You are on page 1of 15

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/332734981

Fire effects on tree physiology

Article  in  New Phytologist · April 2019


DOI: 10.1111/nph.15871

CITATIONS READS

27 1,012

3 authors, including:

Andreas Bär Sean Michaletz


University of Innsbruck University of British Columbia - Vancouver
12 PUBLICATIONS   129 CITATIONS    121 PUBLICATIONS   2,082 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Leaf temperatures, traits, and rates View project

Timberline - Legacy effects after summer and winter drought View project

All content following this page was uploaded by Andreas Bär on 06 August 2019.

The user has requested enhancement of the downloaded file.


Review

Tansley review

Fire effects on tree physiology

Author for correspondence: Andreas B€ar1 , Sean T. Michaletz2 and Stefan Mayr1
Andreas B€ar 1
Tel.: +43 650 8709399 Department of Botany, University of Innsbruck, Sternwartestraße 15, Innsbruck 6020, Austria; 2Department of Botany and
Email: andreas.baer@uibk.ac.at Biodiversity Research Centre, University of British Columbia, Vancouver, BC V6T 1Z4, Canada
Received: 21 January 2019
Accepted: 7 April 2019

Contents

Summary 1728 V. Biotic attacks 1736

I. Introduction 1728 VI. Conclusion 1737

II. First- and second-order fire effects 1729 Acknowledgements 1737

III. Cambium necrosis 1732 References 1737

IV. Hydraulic dysfunction 1733

Summary
New Phytologist (2019) 223: 1728–1741 Heat injuries sustained in a fire can initiate a cascade of complex mechanisms that affect the
doi: 10.1111/nph.15871 physiology of trees after fires. Uncovering the exact physiological mechanisms and relating
specific injuries to whole-plant and ecosystem functioning is the focus of intense current
Key words: biotic attacks, cambium necrosis, research. Recent studies have made critical steps forward in our understanding of tree
forest fires, hydraulic dysfunction, postfire physiological processes after fires, and have suggested mechanisms by which fire injuries may
effects, tree physiology. interact with disturbances such as drought, insects and pathogens. We outline a conceptual
framework that unifies the involved processes, their interconnections, and possible feedbacks,
and contextualizes these responses with existing hypotheses for disturbance effects on plants
and ecosystems. By focusing on carbon and water as currencies of plant functioning, we
demonstrate fire-induced cambium/phloem necrosis and xylem damage to be main disturbance
effects. The resulting carbon starvation and hydraulic dysfunction are linked with drought and
insect impacts. Evaluating the precise process relationships will be crucial for fully understanding
how fires can affect tree functionality, and will help improve fire risk assessment and mortality
model predictions. Especially considering future climate-driven increases in fire frequency and
intensity, knowledge of the physiological tree responses is important to better estimate postfire
ecosystem dynamics and interactions with climate disturbances.

behavior and effects (Bond & Keeley, 2005; Keeley, 2009). High-
I. Introduction
intensity crown fires consume live and dead canopy fuels, and the
Wildfires are one of the most important natural disturbances combustion of all foliage and meristems in a tree crown can cause
acting on plant ecosystems world-wide. The degree to which immediate mortality, unless the tree is able to resprout from heat-
vegetation is impacted by fire depends on the heat fluxes incident resistant organs (Clarke et al., 2013; Pausas & Keeley, 2017). By
on various plant parts, which is an outcome of various fire contrast, low- to moderate-intensity fires often do not constitute a
behavior variables such as fireline intensity and residence time direct lethal threat to mature trees, but rather, may cause a variety
(Michaletz & Johnson, 2007; O’Brien et al., 2018). The water of injuries that can subsequently interact to impact whole-tree
content, arrangement and loading of fuels strongly influence fire functioning.

