment sought by astigmatans. Quintero and
Acevedo (1991) found Carpoglyphus in 33 % of 84 samples of pulque, a beverage
made from the fermented sap of agave.
As for arthropods living in nests of non-human mammals (Durden 1987), not all
mites associated with human habitations and stores are astigmatans interested in
feeding on our foodstuffs. Sinha (1979) notes that approximately 50 species of
prostigmatans, mesostigmatans and astigmatans are associated with stored grains.
These mites may include granivores (although only a few acarids directly eat the
kernels), others feed on non-grain parts such as leaves and stems in hay, and some
Life with Vertebrates 375

are predators and parasitoids of other grain inhabitants. Likewise, van Bronswijk
(1979) views domestic dust as an ecosystem. House dust is composed of skin flakes,
cotton fibres, paper fibres, wool, synthetic fibres, mite droppings and exuviae, etc.
between 0.01 and 1 mm in diameter. Dust together with dust-associated fungi are
eaten by a number of astigmatan species in the genera Dermatophagoides, Hirstia
and Euroglyphus (Pyroglyphidae). These mites are preyed upon by Cheyletus,
pseudoscorpions and probably silverfish and psocopterans (‘book lice’). Other
mites common in house dust include Glycyphagus, Chortoglyphus, Blomia,
Cosmochthonius and Amnemochthonius, which probably feed on fungus or food
scraps rather than dust per se. Van Bronswijk (1979) notes that the structure of the
arthropod community in bird nests is similar to that of house dust except that
species richness is higher in bird nests than in human dwellings.
Dust mites are the most infamous of nidicolous acarines because of their asso-
ciation with ‘dust allergies’ (Colloff 2009). Arlian (1989a) describes the biology of these
animals. Dust mites belong to the astigmatan family Pyroglyphidae, with 14 species
reported from house dust in various parts of the world. The three most common
species are Dermatophagoides pteronyssinus, D. farinae and Euroglyphus maynei.
They live in our rugs, furniture, mattresses, bedding and other areas that accumulate
organic detritus and maintain a high level of humidity. Pyroglyphid mites feed on
cast skin flakes (derma = skin, phagos = eat), hair and other detritus, together with
the microbes that grow on this minute refuse. They are the most important source of
allergens in house dust. Allergens produced by these mites cause severe allergies,
rhinitis, eczema and asthma that affect 50–100 million people worldwide. These
allergens are contained primarily in dust mite faeces. These are dry pellets 20–50 μm
in diameter that are covered with a peritrophic membrane (a membrane produced in
the gut of the mite). On average, about 20 pellets are produced by each mite each
day and they readily become airborne. Other allergens are present in the skin of the
mite; so exuviae and bits of dead mites also cause reactions.
The phrase ‘dry as dust’ suggests that house dust may be an arid environment for
tiny animals. Dust mites have an ingenious solution to the problem of water balance
(shared by numerous other Astigmata). A small gland opening at the base of the first
pair of legs (supracoxal gland) secretes a solution of sodium and potassium chloride
into a gutter that runs to the mouth opening. The salt solution is hygroscopic
and absorbs water from the atmosphere, down to a species-specific critical
threshold of relative humidity (often 70–75 % RH at 25 °C). At lower humidities,
the solution crystallises and blocks the gland opening, preventing water loss. So,
dust mites obtain water from the atmosphere at relative humidities above the critical
threshold but are unable to ‘drink’ at lower relative humidities and will desiccate
and die if low RH is maintained for very long. Any conditions that help to maintain
high humidities may encourage house dust mite populations. Poorly ventilated
homes, those with furnishings that retain humidity (deep rugs, upholstered
furniture) or those that use humidifiers or evaporative coolers will tend to have high
average relative humidities, especially in the areas where the mites live. In extreme
circumstances, densities can reach as high as 3,500 mites per gram of dust
(Kettle 1995).
376 9 Animals as Habitats

Fig. 9.6 Spinturnicids, such


as this male from a bent-wing
bat (Vespertilionidae:
Miniopterus sp.), are one of
several families of mites that
occur only on and in the
bodies of bats (SEM by
DE Walter & M Shaw)

Mites on and in Mammals

An intimidating diversity of mites is permanently associated with the bodies of


mammals (Kim 1985a; OConnor 1994). No group of mammals, nor any part of the
body, is excepted from this acarine onslaught. For example, the Halarachnidae live in
respiratory organs of pinnipeds and primates, Hystrichonyssidae in ears of porcupines
and the Demodicidae in the hair follicles of just about all mammals. Perhaps because
of their diversity of skin glands and tendency to roost in large aggregations, bats are
host to a particularly large array of parasitic mites. In the Parasitiformes, many ticks
and members of several families of Mesostigmata, Spinturnicidae (Fig. 9.6),
Spelaeorhynchidae and Macronyssidae, have bat hosts. Among the Prostigmata,
representatives of the Myobiidae and Trombiculidae attack chiropterans. In the
Astigmata, four families – Chirorhynchobiidae, Gastronyssidae, Rosentsteiniidae
and Teinocoptidae – are exclusive to bats. In fact, epizoic mites of bats include 18
families and more than 1,000 species (Morales-Malacara 1996).
Various parts of the mammalian skin are consumed by the different groups of
parasitic mites. Kim (1985a) states that parasitic mesostigmatans, such as laelapids,
Life with Vertebrates 377

dermanyssids, macronyssids and spinturnicids, are basically polyphagous.


Haemogamasus ambulans and Brevisterna utahensis utilise both fluid and dried
blood of vertebrates, flea faeces and living or dead arthropods but rarely do they
penetrate the skin. Echinolaelaps echidninus and Haemolaelaps glasgowi are also
general feeders but usually penetrate the skin. Ticks feed on blood and lymph,
halarachnids feed on mucous tissue. Sarcoptids, psoroptids and listrophorids feed
on epidermal detritus and secretions of dermal glands, and some may feed on blood.
Sarcoptes burrows into the cornified epidermis by attaching itself to the substratum
with ambulacral suckers and cutting a channel into the skin with its chelicerae and
edges of the foretibiae. Nutting (1985) mentions that prostigmatans on mammals can
be found clinging to hairs (e.g. myobiids), penetrating the epidermis (e.g. trombiculids)
or dermis (e.g. psorergatids) or be semi-endoparasites (e.g. demodicids) in the
sebaceous glands of hair follicles.
Humans are fascinated and appalled by the idea that they may be infested with
almost invisible mites. Demodex follicularis and D. brevis (Demodicidae) inhabit
the follicles and glands of the human face where they feed on subcutaneous secre-
tions and interstitial fluid. Although widespread, Demodex mites are not necessarily
found on every individual. Sengbusch and Hauswirth (1986) examined 370 human
volunteers for the two species of anthropophilic Demodex. One or both species
were detected on about 55 % of volunteers, 31 % had D. brevis only, 11 % had
D. folliculorum only and 14 % had both. Humans appear to accumulate mites over
their lifetimes or at least their populations rise to detectable levels with age: 29 % of
volunteers aged 25 and younger had mites, 53 % of those aged 26–50 had them and
67 % of those 51–90 had them. There appears to be an effect of host sex on Demodex
infestation, as 198 of 327 males (61 %) were infested vs. 5 of 43 females (12 %)
(X2 = 13.16, P = 0.0003). The authors note that sebaceous glands, the site of D. brevis,
are differentially affected by hormones.
Humans are not alone in this infestation. Approximately 150 species of Demodex
have been collected from more than 35 families of mammals (Nutting 1985).
Hosts include both marsupials and placental mammals but demodicids have not
been collected from monotremes, despite much searching. Every species of
(non-monotreme) mammal may host at least one specific demodicid and myobiid
( > 300 spp. described) (Nutting 1985). In contemplating the possession of numer-
ous species of host-specific mites on most mammals, Nutting (1985) comments:
“[b]ecause of the variety and often disjunct distribution of habitats on one mammal,
we can look on the mammal as a wandering Galapagos archipelago with each island
(e.g., an eyelid) having several differing habitats.”
The variety of ‘habitats’ occupied by astigmatan associates of mammals serves
as a good example (Fain and Hyland 1985). Gastronyssidae live attached to the
gastric mucosa, corneae and nasal cavities of bats. Rhyncoptidae are fixed in the
hair follicles of their hosts (hystricid rodents and both neotropical and afrotropical
monkeys). Audycoptidae are in hair follicles of cebid monkeys, ursids (bears) and
procyonids (raccoons). The Listrophoridae and Atopomelidae – ‘fur mites’ – are
permanently attached to the bases of hairs throughout most of their life cycle using
an attachment structure on their ‘sternum’ (Fig. 9.7). They probably eat fatty
378 9 Animals as Habitats

Fig. 9.7 The Koala Fur Mite (Koalachirus perkinsi, Atopomelidae): (a) Two mites clinging to a
koala hair (scale bar = 100 μm). (b) As well as having the first two pairs of legs modified for grasp-
ing hairs, this astigmatan also has a ventral groove that clasps the hair shaft (scale bar = 10 μm)
(SEMs by DE Walter)

substances produced by hair follicles. The Chirodiscidae are also known as ‘fur mites’.
They are laterally compressed and attach to the hair shafts of their hosts with modified
first and second pairs of legs. The Myocoptidae also attach to the hairs of their hosts
using their legs, but use their second and four pairs. The above four families of fur
mites belong to the Astigmata, but the Myobiidae (Prostigmata) have independently
colonized this habitat. Morphological adaptations for living in mammalian fur are
discussed by Labrzycka (2006). The Psoroptidae are astigmatan skin mites that live
on the surface of the skin at the base of thick crusts of exudates produced by the host
in response to irritation caused by the mites. Skin ailments resulting from infesta-
tion by these mites are known as psoroptic mange. The most infamous mange mites
are in another astigmatan family, the Sarcoptidae, or scabies mites.
Sarcoptids have chelate-dentate chelicerae and spines on their fore tarsi that
allow them to burrow in the corneous layers of the skin. The skin disease that
Sarcoptes scabiei causes in humans is termed ‘scabies’. Sarcoptes is thought to
contain only one species – the human itch mite, S. scabiei – and the other 30 species
previously described in this genus are considered host races or varieties (Fain and
Hyland 1985; Arlian 1989b). This mite has been collected from the epidermis of
more than 40 different hosts representing seven orders of mammals. Fain and
Hyland (1985) believe that the variability of S. scabiei is the result of continuous
interbreeding of the strains living on man and on domestic and wild mammals. Even
such recently encountered mammals as koalas and wombats have become infected
with scabies. Morphological variation between the strains is small or absent, and
ITS-2 also shows no host-related variation (Alasaad et al. 2012). However, there is
some evidence of host-related population genetic differentiation in microsatellites
(Rasero et al. 2010) as well as physiological differences among scabies mites from
Life with Vertebrates 379

different hosts. Experimental attempts to transfer scabies from dogs to mice, rats,
guinea pigs, pigs, cattle, cats, goats and sheep were unsuccessful, although most of
these hosts are parasitised naturally by S. scabiei. Oddly, scabies from dogs could
be transferred to New Zealand white laboratory rabbits, although scabies from
humans and pigs could not. Humans often contract transient scabies through infested
dogs, pigs and horses (Kettle 1995). Mites may live and reproduce for up to 13 weeks
in the ‘wrong’ host but these infections are usually self-limiting. However, appropriate
control studies are lacking and one wonders how often infections on the ‘right’ host
are also self-limiting.
Sickness and poor growth of the host appear not to be due to the direct loss of energy
to the mites but to toxic secretions from the mites together with the host’s immune
responses. Secondary bacterial infections are also common. In immune-compromised
humans, untreated scabies eventually leads to scaly skin and the production of skin
crusts from hyperkeratosis. Arlian (1989b) notes that hyperkeratotic crusts may be
up to 2 cm thick and can be easily pried off the host’s skin. Mites are found in their
greatest densities on the undersurface of the crusts; in an infected laboratory rabbit,
there were more than 1,400 mites per cm2 of skin surface.

Mites on, in and Around Birds

Mites are associated both with the homes and the bodies of many birds (reviewed in
Proctor and Owens 2000). Avian mites are currently of great interest to behavioural
ecologists as potential agents of sexual selection (see section “Parasitic Mites and
Mate Choice by Hosts”); however, in this section we will discuss bird mites, not as
evolutionary intermediaries, but as products of selection themselves.

Mites of Birds’ Nests

Many of these mites are also found in nests of other vertebrates, including those
of humans. The most obvious of the shared taxa are species in the astigmatan
family Pyroglyphidae, which includes the human ‘house-dust’ mites (OConnor 1982;
Colloff 2009). The best-studied nest mites are members of the mesostigmatan
family Dermanyssidae. Dermanyssus gallinae is a ubiquitous parasite of domestic
fowl and is also commonly found on wild birds such as pigeons, starlings and
sparrows (Kettle 1995). When these hosts are not available D. gallinae will bite
mammals and, as these birds are often synanthropes, this often includes humans.
The mites feed on roosting birds at night and hide in crevices in and around the nest
by day, where they also oviposit. Masan (1997) describes the population dynamics
of the Dermanyssus hirundinis in nests of the penduline tit, Remiz pendulinus.
As in house dust (see section “Mites in Mammalian Nests and Larders”), whole
food webs of mites can inhabit nests. Burtt et al. (1991) used Tullgren funnels to
extract the arthropods from nests of tree swallows, house wrens and eastern
380 9 Animals as Habitats

bluebirds. A species of detritus-feeding dust mite, Dermatophagoides evansi


(Pyroglyphidae), occurred in all nests of all three species. The predatory mite
Cheletomorpha lepidopterorum (Prostigmata: Cheyletidae) was the least widely
distributed (none in bluebird nests, and inhabiting 33–46 % of the other species’
nests); however, it sometimes occurred in large numbers (up to 4,000 per nest). It
preys on mites and other arthropod nest inhabitants. The blood-feeding Dermanyssus
hirundinis (Mesostigmata: Dermanyssidae) was ubiquitous in nests of all three
birds. The mean number of D. hirundinis per nest ranged from 5,384 for eastern
bluebird to 12,675 for the house wren. In nests with C. lepidopterorum, mean popu-
lation sizes of other mites were typically lower than in nests without this predatory
mite, similar to what Durden (1987) reports for mammals with and without associ-
ated predatory arthropods. Lesna et al. (2012) discuss the possibility of using preda-
tory mites to control the haematophagous nest mite D. gallinae in poultry barns.
Argasid ticks are also nest parasites, visiting the host at night, feeding for a few
minutes and then retreating to a nearby nook in or near the nest. Ticks are the most
conspicuous group of ectoparasites of seabird colonies (Duffy 1991) and also are
abundant around colonies of birds such as swallows (Chapman and George 1991).

Mites Permanently Parasitic on and in the Bodies of Birds

There are many mesostigmatan, astigmatan and prostigmatan species that live on
and in the skin and in the respiratory systems of birds. The northern fowl mite,
Ornithonyssus sylviarum (Mesostigmata: Macronyssidae), is a haematophagous
pest of poultry and a great many species of wild birds throughout the world. The
adults remain on the host except during oviposition and in heavy infestations can
give the bird’s plumage a blackish appearance. Members of the mesostigmatan fam-
ily Rhinonyssidae occupy the respiratory passages of almost all species of birds and
are morphologically simplified compared to free-living Mesostigmata.
Birds are infested with many species of astigmatans that feed on various bodily
parts (OConnor 1982, 1994). Most Hypoderatidae are subcutaneous parasites as
deutonymphs but are nidicoles in other stages (making them protelean parasites).
The life cycle of the best known species, the pigeon-associated Hypodectes
propus, is highly compressed, with only eggs, deutonymphs and adults known.
In this species the females have reduced mouthparts and do not feed, while males
have huge chelicerae that likely have a sexual rather than a feeding function.
The deutonymph, like all astigmatan deutonymphs, has no obviously functional
mouth. Nevertheless, once embedded in the skin of a pigeon, it increases hugely in size.
Presumably, nutrients are absorbed directly through the deutonymph’s integument,
as has been proposed for deutonymphs of Hemisarcoptes cooremani on ladybird
beetles (see section “Phoresy Precedes Parasitism”). Thus in this species, the instar
that lacks both mouth and gut obtains the nutrition for the entire life cycle.
The Knemidokoptidae are also astigmatan burrowers in the skin of birds. They
often cause a pathological condition in domestic poultry called ‘scaly-leg’, in which
the skin of the legs swells and crusts over. Laminisoptidae are also skin burrowers
Life with Vertebrates 381

Fig. 9.8 This long-toed


Turbinoptes strandtmanni
(Turbinoptidae) was living
in the respiratory passages
of a ring-billed gull (Laridae:
Larus delawarensis)
(SEM by HC Proctor)

and cause nodules in the flesh of the host. The respiratory passages of birds contain
the Turbinoptidae (Fig. 9.8), which inhabit the dry regions of the upper respiratory
tract. The Cytoditidae inhabit the lungs, air sacs and tracheae where they can cause
some pathology.
Prostigmatans are also represented among the avian parasites. The chigger
Neoschongastia americana (Trombiculidae) is a serious pest of domestic turkeys
(Kettle 1995). An interesting prostigmatan parasite is Pneumophagus bubonis
(Epimyodicidae) from the lungs of the great horned owl (Fain and Smiley 1989).
This mite was originally placed in the Cloacaridae, a family named for a number of
genera found in the cloacas of freshwater and marine turtles. Most epimyodicids
occur in the internal tissues of shrews, moles and mice, and their similarity to the
Cloacaridae could be interpreted as their having had a common origin in an ances-
tral synaspid reptile (Bochkov 2002), though this would require a great deal of
extinction of this mite lineage in phylogenetically intermediate hosts.

Mites on and in Feathers

The mites that live among feathers are perhaps the most fabulously ornamented of
all Acari (Fig. 9.9). Their bizarre beauty has led to much taxonomic study (see Gaud
and Atyeo 1996) but surprisingly little is known of their basic biology (see review
in Proctor 2003). Although members of many major taxa occupy this habitat,
‘feather mites’ is often used to refer to certain feather-dwelling taxa in the Astigmata
and it is on this diverse group that we concentrate.
Gaud and Atyeo (1996) have produced a massive two-volume work that
summarises the biology and taxonomy of the astigmatan feather mites of the
world. They note that plumage has been invaded by a wide range of both insects
(fleas, bedbugs, hippoboscids, mallophagans) and mites (ticks, mesostigmatans,
prostigmatans, astigmatans). Some are in the plumage for only a short time; others
live there for all of their lives. While Nutting (1985) draws an analogy between the
382 9 Animals as Habitats

Fig. 9.9 An elaborate feather


mite (Opisthocomacarus
umbellifer, Pterolichidae)
(after Gaud and Atyeo 1996)
(Image by HC Proctor)

surface of mammalian skin and the Galapagos Islands, Gaud & Atyeo state that
“[i]n the jungle of feathers there swarms a complete fauna.” Some inhabitants of
this jungle live in the branches (feather surface), some on the ground (skin), others
in the tree trunks (feather rachis) and finally, others infest the roots (the feather
follicles). Some arthropods eat the plumes, some graze on the detrital ‘epiphytes’ of
the feathers, some suck blood and lymph, and others act as predators.
Astigmatans constitute the most numerically important inhabitants of the plumage.
There are more than 2,000 named species in 33 families in two superfamilies that
are collectively termed ‘feather mites’ sensu stricto. They inhabit the surface of the
feathers (plumicoles), the skin under the feathers (dermicoles) and even the interior
of feathers (syringicoles). Feather mites have been recorded from every order of
birds except for Rheiformes (rheas). It not clear why rheas are so far mite-free. It is not
because they are ratites (flightless, running birds) because there are feather mites on
Struthioniformes (ostriches), Casuariiformes (cassowaries) and Apterygiformes
(kiwis). So, perhaps we just have not looked hard enough for mites of rheas.
Living on or in feathers requires some special morphological adaptations (Dabert
and Mironov 1999). For plumicoles, especially those on flight feathers, there is a
great danger of being dislodged by the bird’s movements. So feather mites often
have bizarre body shapes and strange setal appurtenances that allow them to wedge
themselves securely between the barbules. This seems to be especially true for male
feather mites of diving birds, which may require extra anchoring devices in order to
combat hydrodynamic forces while holding onto females with their anal suckers
and hind legs; such males are frequently strongly asymmetrical (see also Chap. 5:
section “The Adventurous Acariformes: Astigmata”) (Fig. 9.10). Atyeo and Gaud
(1979) note that unhealthy birds unable to flap their wings may take on a brown cast
because feather mites, normally displaced from much of the plumage by flight, can
swarm with impunity over all the feather surfaces. Eggs are safely glued to the
feather substrate (Pérez 1996).
Moulting often occurs at particular sites of safety on the feather or even within
the cast skins or egg shells of other feather inhabitants, a behaviour that Pérez
and Atyeo (1984a) call thanatochresis: “the utilization of cadavers, secretions,
Life with Vertebrates 383

Fig. 9.10 Male feather


mites, (a) Michaelia sp.
(b) Dinalloptes sp.,
associated with the
double-crested cormorant
Phalacrocorax auritus.
The asymmetrical legs and
bodies of these males may
allow them to better brace
themselves between the
feather barbs of the wings
when the host bird dives
beneath the water
(photos by HC Proctor)

skeletal pieces, excrements and other products of one species by living individuals
of a second species, but not for food…”. Sometimes feather mites moult inside the
cast skins of other members of their species, building up long exuvial chains.
Feather mites also moult inside the empty eggshells of feather lice. Why engage in
thanatochresis? Perhaps as protection from predation, as fragments of feather mites
can often be observed in the guts of lice. It is also possible that moulting inside
the cast skin of another arthropod reduces loss of moisture or provides a frictional
surface to moult against (like snakes rubbing off skins against rocks).
Are all feather mites parasites? Although they may be referred to as such in the
literature, there is actually little evidence that mites living on feathers damage the
host. In fact, the weight of correlative evidence is that these mites are harmless
commensals (Galván et al. 2012). The generic name of one group of plumicolous
species, Analges (Analgidae), means ‘no pain’ and is based on the observation
that a heavily infested bird usually appears quite healthy and happy. But another
analgid, Megninia, may cause financial losses in poultry industry because heavy
infestations cause birds to pull out their feathers and leads to a decline in weight and
egg production. Similarly, large populations of Dubininia melopsittaci (Xolalgidae)
may cause the host budgerigar to pull out its feathers. Syringicolous mites often
feed on the medulla, or pith, of the feather shaft. Perhaps this weakens the feather
but it is not clear.
384 9 Animals as Habitats

So, if most feather mites do not harm their hosts, how do they make their living?
In his monumental work on feather mite biology, Dubinin (1951) reports only
occasionally having observed tiny feather fragments inside feather mites. Instead, the
guts of a large number of species contained fungal spores. In our own preparations
of feather mites we see spores, hyphae, and also pollen (especially in mites from
nectarivorous birds such as lorikeets) (Proctor 2003). It may be that feather mites
deliberately pick off fungal hyphae and spores from the host’s feathers or they may
feed on uropygial oils smeared on feathers and only accidentally ingest these particles.
An unpublished experiment by John Clark in New Zealand (pers. comm. Proctor)
provides tantalising evidence of deliberate fungivory in one species of feather mite.
He dusted the wing feathers of a chaffinch with spores of the inky top mushroom
(Coprinus sp.) and later removed the feather mites and observed that their guts were
black with the fungal particles.
Cross-species studies indicate that feather mite load is positively correlated with
the size of uropygial glands of their host species (Galván et al. 2008). Some obser-
vations suggest that feather mites may benefit their hosts by cleaning feathers of
old uropygial oil and fungi. Blanco et al. (1997) looked at Gabucinia delibata
(Gabuciniidae) on the red-billed chough, Pyrrhocorax pyrrhocorax. Rather to their
surprise, they found a positive relationship between the number of mites per bird
and the bird’s body condition. The authors concluded that mites had no detrimental
and possibly had positive effects on the hosts. This seems the general relationship
between feather mite load and host condition in passerines (Galván et al. 2012).
In another comparative study, Soler et al. (2012) found that birds with high feather
mite loads tended to have lower bacterial loads on shells of brooded eggs. So most
feather mites would appear to be commensals at worst and microbe-eating
mutualists at best; however, this hypothesis has rarely been tested experimentally.
A study by Pap et al. (2005) is a rare exception. They removed vane-dwelling feather
mites from some barn swallows using an insecticide and left other swallows with
their usual mite load. The commensal nature of the mites was supported by their
observing no differences in the two groups of birds in several measures of host
fitness, including breeding performance and survival. These authors also noted that
female swallows with higher natural mite loads tended to start breeding earlier. It is
thus possible that hosts in very good condition provide more or better quality food
to mites and so support larger numbers of them.
Transmission of feather mites appears to be through direct physical contact
between mite-bearing birds and their offspring or mates, or through other social
contact (e.g. communal roosting). Mite transfer usually occurs in the nest. When the
flight feathers of young birds are 50–75 % developed, mites transfer from the par-
ents to the fledglings (Atyeo and Gaud 1979). According to Gaud and Atyeo (1996),
parasitic cuckoos whose young are raised by small passerines do not acquire the
feather mites of the foster parents. Rather, they pick up their acarofauna later in life
through sexual and social contact with other cuckoos. However, the African diederic
cuckoo, Chrysococcyx caprius, has been recently found to acquire feather mites and
lice typical of their ploceid weaver foster parents (Lindholm et al. 1998). Adult
cuckoos additionally host four species of mites and lice specific to them that are
Life with Vertebrates 385

Fig. 9.11 The array of feather mites living on the external surface of brush turkeys (Alectura
lathami, Megapodiidae). From left to right: Ascetolichus ruidus, Goniodurus sp., Leipobius sp.,
Echinozonus leurophyllus, E. longisetosus (all Pterolichidae) (Image by HC Proctor)

presumably contracted socially. Birds from the galliform family Megapodiidae also
have no contact between parents and hatchlings, though in this case it is because the
eggs are incubated by the heat of rotting vegetation or volcanically heated soils and
not by the parent birds. Megapodes host a rich diversity of feather mites and lice, all
of which must be acquired through horizontal transfer during mating or communal
roosting (Proctor and Jones 2004).
Plumicolous feather mites can move among feathers on the same host with
relative ease. However, many syringicolous species are too large as adults to move
out of the quill of a feather. Rather, they have modified migratory stages that move
from their natal feathers to colonise new sites. In the Ascouracaridae, the skinny
larvae are the dispersal form. They escape from the quill of wing and tail feathers by
chewing through the feather wall, travel on external feather surfaces and enter the
superior umbilicus (a small hole partway up the feather) of developing feathers
(Gaud and Atyeo 1996). Adult ascouracarids, in contrast, are dumpy, weakly
sclerotised and have no means to escape the quill. Not all bird-dwelling mites
travel between hosts under their own power. Some species in the dermicolous
Epidermoptidae have females that are parasites of hippoboscid flies, themselves
parasites of birds: nymphs and males remain on the host.
Atyeo and Gaud (1979) calculated that if every species of bird has an average of
two mite species there must be at least 18,000 species of feather mites. This may be
a conservative estimate, as some species of birds have many more than two feather
mites. The green conure, Aratinga holochlora, is an extreme example, hosting at
least 25 species of mites (Pérez 1997). Although there may be many species of
mites on one species of bird (Fig. 9.11), any given species of feather mite tends to
be specific to a single host species or genus. Part of the reason for host-specificity
386 9 Animals as Habitats

must be the mode of transmission, which is typically between socially or genetically


related individuals. But another answer lies in morphological adaptation to feather
structure, which differs greatly between bird species. Feather structure also varies
within an individual bird. So, it is not surprising to find that when different species
of feather mites share a host, they each occupy a limited area of the plumage.
This partitioning into microhabitats allows many species (even congenerics) to
co-occur on the same individual bird (Pérez and Atyeo 1984b).
Atyeo and Pérez (1988) examined site specificity in three species of Rhytidelasma
(Pterolichidae) from the green conure. They found that R. mesomexicana was most
common on primaries seven to ten and had smaller ‘populations’ that coexisted with
R. cornigera on the inner primaries and secondaries; R. urophila was restricted to
the retrices of the tail. Eggs are glued to species-specific locations on feathers
(Pérez 1996). Likewise, Choe and Kim (1989) examined how microhabitat selection
allowed coexistence of analgoid feather mites on seabirds. Trouessartiid and
proctophyllodid feather mites appear to partition flight feathers by occupying the
dorsal and ventral surfaces, respectively (Mestre et al. 2011). Different instars and
sexes of a given species also occupy different microhabitats, e.g. in kiwi-dwelling
Kiwialges (Analgidae), male nymphs live on the feathers and female nymphs live in
cutaneous pores.
Habitat partitioning occurs inside feathers as well as on them. Kethley (1971)
was intrigued by the fact that a single bird species may host four species of quill
mites (Prostigmata: Syringophilidae). Syringophilids live inside quills and feed by
piercing the wall of the quill with their long needle-like chelicerae to feed on tissue
fluids on the other side. In small feathers of the house sparrow, Passer domesticus,
Kethley found on average 120–130 Syringophiloides minor. In the large covert
feathers there were only 90–95 mites, although one would expect more because
there is more room. Primary feathers are even bigger and they contained no mites or
only dead ones. Kethley felt that the answer to this mystery and the explanation for
multiple species of syringophilids on one host lay in the thickness of the quills.
Primary feathers have the thickest walls and, if the cheliceral stylets of the mites are
too short to reach through the walls, then the population will fail, leaving only
corpses to record their attempt. However, quills too thick for minor would be
suitable for syringophilids of a different size. Kethley noted that across different
species of birds, large mites live in large quills and small mites in small quills; if
large and small quills occurred on the same host, there were often two sizes of quill
mites. Kethley concluded that variation in quill wall thickness probably accounts for
the diversity of quill mite sizes, which range from 0.0014 to 0.0866 mm3 in volume:
a 61-fold range.
Although some feather mites leave the bodies of dead hosts, many remain on and
in the feathers until they themselves expire. Because of this tendency to stay on the
host’s body, much taxonomic work on feather mites has been accomplished through
the examination of museum skins. To collect syringicolous species one must cut
open the feather shaft. As museums frown on the destruction of feathers, it is typi-
cally only the plumicolous mites that are extracted from prepared skins. Gaud and
Atyeo (1996) mention that plumicolous mites can be collected from live birds by
Life with Vertebrates 387

placing the bird in a bag and shaking it together with diatomaceous earth or by
dipping individual feathers in ethanol. Walther and Clayton (1997) suggest dusting
birds with pyrethrin to agitate their arthropod symbionts and then ruffling the bird’s
feathers over a collecting surface. One feather mite worker secures living birds on
their backs and examines the wings under a portable dissecting scope (S.V. Mironov,
pers. comm.).

Fish, Amphibians, Reptiles and the Mystery of Mite Pockets

In contrast to warm-blooded vertebrates, fish, amphibians and reptiles do not host a


huge diversity of mites. Perhaps this is a reflection of the watery habitat of the first
two groups and the relatively low diversity of integumental modifications of all three.

Mites Associated with Fish

Most, if not all, reports of mites on fish represent incidental associations. Water
mites sometimes survive ingestion by fish (Cupp and Willis 1982) and may be
mistaken for internal parasites. Water mite larvae have been illustrated attached to
young fish (Soar and Williamson 1925, plate III) but this appears to be a very rare
occurrence, as Smith and Oliver (1986) do not include fish in their exhaustive list of
host–larval associations. Proctor et al. (1997) discuss the numerous observations of
fish-associated acarids and histiostomatids (Astigmata), spurred by their discovery
of Histiostoma anguillarum and a member of the acarid genus Schwiebea on leeches
suffering from parasitic water mould (Oomycota). Histiostoma anguillarum had
also been recovered from eel tanks in Europe and some other members of the same
subgenus (Ichtanoetus) from fish aquaria. Schwiebea estradai was collected from
wild and farmed trout in Spain. Proctor et al. (1997) suggest that fish-associated
astigmatans may be attracted to animals of any sort that are parasitised by water moulds.
However, it may also be that mites turn to fish when their normal food supply
runs low. Untergasser (1989) notes that the aquatic oribatid Trimalaconothrus
(Malaconothridae) is found on the skin of aquarium fish when their normal food of
algae, fungi and detritus has been depleted for several weeks.

Mites Associated with Amphibians

There are a number of mite associates of amphibians. Species of the ixodid tick
genus Amblyomma and argasid genus Ornithodoros will feed on amphibians (Krantz
1978; Klompen et al. 1996). Larvae of the water mite Thermacarus nevadensis
(Thermacaridae) parasitise toads (Smith and Cook 1991), making it the only larval
parasitengone other than chiggers (Trombiculidae and Leeuwenhoekiidae) to feed
on vertebrates. As mentioned in Chap. 7, post-larval stages of some species of
Hygrobates water mites parasitize salamanders. Among terrestrial parasitengones,
388 9 Animals as Habitats

larvae of Hannemania (Trombiculidae) are intradermal parasites of frogs and


salamanders (Krantz 1978). Ereynetidae (Prostigmata) occur in the respiratory
passages of toads and frogs as well as of birds and mammals (Krantz 1978).

Mites Associated with Reptiles

Snakes, lizards and turtles have an interesting array of ecto- and endoparasitic mites
(reviewed by Fajfer 2012). Terrestrial reptiles are often heavily parasitised by ticks
and there are some ticks on marine snakes and iguanas (see Chap. 7). Hoogstraal
and Kim (1985) feel that the Ixodida evolved as parasites of Reptilia in the late
Paleozoic or early Mesozoic era, about 225 mya; however, this is questioned by
Klompen et al. (1996). The fact that some ticks are specific to particular species of
reptiles provoked a lament from Durden and Keirans (1996) on the diminishment
of tick diversity concomitant with the potential extinction of Komodo dragons
and Galapagos tortoises. One of these tortoise ticks, Ornithodoros transversus,
oviposits on its host, making it the only member of the Ixodida to spend its entire
life cycle on the host (Klompen et al. 1996).
A number of families of Mesostigmata specialise on reptiles. Members of the
Ixodorhynchidae, Omentolaelapidae and Entonyssidae are parasites of snakes
(Domrow 1987). The first two are external parasites, while the entonyssids live in the
lungs of their hosts. Ophionyssus (Macronyssidae) and Ophiomegistus (Paramegistidae)
feed on the blood of snakes and lizards.
Within the Prostigmata, mites in the family Pterygosomatidae include lizards
among their hosts (Paredes-León et al. 2012). Pterygosoma mutabilis, a parasite of
agamid lizards, has all of its life stages on the host except for eggs, which are depos-
ited on the lizard but roll off (Mostafa 1974). Newly hatched P. mutabilis larvae
climb aboard a passing lizard and race randomly over its body, eventually crawling
under a scale and feeding on blood. Hirstiella pyriformis, a pterygosomatid parasite
of the chuckwalla lizard Sauromalus varius, may transmit haemogregarine blood
parasites (Newell and Ryckman 1964). A number of genera in the Cloacaridae inhabit
the cloacal mucosae of freshwater turtles and marine turtles (Camin et al. 1967;
Bochkov 2002). Lizards and snakes are popular hosts of chiggers (Trombiculidae).
Vatacarus specialises on the respiratory passages of marine iguanas and sea
snakes, where it undergoes remarkable neosomatic growth.
Lizard-associated chiggers have been the subjects of heated debate. ‘Mite pockets’
are small invaginations in the skin around the neck, axillae, groin and postfemora of
lizards that are often packed with chiggers (Arnold 1986). These pockets are present
in a wide range of taxa including the Iguanidae, Chamaeleonidae, Gekkonidae,
Lacertidae and Scincidae. Arnold (1986) noted that the skin of the mite pockets is
elastic, rapidly healing and is subtended by dense concentrations of lymph cells.
He felt that these skin folds evolved in taxa prone to chigger infestation in order to
concentrate and minimise the damage caused by the feeding mites.
Bauer et al. (1990) tested Arnold’s hypothesis with the gecko genus Rhacodactylus.
These lizards have pockets at the backs of their thighs termed popliteal folds.
Effects of Parasitic Mites on Their Hosts 389

Bauer et al. examined four Rhacodactylus species, one from the forest floor and three
that were arboreal. All had popliteal folds but only the terrestrial species had chiggers
in them. They argued that the folds were phylogenetic baggage (or spandrels sensu
Gould and Lewontin 1978) that were opportunistically colonised by the chiggers
when geckos inhabited mite-infested terrestrial habitats. Arnold (1993) responded
to their criticism by suggesting that Bauer et al.’s sample sizes were far too small to
demonstrate the absence of chiggers in arboreal species of Rhacodactylus. Although
Bauer et al. (1993) admitted that sample sizes of one to five individuals were
indeed too small to statistically back up their claim, they argued that phylogenetic
constraint was still the best explanation for mite pockets in lizards. They point out
that Arnold has no evidence that concentration of chiggers reduces total damage
incurred by infested lizards and ask whether he would consider that scales evolved
to shelter the pterygosomatid mites that also infest lizards.
Humans who have experienced chigger infestations will have observed that the
subsequent itchy welts are concentrated around waistbands, cuffs of socks and
elasticised straps of undergarments. Like mite pockets, these areas are protected from
abrasion, sheltered from direct sunlight and maintain high humidity. So, although it
is possible that sequestering chiggers in a rapidly healing region may be better than
having them attach everywhere, this idea is unproven; it seems more likely that
‘mite pockets’ of lizards have evolved for structural reasons unrelated to mite
parasitism and only coincidentally provide an ideal habitat for the larval mites.

Effects of Parasitic Mites on Their Hosts

Parasites gain their livings from their hosts and, if the cost to the host is high enough,
reciprocal changes in both host and parasite over evolutionary time (coevolution)
may take place. Before dealing with any evidence of coevolution, however, we
present an overview of host susceptibility and the damage that mites can inflict.

Differential Host Susceptibility to Parasitism

Hosts may vary in their resistance to parasitic mites for a range of reasons; some
representing adaptations and some that may be pure happenstance. Mites them-
selves may avoid certain individuals or some potential hosts may escape parasitism
through behavioural or morphological characteristics that discourage mites.

Mite Choice of Hosts

It would obviously be advantageous for a mite to choose the best host. In the most
general and trivial sense, this involves choosing a host of the correct taxon. More
specifically, a mite may be faced with choosing among a number of conspecific hosts,
390 9 Animals as Habitats

some of which may be ‘better’ than others. We have already discussed how water
mite larvae rely on their host returning to water so that they may drop off into an
aqueous medium. Female hosts are thought to be more likely to return because they
must oviposit on or in water; however, there is no clear evidence that water mite
larvae discriminate between host sexes (Smith 1988). However, in both water mites
and their terrestrial parasitengone relatives, the distribution of larvae on hosts is
typically a negative binomial: many hosts have no mites and a few hosts have many.
How does this clumping come about? Although much of it is likely to be phenological
or environmental (see section “Ecological Avoidance of Mites”), there is laboratory
evidence that parasitengone larvae prefer already parasitised hosts. This congruence
of choice may occur because such hosts are close to moulting (Smith 1988) and so
reducing the waiting time until a new, soft integument is presented. It may also be
that larvae are trading off a possible loss in nutrition due to overcrowding with
the increased probability of finding a mate at the end of the journey (Zhang 1992).
In this latter case, it is the fact that the host already carries a conspecific mite that
makes it attractive to other larvae.
Stronger evidence for adaptive host choice comes from mesostigmatans phoretic
on bumble bees. All stages of Parasitellus fucorum (Parasitidae) live in nests of
Bombus species (Apidae). Huck et al. (1998) found that phoretic deutonymphs
would switch from male bees to queens but never from queens to males. Deutonymphs
also appeared to prefer queens over worker-caste females. Because bumble bee
colonies are annual and young fertilised queens are the only overwintering stage,
this preferential attachment makes sense. Likewise, Campbell et al. (2012) found
that mites associated with ants in Ohio were much more likely to be found on the
bodies of winged female reproductives than on winged males.

Ecological Avoidance of Mites

Although tempting, it is often specious to postulate that ecological characteristics of


a host species (habitat preference, phenology) are adaptations to avoid mite parasitism.
The following examples involve characters that affect the probability of encountering
parasitic mites, with no imputation of adaptation.
Pruett-Jones and Pruett-Jones (1991) examined Ixodes (Ixodidae) tick infestation
of 115 species of New Guinean birds. Non-passerines (161 individuals) had a very
low infestation rate (only one individual of one species) but passerines were higher
(385 of 2,496 individuals). Those passerine species foraging on or near the ground
were harder hit, suggesting that species foraging higher in the canopy were able to
avoid questing ticks, which tend to stay on the ground or in low vegetation. Smith
and McIver (1984a) examined factors influencing host selection of Aedes mosqui-
toes by Arrenurus water mites. As well as behavioural factors (see section
“Behavioural Defences Against Mites”), they found that two species, A. communis
and A. punctor, were much more susceptible to parasitism in the lab than they were
in the field; this was because in the field, the mites and mosquitoes did not overlap
temporally. Smith (1988) describes patterns of parasitism by larvae of Eylais water
Effects of Parasitic Mites on Their Hosts 391

mites on water boatmen (Hemiptera: Corixidae). Cenocorixa expleta and C. bifida


inhabit inland saline waters, the first species occurring in saltier water. Smith sug-
gests that C. expleta may be restricted to these habitats because they suffer more
when parasitised by Eylais euryhalina and the mite cannot deal with very high
salinities. In a rare test of ecological avoidance of parasitism, Sorci et al. (1997)
examined whether the lizard Lacerta vivipara avoided either conspecifics parasit-
ised by Ophionyssus or resting sites previously occupied by parasitised lizards.
Contrary to their expectations, they found that unparasitised lizards did not dis-
criminate against individuals or sites associated with mites.

Physiological Defences Against Mites

Many animals, both vertebrate and invertebrate, mount protective immune responses
to parasitic mites. The intense inflammatory response of domestic chickens to the
bites of the Northern Fowl Mite Ornithonyssus sylviarum increases the thickness of
the host’s skin and hence the distance between the mites’ mouthparts and the bird’s
blood supply, reducing growth and reproductive rate of the mites (Owen et al. 2009).
Davids (1973) noted that when larvae of the water mite Hydrachna conjecta bite
backswimmers (Notonectidae) that are not their normal host, the bug’s haemolymph
reacts so strongly to the saliva that the mites’ mouthparts are obstructed, they cannot
feed and soon die. Lestes forcipatus damselflies emerging later in the season are
more likely to mount melanistic immune reactions against the mouthparts and
stylostomes of larval Arrenurus than are early-season emergers (Yourth et al. 2002).
Pruett-Jones and Pruett-Jones (1991) found that non-passerine birds had lower
infestation rates by ticks than passerines. They believe that non-passerines, being on
average longer lived than passerines, acquire a greater immunity to ticks and thus
suffer lower infestations. This is not a far-fetched hypothesis. Brossard and Wikel
(1997) and Wikel and Bergman (1997) discuss innate and acquired host defences
against ticks and the countermeasures mounted by these parasites. There is good
evidence that vaccination with protein antigens induces immunity to tick infestation
in livestock (Willadsen 2004).
As well as their obvious practical value to graziers and veterinarians, interactions
between the host immune system and ectoparasites are of great interest to behav-
ioural ecologists working on tradeoffs between sexiness and survival. The ‘immu-
nocompetence hypothesis’ predicts that males producing a high level of testosterone
suffer a tradeoff between expression of secondary sexual characters and exposure to
parasite infestations. Salvador et al. (1996) tested this hypothesis in a population of
the lizard Psammodromus algirus during its mating season. They found that males
implanted with testosterone developed larger patches of orange colour and behaved
more aggressively than did control males. Although the tick load of all individuals
increased over the season, testosterone-implanted males suffered a greater increase
than did control males. Likewise, testosterone implanted males of the lizard
Sceloporus jarrovi had greater mean loads of trombiculid mites than did castrated
and sham-operated males (Fuxjager et al. 2010).
392 9 Animals as Habitats

Weatherhead et al. (1993) found a positive correlation between the number of


Dermanyssus (Dermanyssidae) mites on male red-winged blackbirds (Agelaius
phoeniceus) and both testosterone level and the length of the males’ epaulets (the
red shoulder patches that give the bird its name). However, similar correlations were
not found with any of the other parasites (lice, gut flukes, blood protozoans), nor was
there an overall effect of testosterone level on male secondary sexual characters.
Saino et al. (1995) manipulated testosterone levels in male barn swallows
(Hirundo rustica) to determine its effect on tail length and burden of parasites,
including the nest mite Ornithonyssus bursa (Macronyssidae). Results were unclear
due to small sample sizes but testosterone implants did not seem to affect nest mite
numbers. This is not surprising, considering that O. bursa probably feeds primarily
on chicks rather than on male parents. One major problem with studies linking para-
sites, testosterone and plumage in male birds is that testosterone does not determine
showy male feathers. A castrated peacock is just as elaborately plumed as an entire
one (Owens and Short 1995); however, spurs, wattles and combs are under testos-
terone’s control and would make better subjects for such studies.
Another possible effect of hormones on ectoparasites is mediated through
differential motivation to groom. Mooring et al. (1996) found that territorial male
impala had higher tick loads than either females or non-territorial males and that they
groomed themselves less assiduously. They were not sure whether the reduction in
grooming was the result of territorial males having less time to groom or if higher
testosterone levels physiologically suppressed oral grooming. Schalk and Forbes
(1997) tested the testosterone hypothesis by surveying the literature for male biases
in parasitism of mammals. Their meta-analysis found that it seemed to hold for
arthropod ectoparasites. However, they did not appear to control for body size
and males may often be larger and hence have greater surface area for parasites
than females.

Behavioural Defences Against Mites

Anti-ectoparasite behaviour is well known in vertebrates. Mites, especially ticks,


are removed by auto- and allogrooming among conspecifics (e.g. Mooring et al.
1996; Clayton et al. 2010), and relationships with parasite-eating tick-birds. Certain
avian behaviours such as sun-bathing and anting, in which a bird flops on top of an
anthill and allows the insects to patrol through its feathers, are thought to help rid
the birds of ectoparasites (Brooke and Birkhead 1991; but see Clayton et al. 2010).
Nesting behaviour may also have its anti-mite aspects. Burtt et al. (1991) suggest
that the tight weave of bluebird nests may limit occupation by nidicolous mites.
Clark (1991) discusses the ‘nest sanitation hypothesis’ for the commonly observed
inclusion of fronds of fresh plants in the nests of many species of birds. Clark
manipulated starling nests by removing already included green vegetation from
some nests (which he replaced with equivalent volume of dry grass) and by adding
some vegetation (wild carrot, which was commonly chosen by starlings) to others.
By the end of the first breeding attempt, nests with wild carrot contained an average
Effects of Parasitic Mites on Their Hosts 393

of 3,000 Ornithonyssus sylviarum (Macronyssidae), whereas nests without contained


an average of 80,000 mites. By the end of the second breeding attempt in June, nests
with wild carrot contained an average of 11,000 mites, while nests without
contained approximately 500,000 mites; this is strong evidence in favour of the
nest-protection hypothesis. Birds in cities may make use of discarded cigarettes as
nest fumigants (Suárez-Rodríguez et al. 2013).
Anti-mite behaviours also occur in invertebrates. Cowan (1984) notes that female
larvae of the eumenine wasp Ancistrocerus antilope kill the Kennethiella trisetosa
(Winterschmidtiidae) in their brood cells prior to pupation so that they emerge as
mite-free adults. However, as the male larvae do not kill their mites, female wasps
end up being infested by K. setosa through venereal transmission during mating.
Cunliffe (1952) discusses the interactions of the parasite Pimeliaphilus podapoli-
pophagus (Pterygosomatidae) with its cockroach hosts. Cunliffe observed up to
25 mites, each up to 1 mm long, on the body of a single roach. After about 1 h of
such heavy parasitism, the roach falls over on its back and thrashes about for several
hours before dying. He notes ‘[i]t has been observed that roaches engage in apparent
combat with the mites. Fifteen or so mites have been seen harassing or riding along
on a roach when suddenly the roach will succeed in grasping some of the mites with
its front legs and mouthparts. One or two, or sometimes three, mites may become
involved, until either they escape or the roach finally manages to swallow them.’
Water mite larvae are often the target of anti-parasite behaviour. Wiles (1987)
observed adult Chaoborus flavicans (Diptera: Chaoboridae) under attack by
Hydrodroma despiciens larvae. The flies kicked, stamped, jumped, combed and
buzzed their wings. Once female flies began ovipositing, however, they ceased their
anti-parasite behaviour and mite larvae mounted them with impunity. Baker and
Smith (1997) observed that larval damselflies may resist parasitism by Arrenurus
water mite larvae by grooming, crawling, swimming and striking at the mites.
Smith and McIver (1984a) found that pupae of the culicid Aedes cinereus were
able to avoid or dislodge mite larvae before pre-parasitic contact was established by
vigorously straightening their abdomens or by swimming. When these pupae were
anaesthetised with CO2, the proportion of successful contacts rose from 0.013 to
0.499. Other culicid pupae which did not exhibit avoidance movements were more
likely to be parasitised.

The Evil That Mites Do: Adverse Effects of Acarine Symbionts

It is often stated that a well-adapted parasite does not harm its host (Toft and
Aeschlimann 1991). Numerous studies of the effects of mite symbionts on their
hosts have shown no or little negative impact. This has led some to claim that
such mites are the ones ‘best adapted’ or ‘longest associated’ with the host: in
Nutting’s (1985) words, ‘least damage, longest domiciled’. Generally benign
symbionts include demodicid mites of mammals (Nutting 1985) and most mites
dwelling on the external surface of feathers (e.g. Gaud and Atyeo 1996). However,
394 9 Animals as Habitats

if a symbiont’s goals are best met by an action that harms the host, then there will
be little selection for benign behaviour (Herre 1993). Some of the negative effects
of symbiotic mites are easily predicted, e.g. one might expect that a heavy load of
blood-sucking ticks would cause anaemia in their host. But many of the side effects
of carrying mites are not as obvious. It is not only the nasty parasites that can harm
their hosts but also seemingly benign phoretic and nidicolous mites (Table 9.4). In
many of these latter cases, it seems likely that the mites receive no benefit from their
inadvertent negative impacts.

Effects of Non-Parasitic Nidicolous Mites

Nidicoles associated with stored products may spoil food through their excretions,
increased humidity or dispersal of fungal spores. The faeces and bodies of these
pests can result in allergic skin and respiratory reactions in humans (Colloff 2009).

Effects of Phoretic Mites

Carrying one or two phoretic mites is probably no great hardship for any animal;
however, heavy loads or mites lodged in inconvenient locations can adversely affect
the carrier. Inhibition of movement – walking or flight – due to mite burdens has
been documented for beetles, gnats and a lizard (Table 9.4). In the lizard’s case, an
encrustation of hundreds of uropodine deutonymphs resulted in its eventual death
(Domrow 1981).
Other activities may be made more difficult by the presence of phoretic mites.
Moya Borja (1981) documents the effects of phoresy by Macrocheles muscaedo-
mesticae (Macrochelidae) on the human bot fly Dermatobia hominis (Cuterebridae)
in Brazil. Larvae of this fly are serious parasites of mammals, including humans;
however, they do not oviposit directly on their hosts. Rather, female D. hominis
capture other mammal-associated flies (mosquitoes, houseflies) and oviposit on
the belly of the press-ganged insects. When the egg-laden fly lands on warm
mammalian skin, the eggs hatch and little D. hominis larvae drop onto the skin of
the host. Moya Borja found that hitchhiking M. muscaedomesticae interfered
with the reproductive success of D. hominis in two ways. First, because the mites
tended to cluster around the venter and genital regions of both male and female,
they hindered copulation. Secondly, they affected the female fly’s ability to
capture and oviposit on other flies. Similarly, Iverson et al. (1996) found that
deutonymphs of Lardoglyphus zacheri (Lardoglyphidae) attached to dermestid
beetles in such huge numbers ( > 500/beetle) that the hosts were incapable either
of mating or of feeding, resulting in the eventual extirpation of beetle colonies.
A more indirect negative effect of phoretic mites on their hosts is the competition
for food between hummingbirds and the mites (Melicharidae) they ferry between
flowers (Colwell 1995).
Table 9.4 Adverse effects of symbiotic mites on their hosts
Effect of mites on host Host (life history stage) Mite (type of associate) Source of information
Movement hindered Moths Trombidiid larvae (Pa) Treat (1975)
Black flies Sperchon setiger larvae (Pa) Ullrich (1976)
Burying beetles Poecilochirus necrophon (Ph) Hunter and Rosario (1988)
Fungus gnats Ereynetes (Ph) Zhang and Sanderson (1993)
Damselflies Arrenurus (Pa) Reinhardt (1996)
Lizard Uroactinia (Ph) Domrow (1981)
Copulation physically hindered Cuterebrid fly Macrocheles muscadomesticae (Ph) Moya Borja (1981)
Drosophilid fly Macrocheles subbadius (Pa) Polak and Markow (1995)
Oviposition physically hindered Cuterebrid fly Macrocheles muscadomesticae (Ph) Moya Borja (1981)
Feeding physically hindered Dermestid beetle Lardoglyphus zacheri (Ph) Iverson et al. (1996)
Effects of Parasitic Mites on Their Hosts

Breathing hindered Fur seal Orthohalarachne spp. (Pa) Kim (1985a)


Allergic reactions Humans Acarids and glycyphagids (N) Fain and Hyland (1985)
Humans Pyroglyphids (N) Arlian (1989a)
Integument damaged Domestic chickens Knemidocoptidae (Pa) OConnor (1982)
Parrots Cystoidosoma (Pa?) Pérez and Atyeo (1984b)
Humans Sarcoptes scabei (Pa) Arlian (1989b)
Slugs Fuscuropoda marginata (Pa) Raut and Panigrahi (1991)
Damselflies Arrenurus (Pa) Åbro (1982)
Deterioration of teeth and palate Long-nosed bat Radfordiella (Pa) Kim (1985a)
Increased asymmetry Damselflies Arrenurus (Pa) Bonn et al. (1996)
Diseases transmitted Colonial birds Ticks (Pa) Chapman and George (1991);
Duffy (1991)
Mammals, including humans Ticks (Pa) Aeschlimann (1991)
Chuckwalla lizard Hirstiella pyriformis (Pa) Newell and Ryckman (1964)
Fecundity reduced (for female hosts) Domestic chickens Ornithonyssus sylviarum (Pa) Clark (1991)
Scale insects Hemisarcoptes sp. (Pa) Eickwort (1983)
Ladybird beetle Coccipolipus (Pa) Hurst et al. (1995)
395

(continued)
Table 9.4 (continued)
396

Effect of mites on host Host (life history stage) Mite (type of associate) Source of information
Mexican bean beetle Coccipolipus (Pa) Schroder (1982)
Mosquitoes Arrenurus (Pa) Lanciani and Boyt (1977)
Productivity reduced Honey bees Varrroa jacobsoni (Pa) de Jong et al. (1982)
Virility reduced (for males) Domestic chickens Ornithonyssus sylviarum (Pa) Clark (1991)
Mate-finding ability reduced (for males) Damselflies Arrenurus, Limnochares (Pa) Forbes (1991)
Greater susceptibility to predators Damselfly larvae Arrenurus larvae (Pa) Baker and Smith (1997)
Nest abandonment Colonial birds Ticks (Pa) Chapman and George (1991)
Competition for food Hummingbirds Proctolaelaps kirmsei (Ph) Colwell (1995)
Growth rate reduced Rock doves (pigeons) Dermanyssus gallinae (Pa) Clayton and Tompkins
(for juvenile hosts) (1995)
Water striders Limnochares aquatica (Pa) Smith (1989)
Moulting prevented Chironomid midges Hygrobates larvae (Pa) Ullrich (1976)
Lifespan reduced Black flies Sperchon setiger larvae (Pa) Ullrich (1976)
Water boatman Eylais larvae (Pa) Böttger (1962)
Mosquitoes, marsh treaders, Water mite larvae (Pa) Smith (1988)
ceratopogonids
Mexican bean beetle Coccipolipus (Pa) Schroder (1982)
Note: Obvious effects such as ‘loss of body fluids or tissues’ for hosts of parasites are not included. Types of mite associates are N nidicoles, Ph phoretics,
Pa parasites. The list is not exhaustive
9 Animals as Habitats
Effects of Parasitic Mites on Their Hosts 397

Effects of Parasitic Mites

These range from the mildly irritating to the horrific (Table 9.4). Like phoretic
mites, a heavy load of parasites can hinder host movement. This reduced motility
can affect mate-searching ability (Forbes 1991). It may behove parasites not to
interfere with host movement if they rely on hosts for both food and transport.
McLachlan et al. (2008) tested whether the distribution of larval water mites
(Unionicola ypsilophora) on their chironomid midge hosts was more symmetrical
than expected by chance, but did not find strong evidence that the larval mites were
attempting to balance their load on the host. Efforts to avoid parasitism can also
have ill effects. Baker and Smith (1997) found that the anti-mite grooming activity
of larval damselflies put them at greater risk of predation from visually hunting fish.
Feeding activity can cause local tissue damage. For example, pedipalps of the water
mite Unionicola intermedia are sunk deeply into the underlying connective tissues
of the gills of its mussel host causing edema of the gill filaments (Baker 1977).
Mitchell (1965) mentions that Najadicola ingens, which can grow up to 64 mm3 in
volume, commonly destroys host mussel tissues and causes papillate growths.
Feeding by the uropodid Fuscouropoda marginata on the epidermis of slugs reput-
edly causes wounds that become necrotic, eventually killing the molluscs (Raut and
Panigrahi 1991).
Parasitic mites may remove a large amount of nutrition from their hosts, as
exemplified by the degree of engorgement attained by some larval water mites
(Fig. 9.12). Parasitism of juvenile hosts may be sufficiently intense as to slow their
development. This has been demonstrated in a variety of aquatic insects parasitised
by water mites and in birds besieged by nest mites (Table 9.4). However, other stud-
ies of nest mites have shown no clear effects of parasitic mites on nestling growth or
survival (e.g. Bauchau 1997; Pacejka et al. 1998). Water mites in the genus
Hygrobates attach to the skin of adult midges by biting through the integument of
the pupa into the skin of the developing adult and are pulled through the exuviae
when the adult emerges. Ullrich (1976) noted that heavy loads prevented moulting
because the midge lacked the strength to pull so many mites through its skin.
Fecundity of adult female hosts is reduced by parasitism in a wide variety of host
taxa from chickens to mosquitoes. Male reproduction may also be affected, through
lowered sperm count (Clark 1991), smaller testes (Polak 1998) or reduced mate-
searching ability (Forbes 1991). Adult lifespan has been shown to be reduced by
mite parasitism for many groups of insects (Table 9.4). The effect is usually depen-
dent on mite load. Larvae of the trombidiid Allothrombium pulvinum kill their aphid
hosts when the mite load is two or more (Zhang 1998). With only one mite larva per
aphid, development is halted for nymphal hosts and reproductive rate slowed for
adult hosts.
Finally, other mites on the same host can affect parasitic mites. The ‘crowding
effect’, first noticed in tapeworms (Read 1951), is the reduction in mass and size of
individual parasites with the increasing number of parasites per host. The crowding
effect has been observed in water mite larvae. Davies (1959) noted that Sperchon
larvae are larger when fewer are present per black fly host. Lanciani (1984) also
398 9 Animals as Habitats

Fig. 9.12 Larvae of the


water mite Hydrachna
conjecta (Hydrachnidae)
before and after engorgement
on fluids from the host
insect (Corixidae). The
branched structure attached
to the mouthparts is a
ramified stylostome
(After Davids 1973)

observed this in Hydrachna virella larvae parasitising backswimmers. The size of


the host affected this relationship in two ways. First, the larger the host, the smaller
was the crowding-induced rate of decline in mite size. Secondly, the larger the host,
the larger was a given larva at any number of mites per host. It seems likely that
crowding-induced decreases in nymphal size result in a smaller reproductive output
when a mite reaches adulthood. However, crowding in the larval stage may not be
selected against if the loss of fitness due to the inability to find a host is greater than
that caused by feeding from a crowded host and reaching less than maximum size
(Lanciani 1976).

Parasitic Mites and Mate Choice by Hosts

Female animals often appear very particular about which male they will accept as a
mate. The rationale behind this choosiness is likely to vary among taxa. If a male
provides sole parental care for the young (as in many fish), females may choose
on the basis of the quality of the male’s spawning grounds. If males provide tasty
nuptial gifts as a mating incentive (e.g. scorpionflies, many orthopterans), females
Effects of Parasitic Mites on Their Hosts 399

may choose on the basis of the gift’s size. But if the male offers no obvious service
or goods (most invertebrates and many vertebrates), why should females care about
whom they mate with?
The ‘good genes’ hypothesis suggests that females may choose among males on
the basis of the heritable endowment passed on to their joint offspring, i.e. high-
quality males should sire high-quality offspring. Many studies have shown that
females tend to choose the brightest and most ornamented of males but it is unclear
how this reflects good genes. Hamilton and Zuk (1982) suggested that showy orna-
ments in male birds reflect an individual’s genetically based resistance to the current
array of parasites affecting a population. By choosing the brightest male, a female
selects a genotype that, when conferred to her offspring, will protect them from
parasites. Although originally proposed for internal parasites, the Hamilton-Zuk
hypothesis has been extended to include ectoparasitic arthropods.
Tests of the parasite hypothesis have involved both comparative multi-species
approaches and single-species studies. Pruett-Jones and Pruett-Jones (1991) surveyed
115 species of birds from New Guinea to determine whether there was a correlation
between male showiness, blood parasite count and infestation by ticks (Ixodes spp.).
Neither factor correlated with tick load and they noted that sexually dichromatic and
monochromatic species suffered similar levels of parasitism. This is not an uncommon
finding in comparative tests of the Hamilton-Zuk hypothesis and may stem from a
basic uncertainty on what the hypothesis predicts (Harvey et al. 1991). Should species
with showy males have more associated parasites (indicating currently acting selec-
tion for resistance) or should they have fewer (indicating successful past selection for
an effective immune system)? A clearer prediction is that within a given species,
highly ornamented males should have fewer parasites; conversely, heavily parasitised
males should show poor development of secondary sexual characters.
This expectation was supported in a study of the lizard Lacerta viridis, in which
the intensity of blue throat colour in males was negatively correlated with loads of
the tick Ixodes ricinus (Václav et al. 2007); however, somewhat confusingly, satura-
tion of chest colour in females was positively associated with tick load. Møller
(1991) studied the relationship between Ornithonyssus bursa (Macronyssidae) and
‘attractiveness’ of male barn swallows. One measure of attractiveness is whether a
male is successful in attaining a mate. Møller found that mean mite load for males
with mates was 5.7 compared to mean of 10.7 for unmated males. Male barn swal-
lows display a presumably sexually selected trait: tail streamers longer than those of
the female. Møller also states that mite loads were negatively correlated with male
tail length. Darolova et al. (1997) observed a negative relationship between mask
width of male penduline tits and loads of mites in their nests (Dermanyssus hirun-
dinis and Ornithonyssus sylviarum). Mask width expresses age and dominance, and
female tits tend to choose wider masked males. Darolova et al. (1997) conclude that
the sexually dimorphic trait may indicate male quality, including the ability to avoid
parasites. In contrast, Burley et al. (1991) found a positive relationship between
brightness of bills in zebra finches and their loads of lice and mites. They suggest
that birds with brighter colours may have more social contacts and get more para-
sites or that parasites do better on brighter birds for unknown reasons.
400 9 Animals as Habitats

Females may be interested in knowing male parasite load, not for the benefit of
their offspring, but to avoid contagion themselves. Sexual transmission of mites has
been observed in wasps (e.g. Cowan 1984) and beetles (e.g., Seeman 2008). Hurst
et al. (1995) suggest that the often vigorous male-spurning behaviour of female
ladybird beetles (Coccinellidae) is to avoid transmission of the parasitic mite
Coccipolipus (Podapolipidae). Thompson et al. (1997) bring up an interesting point
in their study of Proctophyllodes feather mites and showiness of male house finches.
The authors examined mite load and plumage brightness in males before and after a
moult. Males with no or few mites pre-moult increased in rank of plumage bright-
ness after moult, while heavily laden males decreased in rank. They argue that many
studies that correlate parasite load with current male condition are flawed because
the plumage developed under previous, not current, mite conditions: ‘[t]hus, females
that avoid males for a direct reason such as fear of contagion would not find male
plumage colour at the time of mate choice in spring to be an accurate indicator of
parasite load at that time’. As mentioned previously, it is not clear whether feather
mites like Proctophyllodes are parasites, commensals or mutualists, although there
is more evidence for the latter two relationships (see section “Mites on and in
Feathers”). As Thompson et al. (1997) found that mite load was positively corre-
lated with avian pox, it may be that the pox rather than the mites were responsible
for altered plumage condition.
A negative relationship between mite load and mating frequency may not
indicate that parasitised males are chosen less often by females. Rather, it could be
that parasitised males have lowered testosterone (e.g. roosters, Clark 1991) and
hence are less motivated to mate, they may be stressed and have less energy to
expend on mate-searching (e.g. damselflies, Forbes 1991) or the presence of mites
may physically hinder copulation (e.g. drosophilid flies, Polak and Markow 1995).
A fading fad in evolutionary biology is the use of morphological symmetry as a
measure of an individual’s genetic health and hence desirability as a mate.
Fluctuating asymmetry (FA) is a measure of small, random deviations from
perfect symmetry in a bilaterally or radially symmetric character. Asymmetry is
presumed to be the result of ‘errors’ during the development of a structure. The ‘less
fit’ an individual’s genome, the more errors it will make during development and the
more asymmetrical the organism will be. FA is a population-level measurement but
is often mistakenly used to refer to asymmetry in an individual organism.
A number of studies have indicated that males with more symmetrical ornaments
are more successful in attaining mates (summarised in Polak 1997). Parasitic mites
may have two roles in asymmetry-related sexual selection. First, less symmetrical
( = less fit) individuals may be less able to prevent infestation by parasitic mites.
Second, parasitism by mites may affect the development of a juvenile so that it
grows into an asymmetrical adult. Møller (1992) manipulated loads of tropical fowl
mites (Ornithonyssus bursa) by adding or removing mites from nests of the swallow
Hirundo rustica. Asymmetry was measured in the subsequent year after the swal-
lows had grown new tail ornaments under the altered parasite regime. Asymmetry
was larger at increasing levels of parasites for male tail length but not for the length
of the shortest tail feather or for wing length. Parasite load did not affect tail or wing
Mite–Host Coevolution: Any Evidence? 401

length in females. Møller suggests that the degree of asymmetry in tail ornaments
reveals the level of parasite infestation in males and so it can serve as a vehicle for
female choice (but see the comment on moulting phenology by Thompson et al.
(1997) above). However, Dufour and Weatherhead (1998) found no relationship
between symmetry and quality of male red-winged blackbirds, except that sym-
metrical males carried a higher load of Ornithonyssus sylviarum (Macronyssidae)
than did less symmetrical males, exactly the opposite of what Møller (1992) predicts.
Bonn et al. (1996) found a correlation between asymmetry of forewing length in
damselflies and load of parasitic Arrenurus mites. There was no correlation between
mite load and cell number in the fore wings. They found no evidence that ‘fitter’
( = larger) individuals had lower mite loads. The authors suggest that stylostomes
may interfere with blood pumping into the wings at adult eclosion but that mite
parasitism would not alter cell number, which is determined before the parasitic
impact. They point out that while most explanations for increases in ‘FA’ implicate
environmental stress combined with the inability of the genome to stabilise the phe-
notype, their results suggest that asymmetry could result from very short-term
impacts due to chance parasitism. However, there is evidence that mites affect
symmetry through very indirect routes. Polak (1997) found that female drosophilid
flies parasitised by Macrocheles subbadius produced sons with more asymmetrical
bristle-placement relative to sons of non-parasitised females. He suggests that mite
parasitism may alter provisioning to eggs during ova development. However, Polak
found no evidence that bristle symmetry affected male mating success.
The effect of parasites on sexual selection, host development and host immunol-
ogy are fascinating to contemplate (e.g. Møller and Saino 1994). However, because
parasite load may be the result, rather than the cause, of differences in host morphol-
ogy and physiology, and may even result from host mating success (e.g. transfer of
mites during mating in birds and insects), these factors seem impossible to tease
apart purely by correlational studies. More strictly controlled experimental
approaches, such as those of Clayton and Tompkins (1995) on nest mites of rock
doves (Columba livia), are essential.

Mite–Host Coevolution: Any Evidence?

‘Coevolution’ was coined by Ehrlich and Raven in 1964 in a discussion of the influ-
ences that plants and insects have exerted on each other. Their original definition
was somewhat vague and authors using the word thereafter have given it their own
spins. To forestall semantic arguments, Janzen (1980) defined coevolution specifi-
cally: coevolution occurs when a trait in one species evolves in response to a trait in
another species, whose trait itself evolved in response to a trait in the first species.
Thus coevolution, or more precisely, coadaptation, is a reciprocal microevolution-
ary process involving a pair of species each adapting to the other. This pairwise
interaction appears to be very rare among animals, or at least hard to prove; how-
ever, there are numerous examples of gene-for-gene coevolution among plants and
402 9 Animals as Habitats

their pests (Thompson 1989). A more relaxed ‘diffuse coevolution’ occurs when a
trait in one or more species evolves in response to a trait in one or more others. For
example, a group of flowering plants and a group of bees may be mutually adapted
for pollination interactions. The important thing is the mutual nature of the changes.
If one taxon shows adaptations to characteristics of another, but not vice versa, this
is not coevolution but simply evolution (Futuyma and Slatkin 1983).
Another use of coevolution has been to describe matched speciation events
between two ecologically associated taxa. For example, speciation in fig trees may
be matched with that of their pollinating fig wasps (Feinsinger 1983) or that of
pocket gophers with their lice (Hafner et al. 1994). This phenomenon is more pre-
cisely termed ‘cospeciation’.
Among parasitologists, the expectation of a match between host and parasite
phylogenies is termed Fahrenholz’ Rule (Brooks and McLennan 1993). Like dif-
fuse coevolution, cospeciation may not occur on a one-to-one basis. Hosts may
speciate first, carrying their associated symbionts with them, and symbionts may or
may not develop into new species subsequently. This, too, has a parasitologist’s
moniker: Manter’s First Rule is that parasites evolve more slowly then their hosts
(Brooks and McLennan 1993). As Brooks & McLennan point out, many of these
‘rules’ are statements of belief that often have little empirical support. For example,
Klimov and OConnor (2008) found that early-derivative chaetodactylid mites were
associated with derived species of host bees, not primitive ones, and that the oppo-
site was true for more derived species of mites.

Coevolution by Mutual Adaptation

There is no evidence of gene-for-gene coevolution in arthropod–host relationships


as there are for some plant–pathogen interactions (Kim 1985c). However, evidence
of coadaptation at the microevolutionary level may be the endocrinological battle
between ticks and their hosts. Wikel and Bergman (1997) describe how acquired
resistance to tick infestation can be countered if ticks suppress antibody production
in the host. More tentative support for physiological coadaptation is the often
violent reactions exhibited by hosts when they are infested by mites they would not
normally encounter. Nutting (1985) notes that bird and reptile chiggers may cause
more severe tissue reaction in humans than do mammal chiggers, suggesting that
diffuse coevolution may be at work. Eickwort (1983) states that the milkweed leaf
beetle, Labidomera clivicollis (Chrysomelidae), the normal host of Chrysomelobia
labidomerae (Podapolipidae), can carry large numbers of mites without ill effect;
however, the Colorado potato beetle, Leptinotarsa decemlineata, has higher mortal-
ity and lower fecundity when infested. Delfinado-Baker et al. (1992) note that the
honey bee, Apis mellifera, has much greater difficulty grooming off the mite Varroa
jacobsoni than does its Asian congener, A. cerana, which likely has had a much
longer evolutionary history of contact with V. jacobsoni. Mosquitoes that are rarely
parasitised by Arrenurus in nature because of phenological separation are hardest
Mite–Host Coevolution: Any Evidence? 403

hit in the laboratory, apparently because they lack defence mechanisms against
these water mite larvae (Smith 1983). Conversely, notonectid bugs may be better
able to fend off attacks from water mites that don’t usually feed on them (see section
“Physiological Defences Against Mites”, above). Eickwort (1983) suggests that
transplanting mites from one region where they are apparently harmless to sites
where their hosts have not adapted to them may provide better biocontrol.
Such observations lead to the prediction that the greater the pathology caused by
the parasite, the more recent the association. For example, Nutting (1985) suggests
that myobiid mites are likely to be recently or weakly coevolved with their hosts
because they tend to be more pathogenic than demodicid mites. Considering haema-
tophagous mites in general, Radovsky (1985) contends that trauma associated
with the site of puncture is generally not found in more highly adapted groups of
ectoparasites. Kim (1985c) qualifies the value of this generalisation, stating that the
“[i]ntimacy of host associations indicates that they are either very old in the evolu-
tionary sense or relatively recent but with rigorous mutual selection and stepwise
coadaptation.”
While it is easy to show that a parasite is adapted to its host, for example by the
structure of its grasping appendages or general body shape, it is difficult to show the
opposite. A single species of host often has many different parasites; so an anti-
parasite behaviour like ‘scratching’ is unlikely to be a specific adaptation against
one parasite (Kim 1985c; Timm and Clauson 1985). As well, some species of mites
have more than one host, so their morphological or behavioural adaptations for
parasitism or phoresy are unlikely to be strictly coevolved. Examples of the first
situation include the green conure with at least 25 species of feather mites (Pérez
1997) and the scarab Copris hispanus with 19 species of acarine symbionts (Hunter
and Rosario 1988). Examples of multi-host mites include: the water mite Protzia
eximia on 25 species of insect hosts, including Simuliidae, Chironomidae and various
Trichoptera (Ullrich 1976); Sarcoptes scabiei complex on 40 species belonging to
seven orders of mammals (Arlian 1989b; but see Rasero et al. 2010 for evidence of
host-associated differentiation of S. scabiei); and Pyemotes ventricosus on 139
species of insects in six orders (Eickwort 1983). Any coevolution in these cases
would have to be very diffuse indeed.

Cospeciation

Kim (1985c) notes that two processes can result in association of groups of related
mites with groups of related hosts. The resource-tracking model suggests that the
symbionts track resources (types of food, types of feeding sites) regardless of the
relatedness of hosts. Because related species often share morphological, behav-
ioural and ecological characteristics, particular clades of mites may appear to track
particular clades of hosts. Phylogenetic tracking, or cospeciation, suggests that
symbiont cladogenesis occurs in parallel with host phylogeny. To differentiate
between the processes, one would need to construct independent phylogenetic trees
404 9 Animals as Habitats

for host and mites and compare the branching patterns. If they show a great deal of
congruence, as in the well-known example of pocket gophers and their lice (Hafner
et al. 1994), this supports the cospeciation hypothesis. If not, or if obvious host-
switching across unrelated host taxa is apparent, then the resource-tracking hypoth-
esis is supported. Of course, a mixture of phylogenetic-tracking, host-switching,
duplication of symbiont lineages in the absence of host speciation, and losses of
symbionts due to extinction can co-occur within a set of hosts and symbionts,
resulting in very complex co-phylogenetic patterns (Paterson and Banks 2001; Page
2003). Loss of symbionts may be a random ‘missing-the-boat’ event or it may be
strongly related to host ecology. Felsõ and Rózsa (2006, 2007) found that birds
and mammals that engaged in diving and swimming below the surface had lower
richnesses of lice than related non-diving clades, presumably because adaptation to
this submerged lifestyle is difficult for lice. Diving birds of the family Alcidae
(auks, puffins and their relatives) also have a depauperate feather mite fauna
(S.V. Mironov, pers. comm.).
When the first edition of this book was published in 1999, we could find few
rigorous phylogenetic tests of cospeciation using mites and their hosts. Claims of
parallel histories at that time were usually supported by observations of host speci-
ficity at the species level or by higher-taxon congruity (e.g. ‘all mites of genus
A are found on hosts of family B’); neither rules out resource-tracking as a cause.
Sometimes authors appeared to be selective in their observation of relationships.
Fain and Hyland (1985) note that mites of the family Lemurnyssidae are found in
the nasal cavity of the galago in Madagascar and in noses of South American mon-
keys: “The presence of very closely related mites … suggests the existence of
some relationship between these primates.” But when they describe how
Rhyncoptidae are found on hairs of hystricid rodents, neotropical monkeys and
afrotropical monkeys, and that Audycoptidae are in hair follicles of cebids, ursids
and procyonids, there is no suggestion that monkeys are related to porcupines,
bears or raccoons.
In fact, it was easier to find strikingly disjunct patterns of host and mite phylog-
eny than to find parallel ones. In host-switching (or -jumping), symbiotic mites are
transferred from their normal host taxon to an ecologically associated but phyloge-
netically distant one. After transfer, the mite may develop into a new species and
cospeciation with the new host taxon may occur. For example, larvae of the water
mite genus Arrenurus (Arrenuridae) parasitise Odonata if in the subgenus Arrenurus
but parasitise Diptera if they are members of other subgenera. As dragonflies and
damselflies are about as phylogenetically distant from mosquitoes and midges as
insects can be, the association must be an example of host-switching. An even more
extreme example is pterygosomatid mites parasitic on cockroaches, reduviid bugs
and lizards (Cunliffe 1952). Living in the same habitat might have facilitated the
above host-jumps. Another way to achieve such phylogenetic leaps would be
through transfer from a prey animal to a predator. Gaud and Atyeo (1996) sug-
gest that the feather mite genus Falcolichus, which is found on five species of
African Falco (falcons) and is morphologically similar to pterolichines that
inhabit galliforms, is derived from a prey-to-predator transfer. Similarly, predation
Mite–Host Coevolution: Any Evidence? 405

Table 9.5 A sample of co-phylogenetic studies of mites and their hosts, indicating whether the
focus is on microevolutionary patterns within a species or macroevolutionary ones at or above
the species level
Micro- or
macroevolutionary
Host taxa Mite taxa focus? References
Coleoptera: Mesostigmata: Micro and macro Knee et al. (2012a)
Silphidae Urodinychidae:
Uroobovella
Hymenoptera: Astigmata: Macro Klimov and OConnor
Apoidea Chaetodactylidae (2008)
Reptilia and Insecta Prostigmata: Macro Paredes-León et al.
Pterygosomatidae (2012)
Aves: Charadriiformes Astigmata: Syringobiidae Macro Dabert (2003)
Aves: Mimidae Astigmata: Analgidae: Micro and macro Štefka et al. (2011)
Analges
Aves: Accipitridae and Astigmata: Micro Whiteman et al. (2006)
Phalacrocoracidae Epidermoptidae:
Myialges caulotoon
Aves: Passeriformes Mesostigmata: Micro Morelli and Spicer
Rhinonyssidae: (2007)
Ptilonyssus sairae
Mammalia Astigmata: Sarcoptidae: Micro Alasaad et al. (2012)
Sarcoptes scabiei
Mammalia Astigmata: Psoroptidia Macro Bochkov and Mironov
(2011)
Mammalia: Primates Astigmata: Psoroptidae: Macro Bochkov et al. (2011)
Makialginae
Mammalia: Chiroptera Mesostigmata: Macro Bruyndonckx et al.
Spinturnicidae: (2009)
Spinturnix

related horizontal transfer is the most likely explanation for the discovery of the
finch-associated quill mite Syringophilopsis kirgizorum on an owl (Nattress 2011).
But what a difference a decade makes! Since 1999, there has been an explosion
of studies examining host-mite co-phylogenetic patterns, likely aided both by
improvements in molecular phylogenetic methods and statistical approaches to
testing hypotheses of co-diversification. In most cases, a rigorous phylogenetic
analysis of the mites is compared to previously published phylogenites/taxonomies
of the hosts, but sometimes the authors include their own phylogenetic analysis of
the hosts (e.g., Morelli and Spicer 2007; Štefka et al. 2011). These studies have
been at both the macro- and microevolutionary scales (Table 9.5). Macroevolutionary
studies have provided evidence for both cospeciation (e.g., rhinonyssid mites of
birds, Morelli and Spicer 2007) and for host-switching (e.g., chaetodactylid mites
of bees, Klimov and OConnor 2008; pterygosomatid mites moving from lizards to
arthropods, Paredes-León et al. 2012), often within the same taxon. Host-switching
can sometimes be tied in with ecology. Instances of host-switching in Spinturnix
mites seemed to be associated with similar roosting habits of their bat hosts
406 9 Animals as Habitats

(Bruyndonckx et al. 2009). Whether the mite tree matches the host tree may
depend on the data used to create the host phylogeny. Bochkov et al. (2011)
observed that a morphological phylogeny of primates fit the makialgine mite
(Psoroptidae: Makialginae) tree well, but a molecular phylogeny provided no
co-phylogenetic signal.
There is also some microevolutionary evidence of codiversification of mites and
their hosts, both morphological and molecular. The mite Poecilochirus carabi
(Parasitidae) appears to be in the process of speciating on the basis of host prefer-
ence. Mites at one location in Michigan show no host preferences while at another
site they exhibit clear genetically based preferences between beetle species (Brown
and Wilson 1994). This difference between nearby populations suggests that the
preferences are evolutionarily recent. A similar situation appears to exist in Germany
(Schwarz 1996). Athias-Binche et al. (1993) report potential host-based sympatric
speciation in the uropodine mite Neoseius novus, also phoretic on silphid beetles.
However, these authors do not imply that cospeciation is occurring, as they do not
state whether the silphid beetles are each other’s closest relatives. So resource-
tracking rather than phylogenetic-tracking may be the driving force. In a molecular
phylogenetic analysis of Uroobovella mites on silphids, Knee et al. (2012a) found
that the nominate species U. nova was composed of several host-specific lineages
sufficiently genetically and morphologically distinctive as to represent different
species. Moving to birds as hosts, in a study of morphological differentiation in
subspecies of the silvereye, Zosterops lateralis, Lombert (1988) also performed a
multivariate analysis of the bird’s feather mite, Trouessartia megadisca. The mites
showed morphological differences associated with host subspecies, suggesting
cospeciation may be in progress. Likewise, Whiteman et al. (2006) observed host-
related genetic differentiation among Myialges caulotoon skin mites in the
Galapagos. Also in the Galapagos, Štefka et al. (2011) found that differentiation
among populations of a species of Analges feather mite showed stronger associa-
tions with geography than with the phylogeny of the host mockingbirds.
If one were to embark on a quest for cospeciation, what groups of symbiotic
mites would be the best candidates? Taxa containing species that have broad host
tolerances are probably poor choices (e.g. Pyemotes ventricosus on 139 species of
insects [Eickwort 1983]). Likewise, all 155 Haemaphysalis species (Ixodida) have
a life cycle in which hosts of immature ticks and adults are often hugely different,
e.g. hedgehog and fox, shrew and yak, mouse and lion, rat and boar, lizard and ibex,
bird and bison (Hoogstraal and Kim 1985). Although these associations might make
for tangled cladograms, they are good names for pubs (which are possibly the best
places for such speculation).
Mites with less catholic tastes are more likely to show cospeciation. Terminology
for host specificity is prolific. Monoxeny means that a symbiotic species frequents
only one host species, whereas polyxeny means it infests two or more hosts. If the
hosts are in the same genus this may be termed oligoxeny, and if in different genera
in the same family termed pleioxeny. Kim (1985a) states that host associations of
mammal-parasitic acarines are usually oligoxenous or pleioxenous but that three
families – Demodicidae, Myobiidae and Dermanyssidae – are usually monoxenous.
Mite–Host Coevolution: Any Evidence? 407

Fig. 9.13 Relationships


between species richness of
host taxa and the diversity of
their mite associates: (top)
bee families (or subfamilies
for the Apidae); (bottom)
mammalian orders (Data
from Eickwort 1994 and
Appendix B in Kim 1985b)

Even in these groups, host-switching may occur. For example, predator–prey transfer
of Protomyobia claparedei (Myobiidae) may have occurred between Sorex cinereus
and its predator Blarina brevicauda (both shrews) (Nutting 1985).
Another potential clue is in the intimacy of association between host and symbiont.
Temporary symbionts are associated with the host for only a short period in the life
cycle (e.g. parasitengone larvae), while permanent symbionts spend their entire life
cycle on the host. Permanent symbionts are more likely to be monoxenous (Kim
1985a). Athias-Binche (1991) hypothesised that in mites such as Adactylidium and
Acarophenax, in which males tend to mate with their sisters, genetic drift by founder
effect should result in monoxenous specificity.
Deriving and comparing cladograms both for host taxa and for their symbiotic
mites is a daunting task, as systematic expertise in different classes or phyla would
be required. A single person is unlikely to be, for example, both an avian and an
astigmatan systematist. Is there any easy way of testing some broad prediction of
phylogenetic tracking? One might predict that if mites cospeciate with their hosts,
the species richness of a host taxon should correlate with that of its associated mites.
If mites evolve more slowly than their hosts (Manter’s First Rule), then the
408 9 Animals as Habitats

correlation will not be one-to-one but the pattern of high host richness and high mite
richness should still occur. Data from Eickwort (1994) allow for a simple examina-
tion of numerical correlates between numbers of bees and their associated mites. It
does not support a cospeciation relationship (Fig. 9.13). Eickwort (1994) suggests
that differences in bee ecology and nest structure may explain these discrepan-
cies. Of course, this may also be a reflection of the relative degree of effort expended
on the bee taxa; honey bees, with eight species, have 14 associated mite genera,
while the economically less important Halictidae, with 5,000 species, have only 13
known mite genera. In contrast, if one looks at the number of mite taxa associated
with mammalian orders, there is a positive relationship with the number of species
per order (Fig. 9.13). Perhaps this is because, being mammals ourselves, we have
searched even the most obscure species for mite associates.

Summary

Because of their relatively small size, mites tend to treat other animals as habitats.
All taxa larger than mites have been colonised: insects, arachnids (including other
mites), myriapods, crustaceans, molluscs, annelids and all orders of terrestrial
vertebrates host symbiotic mites. Mites may be temporary or permanent symbionts
and may act as commensals, mutualists, parasites or parasitoids. The causes and
effects of evolutionary transitions between free-living existence, phoresy and para-
sitism are subjects of much interest. The parasitic habit appears easy to evolve but
returning to a free-living lifestyle is a more difficult affair.
Among invertebrate hosts, beetles associated with rotting wood and social insects
are particularly popular with mites. It is easy to understand why the food- and
host-rich nests of social bees, ants and termites have been colonised by numerous
lineages of mites but the attraction of the beetle family Passalidae – which hosts
more than 20 families of acarine symbionts (O. Seeman, pers. comm.) – is not
immediately obvious. Mammals and birds are host to a greater diversity of mites
than are the other terrestrial vertebrate classes. This may be because of the greater
diversity of habitats offered by fur and feathers in comparison to reptilian scales and
amphibian skin. Certain individual hosts may end up more laden by mites than
others, either through direct selection by the mites or ecological overlap between
the two taxa (the ‘right place in the right time’ phenomenon). Active avoidance of
mites through grooming or evasive actions has evolved in a number of host groups.
Effects of mites on their hosts range from beneficial (a seldom-proven interaction)
to neutral, to mildly deleterious, to deadly.
Behavioural ecologists are turning to mites as potential agents of sexual selection
on their hosts and as causes or correlates of developmental asymmetry. The ideas of
coadaptation and cospeciation pervade literature on mite–host associations but
support for either phenomenon is circumstantial at best. At the microevolutionary
level, endocrinological battles between ticks and their vertebrate hosts provide
evidence of a coevolutionary arms race, and there is increasing evidence of
References 409

intraspecific host-race formation among mites associated with many types of hosts.
At the macroevolutionary scale, progress has been made towards phylogenetic tests
of host-mite cospeciation, with several studies showing a diversity of patterns
supporting both host-tracking and host-switching. There are many relatively host-
specific mite taxa that would be candidates for detailed host–parasite cophylogeny
studies. These include Myobiidae and Demodicidae on mammals and numerous
lineages of astigmatan feather mites on their avian hosts.