1728 New Phytologist (2019) 223: 1728–1741 Ó 2019 The Authors


www.newphytologist.com New Phytologist Ó 2019 New Phytologist Trust
New
Phytologist Tansley review Review 1729

Fire effects on trees can be classified as two types: first- and considered to be completed at 60°C. However, cell necrosis rates
second-order effects. First-order effects comprise the immediate increase exponentially with temperature and lower temperatures
impacts of heat transfer on plant tissues (for a review, see Michaletz also can result in cell death if exposed for longer periods (Hare,
& Johnson, 2007). Nonlethal first-order heat injuries can trigger 1961; Dickinson & Johnson, 2004). Cambium and phloem
second-order effects, such as physiological limitations in carbon necrosis in roots of upper soil layers and close to the root crown are
(C) and water relations or increased susceptibility to insect attacks most likely to occur during ground fires, where smoldering
and pathogenic infections (Michaletz & Johnson, 2008). Altered combustion of organic soil (duff) over hours or days can lead to
physiology and a weakened defense against biotic agents can lead to substantial heating of soils and roots (Ryan & Frandsen, 1991;
long-term restrictions of whole-plant function and may ultimately Swezy & Agee, 1991; Smirnova et al., 2008).
cause latent postfire tree mortality (Figs 1, 2). Rates of heat transfer from the fire into tree tissues are mediated
This review briefly summarizes first- and second-order effects in by plant functional traits (Keeley et al., 2011). In tree stems heat
roots, stems and crowns (section II), before outlining a framework first has to be conducted through the bark to affect other tissues.
to advance a better understanding of physiological responses to fire The thermal insulation ability of bark, which is determined mainly
in context of tree mortality hypotheses (sections III – V). The main by bark thickness, density and moisture content (van Mantgem &
emphasis is necessarily on stems as these have been studied the most. Schwartz, 2003; Lawes et al., 2011; Brando et al., 2012; Pausas,
Focusing on cambium necrosis, hydraulic dysfunction and biotic 2015), is therefore crucial for controlling heat fluxes through bark
attacks, we outline the physiological processes involved, their and potential injuries of underlying tissues. Without sufficient bark
consequences, their mechanistic connections and possible feedback insulation, phloem and vascular cambium can exceed critical
loops. The detailed process steps are described individually, and temperatures for necrosis. If the insulation capability is too low,
illustrated in a conceptual framework (Fig. 3). lethal temperatures can be attained not only in phloem and vascular
cambium tissues, but also in the underlying xylem (Chatziefstratiou
et al., 2013). We are unaware of any empirical data for sapwood
II. First- and second-order fire effects
temperatures during forest fires, but simulations based on
measured cambium data suggest that xylem often can achieve
1. First-order fire effects
critical necrosis temperatures (Michaletz et al., 2012).
First order fire effects result from heat transfer from combusting Heat transfer into the crown can cause immediate bud and/or
fuels into tissues within the roots, stem or crown (Fig. 2; Michaletz foliage necrosis, as well as cambium and phloem damage in
& Johnson, 2007; Bergman & Incropera, 2011). While convection branches. The degree of injuries on crown components is related to
and radiation processes transfer heat to tissue and soil surfaces, heat their functional traits (thermophysical properties), fireline intensity
is further transferred within solid materials (soil particles and/or and residence time, as well as to crown base height (Van Wagner,
plant parts) via conduction. Heat-induced injuries in roots occur as 1973; Michaletz & Johnson, 2006, 2007). Crown component
a result of heat transfer through soil and successive heat conduction traits such as width, surface area, shape, orientation and degree of
into the root interior (Michaletz & Johnson, 2007). They can shielding by foliage control heat fluxes, whereas other traits such as
manifest directly in root death, especially in fine roots, or cause mass, water content, and specific heat capacity determine how
cambium and phloem necrosis. Tissue mortality, which is caused much energy is required to cause a temperature increase. For
by protein denaturation (Rosenberg et al., 1971), is generally example, species with relatively large buds such as ponderosa pine

(a) (b)

Fig. 1 Example of delayed postfire mortality in


a Norway spruce (Picea abies) specimen. The
tree survived a forest fire in March 2014
(Absam, Austria) with a scorched stem and leaf
and bud necroses in the lower crown portion,
but eventually died in the following year.
Although mortality occurred from interactions
of potential injuries to cambial, vascular, and
resource acquisition tissues, it is not known
whether carbon starvation or hydraulic
dysfunction was the ultimate cause. Pictures of
the same tree were taken shortly before (a,
June 2015) and after (b, September 2015) its
death.