References

Åbro, A. (1982). The effects of parasitic water mite larvae (Arrenurus spp.) on zygopteran imagoes
(Odonata). Journal of Invertebrate Pathology, 39, 373–381.
Åbro, A. (1984). The initial stylostome formation by parasitic larvae of the water-mite genus
Arrenurus on zygopteran imagines. Acarologia, 25, 33–44.
Åbro, A. (1988). The mode of attachment of mite larvae (Leptus sp.) to harvestmen (Opiliones).
Journal of Natural History, 22, 123–130.
Åbro, A. (1991). The incipient stylostome of parasitic water mite larvae (Arrenurus spp.). Journal
of Parasitology, 77, 313–314.
Aeschlimann, A. (1991). Ticks and disease: Susceptible hosts, reservoir hosts, and vectors. In
A. Toft, A. Aeschlimann, & L. Bolis (Eds.), Parasite–host associations: Coexistence or
conflict? (pp. 148–156). Oxford: Oxford Science.
Alasaad, S., Soglia, D., Spalenza, V., Maione, S., Soriguer, R. C., Pérez, J. M., Rasero, R., Ryser
Degiorgis, M. P., Nimmervoll, H., Zhu, X. Q., & Rossi, L. (2012). Is ITS-2 rDNA suitable
marker for genetic characterization of Sarcoptes mites from different wild animals in different
geographic areas? Veterinary Parasitology, 159, 181–185.
Aoki, J.-I., Takaku, G., & Ito, F. (1994). Aribatidae, a new myrmecophilous oribatid mite family
from Java. International Journal of Acarology, 20, 3–10.
Arlian, L. G. (1989a). Biology and ecology of house dust mites, Dermatophagoides spp. and
Euroglyphus spp. Immunological Allergy Clinician North America, 9, 2335–2339.
Arlian, L. G. (1989b). Biology, host relations, and epidemiology of Sarcoptes scabiei. Annual
Review of Entomology, 34, 139–161.
Arnold, E. N. (1986). Mite pockets of lizards, a possible means of reducing damage by ectopara-
sites. Biological Journal of the Linnean Society, 29, 1–21.
Arnold, E. N. (1993). Comment – function of the mite pockets of lizards: An assessment of a recent
attempted test. Canadian Journal of Zoology, 71, 862–864.
Athias-Binche, F. (1991). Evolutionary ecology of dispersal in mites. In F. Dusbabek & V. Bukva
(Eds.), Modern acarology 1 (pp. 27–41). Prague: SPB Academic.
Athias-Binche, F. (1995). Phenotypic plasticity, polymorphisms in variable environments
and some evolutionary consequences in phoretic mites (Acarina): A review. Ecologie, 26,
225–241.
Athias-Binche, F., Schwarz, H. H., & Meierhofer, I. (1993). Phoretic association of Neoseius novus
(Ouds., 1902) (Acari: Uropodina) with Nicrophorus spp. (Coleoptera: Silphidae): A case of
sympatric speciation? International Journal of Acarology, 19, 75–86.
Atyeo, W. T., & Gaud, J. (1979). Feather mites and their hosts. In J. G. Rodriguez (Ed.), Recent
advances in acarology II (pp. 355–361). New York: Academic Press.
Atyeo, W. T., Kethley, J. B., & Perez, T. M. (1984). Paedomorphosis in Metacheyletia (Acari:
Cheyletidae), with the description of a new species. Journal of Medical Entomology, 21,
125–131.
410 9 Animals as Habitats

Atyeo, W. T., & Pérez, T. M. (1988). Species in the genus Rhytidelasma Gaud (Acarina:
Pterolichidae) from the green conure, Aratinga holochlora (Sclater) (Aves: Psittacidae).
Systematic Parasitology, 11, 85–96.
Bajerlein, D., & Przewoźny, M. (2012). When a beetle is too small to carry phoretic mites? A case
of hydrophilid beetles (Coleoptera: Hydrophilidae) and Uropoda orbicularis (Acari:
Mesostigmata). Canadian Journal of Zoology, 90, 368–375.
Bajerlein, D., & Witaliński, W. (2012). Anatomy and fine structure of pedicellar glands in phoretic
deutonymphs of uropodid mites (Acari: Mesostigmata). Arthropod Structure & Development, 41,
245–257.
Baker, R. A. (1970). Studies on the life history of Ricardoella limacum (Schrank) (Acari:
Trombidiformes). Journal of Natural History, 4, 511–519.
Baker, R. A. (1977). Nutrition of the mite Unionicola intermedia, Koenike and its relationship to
the inflammatory response induced in its molluscan host Anodonta anatina, L. Parasitology,
75, 301–308.
Baker, R. L., & Smith, B. P. (1997). Conflict between antipredator and antiparasite behaviour in
larval damselflies. Oecologia, 109, 622–628.
Bartsch, I. (1987). Australacarus inexpectatus gen. et spec. nov. (Halacaroidea, Acari), mit einer
Übersicht über parasitisch lebende Halacariden. Zoologischer Anzeiger, 218, 17–24.
Bauchau, B. (1997). Do parasitic mites decrease growth of nestling Pied Flycatchers Ficedula
hypoleuca? Ardea, 85, 243–247.
Bauer, A. M., Russell, A. P., & Dollahon, N. R. (1990). Skin folds in the gekkonid genus
Rhacodactylus: A natural test of the damage limitation hypothesis of mite pocket function.
Canadian Journal of Zoology, 68, 1196–1201.
Bauer, A. M., Russell, A. P., & Dollahon, N. R. (1993). Function of the mite pockets of lizards:
A reply to E.N. Arnold. Canadian Journal of Zoology, 71, 865–868.
Beaulieu, F., Déchêne, A. D., & Walter, D. E. (2008). Phase morphs and phoresy: New species of
Antennoseius (Vitzthumia) mites (Acari: Mesostigmata: Ascidae) associated with pyrophilous
carabids (Carabidae: Sericoda spp.) in Alberta, Canada. Zootaxa, 1961, 37–57.
Binns, E. S. (1982). Phoresy as migration – some functional aspects of phoresy in mites. Biological
Reviews, 57, 571–620.
Blanco, G., Tella, J. L., & Potti, J. (1997). Feather mites on group-living Red-billed Choughs:
A non-parasitic interaction? Journal of Avian Biology, 28, 197–206.
Bochkov, A. V. (2002). The classification and phylogeny of the mite superfamily Cheyletoidea
(Acari: Prostigmata). Entomological Review, 81, 488–513 (in Russian).
Bochkov, A. V., & Mironov, S. V. (2011). Phylogeny and systematics of mammal-associated pso-
roptidian mites (Acariformes: Astigmata: Psoroptidia) derived from external morphology.
Invertebrate Systematics, 25, 22–59.
Bochkov, A. V., Klimov, P. B., & Wauthy, G. (2011). Phylogeny and coevolutionary associations of
makialgine mites (Acari, Psoroptidae, Makialginae) provide insight into evolutionary history
of their hosts, strepsirrhine primates. Zoological Journal of the Linnean Society, 162, 1–14.
Bohonak, A.J. (1998). Dispersal and gene flow in freshwater invertebrates. Ph.D. thesis, Cornell
University, Ithaca.
Bohonak, A. J., Smith, B. P., & Thornton, M. (2004). Distributional, morphological and genetic
consequences of dispersal for temporary pond mites. Freshwater Biology, 49, 170–180.
Bonn, A., Gasse, M., Rolff, J., & Martens, A. (1996). Increased fluctuating asymmetry in the dam-
selfly Coenagrion puella is correlated with ectoparasitic water mites: Implications for fluctuat-
ing asymmetry theory. Oecologia, 108, 596–598.
Booth, J. P. (1978). The parasitization of chironomid midges by water-mite larvae in a eutrophic
reservoir in South Wales. Archiv für Hydrobiologie, 84, 1–28.
Böttger, K. (1962). Zur Biologie und Ethologie der einheimischen Wassermilben Arrenurus
(Megaluracarus) globator (Müll.), 1776 Piona nodata nodata (Müll.), 1776 und Eylais infun-
dibulifera meridionalis (Thon), 1899 (Hydrachnellae, Acari). Zoologische Jahrbücher.
Abteilung für Systematik, 89, 501–584.
Brooke, M., & Birkhead, T. (1991). The Cambridge encyclopedia of ornithology. New York:
Cambridge University Press.
References 411

Brooks, D. R., & McLennan, D. A. (1993). Parascript: Parasites and the language of evolution.
Washington: Smithsonian Institution Press.
Brossard, M., & Wikel, S. K. (1997). Immunology of interactions between ticks and hosts. Medical
and Veterinary Entomology, 11, 270–276.
Brown, J. M., & Wilson, D. S. (1994). Poecilochirus carabi: Behavioral and life-history adapta-
tions to different hosts and the consequences of geographical shifts in host communities.
In M. A. Houck (Ed.), Mites: Ecological and evolutionary analyses of life-history patterns
(pp. 1–22). New York: Chapman & Hall.
Bruyndonckx, N., Dubey, S., Ruedi, M., & Christe, P. (2009). Molecular cophylogenetic relationships
between European bats and their ectoparasitic mites (Acari, Spinturnicidae). Molecular
Phylogenetics and Evolution, 51, 227–237.
Burley, N., Tidemann, S. C., & Halupka, K. (1991). Bill colour and parasite levels of zebra finches.
In J. E. Loye & M. Zuk (Eds.), Bird-parasite interactions: Ecology, evolution and behaviour
(pp. 359–376). Oxford: Oxford University Press.
Burtt, E. H., Jr., Chow, W., & Babbitt, G. A. (1991). Occurrence and demography of mites of
tree swallow, house wren, and eastern bluebird boxes. In J. E. Loye & M. Zuk (Eds.), Bird–
parasite interactions: Ecology, evolution and behaviour (pp. 104–122). Oxford: Oxford
University Press.
Camin, J. H., Moss, W. W., Oliver, J. H., Jr., & Singer, G. (1967). Cloacaridae, a new family of
cheyletoid mites from the cloaca of aquatic turtles. Journal of Medical Entomology, 4,
261–272.
Campbell, K. U., Klompen, H., & Crist, T. O. (2012). The diversity and host specificity of mites
associated with ants: The roles of ecological and life history traits of ant hosts. Insectes Sociaux,
60, 31–41. doi:10.1007/s00040-012-0262-6.
Chapman, B. R., & George, J. E. (1991). The effects of ectoparasites on cliff swallow growth and
survival. In J. E. Loye & M. Zuk (Eds.), Bird–parasite interactions: Ecology, evolution and
behaviour (pp. 69–92). Oxford: Oxford University Press.
Cheng, T. C. (1991). Is parasitism symbiosis? A definition of terms and the evolution of concepts.
In A. Toft, A. Aeschlimann, & L. Bolis (Eds.), Parasite–host associations: Coexistence or
conflict? (pp. 16–36). Oxford: Oxford Science Publications.
Choe, J. C., & Kim, K. C. (1989). Microhabitat selection and coexistence in feather mites (Acari:
Analgoidea) on Alaskan seabirds. Oecologia, 79, 10–14.
Clark, L. (1991). The nest protection hypothesis: The adaptive use of plant secondary compounds
by European starlings. In J. E. Loye & M. Zuk (Eds.), Bird–parasite interactions: Ecology,
evolution and behaviour (pp. 205–221). Oxford: Oxford University Press.
Clayton, D. H., & Tompkins, D. M. (1995). Comparative effects of mites and lice on the reproduc-
tive success of rock doves (Columba livia). Parasitology, 110, 195–206.
Clayton, D. H., Koop, J. A. H., Harbison, C. H., Moyer, B. R., & Bush, S. E. (2010). How birds
combat ectoparasites. The Open Ornithology Journal, 3, 41–71.
Clift, A. D., & Larsson, S. F. (1987). Phoretic dispersal of Brennandania lambi (Kcrzal)
(Acari: Tarsonemida: Pygmephoridae) by mushroom flies (Diptera: Sciaridae and Phoridae)
in New South Wales, Australia. Experimental and Applied Acarology, 3, 11–20.
Collins, N. C. (1975). Tactics of host exploitation by a thermophilic water mite. Miscellaneous
Publications of the Entomological Society of America, 9, 250–254.
Colloff, M. J. (2009). Dust mites (p. 583). Dordrecht: CSIRO Publishing and Springer.
Colwell, R. K. (1995). Effects of nectar consumption by the hummingbird flower mite Proctolaelaps
kirmsei on nectar availability in Hamelia patens. Biotropica, 27, 206–217.
Colwell, R. K., & Naeem, S. (1994). Life-history patterns of hummingbird flower mites in relation
to host phenology and morphology. In M. A. Houck (Ed.), Mites: Ecological and evolutionary
analyses of life-history patterns (pp. 23–44). New York: Chapman & Hall.
Cook, W. J., Smith, B. P., & Brooks, R. J. (1989). Allocation of reproductive effort in female
Arrenurus spp. water mites (Acari: Hydrachnidia; Arrenuridae). Oecologia, 79, 184–188.
Cowan, D. P. (1984). Life history and male dimorphism in the mite Kennethiella trisetosa (Acarina:
Winterschmidtiidae), and its sympiotic relationship with the wasp Ancistrocerus antilope.
Annals of the Entomological Society of America, 77, 725–732.
412 9 Animals as Habitats

Cowan, D. P. (1985). Life history and male dimorphism in the mite Kennethiella trisetosa (Acarina:
Winterschmidtiidae), and its sympiotic relationship with the wasp Ancistrocerus antilope.
Annals of the Entomological Society of America, 77, 725–732.
Cross, E. A., & Kaliszewski, M. J. (1988). The life history of a mushroom pest mite, Pediculaster
flechtmanni (Wicht) (Acari: Pygmephoroidea), with studies of alternate morph formation.
Environmental Entomology, 17, 309–315.
Crowell, R. M. (1960). The taxonomy, distribution and developmental stages of Ohio water mites.
Bulletin of the Ohio Biological Survey, 1, 1–77.
Cunliffe, F. (1952). Biology of the cockroach parasite, Pimeliaphilus podapolipophagus Tragardh,
with a discussion of the genera Pimeliaphilus and Hirstiella (Acarina, Pterygosomidae).
Proceedings of the Entomological Society of Washington, 54, 153–169.
Cupp, J. K., & Willis, D. W. (1982). Occurrence of the mite Lebertia in a green sunfish (Lepomis
syanellus). Journal of Parasitology, 68, 876.
Dabert, J. (2003). The feather mite family Syringobiidae Trouessart, 1896 (Acari, Astigmata,
Pterolichoidea). II. Phylogeny and host-parasite evolutionary relationships. Acta Parasitologica,
48 (Supplement), 185–233.
Dabert, J., & Mironov, S. V. (1999). Origin and evolution of feather mites (Astigmata). Experimental
and Applied Acarology, 23, 437–454.
Darolova, A., Hoi, H., & Schleicher, B. (1997). The effect of ectoparasite nest load on the breeding
biology of the penduline tit Remiz pendulinis. Ibis, 139, 115–120.
Davids, C. (1973). The water mite Hydrachna conjecta, bionomics and relation to species of
Corixidae. Netherlands Journal of Zoology, 23, 363–429.
Davids, C. (1991). Water mites: The impact of larvae and adults on their host and prey populations.
In F. Dusbábek & V. Bukva (Eds.), Modern acarology (Vol. 1, pp. 497–501). The Hague:
Academia, Prague and SPB Academic.
Davids, C. (1997a). A new water mite (Acari, Hydrachnidia: Limnesiidae) split off from Limnesia
undulata. Entomologische Berichten Amsterdam, 57, 157–160.
Davids, C. (1997b). The influence of larval parasitism on life history strategies in water mites
(Acari, Hydrachnidia). Archiv für Hydrobiologie, 141, 35–43.
Davies, D. M. (1959). The parasitism of black flies by larval water mites, mainly of the genus
Sperchon. Canadian Journal of Zoology, 37, 353–369.
de Jong, D., Morse, R. A., & Eickwort, G. C. (1982). Mite pests of honey bees. Annual Review of
Entomology, 27, 229–252.
Delfinado-Baker, M., Rath, W., & Boecking, O. (1992). Phoretic bee mites and honeybee grooming
behavior. International Journal of Acarology, 18, 315–320.
Domrow, R. (1981). A small lizard stifled by phoretic deutonymphal mites (Uropodina).
Acarologia, 22, 247–252.
Domrow, R. (1987). Acari Mesostigmata parasitic on Australian vertebrates: An annotated checklist,
keys and bibliography. Invertebrate Taxonomy, 1, 817–948.
Dubinin, V. B. (1951). Feather mites (Analgesoidea). Part I. Introduction to their study. Fauna
U.S.S.R., 6, 1–363 (in Russian).
Duffy, D. C. (1991). Ants, ticks, and nesting seabirds: dynamic interactions? In J. E. Loye &
M. Zuk (Eds.), Bird–parasite interactions: Ecology, evolution and behaviour (pp. 243–257).
Oxford: Oxford University Press.
Dufour, K. W., & Weatherhead, P. J. (1998). Bilateral symmetry as an indicator of male quality in
red-winged blackbirds: Associations with measures of health, viability, and parental effort.
Behavioral Ecology, 9, 220–231.
Dunlop, J. A., Wirth, S., Penney, D., McNeil, A., Bradley, R. S., Withers, P. J., & Preziosi, R. F.
(2012). A minute fossil phoretic mite recovered by phase-contrast X-ray computed tomogra-
phy. Biology Letters, 8, 457–460.
Durden, L. A. (1987). Predator–prey interactions between ectoparasites. Parasitology Today, 3,
306–308.
Durden, L. A., & Keirans, J. E. (1996). Host–parasite coextinction and the plight of tick conserva-
tion. American Entomologist, 42, 87–91.
References 413

Ebermann, E. (1991a). Das Phanomenon Polymorphismus in der Milbenbfamilie Scutacaridae


(Acari, Heterostigmata, Tarsonemina, Scutacaridae). Zoologica, 47, 1–76.
Ebermann, E. (1991b). Records of polymorphism in the mite family Scutacaridae (Acari,
Tarsonemina). Acarologia, 32, 119–138.
Ehrlich, P. R., & Raven, P. H. (1964). Butterflies and plants: A study in coevolution. Evolution, 18,
586–608.
Eickwort, G. C. (1983). Potential use of mites as biological control agents of leaf-feeding insects.
In M. A. Hoy, G. L. Cunningham, & L. Knutson (Eds.), Biological control of pests by mites
(Special publication, Vol. 3304, pp. 41–52). Berkeley: University of California, Agriculture
Experiment Station.
Eickwort, G. C. (1990). Associations of mites with social insects. Annual Review of Entomology,
35, 469–488.
Eickwort, G. C. (1994). Evolution and life-history patterns of mites associated with bees. In
M. A. Houck (Ed.), Mites: Ecological and evolutionary analyses of life-history patterns
(pp. 218–251). New York: Chapman & Hall.
Ellis-Adam, A. C., & Davids, C. (1970). Oviposition and post-embryonic development of the
watermite Piona alpicola (Neuman, 1880). Netherlands Journal of Zoology, 20, 122–137.
Elzinga, R. J. (1982). The genus Antennequesoma (Acari: Uropodina) and descriptions of four new
species. Acarologia, 23, 319–325.
Elzinga, R. J. (1993). Larvamimidae, a new family of mites (Acari: Dermanyssoidea) associated
with army ants. Acarologia, 34, 95–103.
Fain, A. (1969). Adaptation to parasitism in mites. Acarologia, 11, 429–449.
Fain, A., & Hyland, K. W., Jr. (1985). Evolution of astigmatid mites on mammals. In K. C. Kim
(Ed.), Coevolution of parasitic arthropods and mammals (pp. 641–658). New York: Wiley.
Fain, A., & Smiley, R. L. (1989). A new cloacarid mite (Acari: Cloacaridae) from the lungs of the great
horned owl, Bubo virginanus, from the U.S.A. International Journal of Acarology, 15, 111–115.
Fajfer, M. (2012). Acari (Chelicerata) – parasites of reptiles. Acarina, 20, 108–129.
Farfan, M. A., & Klompen, H. (2012). Phoretic mite associates of millipedes (Diplopoda, Julidae) in the
northern Atlantic region (North America, Europe). International Journal of Myriapodology, 7, 69–91.
Farish, D. J., & Axtell, R. C. (1971). Phoresy redefined and examined in Macrocheles muscaedo-
mesticae (Acarina: Macrochelidae). Acarologia, 13, 1–29.
Fashing, N. J., & Chua, T. H. (2002). Systematics and ecology of Naiadacarus, a new species of
Acaridae (Acari: Astigmata) inhabiting the pitchers of Nepenthes bicalcarata Hook. F. in
Brunei Darussalam. International Journal of Acarology, 28, 157–167.
Feinsinger, P. (1983). Coevolution and pollination. In D. J. Futuyma & M. Slatkin (Eds.),
Coevolution (pp. 282–310). Sunderland: Sinauer Associates.
Felsõ, B., & Rózsa, L. (2006). Reduced taxonomic richness of lice (Insecta: Phthiraptera) in diving
birds. Journal of Parasitology, 92, 867–869.
Felsõ, B., & Rózsa, L. (2007). Diving behavior reduces genera richness of lice (Insecta,
Phthiraptera) of mammals. Acta Parasitologica, 52, 82–85.
Forbes, M. R. L. (1991). Ectoparasites and mating success of male Enallagma ebrium damselflies
(Odonata: Coenagrionidae). Oikos, 60, 336–342.
Futuyma, D. J., & Slatkin, M. (1983). Coevolution. Sunderland: Sinauer Associates.
Fuxjager, M. J., Foufopoulos, J., Diaz-Uriarte, R., & Marler, C. A. (2010). Functionally opposing
effects of testosterone on two different types of parasite: Implications for the immunocompe-
tence handicap hypothesis. Functional Ecology, 25, 132–138.
Galván, I., Barba, E., Piculo, R., Cantó, J. L., Cortés, V., Monrós, J. S., Atiénzar, F., & Proctor, H.
(2008). Feather mites and birds: An interaction mediated by uropygial gland size? Journal of
Evolutionary Biology, 21, 133–145.
Galván, I., Aguilera, E., Atiénzar, F., Barba, E., Blanco, G., Cantó, J. L., Cortés, V., Frías, Ó.,
Kovács, I., Meléndez, L., & Møller, A. P. (2012). Feather mites (Acari: Astigmata) and body
condition of their avian hosts: A large correlative study. Journal of Avian Biology, 43, 1–7.
Gaud, J., & Atyeo, W. T. (1996). Feather mites of the world (Acarina, astigmata): The supraspecific
taxa. Part I. Annalen Zoologische Wetenschappen, 277, 1–193.
414 9 Animals as Habitats

Gledhill, T., Crowley, J., & Gunn, R. J. M. (1982). Some aspects of the host-parasite relationships
between adult blackflies (Diptera; Simuliidae) and larvae of the water-mite Sperchon setiger
(Acari: Hydrachnellae) in a small chalk stream in southern England. Freshwater Biology, 12,
345–357.
Gotwald, W. H., Jr. (1996). Mites that live with army ants: A natural history of some myrmecophi-
lous hitch-hikers, browsers and parasites. Journal of the Kansas Entomological Society,
Supplement 69, 232–237.
Gould, S. J., & Lewontin, R. C. (1978). The spandrels of San Marco and the Panglossian paradigm:
a critique of the adaptationist programme. Proceedings of the Royal Society of London, 205,
581–598.
Grossman, J. D., & Smith, R. J. (2008). Phoretic mite discrimination among male burying beetle
(Nicrophorus investigator) hosts. Annals of the Entomological Society of America, 101, 266–271.
Hafner, M. S., Sudman, P. D., Villablanca, F. X., Spradling, T. A., Demastes, J. W., & Nadler, S. A.
(1994). Disparate rates of molecular evolution in cospeciating hosts and parasites. Science,
265, 1087–1090.
Halliday, R. B. (1993). A new species of Scissuralaelaps Womersley (Acarina: Laelapidae)
associated with large Australian cockroaches. Journal of the Australian Entomological Society,
32, 347–353.
Hamilton, W. D., & Zuk, M. (1982). Heritable true fitness and bright birds: A role for parasites?
Science, 218, 384–387.
Harvey, P. H., Read, A. F., John, J. L., Gregory, R. D., & Keymer, A. E. (1991). An evolutionary
perspective: Using the comparative method. In A. Toft, A. Aeschlimann, & L. Bolis (Eds.),
Parasite–host associations: Coexistence or conflict? (pp. 344–355). Oxford: Oxford
Science.
Herre, E. A. (1993). Population structure and the evolution of virulence in nematode parasites of
fig wasps. Science, 259, 1442–1445.
Hevers, J. (1980). Biologisch-ökologische Untersuchungen zum Entwicklungszyklus der in
Deutschland auftretenden Unionicola-Arten (Hydrachnellae, Acari). Archiv für Hydrobiologie,
Supplementband 57, 324–373.
Hoogstraal, H., & Kim, K. C. (1985). Tick and mammal coevolution, with emphasis on
Haemaphysalis. In K. C. Kim (Ed.), Coevolution of parasitic arthropods and mammals
(pp. 505–568). New York: Wiley.
Houck, M. A. (1994). Adaptation and transition into parasitism from commensalism: A phoretic
model. In M. A. Houck (Ed.), Mites: Ecological and evolutionary analyses of life-history
patterns (pp. 252–281). New York: Chapman & Hall.
Houck, M. A., & OConnor, B. M. (1991). Ecological and evolutionary significance of phoresy in
the Astigmata. Annual Review of Entomology, 36, 611–636.
Huck, K., Schwarz, H. H., & Schmid-Hempel, P. (1998). Host choice in the phoretic mite
Parasitellus fucorum (Mesostigmata: Parasitidae): Which bumblebee caste is the best?
Oecologia, 115, 385–390.
Hughes, A. M. (1976). The mites of stored food and houses (Ministry of agriculture, fisheries and
food, Technical bulletin 2nd ed., Vol. 9). London: H.M.S.O.
Hunter, P. E., & Rosario, R. M. T. (1988). Associations of Mesostigmata with other arthropods.
Annual Review of Entomology, 33, 393–417.
Hurst, G. D. D., Sharpe, R. G., Broomfield, A. H., Walker, L. E., Majerus, T. M. O., Zakharov, I. A.,
& Majerus, M. E. N. (1995). Sexually transmitted disease in a promiscuous insect, Adalia
bipunctata. Ecological Entomology, 20, 230–236.
Itu, F., & Takaku, G. (1994). Obligate myrmecophily in an oribatid mite. Naturwissenschaften, 81,
180–182.
Iverson, K., OConnor, B. M., Ochoa, R., & Heckmann, R. (1996). Lardoglyphus zacheri (Acari:
Lardoglyphidae), a pest of museum dermestid colonies, with observations on its natural ecology
and distribution. Annals of the Entomological Society of America, 89, 544–549.
Janzen, D. (1980). When is it coevolution? Evolution, 34, 611–612.
References 415

Kaliszewski, M., Athias-Binche, F., & Lindquist, E. E. (1995). Parasitism and parasitoidism in
Tarsonemina (Acari: Heterostigmata) and evolutionary considerations. Advances in Parasitology,
35, 335–367.
Kethley, J. (1971). Population regulation in quill mites (Acarina: Syringophilidae). Ecology, 52,
113–118.
Kettle, D. S. (1995). Medical and veterinary entomology (2nd ed.). Wallingford: CAB International.
Kim, K. C. (1985a). Evolutionary relationships of parasitic arthropods and mammals. In K. C. Kim
(Ed.), Coevolution of parasitic arthropods and mammals (pp. 3–82). New York: Wiley.
Kim, K. C. (1985b). Coevolution of parasitic arthropods and mammals. New York: Wiley.
Kim, K. C. (1985c). Parasitism and coevolution: Epilogue. In K. C. Kim (Ed.), Coevolution of
parasitic arthropods and mammals (pp. 661–682). New York: Wiley.
Klimov, P. B., & OConnor, B. M. (2008). Morphology, evolution, and host associations of
bee-associated mites of the family Chaetodactylidae (Acari: Astigmata) with a monographic
revision of North American taxa (Miscellaneous publications, Vol. 199, pp. 1–243). Ann Arbor:
Museum of Zoology, University of Michigan.
Klimov, P. B., & OConnor, B. (2013). Is permanent parasitism reversible? Critical evidence from
early evolution of house dust mites. Systematic Biology. doi:10.1093/sysbio/syt008.
Klompen, J. S. H., Black, W. C., IV, Keirans, J. E., & Oliver, J. H., Jr. (1996). Evolution of ticks.
Annual Review of Entomology, 41, 141–161.
Knee, W., Beaulieu, F., Skevington, J. H., Kelso, S., & Forbes, M. R. (2012a). Cryptic species of mites
(Uropodoidea: Uroobovella spp.) associated with burying beetles (Silphidae: Nicrophorus):
The collapse of a host generalist revealed by molecular and morphological analyses. Molecular
Phylogenetics and Evolution, 65, 276–286.
Knee, W., Beaulieu, F., Skevington, J. H., Kelso, S., Cognato, A. I., & Forbes, M. R. (2012b).
Species boundaries and host range of tortoise mites (Uropodoidea) phoretic on bark beetles
(Scolytinae), using morphometric and molecular markers. PLoS One, 7, e47243.
Knülle, W. (1987). Genetic variability and ecological adaptability of hypopus formation in a stored
product mite. Experimental and Applied Acarology, 3, 31–32.
Knülle, W. (1991). Genetic and environmentl determinants of hypopus duration in the stored-product
mite Lepidoglyphus destructor. Experimental and Applied Acarology, 10, 231–258.
Krantz, G. W. (1978). A manual of acarology. Corvallis: Oregon State University Bookstores.
Krantz, G. W. (1983). Mites as biological control agents of dung-breeding flies, with special refer-
ence to the Macrochelidae. In M. A. Hoy, G. L. Cunningham, & L. Knutson (Eds.), Biological
control of pests by mites (Special publication, Vol. 3304, pp. 91–98). Berkeley: University of
California, Agriculture Experiment Station.
Krantz, G. W., & Walter, D. E. (Eds.). (2009). A manual of acarology (3rd ed.). Lubbock: Texas
Tech University Press.
Labrzycka, A. (2006). A perfect clasp – adaptation of mites to parasitize mammalian fur. Biological
Letters, 43, 109–118.
Lanciani, C. A. (1976). Intraspecific competition in the parasitic water mite Hydryphantes tenuabilis.
American Midland Naturalist, 96, 210–214.
Lanciani, C. A. (1984). Crowding in the parasitic stage of the water mite Hydrachna virella (Acari:
Hydrachnidae). Journal of Parasitology, 70, 270–272.
Lanciani, C. A., & Boyt, A. D. (1977). The effect of a parasitic water mite, Arrenurus pseutotenui-
collis (Acari: Hydrachnellae), on the survival and reproduction of the mosquito Anopheles
crucians. Journal of Medical Entomology, 14, 10–15.
Lanciani, C. A., & Smith, B. P. (1989). Constancy of stylostome form in two water mite species.
Canadian Entomologist, 121, 439–443.
Lesna, I., Sabelis, M. W., van Niekerk, T. G. C. M., & Komdeur, J. (2012). Laboratory tests for
controlling poultry red mites (Dermanyssus gallinae) with predatory mites in small ‘laying
hen’ cages. Experimental and Applied Acarology, 58, 371–383.
Lindholm, A. K., Benter, G. J., & Ueckermann, E. A. (1998). Persistence of passerine ectoparasites
on the diederik cuckoo Chrysococcyx caprius. Journal of Zoology, 244, 145–153.
416 9 Animals as Habitats

Lindquist, E. E. (1969). Review of Holarctic tarsonemid mites (Acarina: Prostigmata) parasitizing


eggs of ipine bark beetles. Memoirs of the Entomological Society of Canada, 60, 1–111.
Lindquist, E. E. (1985). Discovery of sporothecae in adult female Trochometridium Cross, with
notes on analogous structures in Siteroptes Amerling (Acari: Heterostigmata). Experimental
and Applied Acarology, 1, 73–85.
Lindquist, E. E., & Walter, D. E. (1989). Biology and description of Antennoseius janus, new
species (Mesostigmata: Ascidae), a mesostigmatic mite exhibiting adult female dimorphism.
Canadian Journal of Zoology, 67, 1291–1310.
Lombert, H.A.P.M. (1988). Co-evolution of three eastern Australian subspecies of the passerine
bird Zosterops lateralis (Latham, 1801) and their ectoparasitic feather mite Trouessartia
megadisca Gaud, 1962. Ph.D. thesis, Australian School of Environmental Studies, Griffith
University, Nathan, Australia.
Makol, J., Kłosińska, A., & Łaydanowicz, J. (2012). Host–parasite interactions within terrestrial
Parasitengonina (Acari, Trombidiformes, Prostigmata). International Journal of Acarology, 38,
18–22.
Masan, P. (1997). Changes in infestation rate and age structure of Dermanyssus hirundinis and
Ornithonyssus sylviarum (Acarina) during nidification and breeding period of penduline tit.
Journal of Medical Entomology, 34, 609–614.
McLachlan, A., Pike, T., & Thomason, J. (2008). Another kind of symmetry: Are there adaptive
benefits to the arrangement of mites on an insect host? Ethology Ecology and Evolution, 20,
257–270.
Mestre, A., Mesquita-Joanes, F., Proctor, H., & Monrós, J. S. (2011). Different scales of spatial
segregation of two species of feather mites on the wings of a passerine bird. The Journal of
Parasitology, 97, 237–244.
Meyer, E. (1985). Der Entwicklungszyklus von Hydrodroma despiciens (O.F. Müller 1776) (Acari:
Hydrodromidae). Archiv für Hydrobiologie, Supplementband 66, 321–453.
Miko, L., & Stanko, M. (1991). Small mammals as carriers of non-parasitic mites (Oribatida,
Uropodina). In F. Dusbabek & V. Bukva (Eds.), Modern acarology 1 (pp. 395–402). Prague:
SPB Academic.
Mitchell, R. (1960). Behavior of the larvae of Arrenurus fissicornis Marshall, a water mite parasitic
on dragonflies. Journal of Animal Ecology, 9, 220–224.
Mitchell, R. (1965). Population regulation of a water mite parasitic on unionid mussels. Journal of
Parasitology, 51, 990–996.
Mitchell, R. (1967). Host exploitation of two closely related water mites. Evolution, 21, 59–75.
Mitchell, R. (1970). An analysis of dispersal in mites. American Naturalist, 104, 425–431.
Møller, A. P. (1991). Parasites, sexual ornaments, and mate choice in the barn swallow. In
J. E. Loye & M. Zuk (Eds.), Bird–parasite interactions: Ecology, evolution and behaviour
(pp. 328–343). Oxford: Oxford University Press.
Møller, A. P. (1992). Parasites differentially increase the degree of fluctuating asymmetry in sec-
ondary sexual characters. Journal of Evolutionary Biology, 5, 691–699.
Møller, A. P., & Saino, N. (1994). Parasites, immunology of hosts, and host sexual selection.
Journal of Parasitology, 80, 850–858.
Mooring, M. S., McKenzie, A. A., & Hart, B. L. (1996). Role of sex and breeding status in groom-
ing and total tick load of impala. Behavioural Ecology and Sociobiology, 39, 259–266.
Morales-Malacara, J. B. (1996). Mesostigmatid (Mesostigmata) ectoparasites of bats in México. In
R. Mitchell, D. J. Horn, G. R. Needham, & W. C. Welbourn (Eds.), Acarology IX: Volume 1,
proceeding (pp. 105–108). Columbus: Ohio Biological Survey.
Morelli, M., & Spicer, G. S. (2007). Cospeciation between the nasal mite Ptilonyssus sairae
(Acari: Rhinonyssidae) and its bird hosts. Systematic and Applied Acarology, 12, 179–188.
Moser, J. C., & Cross, E. A. (1975). Phoretomorph: A new phoretic phase unique to the Pyemotidae
(Acarina: Tarsonemoidea). Annals of the Entomological Society of America, 68, 820–822.
Moss, W. W. (1979). Patterns of host-specificity and co-evolution in the Harpyrhynchidae. In
J. G. Rodriguez (Ed.), Recent advances in acarology II (pp. 379–384). New York: Academic
Press.
References 417

Mostafa, A. R. (1974). Biological and behavioral aspects of the lizard mite Pterygosoma mutabilis
Jack, 1961 (Acarina: Pterygosomidae). Acarologia, 16, 100–105.
Moya Borja, G. E. (1981). Effects of Macrocheles muscadomesticae (Scopoli) on the sexual
behavior and longevity of Dermatobia hominis. Revista Brasileira de Biologia, 41, 237–241.
Nadchatram, M. (2006). A review of endoparasitic acarines of Malaysia with special reference to
novel endoparasitism of mites in amphibious sea snakes and supplementary notes on ecology
of chiggers. Tropical Biomedicine, 23, 1–22.
Nattress, B. (2011). Horizontal transmission of Syrngophilopsis [sic] kirgizorum (Acari:
Cheyletoidea: Syringophilidae). Acarina, 19, 270.
Newell, I. M. (1963). Feeding habits in the genus Balaustium (Acarina, Erythraeidae), with special
reference to attacks on man. Journal of Parasitology, 49, 498–502.
Newell, I. M., & Ryckman, R. E. (1964). Hirstiella pyriformis sp. n. (Acari, Pterygosomidae), a
new parasite of lizards from Baja California. Journal of Parasitology, 50, 163–171.
Norton, R. A. (1980). Observations on phoresy by oribatid mites (Acari: Oribatei). International
Journal of Acarology, 6, 121–129.
Nutting, W. B. (1985). Prostigmata–Mammalia: Validation of coevolutionary phylogenies. In K. C. Kim
(Ed.), Coevolution of parasitic arthropods and mammals (pp. 569–640). New York: Wiley.
OConnor, B. M. (1979). Evolutionary origins of astigmatid mites inhabiting stored products. In
J. G. Rodriguez (Ed.), Recent advances in acarology I (pp. 273–278). New York: Academic
Press.
OConnor, B. M. (1982). Evolutionary ecology of astigmatid mites. Annual Review of Entomology,
27, 385–409.
OConnor, B.M. (1984). Acarine–fungal relationships: The evolution of symbiotic associations. In
Q. Wheeler, & M. Blackwell (Eds.), Fungus–insect relationships: Perspectives in ecology and
evolution (pp. 354–381). New York: Columbia University Press. [this is cited as 1984a in the
text but there is now no 1984b, so the ‘a’ should be removed] (p. 226)
OConnor, B. M. (1994). Life-history modifications in astigmatid mites. In M. A. Houck (Ed.),
Mites: Ecological and evolutionary analyses of life-history patterns (pp. 136–159). New York:
Chapman & Hall.
Okabe, K., & Makino, S. (2008a). Life cycle and sexual mode adaptations of the parasitic mite
Ensliniella parasitica (Acari: Winterschmidtiidae) to its eumenine wasp host, Allodynerus
delphinalis (Hymenoptera: Vespidae). Canadian Journal of Zoology, 86, 470–478.
Okabe, K., & Makino, S. (2008b). Parasitic mites as part-time bodyguards of a host wasp.
Proceedings of the Royal Society B, 275, 2293–2297.
Okabe, K., & Makino, S. (2010). Conditional mutualism between Allodynerus delphinalis
(Hymenoptera: Vespidae) and Ensliniella parasitica (Astigmata: Winterschmidtiidae) may
determine maximum parasitic mite infestation. Environmental Entomology, 39, 424–429.
Owen, J. P., Delany, M. E., Cardona, C. J., Bickford, A. A., & Mullens, B. A. (2009). Host inflam-
matory response governs fitness in an avian ectoparasite, the northern fowl mite (Ornithonyssus
sylviarum). International Journal of Parasitology, 39, 789–799.
Owens, I. P. F., & Short, R. V. (1995). Hormonal basis of sexual dimorphism in birds: Implications
for new theories of sexual selection. Trends in Ecology & Evolution, 10, 44–47.
Pacejka, A. J., Gratton, C. M., & Thompson, C. F. (1998). Do potentially virulent mites affect
house wren (Troglodytes aedon) reproductive success? Ecology, 79, 1797–1806.
Page, R. D. M. (Ed.). (2003). Tangled trees. Phylogeny, cospeciation and coevolution. Chicago:
The University of Chicago Press.
Pap, P. L., Tökölyi, J., & Szép, T. (2005). Host–symbiont relationship and abundance of feather
mites in relation to age and body condition of the barn swallow (Hirundo rustica): An experi-
mental study. Canadian Journal of Zoology, 83, 1059–1066.
Paredes-León, R., Klompen, H., & Pérez, T. M. (2012). Systematic revision of the genera
Geckobiella Hirst, 1917 and Hirstiella Berlese, 1920 (Acari: Prostigmata: Pterygosomatidae)
with description of a new genus for American species parasites on geckos formerly placed in
Hirstiella. Zootaxa, 3510, 1–40.
418 9 Animals as Habitats

Paterson, A. M., & Banks, J. C. (2001). Analytical approaches to measuring cospeciation of host
and parasites: Through a glass darkly. International Journal of Parasitology, 31, 1012–1022.
Pérez, T. M. (1996). The eggs of seven species of Fainalges Gaud and Berla (Xolalgidae) from the
green conure (Aves, Psittacidae). In R. Mitchell, D. J. Horn, G. R. Needham, & W. C. Welbourn
(Eds.), Acarology IX: Volume 1, proceedings (pp. 297–300). Columbus: Ohio Biological Survey.
Pérez, T. M. (1997). Eggs of feather mite congeners (Acarina: Pterolichidae, Xolalgidae) from
different species of new world parrots (Aves, Psittaciformes). International Journal of
Acarology, 23, 103–106.
Pérez, T. M., & Atyeo, W. T. (1984a). Feather mites, feather lice, and thanatochresis. Journal of
Parasitology, 70, 807–812.
Pérez, T. M., & Atyeo, W. T. (1984b). Site selection of the feather and quill mites of Mexican par-
rots. In D. A. Griffiths & C. E. Bowman (Eds.), Acarology VI (Vol. 1, pp. 563–570). Chichester:
Ellis Horwood Ltd.
Polak, M. (1997). Ectoparasitism in mothers causes higher positional fluctuating asymmetry in
their sons: Implications for sexual selection. American Naturalist, 149, 955–974.
Polak, M. (1998). Effects of ectoparasitism on host condition in the Drosophila–Macrocheles sys-
tem. Ecology, 79, 1807–1817.
Polak, M., & Markow, T. A. (1995). Effect of ectoparasitic mites on sexual selection in a Sonoran
fruit fly. Evolution, 49, 660–669.
Polak, M., Luong, L. T., & Starmer, W. T. (2007). Parasites physically block host copulation:
A potent mechanism of parasite-mediated sexual selection. Behavioral Ecology, 18, 952–957.
Poulin, R. (1995). Clutch size and egg size in free-living and parasitic copepods: A comparative
analysis. Evolution, 49, 325–336.
Proctor, H. C. (2003). Feather mites (Acari: Astigmata): Ecology, behavior and evolution. Annual
Review of Entomology, 48, 185–209.
Proctor, H. C., & Jones, D. N. (2004). Geographical structuring of feather mite assemblages from
the Australian brush-turkey (Aves: Megapodiidae). The Journal of Parasitology, 90, 60–66.
Proctor, H. C., & Owens, I. (2000). Mites and birds: Diversity, parasitism and coevolution. Trends
in Ecology & Evolution, 15, 358–364.
Proctor, H. C., & Pritchard, G. (1989). Neglected predators: Water mites (Acari: Parasitengona:
Hydrachnellae) in freshwater communities. Journal of the North American Benthological
Society, 8, 100–111.
Proctor, H. C., Gray, H. M., & OConnor, B. M. (1997). Subaquatic mites (Acari: Astigmata)
associated with adult freshwater leeches (Hirudinea: Erpobdellidae). Journal of Natural
History, 31, 539–544.
Pruett-Jones, M., & Pruett-Jones, S. (1991). Analysis and ecological correlates of tick burdens in a
New Guinea avifauna. In J. E. Loye & M. Zuk (Eds.), Bird–parasite interactions: Ecology,
evolution and behaviour (pp. 154–176). Oxford: Oxford University Press.
Quintero, M. T., & Acevedo, H. A. (1991). Mites of fermented liquid foods in Mexico. In
F. Dusbabek & V. Bukva (Eds.), Modern acarology 1 (pp. 611–614). Prague: SPB Academic.
Radovsky, F. J. (1985). Evolution of mammalian mesostigmate mites. In K. C. Kim (Ed.),
Coevolution of parasitic arthropods and mammals (pp. 441–568). New York: Wiley.
Radovsky, F. J., Krantz, G. W., & Whitaker, J. O., Jr. (1997). A remarkable example of predation
in the parasitic mite family Macronyssidae. International Journal of Acarology, 23, 3–6.
Rasero, R., Rossi, L., Soglia, D., Maione, S., Sacchi, P., Rambozzi, L., Sartore, S., Soriguer, R. C.,
Spalenza, V., & Alasaad, S. (2010). Host taxon-derived Sarcoptes mite in European wild
animals revealed by microsatellite markers. Biological Conservation, 143, 1269–1277.
Raut, S. K., & Panigrahi, A. (1991). The mite Fuscuropoda marginata (C.L. Koch) for the control
of pest slugs Laevicuaulis alte (Férussac). In F. Dusbabek & V. Bukva (Eds.), Modern
acarology II (pp. 683–687). Prague: SPB Academic.
Read, C. P. (1951). The ‘crowding effect’ in tapeworm infections. Journal of Parasitology, 37,
174–178.
Reinhardt, K. (1996). Negative effects of Arrenurus water mites on the flight distances of the
damselfly Nehalennia speciosa (Odonata: Coenagrionidae). Aquatic Insects, 18, 233–240.
References 419