Ó 2019 The Authors New Phytologist (2019) 223: 1728–1741


New Phytologist Ó 2019 New Phytologist Trust www.newphytologist.com
New
1730 Review Tansley review Phytologist

Heat transfer Recovery

First-order effects Functional and


• Foliage and growth limitations
bud necrosis Carbon and
• Cambium and water relations Fig. 2 Overview of fire effects in trees. Heat
transfer into crown, stem and root tissues is
phloem necrosis
mediated by functional traits and can
• Xylem hydraulic immediately lead to first-order injuries, which
impairments potentially can induce second-order effects.
• Fine root necrosis Both first- and second-order effects can lead to
physiological impairments in tree carbon (C)
and water relations, consequently limiting
functioning and growth. Depending on
postfire environmental conditions and species-
Second-order effects Tree mortality
specific traits (e.g. abilities to balance C and/or
• Photosynthesis and water restrictions), affected trees may either
C-supply limitations recover from postfire limitations or succumb to
• Water uptake and fire legacy effects. All first-order fire effects
(red box) can also be found in Fig. 3, where
transport restrictions links with second-order effects and
• Biotic attacks physiological feedbacks are detailed. Carbon
starvation and reduction of the hydraulic
efficiency represent the key mechanisms in C
and water relations and are indicated with a
key symbol here and in Fig. 3.

and longleaf pine are far less susceptible to bud necrosis than are physiological functionality, show reduced growth and be more
species with relatively small buds such as sugar maple and American likely to succumb to delayed death (Fig. 1; e.g. Lambert &
beech (Michaletz & Johnson, 2006). For larger crown components Stohlgren, 1988; van Mantgem et al., 2003, 2011; Granzow-de la
like branches, fruits and cones, internal temperature gradients exist Cerda et al., 2012; Nesmith et al., 2015; Maringer et al., 2016;
and conduction within the component becomes important. In Thompson et al., 2017). On the other hand, it is known that
these cases, diameter plays an important role in insulating tissues injured trees may also benefit in the short- and mid-term from
against heat necrosis, because conduction rates decrease with radial reduced competition (Pearson et al., 1972; Keyser et al., 2010;
depth. Thus, as for stems, bark can help insulate vascular tissues and Battipaglia et al., 2014b; Valor et al., 2015, 2018).
cambium. Likewise, many cones and fruits can insulate seed Crown injuries are relatively easy to assess and correlations
embryo tissues against heat necrosis. This has been widely between fire-induced foliage loss and tree growth or mortality have
recognized in serotinous species that store seeds in aerial seed been established by many previous studies (Michaletz & Johnson,
banks, but it also can be important for nonserotinous species 2008; Woolley et al., 2012). In theory, necrotic foliage after a fire
provided that the fire occurs during a temporal window when seeds reduces the leaf area of affected trees, which leads to improved water
are germinable but not yet abscised (Michaletz et al., 2013; availability for the remaining foliage. Consequently, available water
Pounden et al., 2014). can be used by undamaged leaves to increase stomatal conductance
and photosynthesis rates (Reich et al., 1990; Ryan, 2000; Wallin
et al., 2003). In combination with decreased competition from
2. Second-order fire effects
herb and shrub layers for soil resources, trees with crown injuries
Second-order fire effects are more complex and their mechanisms can show unaffected or even enhanced postfire production and
are not as well-understood as those of first-order effects. The growth (Pearson et al., 1972; Keyser et al., 2010; Battipaglia et al.,
response of plant function to fire injuries can vary widely. Fire- 2014b; Valor et al., 2015, 2018). However, these beneficial effects
surviving trees, on the one hand, can be compromised in their can be limited when levels of leaf necrosis are high. With extensive