Rettenmeyer, C. W., Rettenmeyer, M. E., Joseph, J., & Berghoff, S. M. (2011). The largest animal
association centered on one species: The army ant Eciton burchellii and its more than 300
associates. Insectes Sociaux, 58, 281–292.
Ribeiro, J. M. C., Labruna, M. B., Mans, B. J., Maruyama, S. R., Francischetti, I. M. B., Barizon,
G. C., & de Miranda Santos, I. K. F. (2012). The sialotranscriptome of Antricola delacruzi
female ticks is compatible with nonhematophagous behavior and an alternative source of food.
Insect Biochemistry and Molecular Biology, 42(5), 332–342. doi:10.1016/j.ibmb.2012.01.003.
Robinson, J. V. (1983). Effects of water mite parasitism on the demographics of an adult population of
Ischnura posita (Hagen)(Odonata: Coenagrionidae). American Midland Naturalist, 109,
169–174.
Rolff, J. (1997). Better hosts dive: Detachment of ectoparasitic water mites (Hydrachnellae: Arrenuridae)
from damselflies (Odonata: Coenagrionidae). Journal of Insect Behaviour, 10, 819–827.
Saino, N., Møller, A. P., & Bolzern, A. M. (1995). Testosterone effects on the immune system
and parasite infestations in the barn swallow (Hirundo rustica): An experimental test of the
immunocompetence hypothesis. Behavioural Ecology, 6, 397–404.
Salvador, A., Veiga, J. P., Martin, J., Lopez, P., Abelenda, M., & Puerta, M. (1996). The cost of
producing a sexual signal: Testosterone increases the susceptibility of male lizards to ectopara-
sitic infestation. Behavioural Ecology, 7, 145–150.
Sammataro, D., & Needham, G. R. (1996). Host-seeking behaviour of tracheal mites (Acari:
Tarsonemidae) on honey bees (Hymenoptera: Apidae). Experimental and Applied Acarology,
20, 121–136.
Schalk, G., & Forbes, M. R. (1997). Male biases in parasitsim of mammals: Effects of study type,
host age, and parasite taxon. Oikos, 78, 67–74.
Schroder, R. F. W. (1982). Effect of infestation with Coccipolipus epilachnae Smiley (Acarina:
Podapolipidae) on fecundity and longevity of the Mexican bean beetle. International Journal
of Acarology, 8, 81–84.
Schwarz, H. H. (1996). Host range and behavioral preferences in German sibling species of the
Poecilochirus carabi complex (Acari: Mesostigmata: Parasitidae). International Journal of
Acarology, 22, 135–140.
Schwarz, H. H., & Müller, J. K. (1992). The dispersal behaviour of the phoretic mite Poecilochirus
carabi (Mesostigmata, Parasitidae): Adaptation to the breeding biology of its carrier
Necrophorus vespilloides (Coleoptera, Silphidae). Oecologia, 89, 487–493.
Seeman, O. D. (2008). Systematics and phylogeny of Chrysomelobia species (Acari: Podapolipidae),
sexually transmitted parasites of chrysomelid beetles. Invertebrate Systematics, 22, 55–84.
Sengbusch, H. G., & Hauswirth, J. W. (1986). Prevalence of hair follicle mites, Demodex fol-
liculorum and D. brevis (Acari: Demodicidae), in a selected human population in western
New York, USA. Journal of Medical Entomology, 23, 384–388.
Siddall, M. E., & Burreson, E. M. (1995). Phylogeny of the Euhirudinea: Independent evolution of
blood feeding by leeches? Canadian Journal of Zoology, 73, 1048–1064.
Sinha, R. N. (1979). Role of Acarina in the stored grain ecosystem. In J. G. Rodriguez (Ed.),
Recent advances in acarology I (pp. 263–272). New York: Academic Press.
Skaife, S. H. (1952). The yellow-banded carpenter bee, Mesotrichia caffra Linn., and its symbiotic
mite, Dinogamasus brausni Vitzthun. Journal of the Entomological Society of South Africa, 15,
63–76.
Smith, B. P. (1983). The potential of mites as biological control agents of mosquitoes. In M. Hoy,
G. Cunningham, & L. Knutson (Eds.), Biological control of pests by mites (Special publication,
Vol. 3304, pp. 79–85). Berkeley: Agricultural Experiment Station, University of California.
Smith, B. P. (1988). Host–parasite interaction and impact of larval water mites on insects. Annual
Review of Entomology, 33, 487–507.
Smith, B. P. (1989). Impact of parasitism by larval Limnochares aquatica (Acari: Hydrachnidia;
Limnocharidae) on juvenile Gerris comatus, Gerris alacris, and Gerris buenoi (Insecta:
Hemiptera; Gerridae). Canadian Journal of Zoology, 67, 2238–2243.
Smith, B. P. (1998). Loss of larval parasitism in parasitengonine mites. Experimental and Applied
Acarology, 22, 187–199.
420 9 Animals as Habitats

Smith, I. M., & Cook, D. R. (1991). Water mites. In J. H. Thorp & A. P. Covich (Eds.), Ecology
and classification of North American freshwater invertebrates (pp. 523–592). San Diego:
Academic Press.
Smith, B. P., & McIver, S. B. (1984a). Factors influencing host selection and successful parasitism of
Aedes spp. mosquitoes by Arrenurus spp. mites. Canadian Journal of Zoology, 62, 1114–1120.
Smith, B. P., & McIver, S. B. (1984b). The patterns of mosquito emergence (Diptera: Culcidae;
Aedes spp.): Their influence on host selection by parasitic mites (Acari: Arrenuridae; Arrenurus
spp.). Canadian Journal of Zoology, 62, 1106–1113.
Smith, I. M., & Oliver, D. R. (1986). Review of parasitic associations of larval water mites (Acari:
Parasitengona: Hydrachnida) with insect hosts. Canadian Entomologist, 118, 407–472.
Soar, C. D., & Williamson, W. (1925). The British Hydracarina (Vol. I). London: Adlard & Son
and West Newman Ltd.
Soler, J. J., Peralta-Sánchez, J. M., Martín-Platero, A. M., Martín-Vivaldi, M., Martínez-Bueno,
M., & Møller, A. P. (2012). The evolution of size of the uropygial gland: Mutualistic feather
mites and uropygial secretion reduce bacterial loads of eggshells and hatching failures of
European birds. Journal of Evolutionary Biology, 25, 1779–1791.
Sorci, G., de Fraipont, M., & Clobert, J. (1997). Host density and ectoparasite avoidance in the
common lizard (Lacerta vivipara). Oecologia, 111, 183–188.
Štefka, J. P. E. A., Hoeck, L. F. K., & Smith, V. S. (2011). A hitchhikers guide to the Galápagos:
Co-phylogeography of Galápagos mockingbirds and their parasites. BMC Evolutionary
Biology, 11, 284. doi:10.1186/1471-2148-11-284.
Strathmann, R. R. (1978). The evolution and loss of feeding larval stages of marine invertebrates.
Evolution, 32, 894–906.
Suárez-Rodríguez, M., López-Rull, I., & Garcia, C. M. (2013). Incorporation of cigarette butts into
nests reduces nest ectoparasite load in urban birds: new ingredients for an old recipe? Biology
Letters, 9(1), 20120931. doi:10.1098/rsbl.2012.0931.
Tedla, S., & Fernando, C. H. (1970). Some aspects of the ecology of parasite fauna of the gills of
yellow perch (Perca flavescens). Journal of the Fisheries Research Board of Canada, 27,
1045–1050.
Thompson, J. N. (1989). Concepts of coevolution. Trends in Ecology & Evolution, 4, 179–183.
Thompson, C. W., Hillgarth, N., Leu, M., & McClure, H. E. (1997). High parasite load in house
finches (Carpodacus mexicanus) is correlated with reduced expression of a sexually selected
trait. American Naturalist, 149, 270–294.
Timm, R. M., & Clauson, B. L. (1985). Mammals as evolutionary partners. In K. C. Kim (Ed.),
Coevolution of parasitic arthropods and mammals (pp. 101–154). New York: Wiley.
Toft, C. A., & Aeschlimann, A. (1991). Introduction: Coexistence or conflict? In C. A. Toft,
A. Aeschlimann, & L. Bolis (Eds.), Parasite–host associations: Coexistence or conflict?
(pp. 1–12). Oxford: Oxford Science.
Treat, A. E. (1975). Mites of moths and butterflies. Ithaca: Cornell University Press.
Ullrich, F. (1976). Biologisch-ökologische Studien an rheophilen Wassermilben (Hydrachnellae,
Acari), unter besonderer Berücksichtigung von Sperchon setiger (Thor 1898). Ph.D. thesis,
Kiel: University of Kiel.
Untergasser, D. (1989). Handbook of fish diseases. Neptune City: T.F.H. Publications. German
edition: (trans: Hirschhorn, H.H.).
Václav, R., Prokop, P., & Fekiač, V. (2007). Expression of breeding coloration in European Green
Lizards (Lacerta viridis): Variation with morphology and tick infestation. Canadian Journal of
Zoology, 85, 1199–1206.
van Bronswijk, J. E. M. H. (1979). House-dust as an ecosystem. In J. G. Rodriguez (Ed.), Recent
advances in acarology II (pp. 167–172). New York: Academic Press.
Waage, J. K. (1979). The evolution of insect/vertebrate associations. Biological Journal of the
Linnean Society, 12, 187–224.
Walther, B. A., & Clayton, D. H. (1997). Dust-ruffling: A simple method for quantifying ectopara-
site loads of live birds. Journal of Field Ornithology, 68, 509–518.
References 421

Weatherhead, P. J., Metz, K. J., Bennett, G. F., & Irwin, R. E. (1993). Parasite faunas, testosterone
and secondary sexual traits in male red-winged blackbirds. Behavioural Ecology and
Sociobiology, 33, 13–23.
Welbourn, W. C. (1983). Potential use of trombidioid and erythraeoid mites as biological control
agents of insect pests. In M. A. Hoy, G. L. Cunningham, & L. Knutson (Eds.), Biological
control of pests by mites (Special publication, Vol. 3304, pp. 103–140). Berkeley: University of
California, Division of Agriculture and Natural Resources, Agricultural Experiment Station.
Wendt, F.-E., Wohltmann, A., Eggers, A., & Otto, J. C. (1994). Studies on parasitism, development
and phenology of Johnstoniana parva n. sp. (Acari: Parasitengonae: Johnstonianidae) includ-
ing a description of all active instars. Acarologia, 35, 49–63.
Wheeler, W. M. (1919). The phoresy of Antherophagus. Psyche, 26, 145–152.
Wheeler, W. M. (1923). Social life among the insects. New York: Harcourt, Brace & World.
Whiteman, N. K., Sánchez, P., Merkel, J., Klompen, H., & Parker, P. G. (2006). Cryptic host
specificity of an avian skin mite (Epidermoptidae) vectored by louseflies (Hippoboscidae)
associated with two endemic Galapagos bird species. Journal of Parasitology, 92, 1218–1228.
Wiggins, G. B., Mackay, R. J., & Smith, I. M. (1980). Evolutionary and ecological strategies of
animals in annual temporary pools. Archiv für Hydrobiologie, Supplementband 58, 97–206.
Wikel, S. K., & Bergman, D. (1997). Tick host immunology: Significant advances and challenging
opportunities. Parasitology Today, 13, 383–389.
Wiles, P. R. (1987). Observations on the parasitic biology of the watermite Hydrodroma despiciens
pilosa Besseling (Acari: Hydrodromidae). Archiv für Hydrobiologie, Supplementband 76,
369–392.
Willadsen, P. (2004). Anti-tick vaccines. Parasitology, 129, S367–S387.
Wilson, D. S. (1983). The effect of population structure on the evolution of mutualism: A field test
involving burying beetles and their phoretic mites. American Naturalist, 121, 851–870.
Wilson, D. S., & Knollenberg, W. G. (1987). Adaptive indirect effects: The fitness of burying
beetles with and without their phoretic mites. Evolutionary Ecology, 1, 139–159.
Wirth, S. (2009). Necromenic lifestyle of Histiostoma polypori (Acari: Histiostomatidae).
Experimental and Applied Acarology, 49, 317–327.
Wohltmann, A. (1996). Parasitism and life cycle strategies in Microtrombidiidae (Prostigmata:
Parasitengonae) from Europe. In R. Mitchell, D. J. Horn, G. R. Needham, & W. C. Welbourn
(Eds.), Acarology IX: Volume 1, proceedings (pp. 101–104). Columbus: Ohio Biological
Survey.
Wohltmann, A., & Wendt, F.-E. (1996). Observations on the biology of two hygrobiotic trombidi-
oid mites (Acari: Prostigmata: Parasitengonae), with special regard to host recognition and
parasitism tactics. Acarologia, 37, 31–44.
Wrensch, D. L., & Bruce, W. A. (1991). Sex ratio, fitness and capacity for population increase in
Pyemotes tritici (L.-F. & M.) (Pyemotidae). In R. Schuster & P. W. Murphy (Eds.), The Acari:
Reproduction, development and life-history strategies (pp. 209–221). New York: Chapman & Hall.
Yankovskaya, A. I., & Fernando, K. G. (1982). On parasitism of hydracarines larvae in the wall of
the stomach of fishes. Parasitologiya, 16, 244–246.
Yoder, J. A. (1996). The Madagascar hissing-cockroach mite (Gromphadorholaelaps schaeferi):
First observation of its larva and ptyalophagy in Acari. International Journal of Acarology, 22,
141–148.
Yourth, C. P., Forbes, M. R., & Smith, B. P. (2002). Immune expression in a damselfly is related to
time of season, not to fluctuating asymmetry or host size. Ecological Entomology, 27, 123–128.
Youson, J. H., & Beamish, R. J. (1991). Comparsion of the internal morphology of adults of a
population of lampreys that contains a nonparastic life-history type, Lampetra richardsoni, and
a potentially parasitic form, L. richardsoni var. marifuga. Canadian Journal of Zoology, 69,
628–637.
Zeh, D. W., & Zeh, J. A. (1992). On the function of harlequin beetle-riding in the pseudoscorpion,
Cordylochernes scorpioides (Pseudoscorpionida: Chernetidae). Journal of Arachnology, 20,
47–51.
422 9 Animals as Habitats

Zhang, Z.-Q. (1992). The adaptive significance of superparasitism in a protelean parasite,


Allothrombium pulvinum (Acari: Trombidiidae). Oikos, 65, 167–168.
Zhang, Z.-Q. (1998). Biology and ecology of trombidiid mites (Acari: Trombidioidea).
Experimental and Applied Acarology, 22, 139–155.
Zhang, Z.-Q., & Sanderson, J. P. (1993). Association of Ereynetes tritonymphs (Acari: Ereynetidae)
with the Fungus Gnat, Bradysia impatiens (Diptera: Sciaridae). International Journal of
Acarology, 19, 179–183.
Chapter 10
Mites That Cause and Transmit Disease

Disease implies a dichotomy: that one’s life can be at ease, but that it can also be
distant from ease. If our minds or bodies are not functioning properly, then we may
have a disease. For example, if one’s skin feels unusually hot, you may have a fever.
We’ve all experienced fevers because they usually occur as a general response by
one’s immune system when it goes after an invasive microbe, but fever also can
result from physiological problems and environmental stress. So, fever is also a
symptom, an indication that something is wrong, but not an especially good indi-
cator of exactly what is wrong.
This is a general truth: symptoms are only a guide to what may be wrong and can
be misleading. For example, if you often feel as if insects are crawling on your skin,
then you may be experiencing a sensation called formication, from a Latin word
for an ant, formica. Like fever, formication may have any number of causes from
adverse reactions to drugs (legal and otherwise) to environmental irritants
(e.g. fibreglass) to manifestations of serious illness such as skin cancer or diabetes.
Strangely, if an insect is actually crawling on your skin, then you aren’t experiencing
formication – you just have a bug to brush off. But how about something so small
that you can’t actually see it – what if a tiny mite is crawling on your skin?
Feeling a mite crawling on your skin is unlikely – if our skin reacted to every
dust mote-sized particle, we’d all always be itchy. However, it is possible to have an
unreasonable fear that tiny bugs are crawling or burrowing in your skin, and this is
called acarophobia. Literally, that means ‘fear of mites’, but even if you are afraid
that tiny insects are crawling on you (and they are not), the same name is applied.
Acarophobia and formication often go together; but in either case, it is not crawling
mites that are causing the problem: the itch has other causes. Real mites, however,
can and do cause disease and transmit pathogens that cause disease. In this chapter
we will briefly review how mites (including the large, blood-sucking mites we call
ticks) themselves can cause disease, discuss some examples of when they transmit
other organisms that cause disease, and then return to the strange phenomenon of
blaming mites for problems of which they are innocent.

D.E. Walter and H.C. Proctor, Mites: Ecology, Evolution & Behaviour: Life at a Microscale, 423
DOI 10.1007/978-94-007-7164-2_10, © Springer Science+Business Media Dordrecht 2013
424 10 Mites That Cause and Transmit Disease

Fig. 10.1 Dermatophagoides


farinae – one of several mites
that can be found living in
house dust (SEM by DE
Walter)

Critical Concepts and Terminology

For a mite to be involved in the causation or transmission of a disease, certain factors


must first be in place. These factors are not unique to mites, but are needed for any
arthropod-caused or arthropod-borne disease. First of all, there must be a source
of the disease. If a mite is the cause of disease, then populations of these mites must
be present (Fig. 10.1). For example, at least a 100 million people worldwide suf-
fer from reactions to allergens produced by house dust mites (see Colloff 2009 for
a thorough review), from runny noses to atopic asthma. In this case the dust mites
themselves (and there are a diversity of species involved – Colloff 2008) cause the
disease, because we can become sensitized to the proteins and enzymes in their skin
and feces. If the mites lacked these allergens, then no disease would result. If the
mites are vectors, i.e. carriers of disease-causing microbes (e.g. the rickettsiae that
cause Rickettsial Pox), then populations of mites infected with a disease-causing
microorganism must be present and capable of biting us. Such populations are
called reservoirs. When the reservoir of a disease is an animal and the disease
spills-over from the animals into a human population, then the disease is called a
zoonosis. For this definition to make sense, one has to pretend that people are not
animals (we are) and understand that mites are animals, just as insects, birds and
mammals are, because sometimes a mite is a reservoir.
The second critical factor that must be present for disease to occur is a population
of susceptible hosts – those whose immune systems are not able to ward off the mite
or its microbes. Naïve hosts, those that have never encountered a disease-causing
organism before are, likely to be especially susceptible to a newly encountered
pathogen, i.e. a biological agent that causes disease. Finally, there must be a means
of transmission of the disease between the reservoir and susceptible host. Typically,
a disease-carrying mite or tick transmits a pathogen by hunting down and biting a
susceptible host. Dust mites, however, cause disease when the remnants of their
cuticle and fecal pellets become airborne and are inhaled into the lungs. It is
Critical Concepts and Terminology 425

Fig. 10.2 Zoonotic disease requires three components: a source of disease, a means of spreading
disease and a susceptible human population. (a) Dermatitis caused by the House Mouse Mite
occurs when the mites infest mice living in or about human habitations. Mite populations increase,
mites wander from nests onto people and bite them. (b) More serious disease can result when the
mice are reservoirs of the pathogen that causes Rickettsial Pox. Infected mites both maintain
and spread the infection in the mouse population and allow the disease to spillover into human
populations. In both cases, solving the problem involves first identifying the source of the disease
and then eliminating the rodents and their mites (which cannot reproduce on people) (Image by
DE Walter)

important to remember that some proportion of the population of concern will show
a gradation in resistance to the pathogen or allergen and will not exhibit symptoms
of disease. For example, reservoirs of a disease are likely to be more or less tolerant
of a pathogen (Fig. 10.2).
The relationship between mites and disease is actually much more complicated
than the above, but these three components (reservoir, means of transmission,
susceptible population) are required. As an example, spirochete bacteria in the genus
426 10 Mites That Cause and Transmit Disease

Borrelia cause Lyme Disease and the primary reservoir of the disease is thought to
be deer mice (Peromyscus spp.) and other small mammals such as chipmunks and
shrews. The primary vectors of Lyme disease are ticks in the genus Ixodes. These
are hard ticks (Ixodidae) with a 3-host life cycle. Larvae feed primarily on deer mice
and some become infected with the Lyme Disease spirochetes. After dropping off
the mice, the tick larvae moult into nymphs that then acquire a new host, but nymphs
are more catholic in their tastes than larvae and will feed on a broader range of
mammals. When an infected nymph feeds on a deer mouse, chipmunk or shrew, they
help maintain the reservoir population of the spirochete. If they feed on a human,
though, the spirochete will be trapped in a host that does not act as a reservoir
(dead-end host). The same goes for an adult female tick – they prefer deer and
when they bite deer they may transmit the Lyme spirochetes, but as when they bite
people, the spirochete usually comes to a dead-end. Neither deer nor people appear
to be good amplifying hosts, i.e. hosts in which the pathogen can multiply to levels
where it is likely to be acquired by a feeding parasite. Humans are the reservoirs for
some mosquito-borne diseases (e.g. malaria, dengue), but not usually for mite-borne
diseases. The importance of deer in the epidemiology of Lyme Disease appears to
be mostly due to their maintenance of large populations of female ticks. Deer are
amplifying hosts for ticks, but not for the Lyme disease spirochete. Other recent
ecological changes such as the invasion of northeastern North America by the coyote
(Canis latrans) and its intraguild predation on smaller carnivores (e.g. foxes) may
be contributing to the expansion of the small rodent reservoir populations (Levi
et al. 2012). Migratory birds may also be important in the spread of the disease
(Radolf et al. 2012).
Other terms that are useful to understand include emerging disease. Diseases
can emerge when a reservoir and a susceptible population first come into contact,
but an apparently emerging disease also may be an older one that was previously
uncommon or undetectable. For example, Lyme Disease has probably been in North
America for a long time, but did not become a noticeable problem until the 1970s.
A primary cause of this emergence was that White-tailed Deer populations boomed
during the latter part of the Twentieth Century as a result of changes in human
behaviour. In earlier times, deer were hunted extensively and their preferred habitat
of early-succession forest had mostly been turned into farmland. In the latter part of
that Century, hunting had declined, much farmland had turned into young forest,
large numbers of deer wandered the land, and many happy female ticks fed on deer
and laid large masses of eggs. The larvae from these eggs helped maintain and
probably also to spread the Lyme spirochete in the small rodent reservoirs. So, in a
sense, Lyme disease is a resurging disease – one that was not very common when
deer and ticks were rare but is now increasingly common (approximately 30,000
diagnosed cases per year in the USA, and probably many more thousands of
undiagnosed ones). If you are interested in learning more about the changes in the
epidemiology of mite-borne and other arthropod-borne diseases see Garrett (1994);
Gratz (1999); Mullen and Durden (2002); Marquardt (2005), and more recent
reviews (e.g. Kilpatrick and Randolph 2012) (Fig. 10.3).
Mite-Caused Diseases 427

Fig. 10.3 The spirochetes that cause Lyme Disease in humans are parasites of small mammals
(deer mice, chipmunks, shrews). Three-host ticks in the genus Ixodes become infected as larvae
feeding on small mammals. Infected nymph and adult ticks pass-on the spirochetes to a new host
when they feed. Factors that increase the number of reproductive ticks (e.g. large deer populations) or
larval hosts (e.g. reduced predation pressure on small mammals) increase the incidence of disease
in human populations. Lyme Disease is spreading across North America in infected ticks (Image
by DE Walter)

Mite-Caused Diseases

When a mite is the actual causative agent of disease, and not just a vector of a microbe
between a reservoir and a susceptible host population, then the mite itself is the
actual pathogen. Of course, things are rarely so simple as our definitions imply: a
mite may both cause disease and vector disease, although usually the latter is more
severe. For example, when a tick bites you it makes a cut in your skin and injects
saliva that often causes an itchy reaction, a simple dermatitis. If you don’t clean the
wound properly or you scratch it with dirty fingers, then more serious disease may
result: a large ulcer or gangrene. If the tick is infected with disease-causing microbes,
however, no matter how well you take care of the bite, you may get a nasty disease.
For the time being, though, lets pretend things are simpler and look at which mites
bite us and why.
428 10 Mites That Cause and Transmit Disease

The Human Itch Mite: A Life in the Skin

Scabies (literally from the Latin ‘to scratch’) is a disease caused by the Human Itch
Mite Sarcoptes scabiei (Sarcoptidae). Chap. 9 and Walter and Shaw (2005) provide
a more detailed overview of the life cycle; Mounsey et al. (2013) provide a recent
review of prevalence, immunology, and diagnosis; Currier et al. (2012) have an
interesting historical perspective and provide an overview of the epidemiology; and
Chap. 9 has information of the spread of the Human Itch Mite to other animals. Here
we will limit ourselves to an overview of the problem in humans and an attempt to
deduce why scabies can be so often missed or misinterpreted. Because, in spite of
the cause of scabies having been known for more than 300 years, and conclusively
demonstrated for more than 150 years (Currier et al. 2012), scabies continues to be
under- and misdiagnosed (Leung and Miller 2011; Mounsey et al. 2013).
Adult female Human Itch Mites create burrows in the upper layers of the skin,
between the layer of mostly dead cells called the stratum corneum and the stratum
granulosum. The burrows are usually a few millimetres to a centimetre or so in
length and within these burrows the mites feed, defecate, and lay 2–3 eggs per day
for about 6 weeks. Our immune systems respond to the antigens in the mite products,
slowly at first (3–6 weeks in a primary infection) but almost immediately upon a
secondary infection. Perhaps successful immune defense is a reason why popula-
tions of Human Itch Mite are usually rather low – often under a dozen burrowing
females in typical scabies. Another reason may be that scabies is one of the few mite
infestations where scratching may be beneficial. Mite burrows are rather shallow
(traditional diagnosis is accomplished by teasing a mite out of a burrow with a
needle or scalpel) and scratching may actually remove female mites as well as
stimulating and directing the immune response to where it is most needed. Being
removed from the burrow is probably fatal for the adult female mites since they
become quite bloated egg-laying machines in their burrows and are not very mobile
when removed. Another reason that populations on an individual tend to be low may
be that the Human Itch Mite is affected by body temperature.
As a general rule, the high body temperatures of endotherms are difficult for
ectoparasitic arthropods to tolerate (e.g. body lice spend most of their time in
clothing and crawl onto the skin only to feed). In a typical infestation, most scabies
burrows are restricted to areas of creased skin on the extremities, especially the
arms from the elbows to the hands or the knees to the feet, where the skin is cooler
than in core regions and where clothing is less often present (at least in the warmer
seasons and warmer parts of the world). The rest of the body tends to be free of
mite burrows except in usual cases (e.g. infants and those with advanced crustose
scabies). Whatever the cause, though, the fact that normal scabies tend to involve
few mite burrows means that it will be easy to miss the causative agents of the itch
and difficult to stop the spread of an infestation.
Scabies is transmitted by close personal contact. Shaking hands may be all that
is needed for a newly emerged and mated female mite (the most likely infective
agent) to move from one host to another; but obviously, the greater the amount of
Mite-Caused Diseases 429

contact, the greater the opportunity for mites to change hosts. Thus, scabies may be
expected to be more common in demographic groups that tend to maximize the area
and frequency of skin-to-skin contact, and also in groups that tend to be handled
more frequently, such as children. This requirement for contact is relaxed when
immune-compromised individuals are infested and develop crustose scabies where
mite populations may number in the tens of thousands. In such situations, often
hospitals or old age homes, even fastidious health care givers may become infected
and act as ‘vectors’ of the disease.
In contrast to the Human Itch Mite, other mite pests of people do not burrow in
our skin: all are restricted to feeding from the surface, although in the case of Human
Follicle and Sebaceous Gland Mites (Demodex spp. – see Chap. 9), they do enter
into openings of the skin. This may also be where early developmental stages of the
Human Itch Mites reside – in our skin pores and hair follicles – although surpris-
ingly, little is known about the development of our most important skin mite. So, let
us now leave behind our true parasitic mites and consider those that make the poor
choice of biting us for no real gain.

Demented Dermanyssoidea: Biting Mites of Birds,


Rodents, and Whatever Else Is Nearby

Years of experience in both hemispheres has taught us that when people complain
about being bitten by mites, and they are not mistaken, the mites usually belong to
one large superfamily of Mesostigmata: the Dermanyssoidea. This is a vast array of
about 17 families (Beaulieu et al. 2011) filled with species parasitic on all kinds
of mammals, birds, snakes, lizards, insects and other arthropods. Chap. 9 reviews
many of these interactions, so here we will concentrate on those mites that directly
impact our lives. Most of these mite-person encounters are accidental and usually
result from two factors: (a) a bird or mammal is nesting near people and (2) the host
animal is not available. Most of the encounters involve only a few species from the
families Macronyssidae, Dermanyssidae, and Laelapidae. These common pests
live in the nests of their hosts for most of their lives and are able to develop and
reproduce only on their normal hosts. So far, so good, but when their normal hosts die
or leave the nest, or if the nest is swarming with too many mites, these little prickers
(Greek nysso = prick) will often wander into human habitations and bite us.
The Starling Mite (also known as the Northern or Nordic Fowl Mite) Ornithonyssus
sylviarum and its close relative the Tropical Fowl Mite Ornithonyssus bursa (both
Macronyssidae and both ‘bird prickers’) have been distributed around the world by
human commerce and favour birds that nest in tree-hollows, bird boxes, and other
protected spots such as swallow nests under eaves. Domestic poultry are attacked
by both species, as are common urban birds such as house sparrows, starlings, and
pigeons. Populations often reach extraordinary levels (tens of thousands per nest),
and some spill-over from nests into human habitations occurs as birds rear their
430 10 Mites That Cause and Transmit Disease

young. The real problem for people, however, occurs after a clutch has fledged or
died. Then thousands of starving mites may wander into homes or offices through
windows, attics or eaves and begin biting people in earnest. The bites are sharp,
painful, and with a little supplementary scratching, develop into large red wheals.
On their natural bird hosts, generation times for Fowl Mites are about 10 days to
2 weeks, but these mites cannot develop on human blood and outbreaks tend to be
self-limiting. Unless you raise poultry or pigeons, preventing the problem from
recurring is as simple as stopping birds from nesting on buildings that people
inhabit. If you like having swallows nesting in your yard, for example, seal off the
eaves of the building with wire mesh and build the swallows a nesting overhang well
away from your home.
The Tropical Fowl Mite spends most of its time off the host and has the broader
geographical and host ranges. The Starling Mite is found in the more temperate
areas of both hemispheres and has a closer relationship with its hosts – feeding
stages are often found resting on the host and eggs are laid there (Walter and Shaw
2005). A closely related mite, Ornithonyssus bacoti, the Tropical Rat Mite, infests
rodents and other small mammals that nest near the ground. Tropical Rat Mite is a
problem in warehouses, laboratory rodent colonies, small-mammal houses in zoos,
and for people who keep rodents as pets (Beck and Fölster-Holst 2009). As well as
spreading diseases of rodents and causing transient dermatitis in people, Tropical Rat
Mite is a suspected or potential vector of disease to people (Walter and Shaw 2005).
Problem species of Ornithonyssus are relatively large mites – about a millimetre
long even before becoming engorged with blood. Yet, the mites are often missed
(they are usually active during the night and hide during the day) and their bites
often confused with scabies, lice infestations, or the bites of insects.
The type family of the Dermanyssoidea, the Dermanyssidae (‘skin prickers’),
also has its share of bird and rodent parasites that sometimes bite people. The House
Mouse Mite Liponyssoides sanguineus not only bites people on occasion but is a
known vector of rickettsial pox (a disease caused by the intracellular bacterium
Rickettsia akari). Rickettsial pox was an emergent disease in New York City in
1946, but is now known worldwide (Renvoisé et al. 2012). The House Mouse (Mus
musculus) is the reservoir of the disease, the House Mouse Mite the vector, and we
are the susceptible population. The type genus of the Dermanyssidae, Dermanyssus,
contains many parasites of wild birds and one important pest of poultry and people,
the Pigeon, Chicken or Poultry Red Mite, Dermanyssus gallinae. Like most of
the dermanyssoid mites we have discussed, Poultry Red Mite spends most of its
time hiding off the host and usually feeds after dark. It is especially bothersome to
poultry workers, but can invade homes or other human habitations when pigeons are
allowed to nest nearby (Walter and Shaw 2005).
The final entry in our litany of troublesome Dermanyssoidea is the family
Laelapidae. About half the species in the family are free-living predators, but most
of the rest are associated with mammals or birds and are often blood-sucking para-
sites that will bite people when the opportunity arises. For example, Spiny Rat Mite
Laelaps echidninus is a common parasite of rats and other small rodents and a fre-
quently reported pest. Even laelapid mites that are not blood-feeders are sometimes
Mite-Caused Diseases 431

reported as pests. For example, Androlaelaps casalis is common in stored products


and in rodent and bird nests, has often been considered a facultative parasite, and is
sometimes reported to bite people. However, Lesna et al. (2012) have shown that
A. casalis is an effective predator of Poultry Red Mite in laboratory mesoscosms.
That doesn’t mean that they don’t sometimes bite people, but this behaviour is likely
more of an aberration in a confused predator than a deliberate attempt to obtain
food. This seems to be the general case in the Dermanyssoidea: people are not their
natural hosts, they cannot develop or reproduce without their natural hosts or prey,
and the solution to any problem with them is to remove the source of the infestation,
whether it be bird nests, rodents, or stored products infested with mites.

Perverse Prostigmata: Whirligigs, Straw Itch,


and Walking Dandruff

The Prostigmata contains some important parasites of birds and mammals including
us (see Chap. 9), but many of those that bite us do so by accident or desperation.
One example is the round red mites that run furiously over vegetation, sidewalks,
and roofs, the whirligig mites (or in Australia, ‘footballers’ in reference to the
frenetic pace of Australian Rules Football) (Anystidae). Anystids are predators of
mites and small insects on exposed and elevated surfaces. When one works in the
garden or walks through vegetation or lays down for a nap in a sunny green spot, a
whirligig mite may run onto you. Since the mite is unlikely to find much prey on a
person’s body, they may eventually get hungry or thirsty enough to stick their
hook-like mouthparts into the skin. When whirligig mites are on a plant, this is
probably a useful behaviour for rehydrating, and even on a person a whirligig mite
can take up a significant amount of fluids (Walter et al. 2009), but as the bite is
painful and the mite relatively large (~ 1 mm in diameter) and brightly coloured, the
result is usually a quick death for the mite.
A similar but more insidious and sometimes serious problem can occur with
mites in the genus Pyemotes (Pyemotidae) that have acquired the common name
straw itch mites. Pyemotes are parasites, or perhaps more correctly parasitoids (the
mite kills its host), of insects: a female mite attaches to an insect, injects a venom
that paralyzes the host, and proceeds to feed on host fluids as her young develop
within her growing body (see Chap. 9). Pyemotes species parasitize a wide variety
of insects, but most species seem to cause us no harm. However, members of the
Pyemotes ventricosus group that wander onto us will bite and a very painful itchy
red wheal with a central blister results from each bite and can persist for several
weeks (especially if assisted by scratching).
Pyemotes mites are very small ( < 0.25 mm long), light tan in colour, and unlikely
to be noticed by their victims. The name ‘straw itch’ comes from the mites biting
people harvesting hay or alfalfa or those working in barns or in warehouses with
stored grains or hay infested with beetles. People who work with laboratory cultures
of insects also are at risk, as are those who encounter any large aggregation of
432 10 Mites That Cause and Transmit Disease

insect hosts that are infested (e.g. stored grains, fallen fruit, or crops infested with
beetles or moths). For example, Del Giudice et al. (2008) document 42 cases of
Pyemotes-induced dermatitis in France associated with furniture infested by the
Furniture Beetle Anobium punctatum. The larvae of this small beetle bore in dry
sapwood, have a very long life cycle, and are attacked by Pyemotes ventricosus.
When homes with infested timber or furniture have large populations of beetles,
large numbers of mites can develop and those sitting in infested chairs or even
walking through an infested room may pick them up. There is no obvious benefit to
the mites from biting people – they must have insects to develop properly – but that
is small consolation to someone covered with numerous large and very itchy bites.
Rarely, a mite has been found embedded in the central blister, presumably from
being rubbed into the wound by a victim, but usually the only way to diagnose the
problem is to find wandering mites or mite-infested insects in a home. Otherwise,
victims, their doctors, and their pest control operators may be unable to explain why
the wheals appear, and misdiagnose the problem, e.g. as environmental irritants
(e.g. fibreglass, chemicals, plant oils).
One particularly interesting outbreak was traced to Pyemotes herfsi, a mite that
parasitizes midges that make galls on oaks (Broce et al. 2006). As is usually the case
with Pyemotes, problems only occur when large populations of the mites have built
up and their normal hosts are rare. This mite has been called the Oak Leaf Itch
Mite, and although it will feed on broader array of insects, most problems have been
associated with oak trees with insect galls. People raking leaves, gardening or other-
wise being in proximity to an oak with galls and mites are especially susceptible.
A veritable rain of mites may occur under such trees and people below or downwind
of trees may develop dermatitis. Pyemotes usually disperse by walking or by hitching
rides on the adult hosts, but P. herfsi may be able to colonise new oaks on the wind
and if they land on a person instead, it is unfortunate for both mite and person.
In contrast to the other prostigmatans that we have discussed so far, species of
Cheyletiella (Cheyletidae) are parasites of vertebrates. They live on the epidermis of
their hosts where they spin silken, cocoon-like webs in which they rest and lay their
eggs. When populations are high or a host is especially sensitive, these mites cause
a disease called mange. ‘Mange’ refers to a collection of symptoms including
itching, hair loss, skin flakes, scabbing, and general deterioration of the epidermis,
rather than to a specific disease, and it can be caused by a the interaction of the host
immune system and variety of parasitic arthropods. Usually, mange is further
defined by referring to the organism that causes it, for example Sarcoptic Mange in
animals is caused by Sarcoptes scabiei. In this case, though, the disease is called
cheyletiellosis. Cheyletiella yasguri is the species found on dogs; C. blakei infests
cats; and C. parasitivorax is found on rabbits. The mites themselves have also
earned a common name: walking dandruff. Seemingly, this is because they are of
a similar colour to the skin flakes produced during an infestation and the fact that the
adult mites move around readily. That is unfortunate for those who have pet dogs,
cats, or rabbits that are infested, especially if the owners like to let the pet sit in
their lap. Mites will wander on to people and take a taste, resulting in a transient
Mite-Caused Diseases 433

dermatitis characterized by red wheals with a central blister (where the needle-like
mouthparts have been inserted). Since walking dandruff mites are large enough to see,
and their silken webs at the base of hairs on the pet are often noticed, cheyletiellosis
is often correctly diagnosed and treated. Only the infested animals need to be treated:
Cheyletiella species cannot survive on pet owners.
A final entry in our catalogue of perverse prostigmatans will also serve as a
transition to mites that transmit disease: the chigger. Also known as harvest
mites, red bugs, berry bugs, colorados, coloradillas, isangos, bêtes rouge,
mocuims, and many other names, chiggers make life intermittently miserable for
people in many parts of the world. Chiggers are the larvae of red velvet mites in the
families Trombiculidae and Leeuwenhoekiidae (usually grouped in the superfamily
Trombiculoidea). Nymphal and adult mites are predators of small arthropods and
their eggs, but the larval stage is ectoparasitic on a vertebrate (see Chap. 4 for the
complete life cycle). Although often bright red or orange in colour (the type genus
name, Trombicula, comes from the Greek thrombos and means ‘little blood clot’),
chiggers are only about a quarter of a millimetre long, and almost impossible to see
or feel until their mouthparts have caused an itchy red wheal to develop – usually
where a belt, sock cuff, or elastic waistband forms a snug furrow on the skin. Scrub
itch mites are not well adapted to feeding on people, and usually the first scratch
will remove the larva (so, we are literally a dead-end host). When attached to a more
appropriate host, the larva begins to swell to many times its original size. Engorged
larvae are easy to see in the ears of small mammals, around the vents of birds, or as
red blebs on the skin of a reptile (and easily confused with other red mite parasites
of lizards in the Pterygosomatidae, see Chap. 9).
Being covered in chigger bites is formally known as trombidiosis, but in
Australia it is described more expressively as scrub itch. For example, scrub itch
mites in the family Leeuwenhoekiidae are pests throughout much of coastal eastern
Australia. Most appear to be generalist parasites of lizards, birds and mammals.
The Sydney Grass Itch Mite, Odontacarus australiensis, occurs as far north as the
Atherton Tablelands in Queensland and causes dermatitis in man, cats, dogs and
horses. Numerous other genera of Trombiculoidea cause scrub itch in various parts
of the world, e.g. species of Eutrombicula in North America and the Harvest
Mite Neotrombicula autumnalis in Europe. The chigger feeds externally (they do
not burrow in the skin), but the mouthparts penetrate the skin and digestive
salivary secretions are injected to form a tube-like stylostome that carries lymph
fluids and cell contents to the mite’s mouth. This itchy reaction between host
immune system and chigger enzymes in the skin appears to be why many people
believe that chiggers are burrowing in their skin. One often hears that the chiggers
that normally feed on reptiles or birds cause the most itchy reactions and that
mammal chiggers are more benign (Nutting 1985). Possibly this story has some
truth to it, but it seems to rely more on anecdote than supporting data (rarely
are chiggers biting people identified). However, bites from mammal chiggers are
potentially much more dangerous, because these mites may vector the rickettsial
disease called Scrub Typhus.
434 10 Mites That Cause and Transmit Disease

Mite- and Tick-Borne Diseases

The most important mite-borne diseases are carried by ticks. Most ticks feed on a
different host individual during each instar (Ixodidae) or repeatedly feed on hosts in
a nest (Argasidae, some Ixodidae). As a result, ticks have repeated opportunities to
acquire pathogens. Ticks also tend to feed for long periods of time, increasing the
chance that they will acquire pathogens from a host’s blood system. While feeding,
ticks inject significant salivary material into the wound and often excrete fluids onto
the host’s skin, again increasing the probability of passing on some acquired infec-
tious agent. All of these factors contribute to ticks being major vectors of disease.
But even the accidental and generally fatal (to the mite) mistake of biting a person
can result in the transmission of disease from mite to human. We have discussed this
above with dermanyssoid mites, but below we cover a much more serious problem.

Trombiculoidea (Chiggers): Scrub Typhus

About one seventh of the World’s population lives in areas where Scrub Typhus is
endemic (Strickman 2001), about one million of these people are infected with this
disease annually, and any one of the other six billion people in the world who might
visit those areas are susceptible to infection (Walker 2003). Scrub Typhus is caused by
a rickettsial bacterium called Orientia tsutsugamushi – the species name comes from
the Japanese words for both ‘River Fever’ and ‘chigger’, suggesting that the associa-
tion between the bite and fever were long understood. Species in a number of genera
of Trombiculidae that normally feed on small mammals are the vectors, especially
species of Leptotrombidium (e.g. L. akamushi, L. deliense) in tropical Australia,
Asia, Japan, and Pacific islands. A fancier name for Scrub Typhus is Chigger-borne
Rickettsiosis (the bacterium was originally placed in the genus Rickettsia) and it has
long been a major disease in rural areas of Japan and on mainland and island South
East Asia. Grassy areas with large populations of rodents tend to be foci of infection.
People contract scrub typhus when they enter a habitat where chiggers carry the
Scrub Typhus bacterium and have the misfortune to be bitten by an infected mite.
The distributions of the mites tend to be patchy: they tend to be closely associated
with the burrows of their hosts. People are more likely to be bitten if they maximize
the potential for contact by spending an extended period of time in an area, e.g. by
working along stream or river margins with rodent populations. Casual hikers are less
likely to pick up a mite if they keep moving, but sacking out along a grassy stream
bank for a nap is an open invitation to the mites. Soldiers on manoeuvres or engaged
in combat are commonly infected. As with many rickettsial diseases, the symptoms
start 1–2 weeks after infection and include fever with headache, profuse sweating,
lethargy, muscle pain, nausea and other non-specific indicators similar to dozens of
other diseases (Walker 2003). After a week or so a rash often develops, but the best
indicator is a primary eschar (ulcerous scab) 4–8 mm in diameter formed at site
of attachment (commonly the genitals, buttocks, lower abdomen, arm, or armpit).
Mite- and Tick-Borne Diseases 435

In many cases, only a single eschar may be present, which complicates diagnosis.
Variable mortality is associated with different strains of Orientia tsutsugamushi.
Lethal strains have mortality rates as high as 60 % (without treatment). Death can
be due to heart failure, circulatory collapse, pneumonia, bleeding or general organ
failure. Mortality is highest in older patients, the chronically ill, and those delaying
treatment.
Rodents, pathogen, mite vector and susceptible human population are the
ingredients needed for the typical zoonosis, but Scrub Typhus has one major devia-
tion from this formula: the mites themselves appear to be the primary reservoir of
disease (Frances 2005; Kuo et al. 2011; Phasomkusolsil et al. 2012). Scrub Typhus
is a disease of chiggers, transmitted from mother to daughter mite and from mite to
rat or human but, at least in Leptotrombidium species, not commonly from rat to
mite (Frances 2005; Phasomkusolsil et al. 2012). Mite larvae can be infected by
feeding on infected rodents, but the infection is usually gone by the time the adult
moult has been reached (Frances 2005). Persistence of a newly acquired rickettsial
infection appears to be rare (presumably because of mite immune defences).
Rodents and other small mammals are critical for maintaining populations of mites,
and may exhibit high levels of exposure to the pathogen (Kuo et al. 2011), but
rodents are not the reservoir of Orientia tsutsugamushi.
When a disease-causing organism can be passed from an arthropod to its eggs it
is called trans-ovarian transmission. If the pathogen can maintain infection of the
arthropod through all of the moults needed to reach adulthood (trans-stadial
transmission), then the new adult is infected without need to feed on an infected host:
the mite has become both vector and reservoir of the disease. Chiggers borne-infected
with Scrub Typhus rickettsiae are able to infect mammalian hosts and to carry the
disease through the moults resulting in the predatory nymphs and adults, and into
the next generation of eggs. This is not without costs to the mites. Phasomkusolsil
et al. (2012) studied two species of Leptotrombidium that are important reservoir/
vectors of Scrub Typhus in Thailand using isofemale lines (colonies reared from a
single female mite) collected from the ears of rats. In general, lineages of mites
infected with Orientia tsutsugamushi had longer generation times, and infected
females laid 20–25 % fewer eggs than females from uninfected lineages. Such
experiments are difficult to carry out and dangerous to the experimenters (minute
mites are can’t easily be contained) and experimental variability can make results
difficult to interpret. However, in this study, infected larvae of one species finished
feeding and dropped off the host 4 days earlier than uninfected larvae. Since attached
larval mortality may be high (rats do scratch their ears), this may be an example of
a pathogen influencing host behaviour/physiology to their mutual benefit.