New Phytologist (2019) 223: 1728–1741 Ó 2019 The Authors


www.newphytologist.com New Phytologist Ó 2019 New Phytologist Trust
New
Phytologist Tansley review Review 1731

c1
Biotic agents

Xylem Phloem Bark


Cambium
c2 c3 b6 b1
Phloeophagous Xylophagous Decrease in Nonstructural
attack attack water potential effects on xylem
hydraulics
b4
Fatal
a1 b2
runaway
Fig. 3 Conceptual diagram illustrating the Cambium and Xylem
embolism
phloem necrosis embolism
cascade of potential physiological responses to
postfire injuries in plant roots, stems, and
crowns. Gray boxes indicate first-order fire a2 b3 b9 b10
effects induced via heat transfer and white Blockage Reduction Reduction Reduction
boxes indicate second-order mechanisms. of carbon of hydraulic of hydraulic of safety
Red-framed boxes (a) describe processes translocation efficiency safety margin
affected by injuries to cambium and phloem
tissues. Blue-frames boxes (b) indicate
b5
processes related to fire effects on xylem Structural effects
hydraulics. Green-framed boxes (c) indicate on xylem hydraulics
postfire biotic agents that may have an
amplifying effect on processes influenced by
both cambium/phloem necrosis and hydraulic
dysfunction. The progressive numbers inside
the boxes refer to the physiological processes,
which are illustrated in the diagram and
discussed in the text. Carbon (C) starvation a5 a6 a7
Reduction of Root Fine root
and reduction of the hydraulic efficiency (key
root growth death necrosis
symbol) constitute crucial physiological
junctures within the carbon and the water
cycle, respectively. The severity and duration a4
of these two key processes depend upon C starvation
postfire environmental conditions, and can
manifest as growth limitations and ultimately a3
determine whether a tree will live or die (see Root C
depletion
also Fig. 2).

leaf necrosis (b11 in Fig. 3), remaining foliage might not be able to resistance to insect colonization (McHugh et al., 2003; Wallin
sustain whole-tree carbohydrate requirements, which consequently et al., 2003). Even if the remaining crown can fully support the
can lead to decreased growth (Pearson et al., 1972; Ryan, 1993; metabolic C demand, less obvious injuries to stem cambial and
Valor et al., 2018), altered root C dynamics (Sword & Haywood, vascular tissues can compromise the tree’s functionality, affecting
1999; Aubrey et al., 2012), or reduced C-based defense and stomatal behavior and tree growth (Battipaglia et al., 2014a;

Ó 2019 The Authors New Phytologist (2019) 223: 1728–1741


New Phytologist Ó 2019 New Phytologist Trust www.newphytologist.com
New
1732 Review Tansley review Phytologist