Ixodoidea (Ticks)

When a chigger bites one of us, it is a fatal error for the chigger. The same is less
often true for ticks. All 900 or so species of ticks are obligate blood-feeders and
usually larvae, nymphs and adult females deliberately seek out a vertebrate animal,
436 10 Mites That Cause and Transmit Disease

clamber onto its body, slice through its skin, inject secretions that interfere with
the host immune response and promote blood flow (Radolf et al. 2012), slash the
capillaries to form a puddle of blood and drink the blood until they are satiated or
are removed by host grooming or killed by the host immune system. Ticks are able
to sense and orient to body heat, carbon dioxide, host odours and secretions, and
vibrations. Although they often sit on vegetation with their front legs outstretched
(questing behaviour) waiting for an animal to walk by, many are not averse to
scurrying towards a stationary meal. Therefore, unlike chiggers where an encounter
tends to be subject to many stochastic variables (e.g. patchy distribution, limited
dispersal from natal area, limited use of vegetation), a tick is more likely to acquire
a person or their pet when the opportunity arises. The best defence against such an
encounter is to use effective repellents (DEET-based repellents are the only ones
proven effective at this time) and to carefully search for ticks on one’s clothing, person
and pets after travelling in tick country. Hard ticks tend to be a bit choosy about
where they settle to start feeding and the feeding process is rather long and intricate,
so vigilance can be an effective method of avoiding tick-borne disease. For example,
the nymphs of Ixodes scapularis that are the primary vectors of Lyme Disease in the
eastern United States do not usually start transmitting the Lyme spirochete until
they have been attached to a host for more than a day (Radolf et al. 2012).
Not all ticks in an area will carry disease-causing pathogens, so not every tick
bite will result in disease. The incidence of disease transmission will depend on
factors such as the prevalence of the pathogen in the reservoir population, how the
pathogen is maintained in the ticks (especially the potential for trans-stadial and
trans-ovarian transmission), and how the tick itself reacts to the pathogen. For
example, Rickettsia rickettsii, the bacterium that causes Rocky Mountain Spotted
Fever in people, also causes significant mortality in its vector the Rocky Mountain
Wood Tick Dermacentor andersoni. In a laboratory experiment using infected
guinea pigs (Niebylski et al. 1999), 94 % of larvae fed on the infected hosts died
during the moult to adults and 88 % of female nymphs fed on infected hosts
died before attaching as adults. In contrast, control mortality was less than 3 %.
The causative agents of Lyme Disease (Borrelia burgdorferi complex) are probably
not trans-ovarially transmitted in Ixodes scapularis, although a related species
Borrelia miyamotoi that may cause a relapsing fever-like disease is (Rollend et al.
2013). This means that the larvae hatching from the eggs of an I. scapularis female
carrying the Lyme Disease spirochete will be not be infected and must acquire
the disease from a rodent reservoir before the nymphs can transmit the pathogen.
Bites from the larvae of I. scapularis do not warrant concern about Lyme Disease
(but they may for other trans-ovarial pathogens).
Ticks are usually considered second in importance as the vectors of disease only
to mosquitoes. This is especially true of three-host ticks, as discussed above, but
the very nature of tick feeding probably contributes to their tendency to acquire
pathogens (see Mehlhorn 2008 for a through overview of tick feeding). Soft ticks
feed for relatively short periods of time and often retain excess fluids until near the
end or after a bout of feeding. However, they still ingest 3–5 times their body weight
in blood if allowed to feed to repletion. If they divest themselves of excess fluids
Diseases That Mites Do Not Cause 437

from the glands on coxae I while still on the host, then disease-causing microbes
may find their way into the host through the bite or other abrasions in the skin. Hard
ticks (which secrete excess fluids from the salivary glands into the wound during
feeding) can ingest more than 100 times their body weight in blood. Blood is stored
in the midgut and its diverticula for extended periods of time and then engulfed by
the digestive cells that line the gut: this can introduce intracellular parasites like
rickettsiae directly into living tick cells.
The number and diversity of emerging tick-borne diseases is horrifying, but this
surge is driven largely by human behaviours: increased world trade, alteration of
habitats and ecologies, and changes in the view of man’s place in nature. For example,
about 20 different species and varieties of Rickettsia are known to cause disease
around the world. Ticks vector all but two of these. Symptoms of Rickettsia infection
tend to be general, distinguishing among these pathogens is difficult, and a steady
flow of infected travellers return to their native countries (Wood and Artsob 2012).
The spread of Lyme Disease in North America is likely partially the result of com-
merce (e.g. stowaway mice and ticks) and travel with tick-infested pets. Trade and
travel are the primary modes by which disease-causing microbes reach new, naïve
populations (pathogen pollution), often causing explosive epidemics (Kilpatrick
and Randolph 2012).
In contrast to pathogen pollution, the resurgence or emergence of endemic
arthropod-vectored disease is primarily a function of environmental changes that
affect vectors and reservoir hosts, or social factors that bring people into contact
with vectors, factors as diverse as warfare or a desire to live closer to nature.
The importance of burgeoning deer populations in the rise of Lyme Disease was
discussed above, but at the same time and seemingly for the same reason, apparently
endemic Ehrlichiosis, Babesiosis, Deer Tick Virus and Powassan Virus are also on
the rise in North America (Kuehn 2013).

Diseases That Mites Do Not Cause

We’ll start this last section of Chap. 10 with a true story. A number of years ago, the
authors (DEW, HCP) took a long drive from Brisbane to Melbourne in eastern
Australia to attend a conference during a semester break. Halfway there we stopped
at a motel for the night. The motel was very charming and had a large Moreton Bay
Fig in the yard. Beneath the tree a crop of figs had fallen and these rotting fruit were
full of interesting insects to ponder during our post-prandial walkabout.
The next morning we both noticed several red wheals with central blisters
on our arms and on DEW’s back (Fig. 10.4). The bites of no-see-ums (Diptera:
Ceratopogonidae) can cause a similar wheal and these near invisible biting flies can
penetrate window screens and will crawl under clothing, so it seemed most likely
we had been bitten while pondering fig bugs or been victimized during the night.
The next day at the conference, however, more bites appeared on DEW (but not
HCP), one appearing on his arm in the middle of dinner! The wheals produced were
438 10 Mites That Cause and Transmit Disease

Fig. 10.4 Five mystery


wheals that appeared
on the senior author’s back
one morning in Australia.
Each large (2 cm diameter)
red wheal has a small blister
in the centre (see text
for details) (Photo by HC
Proctor)

large (about the diameter of an Australian dollar) and extraordinarily itchy, but
without any obvious cause.
When the authors returned home, they gave all their clothes a thorough washing
in case some tiny arthropod in the clothing was the cause of the wheals, but an
allergic reaction or a symptom of a disease were also possibilities, so DEW decided
to visit his General Practitioner before classes started. The good doctor was very
agitated (he knew DEW worked with mites) and kept a good metre away from his
patient’s blistered back, diagnosed scabies, and gave him a referral to a dermatologist.
If you’ve read the earlier parts of this chapter, you know that scabies usually
manifests on the hands and elbows and takes the form of burrows, not wheals. Still,
he was the doctor, atypical scabies does occur, and so off to the dermatologist went
DEW a week or so later with only fading red bumps left to show (it is difficult to
scratch one’s back, and without scratching most bites will recede within a week or so).
The dermatologist stayed a good 2 m away from his patient, never got up from
behind his desk, and diagnosed fleabites! Anyone who has pets with fleas knows
that the ankle to the knee is the most likely place to find fleabites and we certainly
had not taken any pets with us to Melbourne. That was when DEW decided he was
wasting his money with the medical profession and decided to solve the problem
himself.
A thorough examination of DEW’s cleaned clothing with a hand lens found a
few bits of tan cuticle and legs that looked like they came from very small mites.
DEW used some sticky tape to collect the bits, transferred them to a microscope
slide, and was able to convince himself that he had a probable identification (Straw
Itch Mites – Pyemotes sp.). Unlike his doctors, DEW had a good background in
Medical & Veterinary Entomology and an extensive knowledge of biting mites, so
he preferred his self-diagnosis to those of his GP and dermatologist. Without his
special knowledge, however, he would never have had a clue to the cause of his
affliction and if he had continued to scratch at his bites, they never would have gone
away. If he had continued to go to doctors with his complaint, he would have been
labelled a loony and ignored.
Diseases That Mites Do Not Cause 439

Mystery Bites

Mites are small, usually very small, a number of species will bite people (as reviewed
earlier in this chapter and see Table 10.1), and without an alternative culprit, it
would be easy to misdiagnose mites as the cause of a skin problem. Except in the
case of the Human Itch Mite, however, these infestations are transient – without
scratching, the wheals go away in a week or two and the mites themselves cannot
survive for long on human bodies. If people live close to the source of these biting
mites (e.g. pets, furniture, stored grains, bird or rodent nests, gardens), then re-infes-
tation may occur, symptoms may persist for longer periods and a person may become
convinced that the mites are living in their skin, but only the Human Itch Mite can
live in our skin. Proper diagnosis of a biting mite problem requires the capture and
identification of the causal agent, because many other agents and conditions may
cause reactions that resemble mite bites.
In addition to mites, numerous small insects bite people. The extraordinary
resurgence in Bed Bug (Cimicidae) populations over the last few decades has added
an all too common source of mysterious bites and rashes. Mosquitoes, black flies
and other biting flies; head lice; and even fleas are generally large enough to see
during the day, but something as simple as a window without a good screen can
result in mysterious bites appearing during the night. The no-see-ums are minute
biting flies (Ceratopogonidae) that live up to their names: most are no more than a
few millimetres long and very difficult to notice. When biting, the sensation is not

Table 10.1 Systematic synopsis of medically important groups of mites


Parasitiformes
Ixodida (Metastigmata)
Ticks (Ixodidae, Argasidae)
Mesostigmata
Bird mites, poultry mites, nasal mites, rat mites etc. (Dermanyssoidea, especially Dermanyssidae,
Macronyssidae, Laelapidae)
Acariformes
Prostigmata
Footballer and whirligig mites (Anystidae)
Mange and follicle mites, walking dandruff (Cheyletoidea)
Straw itch mites (Pyemotidae)
Scrub itch mites, chiggers (Trombiculoidea)
Suborder Oribatida
Moss or beetle mites that are intermediate hosts of tapeworms (many families, especially in the
Oripodoidea)
Cohort Astigmata
Stored product pests, grocer’s itch mites, baker’s itch, cheese mites, furniture mites, dust mites
(Acaroidea, Glycyphagoidea)
Scabies, mange, scaly leg, fur mites (Sarcoptoidea)
Feather mites (Pterolichoidea)
Feather mites (Analgoidea) including house dust mites (Pyroglyphidae)
440 10 Mites That Cause and Transmit Disease

particularly noticeable (the mouthparts are small), but the immune reaction to the
bite can be extensive and large, red itchy wheals may develop. Since these tiny biting
midges often emerge in great numbers, and unlike most biting flies, no-see-ums
may go under one’s clothing, the victim may become covered in itchy red bite marks
with no obvious cause.
Sometimes, the insects that bite us are not after our blood, but are actually
predators (e.g. Minute Pirate Bugs: Anthocoridae) or plant parasites that are capa-
ble of using their piercing-sucking mouthparts to puncture human skin and with-
draw fluids. Most plant-feeding insects are capable of using their mouthparts to
puncture human skin in defence or in the search for fluids. For example, thrips
(Thysanoptera) are small, slender insects that are commonly found in flowers or
feeding on garden plants. People working in their gardens or admiring a bouquet of
flowers in their homes may be bitten and develop small welts. When only a few
thrips bite, it is annoying but not especially so, because the skin reactions to the
thrips’ saliva tend to be small (~1 mm diameter) and transient. But when large
numbers of biting thrips migrate from crops and land on people or infest clothing
(some thrips are attracted to white or blue colours), then serious rashes may develop
(Childers et al. 2005).
Mystery bites caused by arthropods are common (Kushon et al. 1993).
Additionally, parts of insects (e.g. cockroach body parts, saliva) can cause allergic
reactions that resemble bites, urticating hairs of caterpillars and tarantulas cause
wheals and rashes, and any number of environmental irritants may cause symptoms
that resemble bites from mites. In addition, the simple act of scratching an area of
itchy skin will cause it to turn red and rise into a bump (the ‘itch-scratch cycle’) – a
person can manufacture his or her own ‘mite bites’. Persistent irritation can result in
the belief that mites or other ‘invisible bugs’ are infesting the skin (Illusions of
Parasitosis) when the actual problem is caused by something inorganic as in ‘Sick
Building Syndrome’ (Hinkle 2010). When people with this illusion have the actual
cause determined and explained to them the reaction is usually relief. But this is not
always the case.

Delusions of Mite-Bites

A delusion is an inflexible belief that resists all evidence to the contrary. When
someone is convinced that they are infested with mites or other ‘bugs’, worms,
microbes, or other pathogens that no one else can see, then they may be suffering
from a psychological syndrome called Delusory Parasitosis (DP). In addition to
feeling the infesting bugs, the sufferer often can see them, sometimes by the millions,
and may sometimes convince those around them that their homes and bodies are
also infested. But when samples are presented to objective observers (e.g. doctors,
scientists, pest control operators), no arthropod capable of causing the symptoms
can be found. The sufferer then characteristically reacts with anger, tries harder to
find specimens of their pest, and moves from specialist to specialist repeating their
Diseases That Mites Do Not Cause 441

claims. Because mites are nearly invisible, they often become the focus of these
delusions and anyone who works with mites and is accessible to the public is likely
to encounter people suffering from DP.
DP, also called Ekbom’s Syndrome and a variety of other terms using variants of
‘delusion’ and ‘parasitosis’, has recently been thoroughly reviewed by a medical
veterinary entomologist (Hinkle 2010, 2011) and this discussion will be a simple
summary of the symptoms and possible underlying causes of DP. We will adopt
Hinkle’s convention of referring to the person presenting symptoms and samples as
the ‘sufferer’.
The first and perhaps most important point for an acarologist to consider is their
role in a DP incident: they must be a truthful reporter of the acarological facts. If a
mite is present in the samples, and that mite is capable of producing the symptoms
described, then the sufferer may not have DP and the source and possible elimination
of the problem can be discussed. For example, the simplest solution to a bird mite
problem is to find the nests that are the source of the mites, remove them, and then
secure the area from future nesting. If the birds are considered pests (e.g. as are
house sparrows, starlings and pigeons in many jurisdictions), then this is the simplest
solution. However, many birds are protected by law, so it would be wise to encourage
the person to consult local authorities before acting.
In our experience, though, if no mites are present, then the sufferer is likely to
return again and again with jars or plastic bags full of fluff or pages of paper covered
with bits of dirty sticky tape. When the acarologist demonstrates their incompetence
by repeatedly failing to find the mite pests that the sufferer can clearly see, the result
is typically increasing frustration and often anger and rage. A delusion cannot be
undone by facts, and sufferers often have what they view as confirmatory symptoms:
their ‘bites’. But if no biting arthropods are present then these wheals are result of the
‘itch-scratch cycle’, more determined efforts to eradicate their parasites (see Hinkle
2011), or some underlying organic disease or allergy. The people who have come to
us with invisible mites are often fairly lucid otherwise and some seem remarkably
intelligent. Quite often they have done considerable research on biting mites and
have particular theories as to which mite is the cause and how they acquired it. Even
professional scientists can suffer from DP (Hinkle 2010).
Recent reports from the United States and Europe that summarized about 300
cases of DP (Foster et al. 2012; Freudenmann et al. 2012) have tended to support the
epidemiology and symptomology of DP that previously had been based on smaller
numbers of cases. DP sufferers can be any age (9–91 in these studies), but the
median sufferer was a middle-aged or older female. Most sufferers had previous
psychiatric problems, especially depression, substance abuse and anxiety and also
suffered from a disease that could contribute to the feeling of being infested. As well
as insects, mites, worms and microbes, DP sufferers often thought they were infested
with inanimate objects such as needles or fibres. The latter would have the advan-
tage of being difficult to refute, since fibres are likely to be present on our bodies.
Once an acarologist is convinced that they are dealing with someone who has DP
they face a moral dilemma. Although an acarologist can do nothing more for the
sufferer, there may be a serious underlying cause of the delusion. Physical causes of
442 10 Mites That Cause and Transmit Disease

itchy skin that may induce DP may include environmental irritants (e.g. fibreglass,
formaldehyde), brain disease, pellagra, diabetes mellitus, hypertension, leprosy,
over-the-counter medications and illegal drugs (‘cocaine bugs’, ‘meth mites’)
(Hinkle 2010). Psychological causes of the ‘itch-scratch cycle’ include depression,
anxiety, stress, and psychopathologies (schizophrenia, paranoia, bipolar and
obsessive-compulsive disorders) (Hinkle 2010). None of these are the realm of an
acarologist, but knowing that you are dealing with someone who has DP does leave
an obligation to attempt to inform the sufferer that they need medical help.
Because of the nature of the delusion, this attempt is unlikely to be successful,
but DEW had two instances (out of perhaps 40) where this response may have
helped. In the first case the sufferer was delighted to discover that there was a disorder
that fit his symptoms and went off to talk to his doctor about DP. Since this person
never returned, one would like to think they received appropriate treatment and are
now in remission. In the second case, a family of four appeared with a large number
of samples one of which included one Cheese Mite (Tyrophagus putrescentiae).
Cheese Mites are very common in homes, but do not burrow in our skin and I care-
fully explained the situation to the family. The father and young daughter had some
sores on their arms and claimed they felt itchy (the baby looked fine), but the mother
appeared to be the focus of the delusion and she was covered with scabby sores.
The sores on one leg were associated with the red streaks that indicate incipient
blood poisoning (sepsis) and I pleaded with the family to immediately seek medical
attention. Again, they never returned, so one can be optimistic. However, the other
38 or so cases had very unsatisfactory conclusions. Most DP sufferers simply left in
disgust to find some new and better specialist, but a few left uttering threats.

Summary

Every acarologist needs to be aware of the potential for mites to both cause disease
themselves and to act as vectors of microbes that cause human disease. By far the
most important acarine vectors of disease are the ticks (Ixodida). The reasons that
ticks are major vectors of disease are inherent in their life histories and include such
things as their predilection to feed on multiple hosts, their extended periods of
feeding, the manner in which they shed excess fluids from their blood meals and the
manner in which they digest their dinners. Emerging and resurging tick-borne
disease are significant problems in many parts of the world and many appear to be
spreading. Ticks can also cause diseases ranging from skin irritations from their
bites to paralysis caused by materials they inject into the blood system.
No other group of mites is quite as thirsty for human blood as the ticks, but some
dermanyssoid mites sometimes bite us and vector pathogens that cause disease. One
of the most important non-tick vectors of human disease, however, does not feed on
our blood nor survive on us for very long. A few genera of mammal-feedings
chiggers (larval Trombiculoidea) have species that carry the rickettsial pathogen
that causes Scrub Typhus. We become accidental hosts when we wander into the
References 443

mites’ habitats (typically grassy areas with populations of rodents) and they attach
to our skin, which they try to digest, leaving behind a hard and itchy wheal. If the
chiggers are not infected with rickettsiae, then all that results is an annoying
dermatitis, but about 1/7th of the World’s population lives in areas where chiggers
are infected and a single bite from these mites can put their lives at risk.
People do have mites that live in their skin including the Human Follicle
(Demodex folliculorum) and Sebaceous Gland Mites (Demodex brevis) (both
Demodicidae); however, any relationship between these mites and disease is
obscure. Demodecid mites live in the natural pores of the skin without burrowing.
Like many ectosymbionts of animals, follicle mites usually cause no obvious harm.
However, the Human Itch Mite Sarcoptes scabiei (Sarcoptidae) does burrow in our
skin and does cause harm, primarily because the intense reaction of our immune
system causes an intolerable itch. Because most infestations of scabies involve only
a small number of burrows, the cause of the itch may be missed or misdiagnosed or
itching caused by other arthropods or diseases may be misdiagnosed as the seldom
seen scabies mite.
Most mites that bite people do so not because we are their natural hosts, but
because their natural hosts or prey are unavailable and we are. Solving these
problems requires identification of the biting mite and knowledge of its natural
history. Such problems are the realm of Acarology and acarologists should be
willing to help when they can. However, not all people who think they are being
bitten by mites are – some are simply mistaken and once the real cause of their
misfortune is discovered, they get on with their lives. Others become fixated with
the idea that mites are feeding on them but no mites can ever be found. If all of the
potential causes of a mite problem have been eliminated and the person still claims
there must be mites, sometimes claiming they can see them and becoming infuriated
when the acarologist cannot also see them, then the psychiatric disorder call
Ekbom’s Syndrome, Delusory Parasitosis, Delusional Parasitosis, and a variety of
other names, is likely. People who suffer from this syndrome may have serious
organic disease and need medical treatment, but they are not infested with mites and
an acarologist cannot help them. Because this is a delusion, facts will not sway their
opinion and the sufferers should be told to consult their doctors.

References

Beaulieu, F. A., Dowling, P. G., Klompen, H., de Moraes, G. J., & Walter, D. E. (2011). Superorder
parasitiformes Reuter, 1909. In Z.-Q. Zhang (Ed.), Animal biodiversity: An outline of higher-level
classification and survey of taxonomic richness (pp. 123–128). Auckland: Magnolia Press.
Beck, W., & Fölster-Holst, R. (2009). Tropical rat mites (Ornithonyssus bacoti) – serious
ectoparasites. JDDG (Journal of the German Society of Dermatology), 7, 667–670.
doi:10.1111/j.1610-0387.2009.07140.x.
Broce, A. B., Zurek, L., Kalisch, J. A., Brown, R., Keith, D. L., Gordon, D., Goedeke, J., Welbourn,
C., Moser, J., Ochoa, R., Azziz-Baumgartnee, E., Yip, F., & Weber, J. (2006). Pyemotes herfsi
(Acari: Pyemotidae), a mite new to North America as the cause of bite outbreaks. Journal of
Medical Entomology, 43, 610–613.
444 10 Mites That Cause and Transmit Disease

Childers, C. C., Beshear, R. J., Frantz, G., & Nelms, M. (2005). A review of thrips species biting
man including records in Florida and Georgia between 1986–1997. Florida Entomologist, 88,
447–451.
Colloff, M. J. (2008). Taxonomy and identification of dust mites. Allergy Supplement s48,
7–12(2008). doi: 10.1111/j.1398-9995.1998.tb04989.x
Colloff, M. J. (2009). Dust mites. Dordrecht: CSIRO Publishing and Springer Science. 583 pp.
ISBN 978-9-481-2223-3.
Currier, R. W., Walton, S. F., & Currie, B. J. (2012). Scabies in animals and humans: History, evo-
lutionary perspectives, and modern clinical management. Annals of the New York Academy of
Sciences, 1230, E50–E60. doi:10.1111/j.1749-6632.2011.06364.x.
Del Giudice, P., Blanc-Amrane, V., Bahadoran, P., Caumes, E., Marty, P., Lazar, M., Boissy, C.,
Desruelles, F., Izri, A., Ortonne, J.-P., Counillon, E., Chosidow, O., & Delaunay, P. (2008).
Pyemotes ventricosus dermatitis southeastern France. Emerging Infectious Diseases, 14,
1759–1761.
Foster, A. A., Hylwa, S. A., Bury, J. E., Davis, M. D. P., Pittelkow, M. R., & Bostwick, J. M. (2012).
Delusional infestation: Clinical presentation in 147 patients seen at Mayo Clinic. Journal of the
American Academy of Dermatology, 67, 673.e1–673.e10. doi:10.1016/j.jaad.2011.12.012.
Frances, S. P. (2005). Potential for horizontal transmission of Orientia tsutsugamushi by chigger
mites (Acari: Trombiculidae). International Journal of Acarology, 31, 75–82.
Freudenmann, A. A., Lepping, P., Huber, M., Dieckmann, S., Bauer-Dubau, K., Ignatius, R.,
Misery, L., Schollhammer, M., Harth, W., Taylor, R. E., & Bewley, A. P. (2012). Delusional
infestation and the specimen sign: A European multicentre study in 148 consecutive cases.
Bristish Journal of Dermatology, 167, 24–251. doi:10.1111/j.1365-2133.2012.10995.x.
Garrett, L. (1994). The coming plague, newly emerging diseases in a world out of balance. Victoria:
Penguin.
Gratz, N. G. (1999). Emerging and resurging vector-borne diseases. Annual Review of Entomology,
44, 51–75.
Hinkle, N. C. (2010). Ekbom syndrome: The challenge of “invisible bug” infestations. Annual
Review of Entomology, 55, 77–94.
Hinkle, N. C. (2011). Ekbom syndrome: A delusional condition of “bugs in the skin”. Current
Psychiatry Reports, 13, 178–186.
Kilpatrick, A. M., & Randolph, S. E. (2012). Drivers, dynamics, and control of emerging vector-
borne zoonotic diseases. Lancet, 380, 1946–1955.
Kuehn, B. M. (2013). Emerging tick-borne diseases expand range along with rebounding deer
populations. Journal of the American Medical Association, 309, 124–125.
Kuo, C. C., Huang, C. L., & Wang, H. C. (2011). Identification of potential hosts and vectors of
scrub typhus and tick-borne spotted fever group rickettsiae in eastern Taiwan. Medical and
Veterinary Entomology, 25, 169–177.
Kushon, D. J., Helz, J. W., Williams, J. M., Lau, K. M. K., Pinto, L., & St. Aubin, F. E. (1993).
Delusions of parasitosis: A survey of entomologists from a psychiatric perspective. Bulletin of
the Society for Vector Ecology, 18, 11–15.
Lesna, I., Sabelis, M. W., van Niekerk, T. G. C. M., & Komdeur, J. (2012). Laboratory tests for
controlling poultry red mites (Dermanyssus gallinae) with predatory mites in small ‘laying
hen’ cages. Experimental and Applied Acarology, 58, 371–383.
Leung, V., & Miller, M. (2011). Detection of scabies: A systematic review of diagnostic methods.
Canadian Journal of Infectious Diseases & Medical Microbiology, 22, 143–146.
Levi, T., Kilpatrick, A. M., Mangel, M., & Wilmers, C. C. (2012). Deer, predators, and the
emergence of Lyme disease. Proceedings of the National Academy of Sciences of the United
States of America, 109, 10942–10947. doi:10.1073/pnas.1204536109.
Marquardt, W. (2005). Biology of disease vectors (2nd ed.). Amsterdam: Elsevier.
Mehlhorn, H. (Ed.). (2008). Encyclopedia of parasitology (3rd ed., Vol. 1–2). New York: Springer.
1573 pp.
References 445

Mounsey, K. E., McCarthy, J. S., & Walton, S. F. (2013). Scratching the itch: New tools to advance
understanding of scabies. Trends in Parasitology, 29, 30–42.
Mullen, G., & Durden, L. (Eds.). (2002). Medical and veterinary entomology. New York:
Academic.
Niebylski, M. L., Peacock, M. G., & Schwan, T. G. (1999). Lethal effect of Rickettsia rickettsii on its
tick vector (Dermacentor andersoni). Applied and Environmental Microbiology, 65, 773–778.
Nutting, W. B. (1985). Prostigmata–Mammalia: Validation of coevolutionary phylogenies. In
K. C. Kim (Ed.), Coevolution of parasitic arthropods and mammals (pp. 569–640). New York:
Wiley-Interscience.
Phasomkusolsil, S., Tanskul, P., Ratanatham, S., Watcharapichat, P., Phulsuksombati, D.,
Frances, S. P., Lerdthusnee, K., & Linthicum, K. J. (2012). Influence of Orientia tsutsugamushi
infection on the developmental biology of Leptotrombidium imphalum and Leptotrombidium
chiangraiensis (Acari: Trombiculidae). Journal of Medical Entomology, 49, 1270–1275.
Radolf, J. D., Caimano, M. J., Stevenson, B., & Hu, L. T. (2012). Of ticks, mice and men:
Understanding the dual-host lifestyle of Lyme disease spirochaetes. Nature Reviews
Microbiology, 10, 87–99. doi:10.1038/nrmicro2714.
Renvoisé, A., van’t Wout, J. W., van der Schroeff, J.-G., Beersma, M. F., & Raoult, D. (2012).
A case of rickettsialpox in Northern Europe. International Journal of Infectious Diseases, 16,
e221–e222.
Rollend, L., Fish, D., & Childs, J. E. (2013). Transovarial transmission of Borrelia spirochetes by
Ixodes scapularis: a summary of the literature and recent observations. Ticks and Tick-borne
Diseases, 4, 46–51.
Strickman, D. (2001). Scrub typhus. In M. W. Service (Ed.), The Encyclopedia of Arthropod-
transmitted Infections (pp. 608–462). New York: CABI Publishing.
Walker, D. H. (2003). Rickettsial diseases in travellers. Travel Medicine and Infectious Disease, 1,
35–40.
Walter, D. E., & Shaw, M. (2005). Mites and disease. In W. Marquardt (Ed.), Biology of disease
vectors (2nd ed., pp. 24–44). Amsterdam: Elsevier.
Walter, D. E., Lindquist, E. E., Smith, I. M., Cook, D. R., & Krantz, G. M. (2009). Order
Trombidiformes. In G. W. Krantz & D. E. Walter (Eds.), A manual of acarology (3rd ed.,
pp. 233–420). Lubbock: Texas Tech University Press.
Wood, H., & Artsob, H. (2012). Spotted fever group rickettsiae: A brief review and a Canadian
perspective. Zoonoses and Public Health, 59(Supplement 2), 65–79.
Chapter 11
Mites and Biological Diversity

‘Ubiquitous’ is a much abused adjective that literally means ‘being everywhere’


but in practice it means considerably less. The word is often trotted out by ecolo-
gists and systematists to justify their working on a particular group of plants or
animals (e.g. ‘rodents make good bioindicators because they are ubiquitous in
terrestrial ecosystems’), sometimes with little support for their assertions.
However, mites have a fair claim for being truly omnipresent (Fig. 11.1). With the
exception of the water column of the open ocean, they exist in every sort of
aquatic, terrestrial, arboreal and parasitic habitat. In spite of this, mites rarely
appear in general biodiversity surveys (but see Basset et al. 2012). Data on mite
diversity in tropical ecosystems is especially rare. Two interrelated reasons for
this neglect, shared with other ubiquitous and ‘hyper-diverse’ taxa such as nema-
todes, are small body size and alleged difficulty of identification. Small-bodied
organisms are often overlooked in rapid assessments of biodiversity, particularly
when samples are sorted with the naked eye. Even if mites are seen, it is almost
impossible to identify them to species in the field; instead, maceration, slide-
mounting and microscopic examination are required.
Although the number of described species of mites is similar to that of spiders,
recent estimates of global acarine diversity reach as high as one million species:
about five times as high as the most optimistic spider claims. Why would mite diver-
sity be so much greater than that of their more familiar relatives? Trophic diversity
is certainly one contributing factor. Except for occasional sipping of nectar and
consumption of pollen when re-ingesting their webs, spiders feed only on fluids
from animal prey (Pollard et al. 1995). A few spiders are kleptobionts that steal dead
prey from other spiders but, by and large, spiders capture living prey (Wise 1993).
Mites, as we have documented in previous chapters, eat everything and everyone.
But trophic flexibility alone does not seem enough to explain acarine success.
The species richness of any site results from a complex interplay between evolu-
tionary, ecological and stochastic factors. Despite this, assessing local diversity is a
simple process at first glance – species are identified in collections of organisms
from local habitats. For example, determining the species richness of birds in a

D.E. Walter and H.C. Proctor, Mites: Ecology, Evolution & Behaviour: Life at a Microscale, 447
DOI 10.1007/978-94-007-7164-2_11, © Springer Science+Business Media Dordrecht 2013
448 11 Mites and Biological Diversity

Fig. 11.1 Moose pasture or miteome? Although biological diversity is often portrayed by its
vertebrate components, it is the invertebrates that are the weft and warp of the diversity tapestry
(Image by DE Walter and HC Proctor)

patch of rainforest would not be an especially daunting task. Three major habitats
immediately come to mind: the leafy canopy, the understorey and the ground.
Binocular surveys, mist nets and song identification are obvious assessment meth-
ods. However, attempts to tally mite species in the same forest patch immediately
founder on this question: what are the major habitats containing mites?

Mites and Microhabitats

It is almost an ecological truism that the greater the complexity and diversity of
habitats in an area, the greater is the diversity of species that should be able to live
there (Kovalenko et al. 2012). For very small organisms such as mites, there may be
no simple answer to the question of mite habitats. For example, it can be convenient
to write about ‘rainforest mites’ and this concept may be appropriate for plant para-
sites and for those that feed on leaf litter; however, the immediate habitats of many
‘rainforest’ mites are only tangentially tree-like. For example, any species of bird,
beetle or butterfly inhabiting the rainforest may act as home for one or more species
of mite (e.g. Treat 1975; Colwell and Naeem 1994; Brown and Wilson 1994;
Seeman 2002, 2007, 2001).
Consider birds as one obvious mite habitat in a forest patch. Birds represent not
one but at least three complexes of potential habitats: nests, outer surface of the
Mites and Microhabitats 449

body, internal body. On closer examination each of these may be further divisible,
e.g. the outer body can be divided into skin, leg scales and feathers; feathers into
body contour, wing and tail; each feather into quill and barbules. The more closely
one looks at a presumed mite habitat, the more microhabitats one finds – a kind of
reductio ad absurdum reminiscent of Zeno’s story of Achilles and the Tortoise.
Zeno of Elea was a Greek philosopher of the fifth century BC who was infamous
for setting paradoxes that questioned the reality of motion (Faris 1996). Perhaps the
best known of these is a hypothetical race between a fleet-footed hero, Achilles, and
a plodding tortoise. The tortoise was granted a head start but whenever Achilles
reached the tortoise’s previous position, the tortoise had moved ahead. Therefore,
Achilles might spend an infinity narrowing the distance between himself and the
tortoise, but he could never catch up.
Zeno’s paradox of Achilles and the Tortoise depends on the infinite divisibility of
space and seems an appropriate analogy for any assessment of ‘microhabitats’. The
term itself is ambiguous but then again, so is ‘habitat’. Consider a mite that spends
most of its life under the elytra of a passalid beetle. If we consider the subelytral
space on these beetles as this mite’s microhabitat, then is the beetle the habitat?
What of the log the beetle lives in? The floor of the forest on which the log rests?
The forest itself?
For some mites, such as permanent ectosymbionts, the subelytral space may be
the best habitat designation, but others range through the habitat hierarchy. For
example, more than two-dozen different families of mites make their homes, or at
least their transportation, on Betsy Beetles (Coleoptera: Passalidae, see Seeman
2000, 2002, 2007, 2001). Many of these mites are only known from these beetles –
although little is known about what they may be doing. For some, all stages are found
on the beetles, a commensal or parasitic association is likely and the beetle itself is
the best representation of the microhabitat. For others, though, only one stage is
found on the beetles and all stages can be found in the beetles’ galleries in rotting
logs. For example, the large adults (up to 5 mm long in the genus Megisthanus)
of trigynaspids (Mesostigmata: Trigynaspida) are often found on Betsy Beetles.
Although these mites have fearsome-looking chelicerae, they seem to do the beetles
no harm and when living off the beetles in their galleries in logs, feed on nematodes,
small arthropods and carrion. They may possibly provide a benefit to the beetles by
reducing pests. Immature stages of trigynaspids also are predatory in the galleries of
the beetles. The presence of adult trigynaspid mites on beetles is most likely a pho-
retic association for dispersal from log to log. For some species, adult male mites
appear to use the beetles as a place to find mates (e.g. usually only one male
Megisthanus is found on a beetle, although several females are often present) and the
mated females disperse to new logs on the beetles. Passalid beetles tend to occur in
logs with a high water content found in closed forests where termites are rare, so in
this sense the forest type may be the ‘microhabitat’ for both the beetle and its mites.
In the following discussion we will use ‘microhabitat’ as the smallest habitat unit
and use ‘habitat’ in a general sense that includes the more inclusive components.
Undoubtedly there are limits to the degree that mites can partition habitat space but
these may be artificially truncated by the limits of the investigator’s imagination.
450 11 Mites and Biological Diversity

Until they were first discovered, who would have thought to look for mites inside
feather shafts, turtle cloacas, crab gills, bee tracheae, moth ears, the stink glands of
bugs or any of the myriad other unexpected microhabitats exploited by mites?

Mites and Complementarity

Assessment of the species richness of any group requires two measures: the local
richness (α-diversity) of collections and the complementarity between collections,
where complementarity (β-diversity, or ‘turnover’) refers to the distinctiveness of
species assemblages across two or more sites (Colwell and Coddington 1994). Both
of these components are influenced by recent historical factors (dispersal events,
extremes of weather, etc.) and by longer-term evolutionary factors (age of habitats,
geographical distribution, coevolutionary interactions). If mites are megadiverse on
a global scale, then they should also tend to be highly diverse at a local scale. This
local richness will be a function of both microhabitat diversity and microhabitat
specificity. That is, both a large number of local microhabitats must exist and mite
species must tend to be restricted to particular microhabitats.
Young et al. (2012) evaluated mite diversity at one site using molecular taxo-
nomic units based on the barcode region of the mitochondrial gene CO1. They
found almost 900 apparent species of mites (135 Mesostigmata, 327 Sarcoptiformes,
437 Trombidiformes) at their site and a high species turnover among substrates (e.g.
moss, soil, litter, woody debris and lichens) and between forested and non-forested
sites. Accumulation curves were used to estimate that 423 species (34 %) of
Sarcoptiformes (Oribatida, Endeostigmata), 173 species (14 %) of Mesostigmata,
and 633 species (52 %) of Trombidiformes (Prostigmata, Sphaerolochida) were
likely present, for a total of 1,229 species of Acari. That is a lot of mite diversity, but
the surprising thing is that this study took place not in the tropics, but in the cold
wastes along Hudson Bay near Churchill, Manitoba, Canada in boreal forest, bog,
fen, tundra, marine beach, and rock bluff habitats.
Far to the southwest of Churchill in the aspen parklands of central Alberta, the
authors have a quarter section (160 acres more or less) of land that they have been
sampling for mites for the last 6 years. Our estimates are based on morphological
criteria of one stage (adult, larva or deutonymph) and we have sampled only very
limited areas of aspen woodland, scrubby hillside, pasture, and slough, but we have
identified 256 species of mites from our site: 68 Parasitiformes (Mesostigmata + 1
tick) (27 %), 108 Sarcoptiformes (42 %), and 80 Trombidiformes (31 %). Although
we have done our best to identify all of the mite species collected, the smaller
Trombidiformes are undoubtedly under-sampled relative to the other groups and are
the least tractable taxonomically. Our sampling was not as rigorous as that of Young
et al. (2012) and lacked the molecular component that may tease out cryptic species,
but it seems clear that even in boreal habitats, local mite diversity may be high.
Local microhabitat diversity, however, is not enough to produce global megadi-
versity. There must also be a high level of complementarity from site to site.
Mites and Complementarity 451

Fig. 11.2 Similar sampling effort (number of individual mites identified) produced more than four
times as many species of leaf-inhabiting predatory Mesostigmata in tropical rainforest (Far North
Queensland) compared to temperate rainforest (Victoria) in Australia (Image by DE Walter)

Consider the fauna of predatory mesostigmatans found on the leaves of rainforest


trees in temperate south-eastern Australia (Fig. 11.2). In these accumulation curves
(also called collector’s curves), some measure of collection effort is accumulated on
the x-axis and total species richness is accumulated on the y-axis. Collection effort
is commonly measured in terms of accumulated number of samples, of areas or of
time spent sampling; however, perhaps the clearest measure of effort is the number
of individuals identified to species. Although the logical sequence to accumulate
sampling effort is the order in which samples were taken or processed, multiple
randomisations of collection order (rarefaction) produce a smoother line that is
easier to interpret. The lines in Fig. 11.2 were produced using 50 randomisations of
sample order and were calculated in EstimateS 5.0 (available from R.K. Colwell,
Department of Ecology and Evolutionary Biology, University of Connecticut, U-42,
Storrs, CT 06269–3042, USA).
More than a dozen species of leaf-dwelling mesostigmatans were found in
temperate Australia and most of these were common in many areas. So, the collector's
curve in Fig. 11.2 quickly reaches an asymptote (i.e. flattens out), indicating that
about 14 species of mesostigmatans live on leaves in temperate forests and that
most of these species were collected in the first half dozen samples. The additional
samples added no more information about species richness. However, compare this
to the curve for collecting foliar mesostigmatans from tropical rainforest in Far
North Queensland. In this case the curve shows no sign of levelling off, meaning
that the total number of mite species is much higher than the maximum collected.
It seems that many tropical mite species are distributed over relatively small
geographical areas (Walter and Proctor 1998). Stanton (1979) also found a higher
452 11 Mites and Biological Diversity

rate of turnover among tropical compared to temperate litter mites and that tropical
mites tended to be more habitat restricted and had higher proportions of rare spe-
cies. The image of Achilles endlessly narrowing the gap between himself and the
Tortoise, but never catching up, seems especially appropriate when chasing species
richness in tropical rainforests. But might there be a shortcut to estimating local
richness of mites, in the tropics or elsewhere? There is a vast literature on the value
of taxonomic surrogacy (the use of one taxon to estimate richness of another, espe-
cially with regard to conservation planning (Rodrigues and Brooks 2007). For
example, it seems logical that there should be a positive relationship between local
richness of plants and richness of the fauna associated with the plants, either on their
living tissues (such as the mesostigmatans mentioned above) or on and in their dead
tissues on leaf litter. Proctor et al. (2003) tested this by surveying invertebrates,
particularly mites, from 50 tropical and 54 subtropical sites in Queensland and New
South Wales, Australia. Floristic richness of each site was calculated and structural
attributes such as canopy cover, dead wood, and litter were determined. They found
positive, though weak, correlations between plant species richness and higher-taxon
richness of non-mite invertebrates in both regions; however, mite family/genus
richness showed no significant relationship with plant richness. A pasture site with
only 10 vascular plant species could have more mite taxa than a rainforest site with
> 100 plant species. Structural variables were better predictors of richness of mites
and other invertebrates than was plant richness. They concluded that floristic diversity
should not a priori be considered a surrogate for litter-invertebrate diversity, at least
in Australia.

Size and Biodiversity

Robert May (1978, 1988) speculated that one reason for the high diversity of insects
was their small size. Logically then, very small-bodied taxa such as mites may be
even more diverse than insects. May noted that on a log-log scale, the number of
terrestrial animal species declined sharply above the modal size class (3.16 mm–1 cm)
so that a 10-fold increase in length resulted in a 100-fold decrease in number of spe-
cies. Although it is likely that this simple size–diversity relationship is an artefact
(Loder et al. 1997), a decline in diversity among animals larger than the modal size
class is understandable in terms of habitat theory and common experience (May
1978). We all know that there are more species of mice than of elephants, and more
fish than whales. However, May’s size classes below 1 cm also decline precipitously
in richness. Why this happens is not so obvious, unless this small-species deficit
results from a bias against identifying small species. So, if S ~ L-2 is a reasonable
approximation to the distribution of species by size, then the shortfall of small spe-
cies might be filled by the study of mites and other very small animals.
To test this hypothesis we estimated the size distribution of 415 species of mites
from subtropical forest in the Green Mountains (28°15′S, 153°08′E) section of
Lamington National Park, Queensland. Only adult mites were used, and body size
Size and Biodiversity 453

Fig. 11.3 Body size (log of


body length in millimetres)
of 415 species of mites from
Subtropical Rainforest in
Lamington National Park,
Queensland, Australia
(see text for details)

was estimated to the nearest 5 μm by measuring along the midline of the idiosoma
with an ocular micrometer at ×100. If a series of specimens was available, then the
median length was used. Mites were collected over a 10-year period using a variety
of collecting techniques including pyrethrum knockdown, bark spraying, leaf and
stem collections, rearings from fungal sporocarps, hand collection of large arthro-
pods, kick sampling in a stream, and Tullgren-funnel extraction of litter (Kitching
et al. 1993; Walter and Proctor 1998; Walter et al. 1998).
Parasitiform mites were significantly larger than sarcoptiform or trombidiform
mites (Fig. 11.3), primarily because of the relatively large size of ticks and of
beetle-associated mesostigmatans (right-side histogram bars). For example, the six
largest species of mites were equally divided between the passalid-associated
genus Megisthanus and the tick genus Ixodes. Only one of the 27 species less than
250 μm in length was a mesostigmatan, the remainder were Eriophyoidea,
Tarsonemidae (both Prostigmata) and Oribatida. Overall, more than half (52 %) of
all mite species were < 500 μm in length but the smallest size classes show a relative
decline in diversity.
Although mites are very small (Fig. 11.4), species in the very smallest size class
(< 317 μm in length) are not as common as the next largest size class and the modal
class for mites is even larger. So even though we were careful to identify even the
smallest mite, our data supports a decline in diversity of very small species. Part of
this decline, however, may still result from collecting bias. For example, only five
species of Eriophyoidea (tiny gall mites) were collected, although more than 70 spe-
cies of woody plants occur on just one plot in the Green Mountains (Kitching et al.
1993) and an estimate of one species of Eriophyoidea for each species of woody
plant is conservative. More intensive collecting of plant parasites, however, is also
likely to uncover numerous additional species of Tenuipalpidae, Tuckerellidae
and Tetranychidae – most of which would belong to the next larger size class.
454 11 Mites and Biological Diversity

Fig. 11.4 Although many insects are smaller than some mites, the very smallest mites probably
approach the smallest size at which an arthropod in a terrestrial environment can function. The
smallest known adult mites have body lengths of < 0.1 mm and mobile larvae reach down to about
0.05 mm (SEMs by DE Walter)

Additionally – and conclusively – the smallest size class shown in Fig. 11.3 was not
even considered in May’s original scheme. To raise this size class to the S ~ L-2 line
would require a staggering diversity of the smallest mites.

Host Specificity, Size and Diversity

Because mites are much smaller than most other organisms, larger animals and plants
represent cornucopias of microhabitats to mites. Even the feathers of a single species
of parrot can provide specific microhabitats for more than two-dozen species of
feather mites alone (Pérez 1996). Many host habitats have been colonised repeatedly
by different lineages of mites, which have then undergone extensive radiations.
Therefore, the species richness of symbiotic mites will be largely a function of the
current diversity of hosts, the success that mites have had colonising those hosts
and the host specificity of the mites. For example, most species of bees (Apoidea)
are host to several species of symbiotic mites and at least 31 lineages of mites have
independently colonised the bee habitat. Individual bee–mite species tend to be
geographically restricted in their distribution and to be host specific at the species,
species-complex or generic levels (Eickwort 1994). About 25,000 species of bees
Host Specificity, Size and Diversity 455

Table 11.1 Known families and genera of mites, Walter and Proctor (1999) estimates of described
species diversity, 2013 estimates (described, in press and in hand) and estimates of likely minimum
and maximum species richness of living mites
1999 2013 Low High
Families Genera species species estimate estimate
Opilioacarida 1 10 17 35 100 200
Holothyrida 3 13 32 35 160 320
Ixodida 3 19 880 900 1,000 1,200
Mesostigmata 110 916 11,632 12,017 71,000 138,000
Sejida 6 19 69 88 800 1,200
Trigynaspida 27 112 267 310 3,000 6,000
Monogynaspida 77 785 11,296 12,587 60,000 120,000
Parasitiformes 117 958 12,561 12,985 72,260 150,520
Endeostigmata 10 27 100 110 1,000 2,000
Oribatida 174 1,260 11,000 10,300 33,000 110,000
Astigmata 80 1,133 4,500 6,220 90,000 180,000
Sphaerolichida 2 2 20 25 200 400
Prostigmata 145 2,123 17,050 25,121 317,250 637,500
Acariformes 411 4,545 32,670 41,776 441,450 929,900

have been described and perhaps 40,000 currently exist; therefore, the diversity of
bee–mites alone is likely to exceed the current described diversity of the Acari.
In Table 11.1 we present our hypotheses about the likely diversity of major mite
taxa. Our best estimate of the number of known species of Acari (those formally
described minus synonymies, in press and in hand) is about 55,000 – which is
~10,000 more than in Walter and Proctor (1999). Table 11.1 reflects the current
taxonomic structure (see open access 2011 Zootaxa overview http://www.mapress.
com/zootaxa/list/2011/3148.html) The number of known species of Parasitiformes
has only increased about 3 % in the last 13 years and much of that increase has been
in one family: the Phytoseiidae. Our expectations that other Dermanyssiae and in
particular the Uropodina would show substantial growth in species has not occurred.
Although this poor showing likely reflects the few systematists working in these
groups, and the largely unknown Gondwanan faunas that have yet to be seriously
studied, it has caused us to reconsider our 1999 minimum and maximum estimates
and pull them back a bit. One of the few parasitiform groups that we have increased
our estimate for is the Opilioacarida (treated as the sistergroup of the Parasitiformes
in 1999). These intriguing and seemingly primitive mites appear to be much more
diverse than previously expected. Barring sibling species complexes, most species
of ticks (Ixodida) have probably been described, and although most Holothyrida
await description, this taxon does not seem to be especially diverse. When finally
described, the Sejida, Trigynaspida and Monogynaspida of the southern continents
will add tremendously to the total known Mesostigmata, but we think the likely total
number of species living today is closer to our lower than to our higher estimate.
The Acariformes is by far the more diverse of the two superorders of mites,
primarily because of a number of symbiotic radiations. The Oribatida have only a
456 11 Mites and Biological Diversity

few species associated with other animals and their diversity has changed only
incrementally over the last decade. In the boreal forests, grasslands, and mountains
of Alberta, Canada between 10 % and 15 % of the species discovered in the last 6
years appear to be undescribed (Walter et al. 2013). A host of tropical and temperate
species on the southern continents remain to be described, but we see no need to
change our 1999 estimates of how many species may be alive today. The same goes
for the Endeostigmata (which included the Sphaerolichida in 1999). In contrast, the
derived lineage of oribatid mites that we call the Astigmata includes the common
mange mites of domestic animals, and synanthropes such as the house dust mites,
cheese mites and scabies. However, most astigmatan habitats are not of such imme-
diate economic interest and the bulk of the diversity in the Astigmata consists
largely of undescribed parasites and commensals than live on and in economically
unimportant species of birds, mammals, reptiles, arthropods and their nests.
Numbers of these mites continue to be described and our current estimate of the
number of known species is about 38 % higher than in 1999.
Similarly, the Parasitengonina include a vast array of chiggers, velvet mites and
water mites, most of which require a vertebrate or invertebrate host to complete
their life cycle. Their known diversity has also increased impressively over the last
13 years. So has what is in all likelihood the most diverse functional group of acari-
form mites: the plant parasites, including the spider mites and their relatives in the
Raphignathina, and the gall and rust mites in the Eupodina. We suspect that the final
totals among these groups will be nearer the upper rather than the lower of our
estimates.
Repeated colonisation and diversification by lineages of symbiotic mites has
occurred in association with many species-rich groups, e.g. wasps, ants, beetles,
termites, cockroaches, orthopterans, hemipterans, myriapods, spiders, reptiles,
birds, mammals, conifers, angiosperms and so on. Many of these mites are parasitic
on their hosts but others are commensals or mutualists (Eickwort 1994; Walter and
O’Dowd 1995). Regardless of the nature of the association, a tendency towards host
specificity appears to be common in those mites that spend all of their life on the
host (see Chap. 9).
For example, an undescribed mite in the family Acarophenacidae from Alberta,
Canada, is shown in Fig. 11.5. The 35 described species in this family are host-
specific parasitoids of the eggs of insects such as beetles, thrips and scale insects
(Walter et al. 2009). The short, thick first pair of legs (Fig. 11.5a) are covered in
sensory setae for locating a host, in this case the larva of a hairy fungus beetle
(Fig. 11.5b, top), and terminate in robust nippers used to hold on to the hairs of
beetle larva. When the larva pupates and moults to an adult beetle (Fig. 11.5b), the
mites clamber onto the females and wait for them to begin laying eggs. As an egg is
laid a mite clambers off the beetle and uses the suction-cup like surface at the ante-
rior end (Fig. 11.5c) to attach to the egg as the mouthparts (retracted into the body)
are exerted to puncture the eggshell and begin feeding. Offspring develop within the
growing body of the female (physogastry) and fully mature adults eventually emerge
to seek new hosts. If such host specificity is common in symbiotic mites, then
the maximum guesstimates in Table 11.1 may prove near the true total.
Summary 457

Fig. 11.5 The hidden face of mite-host specificity: an undescribed species of Acarophenacidae
associated with hairy fungus beetles (Mycetophagidae) living in oyster mushrooms on logs in
Alberta, Canada. (a) Legs I are covered with mechanoreceptor and chemosensory setae and the
tips modified as clamps for holding onto a beetle larva as it feeds in decaying fungi. (b) When the
beetle pupates and emerges, the mites transfer to female beetles and then decamp as eggs are laid.
(c) The mite’s mouthparts are withdrawn behind sucker like region with which the mite attaches to
the beetle egg and feeds as a parasitoid (SEMs and photo by DE Walter)

Summary

The spectre of mite megadiversity haunts estimates of global biodiversity but more
noise than hard data exists to support these claims. In temperate regions, the acaro-
faunas are certainly diverse but not exceptionally so. For example, Canada is
estimated to have fewer than 10,000 species of mites (Lindquist et al. 1979), and
America north of Mexico only 30,000 (OConnor 1990). If the Acari is a hyperdi-
verse group (sensu Hammond 1992), then the bulk of acarine biodiversity must
reside in the acarologically unexplored tropics. Based on studies in subtropical
458 11 Mites and Biological Diversity

to tropical rainforests in Australia, this does appear to be true: both rich and
complementary assemblages of mites are characteristic across sites and among
microhabitats within a site. Even extensive collections within strictly defined habi-
tats fail to adequately estimate local species richness.
The belief that small size is positively correlated to high diversity has helped to
sustain the hypothesis of hyperdiversity in the Acari. Microhabitat use and the small
size of most mites (e.g. adults of 52 % of 415 species from a subtropical rainforest
were less than half a millimetre in length) appear to be associated with the success
of tropical mites. Although we have boldly hypothesised high diversity in many
mite taxa (Table 11.1), host and microhabitat specificity in tropical mites must be
addressed before any accurate estimate of global acarine diversity can be attempted.

References

Basset, Y., & 37 others. (2012). Arthropod diversity in a tropical forest. Science, 338, 1481–1484.
doi:10.1126/science.1226727.
Brown, J. M., & Wilson, D. S. (1994). Poecilochirus carabi: Behavioral and life-history adap-
tations to different hosts and the consequences of geographical shifts in host communities.
In M. A. Houck (Ed.), Mites: Ecological and evolutionary analyses of life-history patterns
(pp. 1–22). New York: Chapman & Hall.
Colwell, R. K., & Coddington, J. A. (1994). Estimating terrestrial biodiversity through extrapola-
tion. Philosophical Transactions of the Royal Society of London B, 345, 101–118.
Colwell, R. K., & Naeem, S. (1994). Life-history patterns of hummingbird flower mites in relation
to host phenology and morphology. In M. A. Houck (Ed.), Mites: Ecological and evolutionary
analyses of life-history patterns (pp. 23–44). New York: Chapman & Hall.
Eickwort, G. C. (1994). Evolution and life-history patterns of mites associated with bees. In
M. A. Houck (Ed.), Mites: Ecological and volutionary analyses of life-history patterns
(pp. 218–251). New York: Chapman & Hall.
Faris, J. A. (1996). The paradoxes of Zeno. Sydney: Avebury.
Hammond, P. M. (1992). Species inventory. In B. Groombridge (Ed.), Global biodiversity: Status
of the earth’s living resources (pp. 17–39). London: Chapman & Hall.
Kitching, R. L., Bergelson, J. M., Lowman, M. D., McIntyre, S., & Carruthers, G. (1993). The
biodiversity of arthropods from Australian rainforest canopies: General introduction, methods,
sites and ordinal results. Australian Journal of Ecology, 18, 181–191.
Kovalenko, K. E., Thomaz, S. M., & Warfe, D. M. (2012). Habitat complexity: Approaches and
future directions. Hydrobiologia, 685, 1–17.
Lindquist, E. E., with contributions by Ainscough, B. D., Clulow, F. V., Funk, R. C., Marshall,
V. G., Nesbitt, H. H. J., OConnor, B. M., Smith, I. M., & Wilkinson, P. R. (1979). Acari. In
H. V. Danks (Ed.), Canada and its insect fauna (Memoirs of the Entomological Society of
Canada, Vol. 108, pp. 252–263, 267–284).
Loder, N., Blackburn, T. M., & Gaston, K. J. (1997). The slippery slope: Towards an understanding
of the body size frequency distribution. Oikos, 78, 195–201.
May, R. M. (1978). The dynamics and diversity of insect faunas. In L. A. Mound & N. Waloff
(Eds.), Diversity of insect faunas (pp. 188–204). Oxford: Blackwell Scientific Publications.
May, R. M. (1988). How many species are there on earth? Science, 241, 1441–1449.
OConnor, B. M. (1990). The North American Acari: Current status and future projections. In
M. Kosztarab & C. W. Schaefer (Eds.), Systematics of the North American insects and
References 459

arachnids: Status and needs (Virginia agricultural experiment station information series 90–1,
pp. 21–29). Blacksburg: Virginia Polytechnic Institute.
Pérez, T. M. (1996). The eggs of seven species of Fainalges Gaud and Berla (Xolalgidae) from the
green conure (Aves, Psittacidae). In R. Mitchell, D. J. Horn, G. R. Needham & W. C. Welbourn
(Eds.), Acarology IX: Volume 1, Proceedings (pp. 297–300). Columbus: Ohio Biological
Survey.
Pollard, S. D., Beck, M. W., & Dodson, G. N. (1995). Why do male crab spiders drink nectar?
Animal Behaviour, 49, 1443–1448.
Proctor, H. C., Kanowski, J., Wardell-Johnson, G., Reis, T., & Catterall, C. P. (2003). Does
diversity beget diversity? A comparison between plant and leaf-litter invertebrate richness
from pasture to rainforest. Records of the South Australian Museum Monograph Series, 7,
267–274.
Rodrigues, A. S. L., & Brooks, T. M. (2007). Shortcuts for biodiversity conservation planning: The
effectiveness of surrogates. Annual Review of Ecology, Evolution, and Systematics, 38, 713–737.
Seeman, O. D. (2000). The immature life stages of the Fedrizziidae (Mesostigmata: Fedrizzioidea).
Acarologia, 41, 39–52.
Seeman, O. D. (2002). Mites and passalid beetles: Diversity, taxonomy and biogeography. Ph.D.
thesis. The University of Queensland (p. 339).
Seeman, O. D. (2007). Revision of the Fedrizziidae (Acari: Mesostigmata: Fedrizzioidea).
Zootaxa, 1480, 1–55.
Seeman, O. D., Proctor, H., Norton, R. A., & Colloff, M. (2001). Myriad Mesostigmata associated
with log-inhabiting arthropods. In R. B. Halliday & D. E. Walter (Eds.), Acarology: Proceedings
of the 10th International Congress. Canberra: CSIRO Publishing.
Stanton, N. L. (1979). Patterns of species diversity in temperate and tropical litter mites. Ecology,
60, 295–304.
Tr Treat, A. E. (1975). Mites of moths and butterflies. Ithaca: Cornell University Press.
Walter, D. E., & O’Dowd, D. J. (1995). Beneath biodiversity: Factors influencing the diversity and
abundance of canopy mites. Selbyana, 16, 12–20.
Walter, D. E., & Proctor, H. C. (1998). Predatory mites in tropical Australia: Local species richness
and complementarity. Biotropica, 30, 72–81. [this is cited as 1998b in the text, but should just
be 1998]
Walter, D. E., & Proctor, H. C. (1999). Mites: Ecology, evolution and behaviour (p. 322).
Wallingford: University of NSW Press, Sydney and CABI. ISBN 0 86840 529 9.
Walter, D. E., Latonas S., & Byers, K. (2013). Almanac of Alberta Oribatida. Part 1. Ver. 2.3.
Edmonton: The Royal Alberta Museum. http://www.royalalbertamuseum.ca/natural/insects/
research/research.htm.
Walter, D. E., Seeman, O., Rodgers, D., & Kitching, R. L. (1998b). Mites in the mist: How unique
is a rainforest canopy knockdown fauna? Australian Journal of Ecology, 23, 501–508.
Walter, D. E., Lindquist, E. E., Smith, I. M., Cook, D. R., & Krantz, G. M. (2009). Order
Trombidiformes. In G. W. Krantz & D. E. Walter (Eds.), A manual of acarology (3rd ed.,
pp. 233–420). Lubbock: Texas Tech University Press.
Wise, D. H. (1993). Spiders in ecological webs. Cambridge: Cambridge University Press.
Young, M. R., Behan-Pelletier, V. M., & Hebert, P. D. N. (2012). Revealing the hyperdiverse mite
fauna of subarctic Canada through DNA barcoding. PLoS One, 7, e48755. doi:10.1371/journal.
pone.0048755.
Chapter 12
Mites as Models

In 2010 for the 50th Anniversary issue of Acarologia, the first scientific journal
devoted to the study of mites, Walter and Proctor (yes, that’s us) compared the use
of spiders and mites in the scientific literature. Although there were 2–3 times as
many citations for mites overall, when only a select subset of high profile journals
with broad readerships were searched, spiders came out ahead in every single one,
usually by at least twice as many papers (Walter & Proctor 2010). Ecology, genetics
and agriculture were the dominant topics in papers involving mites: behaviour, mor-
phology and materials science dominated in those involving spiders. In the latter,
the structure of spider webs was the main theme, and there were no papers at all
involving mites or mite silk. Even ticks were poorly represented with most papers
being devoted to the disease-causing microbes they vector and little on the biology
of the ticks themselves. Many of the high-profile journal articles on spiders high-
lighted fascinating aspects of their behaviour and morphology: courtship behaviour,
male ornamentation, male and female genitalic extravagances; maternal care and
social behaviour; predatory behaviour and web structure. If you’ve read the preced-
ing chapters of this book, you know that mites have equally fascinating behaviours
and morphologies, but you would never know that from reading the pages of Nature,
Science, Ecology, Evolution, and the Proceedings of national societies.
To us this seems strange, because although mites are intriguing animals worthy
of study in their own right, they are also excellent models for addressing questions
of more general interest to ecologists and evolutionary biologists (Belliure et al.
2010) (Fig. 12.1). Advantages of using mites as model organisms include their
small size and short generation times (making them good laboratory animals), mul-
tiple independent evolutions of structures and behaviours (making them excellent
test cases for evolutionary theories) and their diversity of interactions with other
organisms. With the exceptions of active flight and maintaining complex societies,
there are few areas of animal behaviour where mites have not ventured. In this final
chapter, we review some of the many theoretical areas in ecology, evolution and
behaviour for which we recommend elucidation through the liberal application of
acarines.

D.E. Walter and H.C. Proctor, Mites: Ecology, Evolution & Behaviour: Life at a Microscale, 461
DOI 10.1007/978-94-007-7164-2_12, © Springer Science+Business Media Dordrecht 2013
462 12 Mites as Models

Fig. 12.1 A model mite

Theoretical and Applied Population Ecology

We start our prospectus of mites as models with population ecology because


55 years ago, Carl Huffaker (1958) proved our point with a few trays of oranges,
some toothpicks, vaseline and mites. Studies of enemy-free space, predator–prey
interactions and the population dynamics of plant-inhabiting mites continue to be a
major emphasis of research. More recently, the work of Maurice Sabelis, Marcel Dicke
and their colleagues has been at the forefront of exploring the complex world of
chemical interactions among plants, their herbivores and their predators (see Chap. 8).
But with rare exceptions (e.g. Lesna et al. 1996), the population dynamics of mites
that do not inhabit the leaves of agricultural plants remain unexplored. We find this
surprising because many currently contentious hypotheses could be put to the test
using mites in simple experimental units.

Microcosms

Given fears of an impending sixth major mass extinction (Barnosky et al. 2011), the
hottest of all ecological topics must be ‘what is biodiversity good for?’. Are most
species redundant, performing the same ‘service’ as dozens of other species, or does
each additional taxon improve ‘ecosystem functioning’? Perhaps the only way to
directly test the value of biodiversity is through the deliberate creation of assem-
blages that vary in species richness and composition. Such experiments are most
readily accomplished in isolated, climate-controlled facilities. A now famous
example is that of Naeem et al. (1994), who used the Ecotron to show that plant
species diversity could directly influence ecosystem productivity.
Although the use of small experimental units (microcosms) that mimic larger
natural units is often questioned, there is no other approach that offers as much
Moss Islands 463

potential for discovery. When considering organisms to use in a microcosm study,


what better inhabitants could one desire than mites? They fill all trophic levels
above primary producers and seem to exhibit a profligate and excessive diversity,
particularly in soil systems (see Chap. 6). And from a pragmatic point of view, the
minute size of mites means that what is a microcosm to a plant is a mesocosm to
them, rendering potentially confounding effects of container size (e.g. Carpenter
1996) less of a concern. It seems likely that mites and other soil microarthropods
have valuable roles in maintaining ecosystem productivity. Oribatid mites have
been shown to enhance the recovery of soil communities after severe disturbances
by dispersing microbial spores (Maraun et al. 1998). Richard Fortey remarks in
his Life: An Unauthorised Biography, ‘But alas for the world if the mites and
their diminutive allies failed to prosper!’ (1997, p. 171). To make this more than
rhetorical one must compare the functioning of microcosms without mites to
those with them.

Moss Islands

The ‘bryosphere’ is a term coined by Lindo and Gonzalez (2010) for the complex
of living and dead mosses, liverworts and their associated organisms that form a
layer of interactions in many terrestrial and freshwater ecosystems. Bryophytes
are not just common, but are also experimentally tractable: it is relatively easy to
take standardized cores of living mosses and manipulate them. The fauna associ-
ated with mosses include a diversity of invertebrates, with mites dominating
the microarthropods and ranging in body mass by at least three orders of magnitude
(0.8–814.84 μg wet mass in Lindo et al. 2012). Thus, even though mites are
small, the cover enough of a range in size to be affected by physical parameters
differently. Oribatid mites are lumbering detritivores and grazers on the micro-
flora and some microfauna (e.g. tardigrades, nematodes) with long generation
times and weak dispersal abilities. These actions are most likely neutral to ben-
eficial (reduced fungal growth, increase nutrient recycling) to the mosses.
In addition, they function as mutualists in being ‘pollinators’ for the mosses
(see Chap. 6). Mesostigmatans are highly vagile predators or (more rarely) fun-
givores with short generation times. Prostigmatans in moss have the most diverse
roles, ranging from herbivores to fungivores to predators to parasitoids of arthro-
pods. Some are highly resistant to desiccation and relatively vagile, but others
are more limited in their dispersal abilities. In general prostigmatans have rela-
tively short generation times.
The diversity of functional roles and dispersal abilities of the mites in the bryo-
sphere make them ideal for testing for effects of habitat fragmentation and manip-
ulations on food webs. The mites also display a broad range of life history
strategies and reproductive modes, further enhancing their value for testing the
effects of environmental changes such as habitat fragmentation on biological
diversity (e.g. Chisholm et al. 2011).
464 12 Mites as Models

Biomonitoring

A final point about mite populations is that they often display high densities and
diversities in small volumes of soil and aquatic substrates, and so are well suited for
ecological assessment and biomonitoring studies. Although soil ecologists have long
included mites in their enumerations of soil fauna, determination of species-specific
habitat preferences is just beginning (e.g. van Straalen and Verhoef 1992; van Straalen
1997; Ruf and Beck 2005). In contrast, most freshwater ecologists ignore mites com-
pletely or lump them together with ostracods, nematodes and miscellaneous small
creatures as ‘others’. This is not a wise move, for there is growing evidence from
European studies that water mites are sensitive bioindicators (see Chap. 7), and pop-
ulation densities of some species may serve as bioassays for particular chemicals or
disturbances. Additionally, life-history characteristics of soil mites are proving useful
as indices of disturbance, with r-selected species dominating on disturbed soils and
K-selected species in stable habitats (Ruf 1998).
The high diversity of mites in natural soil systems, and the strong effects of
disturbance on many species, including effects of invasive species (St. John et al.
2006; Cameron et al. 2012, 2013), suggest that acarine diversity per se may be
used as an indicator of soil health. Impediments to this use include the difficulty
in identifying some mites, patchy distribution in the soil, and the rapid genera-
tion times of some taxa. Boom-and-bust population cycles complicate sampling
regimes and cloud interpretation of data. Oribatid mites, however, have a number
of life history characteristics that make them potentially very useful indicators of
soil health. Adult oribatids are relatively long-lived and tend to have stable adult
populations. They occur wherever organic litter accumulates and achieve high
diversities in habitats with persistent litter layers such as forests. Adult oribatids
are easily sorted into operational taxonomic units, manuals for species-level
identification have been produced for some faunas (for example, see Walter et al.
2013) and taxonomic specialists are less rare than for other mites that are not
considered pests. Perhaps the only major complication to their use as bioindica-
tors is that the life histories are not well studied, especially feeding biologies of
immature stages.

Transgenic Releases

Whether we fear or applaud genetically engineered life, genetically modified crops


are a reality and transgenic organisms will one day be a tool in Integrated Pest
Management. Microbes are the easiest organisms to engineer and mass rear for
release but exploring their fate in the field is far from easy. Determining their effects
on natural systems is even more difficult and there are justifiable worries about
unforeseen effects. Many mites, however, rival microbes for ease of rearing and are
far simpler to monitor. Additionally, many species of predatory phytoseiid mites
have been introduced repeatedly into countries around the world with no apparent
Sexual Selection and Diversification 465

ill effects to agricultural or natural environments. We suggest that these mites would
be excellent models for testing the fate of released transgenics (see Chap. 8). They
are far more vagile than microbes, and hence a better test of a released organism,
but they can be monitored using simple sticky traps or trap plants infested with
preferred prey. Safety, ease of rearing, ease of monitoring and importance all argue
for the use of phytoseiid mites as models for transgenic releases.

The Evolution of Host Specificity and Virulence

Although one might have the gut feeling that symbionts have their evolutionary
origins as generalists and over time become more host-specific, there have been few
tests of this hypothesis. The acarine symbionts of mammals and birds are especially
suited for such studies because they range from generalists to specialists (see Chap. 9).
Extant Dermanyssoidea, for example, run the gamut of possible associations, from
nest-inhabiting predators to obligatory internal parasites (Radovsky 1985, 1994).
Another appropriate model may be gall mites in the Eriophyoidea (Chap. 8). Some
major agricultural pests are found in this superfamily but the vast majority appear to
be host-specific symbionts that cause little or no loss in fitness to their host plants.
At the intraspecific level, experimental manipulation of lab populations of ectoparasitic
Macrocheles subbadius and their host Drosophila nigrospiracula have yielded
insights about maintenance of genetic variation for resistance in the host flies (Luong
and Polak 2007).
Factors that promote the evolution of virulence (the degree of damage caused
by a parasitic organism) are a hot topic given current fears of ‘emergent diseases’
and the sad history of AIDS (Garrett 1994). How many more ‘Lyme diseases’ are
lurking behind the chelicerae of ticks and other biting mites? How important are
neutral symbionts (e.g. phoretic mites) in preventing colonisation by antagonists?
Are novel parasites invariably more harmful? Does a long history of association
between parasite and host guarantee low virulence? Do parasites that are transmit-
ted vertically (from parents to offspring as in many ticks and chiggers) produce
more benign effects than those transmitted horizontally (between unrelated mem-
bers of the same age cohort) (Clayton and Tompkins 1995; Heylen and Matthysen
2011)? These are gaps in parasitological theory (Ebert and Herre 1996; Sober and
Wilson 1998) that could be narrowed through studies of interactions between
mites and their hosts.

Sexual Selection and Diversification

Sexual selection is thought to be a driver of both micro- and macroevolutionary


change. At the microevolutionary end of the scale, Gross and Repka (1998) used
data on fighter and non-fighter morphs of acarid mites (Radwan 1995) to test their
466 12 Mites as Models

model of how alternative male tactics can coexist within populations. Holland and
Rice (1998) suggest that intersexual conflict results in cycles of male ‘attack’ and
female ‘defence’; the shifting copulatory pores in the Dermanyssina (see Chap. 5)
would seem to support their belief. On the macroevolutionary scale, a common but
seldom tested belief among evolutionary biologists is that the intensity of sexual
interactions in a taxon is positively associated with speciation rate (e.g. Kaneshiro
and Boake 1987; Carson 1997). Mites show such an enormous range of contact
between the sexes, from intimate to none at all, that they are ideal subjects for sexual
selection studies at both evolutionary levels.

Sex-Ratio Control and the Devolution of Sex

In certain situations, production of the ‘normal’ 50:50 sex ratio is a waste of eggs
for mothers, for example, when ones sons are likely to have no mates other than
their own sisters and hence would be competing with each other for access to
matings (Local Mate Competition, LMC). Haplodiploid mites are ideal models to
test predictions of LMC theories both across species and within species in
experimental-evolution trials (Macke et al. 2011). In fact, why bother with pro-
ducing males at all? In contrast to the diversity of mating behaviour in some
groups of mites, many others have abandoned sex altogether (see Chap. 5). We
would be hardly going out on a limb in saying that sex has been lost far more than
100 times in the Acari. These repeated losses of sexuality and its potential resur-
gence in the Astigmata (see Chap. 5, Immaculate Conception: Did Sexual
Astigmatans Arise from Asexual Oribatids?) are reflected in numerous closely
related pairs of sexual and asexual taxa. This is an ideal situation for tests of
hypotheses about the costs and benefits of sex. Does sex confer a competitive
advantage in complex environments? Does asexuality provide an advantage for
colonising species? Can asexual species maintain genetic diversity, and if so, how?
The roles of holokinetic chromosomes and inverted meiosis are only two of many
unexplored questions in the evolution of acarine asexuality.
Another unanswered question relates to the influence of the endosymbiotic
Wolbachia and Cardinium groups of alpha proteobacteria on skewed sex ratios and
thelytoky in mites. To date, their effects on host reproduction have been studied
primarily in insects where they cause post-zygotic reproductive incompatibility and
all-female parthenogenesis (O’Neill et al. 1997). Mites, however, are comparable to
the insects in diversity of reproductive systems and many independent origins of
thelytokous parthenogenesis have been demonstrated in various acarine lineages
(Norton et al. 1993). Many examined species have been found to host Cardinium
(Nakamura et al. 2009). Studies of specific mite systems may also eventually help
to clarify the population dynamics of these bacterial infections: for example, if there
are fitness costs associated with various strains, whether multiple infections spread
at the cost of single infections and if Wolbachia-Cardinium dynamics can lead to
speciation events or evolution of asexuality (Weeks et al. 2001).
Selection at More Than One Level 467

Pushing the Limits of Physiology and Morphology

Two of the major themes in mite evolution have been to press developmental rates
and body size to their limits. Extraordinary rates of population increase dependent
on rapid development, mating systems that assure short generation times and high
reproductive output have evolved again and again within the Acari. This wealth of
analogous systems is perfect for testing hypotheses and is matched by the relative
ease of culturing many of these taxa. Recent advances in understanding the devel-
opmental genetics of mites (e.g. Telford and Thomas 1998) suggest that the size of
mites is no impediment to their use as models for understanding much larger
organisms.
All mites are small in an absolute sense but both the largest and the smallest
size classes of Acari appear to be less diverse than intermediate size classes. This
is a general pattern in terrestrial animals (May 1978, 1988) but a confusing one.
Why are the largest mites so small? Why has not at least one lineage of giant
mites evolved? Conversely, what limits the evolution of small body size? Mites
less than 100 μm in length as adults ( ~ 50 μm as larvae) are known but seem to
be very rare. Mites smaller than half this length appear to be impossible — but
why? Is miniaturization of bodies limited by a minimum brain size? Certainly,
mites have very high ratios of brain:biomass, higher than for almost all mammals
(Eberhard and Wcislo 2012).
One intriguing question that combines these evolutionary patterns is that of the
apparent dissociation between body size and developmental time in some
Mesostigmata (see Chap. 4). Given that theory to explain constraints on develop-
ment in most animals appears to be progressing rapidly, this failure of mites to
conform is especially interesting.

Selection at More Than One Level

As well-indoctrinated individual selectionists, many of us become alarmed at the


faintest whiff of group selectionist thinking (Borrello 2010). This hypersensitivity
sometimes prevents us from contemplating how selection can act at different levels
of organisation (Sober and Wilson l998). Mites are often members of tiny commu-
nities demarcated by their location on the bodies of larger organisms (e.g. mites,
fungi and nematodes on burying beetles; mites, fungi, lichen, small insects on
leaves). Their fates are linked with that of their bearers, conditions ripe for selection
for genes within species that support the functioning of a community of species as
an evolutionary unit (e.g. Wilson and Knollenberg 1987; Sober and Wilson 1998).
As well, subdivided populations founded by single females provide opportunities
for group selection for female-biased sex ratios (e.g. Wilson and Colwell 1981).
When moving from theoretical models of multi-level selection to natural systems,
mites probably offer more opportunity for empirical tests than any other taxon.
468 12 Mites as Models

Summary

The average mite is a tiny animal, scarcely visible to the unaided human eye.
Because they are literally beneath our notice, we are often able to forget mites – unless
they cause problems that cannot be ignored. Waiting for a pest outbreak or a new
disease to take mites seriously, however, is a serious mistake. In fields as current,
and as different, as group selection theory, genetic engineering technology and the
effects of climate change mites have proven useful experimental models. We hope
that some readers will realise the potential of mites as experimental organisms in
their fields and take advantage of them. In these lives lived at a microscale there is
much left to discover.

References

Barnosky, A. D., Matzke, N., Tomiya, S., Wogan, G. O. U., Swartz, B., Quental, T. B., Marshall, C.,
McGuire, J. L., Lindsey, E. L., Maguire, K. C., Mersey, B., & Ferrer, E. A. (2011). Has the
Earth’s sixth mass extinction already arrived? Nature, 471, 51–57.
Belliure, B., Montserrat, M., & Magalhães, S. (2010). Mites as models for experimental evolution
studies. Acarologia, 50, 513–529.
Borrello, M. E. (2010). Evolutionary restraints: The contentious history of group selection.
Chicago: University of Chicago Press.
Cameron, E. K., Knysh, K. M., Proctor, H. C., & Bayne, E. M. (2012). Influence of two exotic
earthworm species with different strategies on abundance and composition of boreal microar-
thropods. Soil Biology and Biochemistry, 57, 334–340.
Cameron, E. K., Proctor, H. C., & Bayne, E. M. (2013). Effects of an ecosystem engineer on
belowground movement of microarthropods. PLoS ONE, 8(4), e62796. doi:10.1371/journal.
pone.0062796.
Carpenter, S. R. (1996). Microcosm experiments have limited relevance for community and
ecosystem ecology. Ecology, 77, 677–680.
Carson, H. L. (1997). Sexual selection: A driver of genetic change in Hawaiian Drosophila.
Journal of Heredity, 88, 343–352.
Chisholm, C., Lindo, Z., & Gonzalez, A. (2011). Metacommunity diversity depends on connectivity
and patch arrangement in heterogeneous habitat networks. Ecography, 34, 415–424.
Clayton, D. H., & Tompkins, D. M. (1995). Comparative effects of mites and lice on the reproduc-
tive success of rock doves (Columba livia). Parasitology, 110, 195–206.
Eberhard, W. G., & Wcislo, W. T. (2012). Plenty of room at the bottom? Tiny animals solve
problems of housing and maintaining oversized brains, shedding new light on nervous-system
evolution. American Scientist, 100, 226–233.
Ebert, D., & Herre, E. A. (1996). The evolution of parasitic diseases. Parasitology Today, 12,
96–101.
Fortey, R. (1997). Life: An unauthorised biography. London: Harper Collins.
Garrett, L. (1994). The coming plague, newly emerging diseases in a world out of balance. Victoria:
Penguin.
Gross, M. R., & Repka, J. (1998). Stability with inheritance in the conditional strategy. Journal of
Theoretical Biology, 192, 445–453.
Heylen, D. J. A., & Matthysen, E. (2011). Differential virulence in two congeneric ticks infesting
songbird nestlings. Parasitology, 138, 1011–1021.
References 469

Holland, B., & Rice, W. R. (1998). Perspective: Chase-away sexual selection: Antagonistic seduction
versus resistance. Evolution, 52, 1–7.
Huffaker, C. B. (1958). Experimental studies on predation: Dispersion factors and predator–prey
oscillations. Hilgardia, 27, 343–383.
Kaneshiro, K. Y., & Boake, C. R. B. (1987). Sexual selection and speciation: Issues raised by
Hawaiian Drosophila. Trends in Ecology & Evolution, 7, 207–212.
Lesna, I., Sabelis, M. W., & Conijin, C. G. M. (1996). Biological control of the bulb mite,
Rhizoglyphus robini, by the predatory mite, Hypoaspis aculeifer, on lillies: Predator–prey
interactions at various spatial scales. Journal of Applied Ecology, 33, 369–376.
Lindo, Z., & Gonzalez, A. (2010). The bryosphere: An integral and influential component of the
earth’s biosphere. Ecosystems, 13, 612–627.
Lindo, Z., Whiteley, J., & Gonzalez, A. (2012). Traits explain community disassembly and trophic
contraction following experimental environmental change. Global Change Biology, 18,
2448–2457.
Luong, L. T., & Polak, M. (2007). Environment-dependent trade-offs between ectoparasite
resistance and larval competitive ability in the Drosophila–Macrocheles system. Heredity,
99, 632–640.
Macke, E. S., Magalhães, F. B., & Olivieri, I. (2011). Experimental evolution of reduced sex ratio
adjustment under local mate competition. Science, 334, 1127–1129.
Maraun, M., Visser, S., & Scheu, S. (1998). Oribatid mites enhance the recovery of the microbial
community after a strong disturbance. Applied Soil Ecology, 9, 175–181.
May, R. M. (1978). Diversity of insect faunas. In L. A. Mound & N. Waloff (Eds.), The dynamics
and diversity of insect faunas (pp. 188–204). Oxford: Blackwell Scientific Publications.
May, R. M. (1988). How many species are there on earth? Science, 241, 1441–1449.
Naeem, S., Thompson, L. J., Lawler, S. P., Lawton, J. H., & Woodfin, R. M. (1994). Declining
biodiversity can alter ecosystem performance. Nature, 368, 734–737.
Nakamura, Y., Kawai, S., Yukuhiro, F., Ito, S., Gotoh, T., Kisimoto, R., Yanase, T., Matsumoto, Y.,
Kageyama, D., & Noda, H. (2009). Prevalence of Cardinium bacteria in planthoppers and
spider mites and taxonomic revision of “Candidatus Cardinium hertigii” based on detection
of a new Cardinium group from biting midges. Applied and Environmental Microbiology,
75, 6757–6763.
Norton, R. A., Kethley, J. B., Johnston, D. E., & OConnor, B. M. (1993). Phylogenetic perspec-
tives on genetic systems and reproductive modes of mites. In D. L. Wrensch & M. A. Ebbert
(Eds.), Evolution and diversity of sex ratio in insects and mites (pp. 8–99). New York:
Chapman & Hall.
O’Neill, S. L., Hoffmann, A. A., & Werren, J. H. (1997). Influential passengers inherited microor-
ganisms and arthropod reproduction. New York: Oxford University Press.
Radovsky, F. J. (1985). Evolution of mammalian mesostigmate mites. In K. C. Kim (Ed.), Coevolution
of parasitic arthropods and mammals (pp. 441–568). New York: Wiley-Interscience.
Radovsky, F. J. (1994). The evolution of parasitism and the distribution of some dermanyssoid
mites (Mesostigmata) on vertebrate hosts. In M. A. Houck (Ed.), Mites, ecological and evolu-
tionary analyses of life-history patterns (pp. 186–217). New York: Chapman & Hall.
Radwan, J. (1995). Male morph determination in two species of acarid mites. Heredity, 74,
669–673.
Ruf, A. (1998). A maturity index for predatory soil mites (Mesostigmata: Gamasina) as an indicator
of environmental impacts of pollution on forest soils. Applied Soil Ecology, 9, 447–452.
Ruf, A., & Beck, L. (2005). The use of predatory soil mites in ecological soil classification and
assessment concepts, with perspectives for oribatid mites. Ecotoxicology and Environmental
Safety, 62, 290–299.
Sober, E., & Wilson, D. S. (1998). Unto others, the evolution and psychology of unselfish behaviour.
Cambridge: Harvard University Press.
St. John, M. G., Wall, D. H., & Hunt, H. W. (2006). Are soil mite assemblages structured by the
identity of native and invasive alien grasses? Ecology, 87, 1314–1324.
470 12 Mites as Models

Telford, M. J., & Thomas, R. H. (1998). Expression of homeobox genes shows chelicerate arthropods
retain their deutocerebral segment. Proceedings of the National Academy of Sciences USA,
95, 10671–10675.
van Straalen, N. M. (1997). Community structure of soil arthropods as a bioindicator of soil health.
In C. Pankhurst, B. M. Doube, & V. V. S. R. Gupta (Eds.), Biological indicators of soil health.
Wallingford: CAB International.
van Straalen, M., & Verhoef, H. A. (1992). The development of a bioindicator system for soil acidity
based on arthropod pH preferences. Journal of Applied Ecology, 34, 217–232.
Walter, D. E., & Proctor, H. C. (2010). Mites as modern models: Acarology in the 21st century.
Acarologia, 50, 131–141. doi:10.1051/acarologia/20101955.
Walter, D. E., Latonas, S., & Byers, K. (2013). Almanac of Alberta Oribatida. Part 1. Ver. 2.3.
Edmonton: The Royal Alberta Museum. http://www.royalalbertamuseum.ca/natural/insects/
research/research.htm.
Weeks, A. R., Marec, F., & Breeuwer, J. A. J. (2001). A mite species that consists entirely of
haploid females. Science, 292, 2479–2482.
Wilson, D. S., & Colwell, R. K. (1981). The evolution of sex ratio in structured demes. Evolution,
35, 882–897.
Wilson, D. S., & Knollenberg, W. G. (1987). Adaptive indirect effects: The fitness of burying
beetles with and without their phoretic mites. Evolutionary Ecology, 1, 139–159.
Index