Thompson et al., 2017). Although it has been shown that fires can necrosis and fire scars will be found preferentially on the tree’s
considerably reduce fine root biomass (a7 in Fig. 3; Swezy & Agee, leeward side. If flame temperatures, residence times and mixing/
1991; Smirnova et al., 2008), the knowledge on how root injuries convection rates are sufficiently high, or the bark is sufficiently thin
can compromise postfire tree viability and productivity is scarce. As (e.g. in small plants), cambium necrosis often occurs around the
fine root degradation reduces water uptake, root injuries can entire bole. Necrotic vascular cambium initiates compartmental-
potentially mediate tree decline and mortality via hydraulic ization and wound closure as part of the scar formation process to
limitation. Denaturation of aquaporins, which play a crucial role avoid insect and pathogen infestation, and to restore functionality
in the water uptake process (Groszmann et al., 2017), may disable (Shigo, 1984; Smith & Sutherland, 2001). Depending on tree
radial water transport in roots, even when heat damage of root cells species and fire scar size, it can take years or decades until wound
was not lethal. closure is completed and even longer until the circuit continuity of
There are two main hypotheses for how fire injuries can impact the cambium and full physiological functionality is regained
second-order functionality and mortality (Michaletz, 2018). The (Smith & Sutherland, 1999; Smith et al., 2016; Stambaugh et al.,
cambium necrosis hypothesis assumes that fire-induced cambium 2017).
and phloem necrosis limits carbohydrate translocation and initiates During the heat transfer process (for further details, see
C starvation (see section III Cambium necrosis), whereas the Michaletz & Johnson, 2007) from the bark surface to the vascular
hydraulic dysfunction hypothesis assumes that the heat of forest cambium, heat is conducted through the intermediate phloem
fires can impair the xylem of trees and cause hydraulic failure (see tissue. Heat fluxes and maximum temperatures of phloem will be
section IV Hydraulic dysfunction). Both hypothesized mecha- generally equal to (if not slightly greater than) those of the vascular
nisms are considered to be able to trigger physiological cascades, cambium. Therefore, it can be assumed that cambium necrosis is
which impact whole-plant function and/or eventually cause tree preceded by phloem necrosis (a1 in Fig. 3; Michaletz & Johnson,
mortality, independently or in combination. 2007). As phloem regeneration relies on cambial cell activity,
In general, even without the effects of fire, the complex cambium necrosis causes a permanent interruption of the phloem
physiological processes that often interact with biotic disturbance pathway and blockage of the downward translocation of photo-
agents and lead to vigor loss or tree mortality have not yet been fully synthates (a2 in Fig. 3). This is equivalent to a heat-induced girdle if
disentangled (McDowell et al., 2008, 2011, 2018; Sala et al., 2010; the entire bole circumference is affected (Noel, 1970). It is well
Sevanto et al., 2014; Anderegg et al., 2015; Hartmann, 2015; known from nonheat-related experiments that girdling triggers the
Mencuccini et al., 2015; Gessler et al., 2017). For example, depletion of nonstructural carbohydrates (NSCs; soluble sugars
McDowell et al. (2008) postulated generalized hypotheses of and starch) below the girdling zone (Daudet et al., 2005; De
physiological mortality mechanisms under drought: C starvation Schepper et al., 2010; Oberhuber et al., 2017). Depletion of NSCs
and hydraulic failure. Carbon starvation is hypothesized to occur also was observed after stem heating during controlled fires (Varner
when prolonged stomatal closure during drought limits photosyn- et al., 2009). Depending on the root’s C pool size, it may take
thetic C assimilation and the metabolic demand for carbohydrates several years to decades until the root’s carbohydrate demand can
needs to be covered by the plant’s C reserves. With persistent no longer be sustained and C reserves have been completely
drought, C reserves will become exhausted or unable to be depleted (a3 in Fig. 3; Noel, 1970). Once C reserves are fully
metabolized or translocated (Sala et al., 2010; McDowell, 2011) exhausted, roots are considered to suffer from C starvation (a4 in
and cellular metabolism can no longer be maintained. Hydraulic Fig. 3), which then will lead to a cessation of fine root production
failure is expected to occur when drought intensity induces water (a5 in Fig. 3; Marshall & Waring, 1985).
column tensions, which exceed the plant’s hydraulic safety margin Massive reductions in fine root biomass following a forest fire
and cause substantial embolism formation. Resulting lethal were detected by Smirnova et al. (2008), who attributed the loss
conductivity losses can then lead to irreversible desiccation before mainly to the effect of heat girdling at the stem base (see also Swezy
C starvation occurs. In addition, biotic mortality agents (see section & Agee, 1991). As most water uptake occurs via the fine root surface
V Biotic attacks) can act as amplifiers for both mechanisms, and vice area, a decrease in fine root biomass theoretically leads to curtailed
versa, C starvation and hydraulic failure can facilitate insect attacks water supply to the stem and crown (b3 in Fig. 3). In consequence,
or pathogenic infections by weakening the plant defense system lowered leaf water and xylem potentials (b6 in Fig. 3) can trigger
(McDowell et al., 2011; Anderegg et al., 2015). premature stomatal closure (b7 in Fig. 3) and/or potentially fatal
embolism (b2 in Fig. 3), respectively (Tyree & Zimmermann,
2002). Increased periods of stomatal closure limit C assimilation
III. Cambium necrosis
(b8 in Fig. 3), which can reduce plant growth or lead to whole-plant
During forest fires, the bark insulation capability may not be mortality via C starvation. Potential embolism – ultimately
sufficient to prevent critical increases of cambium temperatures. induced by cambium and phloem necrosis – during the phase of
The extent of cambium necrosis around the circumference of a stem C starvation may cause mortality (McDowell, 2011; Sevanto et al.,
depends on the interaction of the fire with the air-flow patterns 2014), before entering an irreversible state of desiccation caused by
around the tree stem. Interactions of leeward vortices with moving definitive root death (a6 in Fig. 3).
firelines increase flame temperatures and residence times in the lee The idea that heat can alter tree physiology by damaging
of a tree as compared with the windward side (Gill, 1974; Tunstall cambium and phloem, and thereby disrupting carbohydrate
et al., 1976; Gutsell & Johnson, 1996). Therefore, partial cambial transport to the root, was proposed in the early 1960s (Hare,