A Aculops allotrichus, 123


Abacarus hystrix, 323 Aculus fockeui, 123
Aboveground, 161, 163, 167, 183, 209, Adactylidium, 367, 407
320, 321 Adamystidae, 59, 123, 187, 200, 298
Åbro, A., 364, 395 Adanal suckers, 119, 130
Abyssal habitat, 231 Adaptationist explanation, 305
Acanthorhynchus tenuirostris, 296 Adar, E., 303
Acaridae, 110, 118–120, 130, 133, 136, 233, Addington, R.N., 170
234, 236, 244, 254, 315, 354 Adelges tsugae, 318
Acariformes, 2, 5, 27, 28, 30, 34, 35, 39, Adelomyrmex, 204
41–43, 50–55, 57, 65, 74, 75, 77, 79, Adhaesozetes polyphyllos, 313, 314
80, 97, 98, 116, 150, 187, 284, Adlerocystis, 126
353–355, 382, 439, 455 Adoxaceae, 307
Acarinaria, 343, 370–373 Aedeagus, 43, 108, 118–120, 141, 150, 256
Acarinomorpha, 32 Aeschlimann, A., 395
Acarodomatium, Acarodomatia, 304–306, 308 Africa, 32, 43, 46, 83, 87, 204, 236, 292,
Acarology Summer Program, 39 297, 317
Acaromorpha, 32 Afrocypholaelaps, 297
Acaronemus, 316 Afrolinotus, 292
Acarophagy, 199, 202–206 Agaonidae, 295
Acarophenacidae, 367, 454, 455 Agistemus
Acarophenax, 82, 407 A. exsertus, 299
A. mahunkai, 88 A. longisetus, 311
Acarophobia, 423 A. olivi, 300
Acarus, 170, 351, 374 Alasaad, S., 405
A. siro, 120 Alaskozetes antarcticus, 171
Accumulation curves, 450, 451 Alberta, 28, 162, 170, 210, 242, 268, 285,
Aceria 294, 450, 456, 457
A. chondrillae, 324 Alberti, G., 40, 110–112, 114, 116
A. lantanae, 324 Alces alces, 204
Acevedo, H.A., 374 Algae, 27, 52, 84, 105, 171, 172, 174–176,
Achipteriidae, 317 178, 180, 182–185, 232–234, 236, 239,
Aclerogamasus stenocornis, 27 243, 253, 283, 284, 287, 288, 298, 309,
Acremonium zonatum, 293 315, 317, 318, 387
Actinopilin, 52 Algophagidae, 233, 234, 236, 244, 254
Actinotrichida, 27, 32, 39, 40, 52 Algophagus pennsylvanicus, 244

D.E. Walter and H.C. Proctor, Mites: Ecology, Evolution & Behaviour: Life at a Microscale, 471
DOI 10.1007/978-94-007-7164-2, © Springer Science+Business Media Dordrecht 2013
472 Index

Alicorhagia fragilis, 59, 78 Antennophorus, 360


Alicorhagiidae, 59 Antennoseius, 358
Alismobates inexpectatus, 234 A. janus, 194, 197, 357, 359
Allergens, 375, 424 Anthocoridae, 141, 440
Allochaetophora, 292 Antricola
Allochaetophoridae, 292 A. delacruzi, 349
Allodynerus delphinalis, 373 A. marginatus, 71
Allometry, 91 Antrodiaetus pacificus, 19, 22
Allothrombium Anystidae, 59, 133, 187, 190, 193, 282, 298,
A. lerouxi, 120 431, 439
A. pulvinum, 357, 397 Anystides, 41
Allothyridae, 45, 46, 48, 198 Anystina, 59, 78, 122, 187, 259
Allothyrus, 45, 144, 198, 199 Apoderm, 77, 80, 81
Alycosmesis corallium, 87 Apoidea, 405, 454
Alycus roseus, 77 Apomorphies, 15
Amber, 8, 27, 28, 118, 282, 283, 299, 355 Aporrectodea, 210
Amblypygida, 23, 80 Aposematism, Aposematic, 206, 258
Amblyseius, 114, 302 Apotriophtydeus, 172
A. swirskii, 303 Arachnid, 6, 11, 21–35, 40, 47, 61, 69, 79, 89,
Ambush predator, 25, 41, 63, 188, 193, 194, 110, 230, 241, 243
200, 298, 299 origin of, 22–25
Amensalism, 342, 345 Arachnida, 5, 11, 20–22, 24, 29, 34, 39,
Ameronothridae, 233, 234, 236, 317 40, 110
Ameroseiidae, 128, 139, 296, 297, 318, 319 Archaeacarus dubinini, 26
Amphibia, 108 Archegozetes longisetosus, 71, 79
Amphipoda, 22, 24, 180, 353 Archispermy, 111
Amphitoky, 144 Arctic, 164
Amplifying host, 426 Argasidae, 43, 44, 76, 110, 111, 126, 343,
Amrine, J.W. Jr., 285 349, 439
Anactinotrichida, 27, 32, 39, 40 Argyroneta aquatica, 230, 256
Analges, 383, 405, 406 Aribates javensis, 353, 361
Analgidae, 383, 386, 405 Aribatidae, 353, 361
Analgoidea, 439 Arlian, L.G., 375, 379, 395
Anamorphosis, 80 Armascirus, 191
Ancistrocerus antilope, 371, 393 Armoured mite, 62, 201–204, 206, 283
André, H.M., 81 Army ants, 343, 361
Androlaelaps Arnold, E.N., 388, 389
A. casalis, 429 Arnold, S.J., 137
Androlaelaps (=Gromphadorholaelaps) Arrenuroidea, 235, 363
schaeferi, 360 Arrenurus
Andropolymorphism, 133 A. angustilimbatus, 350
Anecic Group, 210 A. fissicornis, 363
Anisitsiellidae, 236, 248 A. manubriator, 139, 149, 270
Anisogamy, 105 A. rufopyriformis, 350
Anobium punctatum, 432 Arrhenotokous parthenogenesis, 74
Anoplocephalid tapeworm, 204 Arrhenotoky, 144, 147, 148, 151, 180
Anoplopalpus, 292 Arthropod-borne disease, 424, 426
Antarctica, 87, 164, 170–177, 288, 365 Arthropod-caused disease, 424
Antarcticola, 172 Asca
Antarctic Peninsula, 171, 172, 176 A. aphidioides, 146
Antarctozetes, 172 A. foliata, 300
Antennolaelaps, 51 Ascidae, 49, 79, 90, 186, 187, 233, 241, 300,
A. aremenae, 115 344, 357
Antennophorina, 42, 112, 113 Ascouracaridae, 385
Index 473

Asia, 46, 297, 434 Bee, 54, 297, 344, 345, 349, 367, 368, 370,
Astacopsiphagus parasiticus, 237, 253 390, 402, 407, 408, 450, 454, 455
Asteraeceae, 323 Beetle, 6, 27, 41, 70, 147, 164, 229, 281, 342,
Astigmata, 41, 52, 53, 74, 80, 81, 89, 91, 94, 431, 448, 461
95, 97, 110, 118–120, 129, 130, 133, Behan-Pelletier, M., 232
136, 141, 144, 149–151, 170, 178, Beklemishevia, 200
180, 199, 202, 230, 233, 234, 236, Belgica antarctica, 171
241, 243–244, 260, 263, 315, 318, Belier, R., 246
320, 346, 348, 352, 354, 357–359, Belowground, 161, 163, 164, 166, 167, 182,
371, 374–376, 378, 381, 382, 387, 183, 321
405, 439, 455, 456, 466 Benson, W.W., 308
Asymmetry, 120, 133, 395, 400, 401, 408 Bergman, D., 391, 402
Athias-Binche, F., 348, 349, 355, 359, 360, Berlese funnel, 162, 168
372, 406, 407 Biesiadka, E., 250
Athiasella dentata, 115 Big bud mite, 285
Athiasiella, 50 Bimorph, 133, 134
Aturidae, 54, 251 Binns, E.S., 358, 359
Atyeo, W.T., 133, 343, 381, 382, 384, 385, Bioindicator, 209, 266, 447, 464
395, 404 Biomonitoring, 271, 464
Audycoptidae, 377, 404 Biramous, 18
Australia, 7, 22, 45, 46, 87, 88, 97, 236, 237, Birds, 5, 41, 83, 108, 170, 260, 296, 343, 424,
253, 271, 287, 291, 297, 300, 304, 308, 447, 465
310, 311, 314, 318, 323, 324, 431, 433, Birefringent setae, 40
434, 437, 438, 451–453, 458 Birobates hepaticolus, 317
Australian Corroboree Frog, 205 Bitumen paint, 309, 311
Austromesocypris, 24 Bivalve, 22, 82, 132, 251
Autapomorphy, 17, 118 Black flies, 356, 363, 365, 395, 396, 439
Autotomy, 199–202 Blanco, G., 384
Axtell, R.C., 359 Blattisociidae, 58, 79, 128, 186, 232, 237, 296,
300, 318, 319, 348, 353, 359
Blattisocius tarsalis, 348, 359
B Blister mite, 285
Babesiosis, 437 Blue oat mite, 287, 288
Bader, C., 83 Bochkov, A.V., 346, 405
Bagge, P., 270 Body size, 72, 73, 84, 87, 88, 91–92, 97, 98,
Baker, G.T., 233 107, 130, 136, 137, 165, 177, 205–208,
Baker, R.A., 242, 252, 344 250, 392, 447, 452, 453, 467
Baker, R.L., 396, 397 Bohonak, A.J., 351
Baker’s itch, 439 Bolus, Boli, 61, 62, 179
Balaustium, 144, 299, 349 Bonn, A., 395, 401
B. medicagoense, 284 Bonomoia opuntiae, 77
Bark beetles, 96, 147, 368 Book lungs, 19, 21, 25, 26, 29–31, 241
Barklouse, 194 Booth, J.P., 365
Barr, D.W., 110, 264, 265 Borrelia
Bartlett, J., 1 B. burgdorferi, 436
Bartsch, I., 233, 237, 253, 343 B. miyamotoi, 436
Bathyhalacarus quadricornis, 240 Böttger, K., 260, 264, 265, 351, 396
Batrachoceps attenuates, 205 Boulton, A., 269
Bats, 43, 71, 91, 297, 343, 344, 349, 365, 366, Bovidromus, 124
373, 374, 376, 377 Boyt, A.D., 396
Bauer, A.M., 388, 389 Brachychthoniidae, 87
Bdellidae, 54, 58, 59, 70, 122, 187, 193, Bracket fungi, 318, 319
231, 298 Brackish water habitat, 237
Bedbugs, 108, 141, 381 Braddy, S.J., 107
474 Index

Bradyagaue, 240, 253 Capsicum annuum, 307


Brenddandania lambi, 359 Carabodidae, 317
Brevipalpus phoenicis, 144 Carapace, 5, 6, 18, 19, 25, 42, 200, 247
Brevisterna utahensis, 377 Carbonifereous, 25, 26, 28, 34, 179
Bright colour (of mites), 206 Cardinium, 146, 466
Brightness (of host colouration), 258 Carinozetes bermudensis, 234
Broadcast spawning, 107 Carotenoid(s), 62, 259, 260, 287, 298, 301
Broad mite, 41, 88, 130, 303, 316 Carotenoid pigments, 259, 287, 298
Brooding, 71, 107 Carpenter bees, 370
Brooks, D.R., 402 Carpoglyphidae, 354
Brossard, M., 391 Carpoglyphus, 374
Brown, W.L., 230, 233 Carrier, 70, 94–98, 197, 345, 355, 369,
Bruce, W.A., 367 394, 424
Bruguiera gymnorhiza, 296 phoretic, 94, 197, 297
Bruyndonckx, N., 405 Cassis, G., 142
Bryobia, 183 Celaenopsoidea, 58, 111, 113
Bryophyte, 172, 184, 284, 287, 317, 463 Centipedes, 42, 202, 229, 253
Bryosphere, 184, 463 Central America, 46
Bryum argenteum, 184 Cephalothorax, 19
Bud mite, 55, 285 Ceratopogoindae, 264, 396, 437, 439
Bulmer, M.G., 147 Ceratotarsonemus, 317
Buphagus, 204 Cercomegistidae, 318, 353
Burgess Shale, 15 Cercomegistina, 42
Burley, N., 399 Cerotegument, 62–64, 200, 203, 313
Bursa copulatrix, 118 Chaetodactylidae, 354, 405
Burtt, E.H. Jr., 379, 392 Chant, D.A., 85
Burying beetles, 369, 395, 467 Chapman, B.R., 395, 396
Butterflies, 7, 297, 299, 365 Chapman, R.F., 141
Butterwort, 313 Character state, 17, 29, 30, 50, 63
Charletonia, 282, 299
Chasmataspidida, 20
C Chaussieria venustissima, 78
Cactoblastis, 323 Cheese mite, 439, 442, 456
Caeculidae, 187, 191, 200, 248 Cheiroseius, 232, 236, 243
Caenorhabditis elegans, 91, 145 Chelate-dentate, 55–57, 64–66, 283, 302, 378
Calcium compounds, 52 Chelate palps, 244
California, 177, 205, 207, 292, 321, 322 Cheletomorpha lepidopterorum, 380
Callaghan, C., 233 Chelicera, Chelicerae, 3, 5, 19, 20, 22, 25,
Calyptostase, 77, 82, 84 27–29, 32–34, 41, 43, 45, 46, 48, 50,
Cambrian, 20 52, 55–57, 59–61, 64–66, 71, 72, 77,
Cambrian explosion, 11–15 78, 111–113, 115, 126, 131, 132, 134,
Camel spider, 23, 34 142, 180–182, 186, 188, 191–194, 202,
Camerobiidae, 28, 59 203, 243, 244, 247–249, 253, 283, 284,
Cameron, E.K., 210 290, 302, 303, 347, 353, 358, 363, 364,
Caminella peraphora, 111, 232, 237 373, 377, 378, 380, 386, 449, 465
Camin, J.H., 343 Chelicerata, 3, 5, 11, 15–19, 24, 30, 80
Camisia biverrucata, 53 Cheyletidae, 54, 59, 71, 123, 130, 133, 282,
Camisiidae, 231, 233, 238, 283, 317 293, 298, 299, 312, 343, 354, 380, 432
Campbell, K.U., 390 Cheyletiella
Campodeid, 202 C. blakei, 432
Canestriniidae, 354 C. parasitivorax, 432
Canis latrans, 426 C.yasguri, 432
Capitulum, 3, 4, 27, 29, 32, 48, 66 Cheyletiellosis, 432, 433
Index 475

Cheyletinae, 187 Colwell, R.K., 135, 147, 344, 396


Cheyletus eruditus, 71 Commensalism, 342, 345, 360–361
Chigger, 41, 54, 55, 75, 260, 362, 364, 381, Comminution, 61, 179, 180
387–389, 402, 433–436, 442, 443, Competition, 106, 120, 128–138, 143, 147,
456, 465 149–151, 163, 177, 178, 194, 207,
Chigger-borne Rickettsiosis, 434 252, 342, 345, 372, 394, 396, 466
Chihuahuan desert, 187 Complementarity, 450–452
Chile, 323 Complete dissociation, 109, 150
Chilean Predatory Mite, 301–304, 323 Compton, G.L., 111
Chilli Pepper, 307 Coniopterygidae, 202
Chilli Thrips, 303 Conroy, J.C., 268
Chilochorus, 96 Convergence, 17, 20, 32, 236
Chipmunk, 426, 427 Cook, D.R., 233, 237–239, 268, 362, 363
Chirodiscidae, 129, 378 Cook, W.J., 72
Chironomidae, 403 Copepod, 23, 24, 121, 139, 182, 245–248,
Chiroptera, 405 250, 253, 259, 263, 346
Chirorhynchobiidae, 376 Copepoda, 23
Chitons, 253 Cope’s rule, 26
Chloride epithelia, 254 Copidognathus, 253
Choe, J.C., 386 C. alvinus, 240
Chondrilla juncea, 323 Coprophagy, 179, 349
Chromotydaeus, 185 Copulation, 107, 108, 116, 119, 122, 126, 130,
Chrysomelobia labidomerae, 402 136, 138, 140–142, 150, 252, 256, 295,
Cianciolo, B., 146 394, 395, 400
Cicolani, B., 270 Copulatory pore, 109, 115, 118, 466
Cimicidae, 141, 439 Coqui, 206
Cinnamomum camphora, 308 Corniculus, Corniculi, 45, 51, 56–58, 303
Cladocera, Cladocerans, 245, 247, 248, 250, Correa, 296
251, 262, 269 Cosmochthonius, 200, 201, 375
Cladogram, 16, 17, 406, 407 Cospeciation, 402–409
Claparède’s organs, 42 Costa Rica, 308, 310
Clark, L., 392, 395, 396 Costa Rican Strawberry Poison Frog, 205
Clausiadinychus, 49 Couplet, 63
Clayton, D.H., 387, 392, 396, 402 Cowan, D.P., 343, 371, 393
Cleptotoxin, 206 Coxa, Coxae, 18, 29, 40, 42, 43, 47, 48, 50,
Cloacaridae, 343, 381, 388 61, 64–66, 72, 111, 115, 118, 254, 437
Cloacas, 381, 448 Coyote, 426
Clutch-coat (of water mite eggs), 249 Crabs, 3, 20, 21, 23, 25, 30, 32, 56, 61, 107,
Coadaptation, 401–403, 408 231, 240, 353, 355
Cocaine bugs, 442 Crayfish, 24, 233, 237, 251, 253, 254,
Coccinellid beetle, 320, 323, 355 354, 363
Coccoidea, 308 Cretaceous, 8, 27, 28, 35, 283
Coccorhagidia, 172 Creutzeria, 130, 263
Cockroaches, 354, 369, 404, 456 Cronberg, N., 184
Coevolution, 341, 389, 401–403 Cross, E.A., 358
Coffea arabica, 308 Crotoniidae, 283, 317
Coleoptera, 262, 271, 362, 405, 449 Crotonioidea, 315
Coleoscirus simplex, 190, 196 Crowell, R.M., 351
Collembola, 109, 117, 139, 170 Crustacean, 3, 5, 15, 22–24, 82, 118, 164, 180,
Collins, N. C., 350 242, 244–248, 263, 408
Collohmannia gigantea, 117, 150 Crustose lichen, 317
Collohmanniidae, 117 Crustose scabies, 428, 429
Colobocerasides, 240 Cryptic female choice, 138, 140
476 Index

Cryptobiosis, 187 Demodex


Cryptocarya, 307 D. brevis, 343, 377, 443
Cryptocellus, 23, 47 D. follicularis, 377
Cryptognathidae, 54, 293 D. folliculorum, 343, 377, 443
Cryptopygus antarcticus, 173, 174 Demodicidae, 87, 343, 376, 377, 406,
Cryptostigmata, 52, 241 409, 443
Cryptozoic, 168 Dendrobates pumilio, 205
Cuckoos, 384 Dendrobatidae, 206
Cucullus, 32 Dendrolaelaps
Culicidae, 393 D. fossilis, 27
Cummins, K.W., 178 D. neodisetus, 368
Cunaxa, 188, 191 Dendroptus, 316
Cunaxidae, 54, 58, 59, 187, 188, 193, 245, Dermacentor
298, 299, 312 D. albipictus, 204
Cunaxinae, 191, 298, 299 D. andersoni, 436
Cunliffe, F., 393 Dermanyssiae, 27, 42, 50, 57, 58, 73, 94, 98,
Cupania vernalis, 308 111–113, 115, 142, 357, 455
Cupp, J.K., 257 Dermanyssidae, 133, 343, 360, 379, 380, 392,
Currier, R.W., 426 406, 429, 430, 439
Cybaeidae, 230 Dermanyssoidea, 48, 73, 429–431, 439, 465
Cymbaeremaeoidea, 315 Dermanyssus
Cyphophthalmi, 31 D. gallinae, 379, 380, 396, 430
Cyr, H., 268 D. hirundinis, 379, 380, 399
Cyta latirostris, 193, 202 Dermatitis, 343, 425, 427, 430, 432, 433, 443
Czenspinskia, 315 Dermatophagoides
C. transversostriata, 315 D. evansi, 290, 322, 380
D. farinae, 53, 375, 424
D. pteronyssinus, 375
D Dermicoles, 382
Dabert, M., 150 Desmonomata, 150
Dactyloscirus, 191 Desmonomatides, 41
Daidalotarsonemus, 317 Detritivore, 8, 27, 52, 162, 163, 169, 178–181,
Damaeidae, 71, 317 211, 320, 463
Damaeus verticillipes, 71 Detritus, 24, 48, 62, 63, 70, 162, 164,
Damselflies, 72, 391, 393, 395–397, 400, 171–173, 176, 179, 182, 184, 200,
401, 404 243, 244, 282, 290, 315, 316, 375,
Daphnia, 245–249 377, 380, 387
Darolova, A., 399 Deuterocerebrum, 15
Darwin, C., 106, 128 Deuteroky, 144
Dauer larva, 186, 197 Deutonymph, 27, 30, 41, 42, 59, 64, 74–76,
Davids, C., 132, 246–250, 260, 363, 365 81–83, 94–98, 129, 131, 136, 182, 239,
Davies, D.M., 397 251, 252, 292, 301, 320, 347, 348, 351,
Davis, M.D.P., 322 352, 354–359, 369, 371, 372, 380, 390,
Dead-end host, 424, 431 394, 450
Decapoda, 24 Developmental rate, 84–86, 88–92, 98, 467
Deer, 136, 169, 426, 427, 437 Devonacarus sellnicki, 26
Deer mice, 424, 427 Devonian, 20, 25–27, 33, 34, 41, 56, 199
Deer Tick Virus, 437 Dew-worm, 210
de Jong, D., 396 Diabetes mellitus, 442
Delfinado-Baker, M., 402 Diapterobates, 318
Del Giudice, P., 432 D. humeralis, 318
Delusions of parasitosis, 438 Dicamaralges dasybrachius, 135
Delusory parasitosis (DP), 440, 442, 443 Dichlorodiphenyltrichloroethane (DDT), 300
Index 477

Dichotomous key, 39, 62, 63 Eastern spinebills, 296


Dichotomy, 15–18, 186, 196, 423 Ebbert, M.A., 148
Dicrocheles Eberhard, W.G., 124, 138–140
D. phalaenodectes, 365, 366 Ecdysis, 11, 363
D. scedastes, 366 Echinolaelaps echidninus, 377
Digamasellidae, 79, 94, 133, 282, 318, 357 Eciton, 361
Digestion, 3, 61–62, 66, 179, 205, 243, E.burchellii, 361
247, 253 Ectoparasite, 124, 253, 392
Dimock, R.V., 132, 252 Ectospermatophore, 115, 125, 126
Dinogamasus braunsi, 370, 371 Ectosymbiosis, 96
Dinothrombium Edwards, D.D., 132, 139
D. magnificum, 259 Edwardzetes, 172
D. tinctorium, 86 Eggs, 16, 69, 105, 171, 230, 282, 344, 426,
Diphyletic, 30, 288 456, 466
Diplodiploid, 144, 149, 180, 197 Egg size, 72–74
Diplodontus, 239 Ehrlich, P.R., 401
D. semiperforatus, 254 Ehrlichiosis, 437
Dipluran, 202 Eichhornia crassipes, 293
Diptilomiopidae, 284 Eickwort, G.C., 360, 367, 368, 395, 402,
Direct sperm transfer, 50, 108, 110, 116, 403, 408
118, 120, 122, 123, 126, 128, 150, Ejaculatory complex, 110, 116
252, 256, 257 Ekbom’s Syndrome, 441, 443
Di Sabatino, A., 270 Elaeocarpaceae, 309
Discozerconidae, 113, 353 Elaeocarpus angustifolius, 306, 309
Disease, 3, 35, 44, 143, 176, 185, 244, 286, Eleutherengonides, 41
292, 295–297, 312, 317, 320, 324, 341, Eleutherodactylus coqui, 206
368, 378, 395, 423–443, 461, 465, 466 Ellatostase, 78, 79
Dispersal, 59, 60, 62, 81, 92–98, 147, 169, Ellis-Adam, A.C., 246–248, 365
182, 210, 269, 350–352, 354–360, 367, Elton, C.S., 249, 257, 260
385, 394, 436, 449, 450, 463 Elzinga, R.J., 343
Dissociation (sperm transfer), 108–109 Emerging disease, 426
Dockovdia cookarum, 251 Enarthronota, 26
Dolichocybe, 82 Enarthronotides, 41, 199
Domatia blocking, 309 Encentridophorus similis, 250
Dometorina, 317 Enchytraeid, 183
Domrow, R., 343, 395 Endeostigmata, 27, 41, 52, 53, 59, 77, 79–81,
Donnelly, M.A., 206, 260 144, 172, 182, 187, 198, 450, 455, 456
Downing, J.A., 268 Endofollicular hypopi/hypopodes, 348
Drosophila nigrospiracula, 465 Endogeic Group, 210
Dubinin, V.B., 120, 384 Endogynum, 112
Dubininia melopsittaci, 383 Endospermatophore, 115, 116, 120, 125, 126
Duckweed, 293, 294 Engorgement, 43, 44, 350, 351, 397, 398
Duffy, D.C., 395 Enicocephalidae, 202
Dufour, K.W., 401 Enigma of soil biodiversity, 161–163
Dunlop, J.A., 33, 40, 107 Entomogenous nematode, 197
Durden, L.A., 380, 426 Entomophilous hypopi/hypopodes, 358
Dust mite, 53, 375, 380 Environmental indicators, 208–211, 266–271
Environmental irritants, 423, 432, 440, 442
Eotarbus jerami, 25
E Eotetranychus, 93
Eamer, B., 232 E. willamettei, 322
Earth mite, 287, 288 Ephemeroptera, 271, 362
Earthworm, 210 Epicorticolous, 317
478 Index

Epicriidae, 60, 111 Exuviae, 12, 74, 176, 182, 200, 364, 375, 397
Epicroseius, 50, 319 Eylaidae, 121, 236, 239, 245, 246, 249, 254,
Epidermoptidae, 135, 385, 405 264, 265, 350, 351
Epigeic Group, 210 Eylaioidea, 235, 363
Epimyodicidae, 381 Eylais
Epiphyll, 309, 317 E.euryhalina, 391
Epiphyte, 165, 282, 315, 317, 382 E.infundibulifera, 264
Epistome, 61
Eremaeidae, 317
Ereynetes, 172, 357, 395 F
E. macquariensis, 172, 176 Facilitation, 178, 342
Erginulus clavotibialis, 77 Fahrenholz’s Rule, 402
Eriksson, M.O.G., 257, 262, 270 Fain, A., 343, 344, 346, 378, 395, 404
Erineum, Erinea, 283 Falcolichus, 404
Erinose mite, 55, 281, 284–287 Falculiferidae, 134
Eriophyidae, 54, 59, 123, 130, 131, 284, 294 Falculifer rostratus, 134
Eriophyioidea, 282 Farish, D.J., 359
Eriorhynchidae, 185, 287, 288 Fashing, N.J., 130, 233, 244, 263
Eriorhynchus, 185, 288 Fatal flatulence, 192
E.walteri, 287, 288, 319 Feather(s), 84, 120, 148, 172, 381–387, 392,
Ernst, H., 230 393, 400, 408, 449, 454
Erogalumna zeucta, 117 Feather mite, 53, 55, 73, 106, 118–120, 129,
Erwin, T., 6, 281 133–135, 381–387, 400, 403, 404, 406,
Erythraeoidea, 187, 193 409, 439, 454
Eschar, 434, 435 Feeding guild, 177–190
Estimates, 451 Felso, B., 404
Euchelicerata, 15, 20, 31 Female choice, 128, 138, 140, 150, 401
Euepicrius filamentosus, 115 Feynman, R.P., 1
Eueremaeus, 317 Fig(s), 294–295, 316, 319, 402, 437
Eumeninae (Vespidae), 371 Fig mite, 294–295
Eupalopsellidae, 188 Filter-feeding microbivore, 182, 211
Eupelops, 181 Fish, 5, 6, 12, 106, 107, 138, 140, 233, 234,
Eupodes, 172, 198, 287 244, 249–251, 257–260, 262, 270, 272,
Eupodidae, 59, 181, 199, 287 363, 387, 397, 398, 452
Eupodides, 41 Fisher, D., 20
Eupodina, 59, 79, 122, 456 Fisher, G.R., 252
Eupodoidea, 284, 287 Fisher, J. R., 184, 193
Euproops danae, 20 Flat mites, 55
Euroglyphus maynei, 375 Florida, 196, 309
Europe, 46, 317, 323, 324, 362, 387, 433, 441 Flowers, M.A., 260
Eurypterid, 20, 24, 25, 32, 107, 240 Fluctuating asymmetry, 400
Eurypterida, 4, 12, 20–24 Fluid-feeding, 33, 48, 56
Euseius, 303 Foliose lichen, 317
E. elinae, 311 Follicle mite, 439, 443
Eustigmaeus frigida, 294 Food chain, 173, 175
Eutrombicula, 433 Food web, 171–176, 180, 194, 211, 212, 250,
Evans, G.O., 109, 120 379, 463
Eviphididae, 187, 353 Footballer mite, 298
Excretion, 61–62, 66, 394 Foraging, 93, 205, 290, 299, 321, 390
Exoskeleton, 11, 41, 62, 234 Forbes, M.R., 392, 396
External fertilisation, 106, 107 Forelia, 268
Extraguild predation, 178 Formicaria, 304
Index 479

Formication, 423 Gehypochthonius, 53


Formicidae, 353 Generation time, 8, 69, 74, 85, 86, 88–91, 98,
Fortey, R., 463 171, 172, 180, 182, 283, 289, 300, 310,
Fortuynia atlantica, 234 318, 324, 430, 435, 461, 463, 464, 467
Fortuyniidae, 234 Genital acetabulum/acetabula, 235, 240,
Fossil, 4, 8, 11–35, 41, 42, 46, 50, 52, 56, 118, 254, 255
199, 211, 282, 283, 286, 299, 355 Genital opening, 43, 50, 60, 65, 72, 108–115,
Fowl mite, 35, 49, 380, 391, 400, 429, 430 117, 118, 121–127, 136, 141, 142,
Free-spawning, 106 151, 230
Freshwater sponges, 82, 267 Genital papillae, 19, 30, 42, 47, 50, 51, 74,
Freundlieb, U., 269 254, 255
Functional group, 162, 175, 177–190, 210, George, J.E., 395, 936
212, 283, 456 Gerecke, R., 233, 237, 248, 254, 266, 270
Fungi, 24, 27, 46, 48, 52, 57, 162, 163, 166, Gil, M.J., 267
167, 171, 172, 175, 176, 179–185, 187, Gilboa Shales, 26
189, 198, 204, 211, 283, 287, 288, 293, Glacier, 8, 164, 202, 210
298, 309, 312, 313, 315–320, 344, 345, Glandularium, Glandularia, 235, 257, 260, 261
356, 375, 384, 387, 457, 467 Gledhill, T., 363, 365
Fungitarsonemus, 316 Gliwicz, Z.M., 250
Fur mite, 55, 377, 378, 439 Glomus mosseae, 167
Furniture mite, 439 Glycyphagidae, 118, 354
Fürstenberg, M.H.F., 28 Glycyphagoidea, 439
Fuscouropoda marginata, 397 Glyptholaspis, 166
Gnathobase, 23, 61
Gnathosoma, 3, 27, 32, 40, 45, 48, 51, 55–59,
G 64–66, 72, 77, 87, 111, 112, 120, 194,
Gabucinia delibata, 384 203, 249, 289, 293, 303
Gabuciniidae, 384 Gnathosomal tectum, 48
Gaeolaelaps aculeifer, 183 Gondwana, 45, 455
Galendromus occidentalis, 87, 149, 303, Gonyleptoidea, 69
311, 312 Gonzalez, A., 463
Gall mite, 54, 55, 122, 127, 286, 316, 324, Good genes hypothesis, 399
325, 451, 463 Gordialycus, 168
Galumnidae, 117, 150, 203, 317 Gorse, 323
Gamasellodes, 79, 196 Gotwald, W.H. Jr., 344, 361
G. vermivorax, 89 Gould, S.J., 84, 147, 305, 389
Gamasellus Grandjean, François, 30, 34, 78, 79
G. racovitzai, 172–174, 195 Grant, J.W.A., 257
G. traghardi, 115 Grape, 110, 308, 309, 311, 321, 322
Gamasina, 42, 50, 51, 57 Grape Powdery Mildew, 309
Gamete(s), 3, 60, 105–107, 295 Grasses, 183, 204, 285, 288, 292, 295, 323
Gannon, J.E., 257, 267 Gratz, N.G., 426
Garga, N., 130, 258 Graves, B.M., 260
Garrett, 426 Greenhouse Whitefly, 303
Gas exchange, 3, 18, 20, 30, 50, 232, 234, Greenland, 211
240–243 Green Mountains, 87, 88, 452, 453
Gastronyssidae, 343, 344, 376, 377 Grocer’s itch, 439
Gastropods, 253 Grooming by hosts, 43, 436
Gaud, J., 133, 381, 382, 384, 385, 404 Gross, M.R., 465
Gaudiellidae, 354 Grostal, P., 311, 312
Geckos, 389 Grylloblatta campodeiformis, 202
Geholaspis, 166 Grylloblattodea, 202
480 Index

Guano, bat, 71, 164, 349 Heteroptera, 254, 308


Guglielmone, A.A, 43 Heterostigmata, 73, 74, 81, 82, 95, 98, 182
Guild, 88, 173, 177, 178, 194 Heterozerconidae, 42, 58, 73, 113, 141, 353
Hexapod, 15, 28, 32, 40, 45, 60, 66, 74, 75,
77, 80–82, 84, 97, 108
H Hibbertopteroidea, 20
Haemaphysalis longicornis, 144 Hinton, H.E., 241, 242
Haemocoelic insemination, 141 Hippoboscid flies, 385
Haemogamasus, 374 Hirstia, 375
H. ambulans, 377 Hirstiella pyriformis, 388, 395
Haemolaelaps glasgowi, 377 Histiostoma
Hagvår, E.B., 170 H. laboratorium, 358
Hagvår, S., 170 H. polypori, 359
Halacarellus basteri, 122, 237 Histiostomatidae, 95, 128, 129, 233, 234, 236,
Halacaridae, 122, 253–255, 343 244, 263, 352, 354, 355, 358, 359, 368
Halacaroidea, 233, 234, 237, 252, 263, Hodgkin, L.A., 96
271, 354 Hoffmann, D., 167
Halik, R.N., 264 Holland, B., 466
Halixodes chitonis, 253 Holocelaeno, 166
Haller’s organ, 43, 44, 49 Holocentric chromosomes, 145
Halliday, R.B., 7 Holokinetic chromosomes, 466
Halotydeus, 288 Holostaspella, 166
H. destructor, 70, 288 Holothyran, 41, 45, 48, 81, 85, 87, 111,
Hamilton, W.D., 147, 148, 399 200, 232
Hamilton-Zuk hypothesis, 399 Holothyrida, 41, 43, 45, 46, 57, 65, 74, 76, 77,
Hammen, L., 30, 32 81, 110, 144, 150, 198, 241, 352, 455
Hammeriellidae, 317 Holothyridae, 45, 46
Haplodiploid, 144, 149, 372, 466 Holotrichy, 85
Haplozetidae, 297 Homeomorph, 42, 74, 120, 133, 134, 358
Harems, 128, 131, 139 Homocaligidae, 54, 231, 293
Harpirhynchidae, 59 Homology, 16, 32, 74, 307
Harvest mite, 433 Homoptera, 320
Harvey, M.S., 233, 253 Honeydew, 298, 303, 312, 315, 316
Hattena, 296 Hoogstraal, H., 233, 388
H. floricola, 296 Hoploseius tenuis, 319
H. panopla, 296 Horlick Mountains, 87
H. rhizophorae, 296 Hormosianoetus mallotae, 244
Haug, J.T., 15 Horseshoe crab, 3, 19–21, 23, 30, 32, 56, 61,
Hauswirth, J.W., 377 69, 107, 240
Hawaii, 310, 323 Host, 8, 41, 72, 111, 178, 231, 281, 341, 424,
Hay-itch mites, 367 454, 465
Head, arthropod, 3, 18, 26–27 Host-finding, host-seeking, 350
Heavy metal, 209 Hosts, choice of by mites, 389–390
Hedlund, K., 139 Host-switching, 404, 405, 407, 409
Hemicheyletia morii, 71 Houck, M.A., 96, 348, 355, 368
Hemipterans, 262, 360, 366, 456 House Mouse Mite, 425, 430
Hemisarcoptes Hrbacek, J., 262
H. cooremani, 96, 348, 355, 380 Huck, K., 389, 390
H. malus, 320 Huffaker, C.B., 462
Hemisarcoptidae, 320, 348, 355 Huitfeldtia, 268
Hermanniidae, 150 Human follicle mite, 429, 443
Heteromorph, 41, 42, 74, 75, 81, 95, 120, 133, Human itch mite, 378, 428–429, 439, 443
134, 182, 348, 358 Humerobates, 318
Index 481

Humerobatidae, 317 Hypodermic insemination, 108, 141


Hummingbird flower mite, 296 Hypopode, 74
Hummingbirds, 135, 296, 297, 344, 365, Hypopus, Hypopi, 62, 74, 81, 87, 94–96, 178,
394, 396 198, 343, 348, 354, 357–360
Hunter, P.E., 353, 357, 367, 395 Hypostomal process, 43, 65
Hurst, G.D.D., 395 Hypotrichous, 85
Hutchinson, G.E., 55, 206, 207 Hysterosoma, 52
Hutchinsonian size differences, 207 Hystrichonyssidae, 376
Hybalicus, 78
Hydracarina, 60, 82, 116, 120–125, 149,
233–237, 244, 259, 263, 268, 271, I
349, 362 Idiosoma, 3, 40, 45, 49, 52, 65, 66, 73, 87,
Hydrachna, 61, 127, 239, 242, 245, 246, 249, 293, 453
260, 268, 364 Illusions of parasitosis, 440
H. conjecta, 246, 250, 391, 398 Immune system, 391, 399, 423, 424, 428, 432,
H. virella, 398 433, 436, 443
Hydrachnellae, 235 Incomplete dissociation (sperm transfer), 108
Hydrachnidiae, 234 Indirect sperm transfer, 127, 146
Hydrachnoidea, 235, 268 Indotritia, 198
Hydrodroma despiciens, 127, 246, 250, 264, Infochemicals, 321
350, 393 Infracapitulum, 32
Hydrodromidae, 127, 230, 244–246, 250 Inquilines, 361
Hydrogamasellus antarcticus, 172, 176 Instar, 28, 29, 74–79, 81, 89, 94, 171, 189,
Hydrogamasus salinus, 195 292, 362, 380, 434
Hydrovolzioidea, 235, 363 Intermediate host, 180, 201, 204, 381, 439
Hydrozetes, 183, 242, 243, 265, 266, 293, 294 Internal fertilisation, 107
Hydrozetidae, 232, 233, 236, 265, 293 Intersexual conflict, 108, 128, 140–142,
Hydrozoans, 240, 253 151, 466
Hydryphantes tenuabilis, 242 Intersexual selection, 128, 138–142
Hydryphantidae, 236, 242, 244, 245, 254, 259, Interstitial, 31, 231, 233, 235, 238–239, 254,
265, 351, 364 264, 265, 268, 269, 377
Hydryphantoidea, 235, 363 Intertidal, 41, 48, 55, 171, 195, 229, 230, 234,
Hygrobates 235, 239, 242, 315
H. aloisii, 252 Intraguild predation, 178, 194–196, 303, 312, 426
H. fluviatilis, 250, 270 Intrasexual competition, 106, 120, 128–138
H. longipalpis, 250 Intrinsic rate (of population growth), 74
H. nigromaculatus, 249, 264 Invertebrates as hosts, 362, 408
H. trigonicus, 249 Iphieus, 303
Hygrobatidae, 249, 250, 252, 264, 270, 351, Iphiseius degenerans, 297, 312
364 Iponemus, 368
Hygrobatoidea, 235, 363 Ips grandicollis, 97
Hyland, K.W. Jr., 343, 346, 378, 395, 404 Isobactrus uniscutatus, 237
Hylocomium splendens, 184 Isogamy, 105
Hymenoptera, 109, 290, 295, 353, 354, Isopoda, 22, 24, 180, 229, 264
371–373, 405 Itch-scratch cycle, 440–442
Hyperdiversity, 456 Iteroparity, 72, 180
Hyperkeratosis, 379 Itu, F., 361
Hypertrichy, 48 Iverson, K., 394, 395
Hypertrophied setae, 46, 78, 120, 200 Ixodes, 111, 390, 399, 426, 427, 453
Hypoaspis, 188, 190 I. dammini, 136
Hypoaspis (= Gaeolaelaps) aculeifer, 139, 183 I. holocyclus, 44
Hypodectes propus, 343, 380 I. ricinus, 399
Hypoderatidae, 380 I. scapularis, 436
482 Index

Ixodida, 41, 43, 44, 46, 65, 74, 110, 111, Laelapidae, 49, 79, 115, 128, 133, 187, 188,
115, 142, 233, 352, 388, 406, 439, 231, 318, 320, 343, 360, 365–366, 369,
442, 453, 455 374, 429, 430, 439
Ixodidae, 44, 76, 111, 232, 233, 390, 426, Laelaps echidninus, 430
434, 439 Laird, M., 250
Lamington national park, 87, 88, 452, 453
Laminisoptidae, 380
J Lanciani, C.A., 396, 397
J1 larva, 76–79 Land, colonisation of, 25–26, 30
Johnstoniana Lantana, 49, 324
J. errans, 127 Lardoglyphidae, 394
J. rapax, 349 Lardoglyphus
Johnstonianidae, 127, 349, 356 L. konoi, 120
Jujeremus foveolatus, 283 L. zacheri, 394, 395
Larvacarus, 60
Larva, Larvae, 28, 40, 70, 130, 173, 229, 285,
K 342, 426, 450, 467
Kairomone, 321 Larvamimidae, 343, 353
Kaliszewski, M., 148, 344, 358, 360, 367 Larviparity, 73, 240
Kampimodromus, 303 Larviposition, 73
K. aberrans, 303, 304 Lasioseius
Kano, Y., 107 L. dentatus, 196
Karban, R., 321 L. porulosus, 300
Keeley, E.R., 257 L. subterraneus, 196, 197
Kennethiella trisetosa, 343, 371, 393 Lasius, 360
Kerfoot, W.C., 258 Lauraceae, 307, 308
Kethley, J., 79, 148, 386 Lead (in key), 63, 93, 111, 116, 151, 199, 211,
Kim, K.C., 344, 373, 376, 386, 388, 395, 289, 349, 350, 403, 466
403, 406 Leaf domatia, 304–312, 316, 325
King, P.E., 195 Leaf hair, 299, 304, 315, 316, 321
Kirchner, W.P., 122 Leaf structure, 231, 293, 305–307
Kitching, R.L., 233 Lebertia
Kiwialges, 386 L. porosa, 250
Kleptobionts, 447 L. quinquemaculosa, 242
Kleptoparasite, 360, 362 Lebertiidae, 242
Klimov, P.B., 347, 349, 402, 405 Lebertioidea, 235, 363
Klompen, J.S.H., 388 Ledermuelleria, 293
Knee, W., 405, 406 Leeches, 108, 118, 141, 233, 234, 244, 346,
Knemidokoptidae, 380 355, 387
Knollenberg,W.G., 369, 373 Lee, D.C., 115
Knülle, W., 360 Leeuwenhoekiidae, 354, 387, 431
Krantz, G.W., 109, 111, 233, 237, 242, Lemna, 183, 293, 294
343, 344 Lentic, 242, 257, 265, 266, 269
Lenticulus, 236, 314
Lepidoglyphus destructor, 360
L Lepidoptera, 348, 353, 354, 365
Labidostomma lutea, 110 Lepidosaphes ulmi, 320
Labidostommatidae, 79, 122, 187, 200, Leptokoenenia, 30
202, 284 Leptotrombidium, 432, 433
Labidostommatidae, 79, 122, 187, 200, L. akamushi, 432
202, 284 L. deliense, 432
Labium, 22, 42, 47 Lesna, I., 139, 380, 431
Labrzycka, A., 378 Levitation, 263, 265–266
Index 483

Lewontin, R.C., 305, 346, 387 Lycosidae, 230


Liacaridae, 317 Lyme disease, 341, 426, 427, 436, 437, 465
Lice, 191, 346, 375, 383–385, 392, 399, 402,
404, 426, 428, 437
Lichenivore, 184 M
Lichens, 52, 163, 167, 171, 172, 174, 183, MacArthur, R., 206
184, 209, 283, 287, 288, 309, 314, 315, MacQuitty, M., 253
317–318, 448, 465 Macrocheles
Lifespan, 84, 91, 141, 149, 250, 345, 350, M. muscaedomesticae, 89, 136, 359, 394
396, 397 M. rettenmeyeri, 344, 353
Light, 15, 40, 42, 52, 93, 107, 168, 193, 211, M. schaeferi, 71, 360
236, 253, 260, 265, 266, 286, 287, 294, M. subbadius, 395, 401, 463
298, 313, 348, 366, 429 M. superbus, 191
Limbs, jointed, 3, 12, 55 Macrochelidae, 79, 89, 113, 129, 133, 136,
Limnesia 166, 187, 191, 318, 344, 353, 359, 394
L. fulgida, 249 Macrolaspis, 166
L. jamurensis, 250 Macronyssidae, 49, 58, 349, 376, 380, 388,
L. maculata, 242, 269 392, 393, 399, 401, 429, 439
Limnesiidae, 118, 127, 242, 245, 249, 250, Macrophytes, 233, 236, 264, 268
261, 265, 269, 351 Magellozetes, 172
Limnochares, 83, 84, 127, 248, 249, 258, 264, Maiorana, V.C., 205
265, 270, 396 Malaconothridae, 234, 387
L. aquatica, 83, 84, 248, 249, 270, 396 Malacostraca, 22, 24
Limnocharidae, 248, 259, 270 Malpighian tubules, 42
Limnohalacarus wackeri, 253 Mammals, 5–7, 41–43, 48, 69, 83, 87, 108,
Limnozetidae, 234 131, 142, 161, 170, 180, 182, 204,
Lindo, Z., 463 341, 346, 354, 358, 373–380, 382,
Lindquist, E.E., 27, 32, 39, 368 388, 392–395, 402–409, 424, 426,
Linopenthaleus, 185 427, 429–431, 433–435, 442, 456,
Linopenthalodes, 185 465, 467
Linotetranidae, 167, 292 Mandibulata, 3, 15–18
Liochthonius, 172 Mange, 55, 378, 432, 439, 456
Liodidae, 317 Mangrove, 296, 297
Liponyssoides sanguineus, 428 Manter’s first rule, 402, 407
Litter layer, 145, 164–166, 198, 206, 210, Marine iguana, 231, 232, 388
211, 462 Marine scorpions, 12, 20, 23, 240
Liverwort, 163, 171, 184, 309, 317, 461 Markow, T.A., 395
Lizard, 41–43, 124, 205, 233, 260, 388, 389, Marquardt, W., 426
391, 394, 395, 399, 404–406, 427, 431 Martin, P., 249, 250
Lobster, 253 Masan, P., 379
Local mate competition (LMC), 147–149, Mate-guarding, 128–136, 140
151, 464 Maternal care, 69, 70, 461
Locomotion, 3, 60, 110, 262–265, 362 Maudheimia, 172
Lombert, H.A.P.M., 406 Mayr, E., 206
Lorryia formosa, 308 May, R.M., 206, 207, 452, 567
Loss of parasitism, 349–351, 363 McIver, S.B., 363, 365, 390
Loss of phoresy, 351–352, 360 McLachlan, A., 397
Lotic, 238, 242, 249, 265, 266, 269 McLennan, D.A., 402
Luciaphorus hauseri, 148 McMartney, M.A., 106
Lumbricus Mediolata, 299
L. rubellus, 210 Megacheira, 15, 18
L. terrestris, 210 Megachilidae, 370
Lundström, A.N., 304, 305, 307 Megadiverse, 450
484 Index

Megafauna, 177 Mite-caused disease, 427–431


Megaloptera, 362, 363 Mite-pockets, 343, 372, 387–389
Megapodiidae, 385 Mite-vectored disease, 324
Megaselia halterata, 190 Mixonomatides, 41
Megeremaeidae, 317 Mixopterus, 20
Megisthanidae, 73, 88 Mochlozetidae, 297, 317
Megisthanus, 50, 88, 449, 453 Modlin, R.F., 257, 267
Megninia, 383 Møller, A.P., 399–401
Melicharidae, 49, 58, 135, 296, 297, 319, 394 Molluscs, 94, 251, 397, 408
Mentum, 42, 50, 51, 56, 65 Monogynaspida, 42, 455
Meriläinen, J.J., 270 Monophyletic, 15, 17, 21, 29, 30, 52
Mesofauna, 161, 177 Monoxeny, 406
Metacheyletia, 84 Montaigne, Michel de, 1, 2
Metamerism, 12, 18–19 Mooring, M.S., 392
Metaseiulus, 322 Mor, 164–166
Metasoma, 18, 20, 21 Morelli, M., 405
Meth mites, 442 Mosquitoes, 7, 35, 72, 236, 250, 347, 351,
Meyer, E., 350 363, 365, 390, 394, 396, 397, 402,
Michalska, K., 122, 123, 127 404, 436, 439
Michener, C.D., 84 Moss, 41, 121, 138, 174, 184–185, 209, 211,
Micreremidae, 317 231, 233, 234, 237, 268, 287, 293,
Microbivore, 35, 52, 89, 175, 179, 181–182, 317–318, 346, 439, 450, 463
211, 283, 320 Moss mats, 184
Microbivore-detritivore, 162, 169, Moss mites, 41, 184, 287, 317
178–181, 211 Moss, W.W., 346
Microcosms, 184, 189, 205, 209, 462–463 Moths, 42, 180, 297, 343, 348, 365, 366,
Microdipsidae, 181 395, 432
Microdispus lambi, 182 Moulting, 11, 43, 44, 59, 76–79, 84, 131, 171,
Microfauna, 161, 177, 463 299, 309, 312, 316, 371, 382, 383, 390,
Microhabitat, 8, 62, 94, 163, 179, 211, 315, 396, 397, 401
321, 386, 448–450, 454, 458 Mounsey, K.E., 428
Microlichus, 135 Mouse-to-elephant curve, 86, 91
Micromegistus, 198 Moya Borja, G.E., 394, 395
Microtrombidiidae, 84, 356 Mt. St. Helens, 169
Microtrombidium maculatum, 84 Mucronothrus nasalis, 237, 238, 243
Mideopsidae, 255, 264 Mull, 165
Mideopsis orbicularis, 264 Mullen, B.A., 426
Midges, 72, 127, 132, 229, 351, 396, 397, 404, Müllerian mimicry, 258, 259
430, 437 Mussels (Mollusca), 61, 122, 128, 131, 132,
Migration, 92–94, 98, 169, 197, 198, 208, 211, 139, 233, 234, 251–253, 347, 354,
285, 318, 356, 357, 359 363, 397
Millipede, 25, 42, 126, 164, 165, 180, 206, Mutualism, 296, 325, 341, 342, 345, 353, 360,
229, 353 368–370, 373
Mineral soil, 87, 93, 164–166, 169, 171, Mwango, J., 250
179, 210 Mycangial pouch, 182
Minute Pirate Bug, 440 Mycetophagidae, 457
Mironov, S.V., 346, 405 Mycetophile, 318
Mitchell, R., 247, 357, 363, 365, 396 Mycobatidae, 231, 371
Mite(s) Mycolaelaps maxinae, 319
on bark, 283, 293, 315, 317, 320, 322 Mycorrhizae, 166, 167, 185
load, 97, 384, 397, 399–401 Myialges macdonaldi, 355
origin of, 11–35 Myobiidae, 376, 378, 406, 407, 409
as plant virus vectors, 286, 289 Myocoptidae, 378
word origin, 1, 2, 5, 25 Myrmecina, 204, 361
Index 485

N Neutralism, 342, 345


Naeem, S., 233, 344, 462 Newell, I.M., 265, 266, 395
Nagelkerke, K., 148 New Guinea, 310, 314, 390, 399
Najadicola ingens, 251, 397 Newt, 108, 205, 252
Nanelli, R., 317 New Zealand, 45, 243, 310, 323, 366,
Nanhermanniidae, 317 379, 384
Nanorchestes Nicolai, V., 317
N. antarcticus, 87 Nicrophorus, 359, 369
N. memelensis, 87 Nidicole, 345, 346, 360, 361, 374, 380,
Nanorchestidae, 52, 78, 87, 122 394, 396
Naso, 42, 52 Nilotonia longipora, 248
Necromenic, 359 Nishida, S., 308, 310
Necrophage/Necrophagy, 52, 178 Nitidulid beetle, 297
Nectariniidae, 297 Node, 17
Needham, G. R., 343 North America, 83, 89, 165, 188, 210, 231,
Nematalycidae, 52, 168 236, 251, 257, 266, 268, 310, 317, 318,
Nematode, 8, 52, 56, 57, 89–92, 94, 108, 141, 324, 362, 426, 427, 433, 437, 457
166, 172, 174–176, 179–182, 184–187, Norton, R.A., 116, 143–146, 149, 150, 207,
189, 190, 193–197, 200, 204, 245, 271, 233, 243, 260
318, 320, 344, 356, 359, 368–370, 447, No-see-um, 437, 439
449, 463, 464, 467 Nothrus silvestris, 209
Nematophagy, 187 Notogaster, 41, 64, 199, 201
Nemertea, 253, 354 Notophthalmus viridescens, 205
Neocaeculus, 191, 192 Notostigmata, 46
Neochitina eichhorniae, 293 Nutalliellidae, 43
Neocypholaelaps, 297 Nuttalliella namaqua, 43, 44
Neomolgus littoralis, 231 Nutting, W.B., 377, 381, 393, 402, 403
Neonidulus, 289–291
Neonidulus tereotus, 289–291
Neopodocinum, 166 O
Neoschongastia americana, 381 Oak gall, 432
Neoscirula, 191, 193, 196 Oak leaf itch mite, 432
Neoseiulus, 58, 302, 303 Ocellus, Ocelli, 13, 41, 52, 65
N. fallacies, 303 OConnor, B., 402, 405
Neoseius novus, 406 OConnor, B.M., 233, 347–349, 351, 352, 355,
Neosomy, 44, 75 368, 374, 395
Neospermy, 111 Octopod, 75, 77, 78, 80, 81, 97
Neotenogyniidae, 353 Ocular segment, 15, 19, 25
Neoteny, 84–86 Odonata, 246, 271, 404
Neothyridae, 45, 46 Odontacarus australiensis, 433
Neotrombicula autumnalis, 433 O’Dowd, D.J., 310–312
Neotropacarus, 303, 315 Oldfield, G.N., 122
Neotropics, 45, 204, 296, 297 Oligonychus, 93, 183, 290
Nepenthes, 352 O. perseae, 290
Nests, 48, 55, 71, 72, 132, 166, 171, 187, 230, Oligoxeny, 406
260, 290, 342, 345, 346, 348, 351, 354, Oliver, D.R., 362, 387
359–361, 370–372, 374–375, 379–380, Ologamasidae, 58, 63, 113, 187, 200, 300
390, 392, 393, 399, 400, 408, 425, Olomski, R., 254
429–431, 434, 439, 441, 448, 456 Omnivory, 175
Nest sanitation hypothesis, 392 Operculum, Opercula, 26
Neumania, 121, 139, 149, 247, 248, 253, 261, Ophiomegistus, 388
264, 268 Ophionyssus, 388, 391
Neuroptera, 202, 363 Opilioacaran, 5, 31, 41, 46–48, 62, 77, 199
486 Index

Opilioacaridae, 31, 46 Oxpecker, 204


Opiliones, 30–32, 80, 108, 110, 243 Oxygen, 241, 242, 251, 265, 266
Opisthonotal gland, 199
Opisthosoma, 3, 5, 6, 18–21, 25, 30, 31,
41, 46, 72, 80, 111, 117, 118, 120, P
123, 200 Pacific spider mite, 322
Oppia, 59, 172 Paedogenesis, 84, 86, 293
Oppia nitens, 59 Paedomorphosis, 84–86
Oppiella nova, 59, 145 Paine, R.T., 163, 176
Oppiidae, 317 Palaeoacarology, 211
Opuntia inermis, 323 Palaeosomata, 81
Ordovician, 20–22, 24, 25, 41 Palaeozoic, 21, 27, 35, 286
Oribatei, 52 Paleofantasy, 21
Oribatida, 27, 28, 41, 50–53, 56, 59, 63, 71, Palm, 297
79, 89, 98, 110, 116–118, 145, 149, Palmer, S.C., 143, 145, 146, 149
163, 167, 178, 181, 183, 199, 200, 202, Palpal thumb-claw complex, 244
210, 211, 230, 232–234, 236, 237, 241, Palpcoxal seta, 47, 48
243–244, 254, 260, 263, 271, 297, 314, Palpigrade, 4, 29–31, 230, 241
353, 354, 439, 450, 453, 455 Palpigradi, 29–31, 80, 110
Oribatula tibialis, 313 Panigrahi, A., 395
Oribatulidae, 283, 317 Panonychus, 93
Orientia tsutsugamushi, 434, 435 Pap, P.L., 384
Oriflammella lutulenta, 62, 63, 200 Parahaploidy, 144, 301
Oripodidae, 317 Paraisotoma, 194
Oripodoidea, 314, 315, 439 Paramegistidae, 353, 388
Orlettes, 305 Parantennulidae, 353
Ornithodoros Paraphagy, 349, 360
O. moubata, 110 Paraphyletic, 17, 52, 89, 149
O. transverses, 388 Parapygmephorus costaricanus, 368
Ornithodorus tholozani, 110 Parasitalbia sumatrensis, 251
Ornithonyssus Parasitellus, 368, 390
O. bacoti, 58, 430 P. fucorum, 390
O. bursa, 49, 392, 399, 400, 429 Parasitengonina (= Parasitengona), 41, 55, 74,
O. sylviarum, 380, 391, 393, 395, 79, 81–83, 89, 98, 116, 120, 122, 144,
396, 399, 401, 429 231, 234, 235, 244, 347, 349, 354, 357,
Orthogalumna terebrantis, 293, 294 358, 362–365, 456
Orthotydeus lambi, 309 Parasitiae, 42, 357
Orthotydeus nr. Lindquisti, 309 Parasitidae, 96, 111, 112, 114, 133, 166, 187,
Osmoregulation, 30, 240, 254–255 356, 357, 359, 368, 369, 390, 406
Ostracoda, 22 Parasitiformes, 2, 5, 27–28, 30, 32–34, 39–51,
Otopheidomenidae, 360, 366 57, 59, 64, 80, 97, 98, 111–116, 232,
Oudemansicheyla coprosomae, 71 263, 352–353, 376, 439, 450, 455
Ovaries, 109, 110, 112–114, 136, 137, 142 Parasitism, 88, 96, 98, 182, 244, 247,
Oviger, 16, 29 251–252, 254, 284, 286, 287, 342,
Oviposit, 126, 299, 359, 363, 365, 370, 379, 345–354, 361–368, 380, 389–393,
390, 394 397, 399–401, 403, 408
Oviposition, 61, 69, 70, 72, 109, 111, 126, Parasitoidism, 342, 345, 347, 353, 354,
128, 141, 171, 267, 293, 301, 303, 304, 361–368
312, 359, 365, 367, 370, 372, 380, 395 Parasitus, 96, 190, 357
Ovipositor, 56, 61, 69, 72, 110, 295, 361 P. coleoptratorum, 96
Ovolarviparity, 70 P. consanguineous, 190
Ovoviviparity, 73 Parastacidae, 24
Owls, 381, 405 Paratydeidae, 187
Index 487

Paredes-León, R., 405 Phoresy, 92–98, 166, 180, 182, 236, 345, 346,
Parhalixodes travei, 253 348–349, 351–353, 355–360, 380, 394,
Parhypochthonius aphidinus, 51 403, 408
Parhypsomatides, 41 Phoretic, 24, 27, 41, 42, 57, 60, 64, 84, 94–98,
Parker, G.A., 106, 136, 137 166, 169, 180, 197, 294–297, 316,
Parrots, 84, 395, 454 318, 320, 345, 348, 349, 351–361,
Parthenogenesis, 60, 74, 105, 116, 142–146, 365, 367–369, 371, 390, 394–397,
150, 196, 288, 466 406, 449, 465
Partitioning (of hosts), 386 Phoretomorph, 95, 357, 358, 360
Passalidae, 353, 408, 449 Phoriont, 95, 96
Passalid beetle, 97, 198, 360, 449 Photoprotection, 259, 260
Passive female choice, 139 Phototaxis, 265
Paternal care, 69 Phthiracaridae, 317
Paterson, C.G., 249 Phthiracarus borealis, 53
Pathogen, 313, 315, 317, 325, 402, 403, Phyllocoptes fructiphilus, 324
423–427, 434–437, 440, 442 Phylogentic tracking, 403, 404, 406, 407
Pathogen pollution, 437 Physical gill, 241, 242
Paulus, H.F., 32 Physogastry, 73, 85, 367, 368, 456
Peacock mite, 292 Physolimnesia australis, 118
Pediculaster flechtmanni, 358, 359 Phytotelmata, 41, 55, 231, 234–236
Pediculaster-Siteroptes dyad, 358 Phytonemus pallidus, 144
Pedipalp, 3, 19, 22, 25, 32–34, 40, 42, 50, 56, Phytoptidae, 284
61, 64–66, 397 Phytoscutus, 303
Pedofossa, Pedofossae, 200 Phytoseiidae, 27, 48, 49, 58, 73, 85, 89, 114,
Pedrocortesellidae, 317 128, 144, 148, 196, 232, 282, 299–303,
Peloppiidae, 317 312, 455
Pengilley, R.K., 205 Phytoseiulus
Pennak, R.W., 260, 270 P. longipes, 89, 290
Pentachlorophenol (PCP), 209 P. persimilis, 167, 301, 302, 321, 323
Penthaleidae, 70, 185, 287–288 Phytoseius oreillyi, 49, 302
Penthaleus, 185, 288 Phytotelm, Phytotelmata, 41, 55, 231,
P. major, 288 234–236
Penthalodes, 185, 288 Pica hudsonia, 204
Penthalodidae, 185, 200, 287, 288 Pieczynski, E., 257, 264, 266
Pérez, T.M., 395 Piercing-sucking microbivore, 181–182,
Peritreme, 45, 64–66, 241, 242, 289 211, 283
Peromyscus, 426 Pilicolous hypopi/hypopodes, 357, 358
Pesticide contamination, 209 Pilogalumna, 117, 118
Peza, 237, 253 Pimeliaphilus podapolipophagus, 393
Peza daps, 237, 253 Pinguicula, 313
Pezidae, 237, 253 Pinus radiata, 207, 208
pH, 209, 270 Piona
Pharate, 81, 85, 123, 128, 130, 131 P. alpicola, 246, 247, 365
Pharynx, 42, 48, 58, 61, 303, 364 P. coccinea, 249
Phasomkusolsil, S., 435 P. conglobata, 264
Phaulodinychus P. constricta, 247, 250, 258, 264, 265
P. miti, 232, 242 P. longipalpis, 249
P. repleta, 242 P. nodata, 260, 264, 265, 351
Pheidole, 204 Pionidae, 122, 236, 245, 246, 249–251, 260,
Phenopelopidae, 317 262–265, 267, 351
Pheromones, 111, 116, 127, 131, 138, Pisauridae, 230
198, 199 Pitcher plants, 55, 130, 231–233, 235, 236,
Philip, C.B., 267 263, 352
488 Index

Phylloplane, 297, 298, 300, 304, 313, 315, Powassan virus, 437
316, 325 Prasad, R.S., 250
Plankton, 8, 77, 97, 233, 246, 250, 263, Prasiola crispa, 171, 173, 174
267, 268, 350 Pre-copulatory guarding, 114, 128, 136
Planodiscus, 353 Predation, 48, 56, 93, 163, 167, 171, 174, 176,
Plant mites, origin of, 236, 298 178, 180, 187, 189–202, 212, 244–251,
Plant-paraasitic, 8, 27, 41, 55, 57, 81, 84, 87, 257–262, 264, 272, 293, 299, 303, 309,
88, 98, 148, 167, 182–185, 211, 312, 342, 345, 349, 367, 383, 397, 404,
284–287, 292, 293, 298–300, 304, 426, 427
308–310, 312, 315, 316, 324, 325, Prelarva, 34, 60, 71, 74–83, 97
440, 448, 453, 456 Preston-Mafham, K.G., 243
Plant virus, 286, 289, 325 Preston-Mafham, R.A., 243
Plastron, 232, 242 Prey capture, 59, 176, 247–249
Plateremaeidae, 283 Prickly pear, 323
Platyglyphidae, 354 Primary genital opening, 109–113, 115,
Platynothrus peltifer, 238 141, 142
Platyseius italicus, 232, 237, 241, 242 Primary production, 166, 173
Plecoptera, 271 Prince Charmings, 169
Pleioxeny, 406 Pritchard, G., 248, 250
Pleomorph, 133, 134 Proarctacarus oregonensis, 50
Plethodon cinereus, 205 Proctolaelaps, 49, 58, 62, 189, 296, 297, 396
Plethodontidae, 205 P. lobatus, 49
Plumicoles, 382 Proctor, H.C., 39, 108, 116, 120, 122, 134,
Pneumophagus bubonis, 381 139, 149, 233, 248, 256–259, 261, 264,
Podapolipidae, 130, 400, 402 265, 271, 387, 452, 455
Podocinum, 192 Proctor, H.J., 173
Podopolipidae, 60 Progenesis, 84–86
Podospermy, 114 Prokoenenia wheeleri, 30
Poecilochirus Pronematulus pyrrohippeus, 360
P. carabi, 356, 359, 406 Prosoma, 3, 5, 6, 18, 19, 21, 25, 31, 32, 61,
P. necrophori, 368, 370 71, 79
Poison dart frogs, 206 Prospermia, 125
Polak, M., 395, 401 Protacaris crani, 26
Pollen, 46, 48, 57, 66, 283, 295–298, 301–304, Protelean parasite, 41, 83, 187, 354, 380
308, 312, 316, 324, 349, 366–368, 370, Protocerebrum, 15
384, 447 Protochthonius gilboa, 26
Pollination, 295, 296, 402 Protococcus, 183
Pollinator, 140, 295–297, 317, 345, 463 Protogamasellus, 79, 87, 146, 189
Pollution, 209, 266, 270–272, 437 P. mica, 87
Polyaspis, 193 Protomyobia claparedei, 407
Polygyny, 132 Protonymph, 74–76, 81–83, 97, 251, 347,
Polyphagotarsonemus latus, 88, 130, 308, 316 349, 372
Polyphagous, 93, 291, 303, 323, 377 Protonymphon, 28, 29
Polyphyletic, 17, 235 Protzia eximia, 403
Polyxeny, 406 Pruett-Jones, M., 390, 391, 399
Pomerantziidae, 168, 187 Pruett-Jones, S., 390, 391, 399
Pontarachnidae, 239, 240, 254 Pselaphinae, 203
Porospermy, 114 Pseudoarrhenotoky, 144, 301
Post-displacement, 86 Pseudocheylidae, 187
Postembryonic, 30, 74–76 Pseudophryne corroboree, 205
Potter, D.A., 131 Pseudoscorpion, 3, 4, 33–34, 69, 80, 94,
Poultry, 142, 380, 383, 429–431, 443 108, 109, 126, 137, 139, 202, 230,
Poultry red mite, 430, 431 355, 356, 375
Index 489

Pseudoscorpionida, 33–34, 80 Red velvet mite, 3, 41, 86, 89, 187, 193, 282,
Psoroptidia, 55, 347, 352, 378, 405, 406 299, 433
Psoroptoididae, 135 Reinhardt,K., 395
Psychotria horizontalis, 309, 312 Relative humidity, 301, 312, 375
Pterolichidae, 106, 382, 385, 386 Rensch’s rule, 135
Pterolichoidea, 439 Repka, J., 465
Pteromorph, 75, 201 Reptilia, 388, 405, 408
Pterygosoma mutabilis, 124, 388 Reservoir, 123, 199, 262, 263, 268, 424–427,
Pterygosomatidae, 123, 124, 354, 388, 393, 430, 435–437
405, 433 Resource-tracking, 403, 404, 406
Pterygotidae, 20 Resurging disease, 426
Ptyalophagy, 360 Retrorse teeth, 43, 44
Ptychoid, 200, 201 Reversal, 17, 347
Pugh, P.J.A., 195, 242 Rhagidia
Pulmonate, 21, 25, 79 R. arenaria, 122
Pulvillus, 353 R. clavicrinita, 122
Puncochar, P., 262, 268 R. gerlachei, 172, 175
Pycnogonid, 28, 29 R. halophila, 122
Pycnogonida, 5, 15, 28 Rhagidiidae, 34, 54, 79, 122, 172, 187, 188,
Pyemotes 284, 287
P. herfsi, 432 Rhinodromus, 124
P. tritici, 148, 367 Rhinonyssidae, 380, 405
P. ventricosus, 403, 406, 431, 432 Rhinoseius, 296, 344
Pyemotidae, 148, 357, 358, 367, 431, 439 Rhizoglyphus robini, 110, 128, 130, 133, 135,
Pygidium, 26 137, 183
Pygmephoridae, 54, 148, 151, 181, 357, 358, Rhizophora mangle, 296
367, 368 Rhizosphere, 165–168, 185, 208
Pyramica mazu, 202 Rhodacarellus, 79
Pyroglyphidae, 346, 347, 349, 375, 379, Rhodacaridae, 79, 113, 187
380, 439 Rhodacarus roseus, 87
Rhodes, A.C., 242
Rhynie Chert, 26
Q Rhysotritia duplicata, 209
Quadripedal, 60 Rhytidelasma, 53, 386
Queensland, 22, 24, 87, 88, 117, 200, 271, R. punctata, 53
287, 288, 296, 299, 304, 309, 314, 317, Ribbon sperm, 112–114
323, 433, 451–453 Rice, W.R., 141, 466
Questing behaviour, 436 Ricinulei, 4, 23, 26, 27, 32–33, 47, 61, 80, 108
Quills, 73, 148, 385, 386, 405, 449 Rickettsia
Quintero, M. T., 374 R. akari, 430
R. rickettsii, 436
Rickettsial pox, 424, 425, 430
R Rieradevall, M., 267
Radovsky, F.J., 344, 347, 349, 403 Riessen, H.P., 247, 248, 250, 258, 264, 268
Radwan, J., 110, 130, 133, 136, 137 Ring segment, 18, 25, 26
Rajendran, R., 250 Robinson, J.V., 365
Raphignathina, 59, 456 Rocky mountain
Raphignathoidea, 116, 188, 293 spotted fever, 436
Rath, W., 115 wood tick, 436
Raut, S.K., 395 Rolff, J., 365
Raven, P.H., 401 Romero, G.Q., 308
Red colour, 257–259 Rootknot nematode, 195, 196
Red-legged earth mite, 70, 287, 288 Root, R.B., 176
490 Index

Rosa multiflora, 324 Schalk, G., 392


Rosario, R.M.T., 353, 357, 367, 395 Schatz, H., 233
Rosella, 53 Scheloribates, 181, 208, 317
Rosentsteiniidae, 376 S. latipes, 317
Ross, C., 201 Scheloribatidae, 297, 314
Rostrozetes, 181 Schizotetranychus
Roth, V.D., 230, 233 S. longus, 132
Rotifers, 24, 52, 143, 145, 172, 176, 180, 243, S. miscanthi, 132
245, 260 Schroder, R.F.W., 396
Rousch, J.M., 270 Schwiebea, 387
Rozario, S., 311 Schwoerbel, J., 266, 270
Rózsa, L., 404 Scissuralaelaps, 369
Rubiaceae, 308, 309 Scorpionida, 5, 20–24, 30, 80
Rubriscirus, 191 Scorpions, 1, 3, 5, 12, 18–21, 23, 24, 34, 69,
Rush skeleton weed, 323 76, 79, 107, 108, 196, 229, 240, 354
Rust mite, 41, 81, 281, 285, 286, 294, Scrub itch, 55, 362, 433, 439
308, 456 Scrub typhus, 55, 433–435, 442
Rutellum, Rutella, 46–48, 51, 52, 56, 57, 61, Scutacaridae, 52, 54, 70, 87, 181, 198, 200,
65, 66 354, 357–359
Ryckman, R.E., 395 Scutacarus acarorum, 359
Scutum, 44
Scydmaeninae, 203
S Seastedt, T.R., 170
Sabelis, M.W., 139, 148 Sea urchins, 253, 343, 354
Saino, N., 392 Sebaceous gland mite, 429, 443
Saito, Y., 132 Secondarily aquatic arthropods, 229–231
Salamander, 108, 137, 205, 206, 260, 354, Secondary genital opening, 109, 113, 114,
363, 387, 388 118, 142, 151
Salamandridae, 205, 252 Segments, 3, 12–16, 18–21, 25, 26, 29, 30, 32,
Saltatory search, 192–193 40–43, 46–48, 50, 52, 56, 57, 64, 65,
Saltiseius, 198 77–80, 84, 141, 200, 245, 343
Salvador, A., 391 Seius bdelloides, 27
Sammataro, D., 343 Sejida, 41, 58, 112, 357, 455
Sancassania, 119, 133, 136, 170, 356 Sejidae, 27, 49, 318
S. berlesei, 119, 133, 136 Sejina, 41, 94
Sancassania (= Caloglyphus) rodionovi, 356 Sejugal furrow, 42
Sanderson, J.P., 395 Selenoribatidae, 234
San Marco Cathedral, 305 Sellnickia caudata, 314
Saproglyphidae, 315 Sellnickiidae, 314
Saprophagy, 179 Semelparity, 72
Sarcoptes scabiei, 378, 379, 403, 405, 428, Seminal receptacle, 109, 120, 126, 136, 141
432, 443 Sengbusch, H.G., 377
Sarcoptidae, 378, 405, 428, 443 Sensillus, Sensilli, 50, 56, 313, 314, 366
Sarcoptiformes, 41, 52, 53, 55, 56, 66, 75, 80, Sensu lato, 15, 188
116, 243, 450 Sequential progenesis, 85, 86
Sarcoptoidea, 439 Sex, 8, 9, 55–59, 74, 95, 105–151, 197, 199,
Sarraceniopus, 352 252, 263, 357, 365, 377, 466, 467
Saxidromidae, 123 Sex ratio, 74, 131, 132, 143, 144, 146–149,
Saxidromus delamarei, 78, 123, 124 151, 197, 252, 357, 466, 467
Scabies, 341, 378, 379, 428–430, 438, 439, Sexual dimorphism, 116, 142, 314, 315
443, 456 Sexual selection, 9, 106, 128–142, 315, 379,
Scaly leg, 380, 439 400, 401, 408, 465–466
Scapheremaeus, 283, 314, 315 Shaw, M., 428
Index 491

Shrew, 381, 406, 407, 426, 427 Sperm access system, 113
Shultz, J.W., 32 Spermaduct, 118, 119
Sick Building Syndrome, 440 Spermatheca, 109, 115, 116, 120, 123, 137
Siemer, F., 237 Spermatodactyl, 42, 57, 58, 60, 113,
Sieve plate, 64 115, 142
Silk, 5, 19, 29, 55–59, 69–72, 93, 131, 193, Spermatophore, 21, 43, 107, 252, 286, 361
194, 230, 249, 286, 288–290, 299, 302, Spermatophoric organ, 116, 117
303, 432, 433, 461 Spermatotreme, 42, 57, 112, 113
Silphidae, 353, 369, 405 Spermatozoa, 43, 105, 110, 114, 122, 123,
Silurian, 20–22, 25, 34 126, 136, 137, 256
Simmons, T.W., 251 Sperm competition, 106, 136–138, 150
Simuliidae, 356, 403 Sperm precedence, 128, 136, 140
Sinha, R.N., 374 Sperm transfer, 43, 46, 48, 50, 60, 106–128,
Sister group, 15, 20, 29–34, 40, 43, 46, 291 131, 132, 136, 137, 139–142, 146, 150,
Sister taxon, 17, 150 230, 240, 252, 256–257, 361
Sit-and-wait predator, 191–192, 245 Sperm-transfer modes, 106–109, 116,
Siteroptes, 82, 84, 85, 356, 358 120, 122
S. graminum, 84, 85 Sphaerolichida, 41, 455, 456
Siteroptidae, 358 Spicer, G.S., 405
Siva-Jothy, M.T., 137 Spider mites, 3, 41, 55, 57–59, 70, 71, 87, 93,
Skaife, S.H., 367 128, 131, 136, 140, 146, 149, 167, 199,
Sleeping Beauty Paradox, 169 281, 289–293, 299–301, 303, 304, 310,
Slugs, 5, 41, 264, 344, 354, 395, 397 311, 321–323, 325, 456
Smallegange, I.M., 135 Spielman, A., 136
Sminthuridae, 117 Spillover, 425, 429
Smith, B.P., 251, 264, 265, 349, 350, 362–365, Spindle gall, 316
390, 393, 396, 397 Spinneret, 5, 6, 19, 21, 59, 289
Smith, I.M., 233, 237, 238, 248, 268, 362, Spiny rat mite, 430
363, 387 Spirochete, 425–427, 436
Snails, 233, 251, 354, 363 Sporocarp, 91, 166, 169, 199, 318–319, 453
Snakes, 42, 79, 231–233, 343, 344, 383, Sporotheca, Sporothecae, 182, 368
388, 429 Springtail, 109, 117, 162, 164, 171, 173–177,
Snout mite, 70, 187, 188, 193, 195, 196 183, 184, 187, 188, 191–196, 198, 200,
Soil 204, 205
organisms, 164, 178 Stage, 19, 29, 40, 69, 131, 163, 229, 285, 347,
profile, 164, 168, 169, 208 429, 449, 464
Solenidion, Solenidia, 42, 52, 321 Stanton, N.L., 165, 451
Soler, J.J., 384 Staphylinidae, 203
Solfugida, 23, 80 Starling mite, 429, 430
Somites, 12 Stase, 77, 82, 84
Sonenshine, D.E., 43, 125, 126 Stasny, T.A., 285
Sooty mould, 315 Štefka, J.P.E.A., 405, 406
Sorci, G., 391 Steiner, W.A., 209
Southern Pine Beetle, 96, 368 Stereotydeus, 56, 172, 173, 176, 185, 288
Spadiseius, 297 Sternal region, 42, 43, 50, 64, 65, 72
Spanandry, 150 Sternite, 25
Spandrel, 305, 389 Stigmaeidae, 185, 188, 200, 284, 293, 294,
Speleorchestes poduroides, 79 298, 299, 312
Sperchon Stigmaeopsis longus, 290
S. glandulosus, 269 Stored products, 55, 120, 351, 367, 374, 394,
S. setiger, 245, 363, 395, 396 431, 439
Sperchonopsis verrucosa, 245 Stratiolaelaps, 188
Sperchontidae, 245, 269, 363 S. scimitus, 49, 190
492 Index

Stratum T. confusus, 130


S. corneum, 428 T. dispar, 360
S. granulosum, 428 Tatarnic, N.J., 142
Straw itch mite, 367, 431, 438, 439 Taxonomic surrogacy, 452
Stygothrombidioidea, 235, 238 Taylor, P.D., 147
Stylophore, 289, 292 Tectocepheidae, 317
Stylostome, 363, 364, 391, 398, 401, 433 Tectum, Tecta, 48, 201, 302
Subcapitulum, 32, 33, 40, 42, 44, 45, 50, 64, Teinocoptidae, 376
65, 186, 285, 303 Temperature, 88–90, 93, 98, 107, 169–171,
Subelytral space, 449 251, 266, 267, 301, 313, 351, 428
Subfossil mite, 211 Temporary aquatic habitats, 235
Subnivean, 170 Teneriffiidae, 187
Subsocial, 71, 290, 315, 325 Tenuipalpidae, 60, 286, 291, 294, 453
Substrate, 59, 60, 108, 116, 117, 121–123, Ten Winkel, E.H., 250
127, 138, 150, 162, 191, 238, 239, Tergite, 18, 25, 26, 95
243–244, 247, 249, 250, 252, 257, Terpnacaridae, 87
260, 263, 265, 267–269, 361, 363, Terpnacarus, 198
382, 450, 464 T. glebulentus, 77
Subtidal, 235, 239–240 Tertiary, 27, 46, 237, 282, 283
Sunbirds, 297 Testes, 110, 397
Superorder, 5, 32, 40, 41, 50, 62–66, 455 Testosterone, 391, 392, 400
Superparasitism, 357 Tetranychidae, 54, 59, 71, 123, 128, 130–132,
Supracoxal gland, 375 136, 141, 146, 148, 289, 291, 321, 453
Swimming, 18, 48, 60, 116, 122, 127, 130, Tetranychoidea, 116, 146, 167, 282, 284, 289,
246, 247, 256, 262–266, 393, 404 291–293, 316, 353
Swimming hairs, 263, 264 Tetranychus
Symbioribates, 117, 314 T. desertorum, 323
S. papuensis, 314 T. evansi, 290, 322
Symbioribatidae, 117 T. lintearius, 323
Symbiosis, 341, 345–347, 352, 354, 355 T. pacificus, 322
Symphylan, 177, 202, 229 T. urticae, 81, 290, 291, 321
Symptom, 423, 425, 432, 434, 437–442 Thailand, 251, 435
Synchthonius crenulatus, 53 Thanatochresis, 382, 383
Synspermia, 124 Thanatosis, 199–202
Syringicoles, 382 Thelytoky, 144–146, 148, 151, 180, 466
Syringophilidae, 148, 386 Thermacaridae, 354, 387
Syringophiloides minor, 148, 386 Thermacarus nevadensis, 267, 387
Syringophilopsis kirgizorum, 405 Thinozerconidae, 242
Thinozercon michaeli, 242
Three-host tick, 44, 427, 436
T Thrips, 147, 188–190, 297, 298, 303, 311,
Tagma, tagmata, 18, 40 312, 367, 440, 456
Tagmosis, 18–19 Thyas
Tail-spine, 18–20 T. barbigera, 364
Takaku, G., 361 T. stolli, 351
Talitridae, 22 Thysanoptera, 308, 440
Tapeworm, 180, 204, 397, 439 Tick, 2, 27, 40, 71, 110, 193, 232, 283, 341,
Tarsonemellini, 295 423, 450, 461
Tarsonemidae, 58, 87, 130, 144, 150, 151, Tickbird, 204
181, 283, 284, 286, 294, 295, 298, 308, Tickspiders, hooded, 32, 33, 47
316, 317, 343, 360, 368, 453 Timms, B.V., 239
Tarsonemus, 130, 316 Timms, S., 133
T. acerbilis, 316 Timonius timon, 307
Index 493

Tiphys, 258 Tuckerellidae, 70, 292, 294, 453


Toads, 205, 387, 388 Tullbergia granulata, 198
Tobacco Whitefly, 303 Tullgren-modified Berlese Funnel, 168
Tocospermy, 111 Turchetti, T., 317
Tomato Russet Mite, 303 Turtle, 52, 200, 343, 381, 388, 450
Tomentum, tomenta, 304 Two-fold cost of sex, 143
Tompkins, D.M., 396, 400, 402 Tydeidae, 59, 87, 116, 122, 123, 130, 181,
Tortoises, 52, 388 187, 283, 284, 298, 308, 309, 360
Trachymolgus purpureus, 193 Tydeus, 298
Transgenic, 322–323, 464–465 Typhlodromus, 114, 303
Transmission, 44, 296, 320, 384, 386, 393, T. pyri, 303, 304, 310
400, 424, 425, 434–436 Tyrophagus, 170, 180, 374
Transovarial transmission, 44 T. putrescentiae, 234, 442
Trans-ovarian transmission, 435, 436 Tyrrellia circularis, 267
Trans-stadial transmission, 44, 435
Traumatic insemination, 108, 141
Treat, A.E., 343, 348, 360, 365, 395 U
Trechaleidae, 230 Ulex europaeus, 323
Treehole, 165, 231, 233, 235, 244, 268, 282 Ullrich, F., 245, 395–397
Trhypochthoniidae, 233, 234, 237 Uncate palps, 244
Trichobothrium, Trichobothria, 21, 30, 33, 35, Uncinula necator, 309
42, 43, 46, 47, 49, 50, 52, 65, 66, 192, Unionicola
235, 313 U. crassipes, 247, 248, 250, 251, 265,
Trichoptera, 266, 271, 403 267, 268
Trichoribates, 317 U. formosa, 132, 139, 352
Trigonotarbid, 23, 25, 26 U. intermedia, 132, 252, 397
Trigonotarbida, 4, 23, 25, 26, 32 U. ypsilophora, 132, 397
Trigynaspida, 42, 50, 58, 111, 357, 449, 455 Unionicolidae, 121, 131, 149, 247, 250, 251,
Tritonymph, 74–76, 81–83, 97, 130, 292, 347, 253, 261, 264, 265, 267, 364
357, 358 Unionidae, 251
Tritosternal laciniae, 45 Uniramous, 18
Tritrophic level theory, 321 Untergasser, D., 387
Trochometridium tribulatum, 367 Uroobovella, 27, 405, 406
Troides richmondia, 299 Uropoda orbicularis, 356
Trombicula, 433 Uropodidae, 187, 200, 203, 232, 233, 300,
Trombiculidae, 75, 231, 260, 344, 354, 362, 318, 356, 357
376, 381, 387, 388, 433, 434 Uropodina, 42, 52, 94, 111, 189, 353,
Trombiculoidea, 54, 433–435, 439, 442 357–359, 455
Trombidiformes, 2, 41, 52, 54, 55, 57, 59, 66, Uropodoidea, 49, 95
75, 450 Uropygial gland/oil, 384
Trombidiidae, 120, 259, 349, 354, 357 Urstigmata, 42, 254
Trombidioidea, 193 Usher, M.B., 162, 172, 173, 176, 180, 190
Trombidiosis, 433
Trophic guild, 173
Tropical Fowl Mite, 49, 400, 429, 430 V
Tropical Rat Mite, 430 Vacuolated sperm, 43, 111, 112
Tropicoseius, 296 van Bronswijk, J.E.M.H., 375
Tropilaelaps clarae, 115 van Impe, G., 81
Trouessartia, 119, 120 Varroa, 35, 349, 402
T. megadisca, 406 Vasates
Tuberostoma, 172 V. aceriscrumena, 316
Tuckerella, 183, 292 V. robiniae, 131
T. knorri, 293 Vatacarus ipoides, 79, 344
494 Index

Vector, 35, 44, 169, 286, 289, 293, 312, 320, Winterschmidtiidae, 315, 343, 354, 355, 371,
324, 424, 426, 427, 429, 430, 433–437, 373, 393
442, 461 Witalinski, W., 129, 137
Veermann, A., 301 Witches’-broom, 285
Veigaia, 49, 193 Witte, H., 116, 120, 127, 256
Veigaiidae, 49, 113 Wolbachia, 146, 466
Veladeadcarus gasconi, 253 Womersia strandtmanni, 192
Venereal disease, of plants, 295–297 World trade, 437
Venereal transmission, 296, 393 Woyke, J., 115
Vertebrates as hosts, 244, 354, 387 Wrensch, D.L., 145, 148, 367
Vertebrocentric, 6 Wuria, 142
Vestigial legs, 45
Viburnum, 307
X
Xanthippe, 297
W Xenillidae, 317
Waage, J. K., 346, 347 Xenocaligonellidae, 293
Walking dandruff, 431–433, 439 Xiphosurida, 20, 22
Walking legs, 3, 19, 25, 29, 50 Xolalgidae, 383
Walter, D.E., 3, 39, 109, 173, 207, 310, 311,
431, 437, 438
Walther, B.A., 387 Y
Wandesia thermalis, 350 Yasui, Y., 136
Wasmannian mimicry, 353 Yoder, J.A., 360
Waterhousea floribunda, 290 Yohoia tenuis, 15
Water hyacinth, 293–294 Young, W.C., 242, 267–270, 450
Water mites, 28, 41, 72, 110, 187, 230, 342, Yund, P.O., 106
456, 464 Yuval, B., 136
Weatherhead, P.J., 401
Weeks, A.R., 144–146
Welbourn, W.C., 362 Z
Werthelloides, 240 Zachvatkin, Aleksey A., 34
Western flower thrips (WFT), 189, 297, Zeh, D.W., 137
303, 312 Zeh, J.A., 137
West, G.B., 91 Zeno of Elea, 449
Weygoldt, P., 32 Zeno’s paradox of Achilles
Wheal, 430–433, 437–441, 443 and the Tortoise, 449
Wheeler, W.M., 6, 94. 355. 361 Zerconidae, 2, 111, 187
Whipscorpion, 3, 19, 23, 25, 33, 47, 69, Zetomimidae, 234
76, 77 Zetorchestes, 198
Whirligig mite, 190, 191, 298, 431–433, 439 Zetzellia mali, 299
Whiteman, N.K., 405, 406 Zhang, Z.-Q., 357, 395
Wiggins, G.B., 347 Zoonosis, 424, 435
Wikel, S.K., 391, 402 Zoonotic, 425
Wiles, P.R., 250, 393 Zuk, M., 399
Willamette spider mite, 322 Zygopachylus, 69
Willis, D.W., 257 Zygoribatula exilis, 209
Wilson, D.S., 147, 369, 370, 373 Zygoseius, 73
Winkel, E.H., 249 Z. furciger, 89, 90

You might also like