New Phytologist (2019) 223: 1728–1741 Ó 2019 The Authors


www.newphytologist.com New Phytologist Ó 2019 New Phytologist Trust
New
Phytologist Tansley review Review 1733

1961, 1965; Wagener, 1961; Fahnestock & Hare, 1964) and soon Smith & Sutherland, 1999, 2001; Schoonenberg et al., 2003; De
was quoted as the standard explanation for postfire tree mortality. Micco et al., 2013; Smith et al., 2016). Fire-induced wounding can
However, direct links between heat-injured cambial tissue and initiate the formation of discolored areas within the sapwood
delayed tree mortality were first examined by Ryan et al. (1988). (Smith et al., 2016). Surrounded by barrier zones, these areas are
They assessed several potential predictors of mortality (e.g. thought to be hydraulically isolated from healthy sapwood, which
diameter at breast height (DBH), average scorch height and the would lead to a reduction of the functional sapwood area. Due to
scorched crown volume), and found the circumferential extent of the exclusion of these discolored areas from the hydraulic pathway,
cambial necrosis to be the best predictor of mortality for injured they remain unstained when branch or stem segments are flushed
Douglas-fir 8 yr following a fire. Mortality occurred in all of the with dye solutions, and they also can cause resistivity shifts in
trees that had cambium necrosis around their entire circumference, electrical resistivity tomograms (Fig. 4). Finally, forest fires have
which is consistent with the cambium necrosis hypothesis of been shown to trigger physiological responses associated with a
postfire mortality (Noel, 1970; Michaletz et al., 2012; Michaletz, compromised hydraulic system. In theory, stomatal conductance
2018). Cambium and phloem necrosis also can occur in roots as a should increase in trees with crown scorch due to improved water
result of conduction heat transfer through soil (Michaletz & supply to the remaining leaf area. However, partially defoliated
Johnson, 2007). Necrosis close to the root crown is most likely to trees surprisingly showed reduced stomatal conductance, no
occur during ground fires, where smoldering combustion of duff change in water-use efficiency, and lowered predawn water
over hours or days can lead to substantial heating of soils and roots potential, which strongly points towards fire-induced impairment
(Ryan & Frandsen, 1991; Swezy & Agee, 1991), causing a fatal of the water transport system (Thompson et al., 2017). Thus,
disruption of carbohydrate downward translocation. substantial evidence suggests a critical role for hydraulic dysfunc-
In later studies, heat girdling experiments were performed to tion in postfire tree physiology. In the following, we summarize the
study the physiological effects of heat-induced cambium injuries in
isolation of crown injuries (Ducrey et al., 1996) or xylem injuries
(Ryan, 2000). When excluding heat-caused disruptions of the Control Fire-exposed
xylem, mortality occurred only in trees that were completely or
almost completely girdled around the bole circumference (Ryan,
2000), indicating that processes other than cambium and phloem
necrosis were responsible for postfire mortality in partially girdled
trees. In girdling experiments, where heat was potentially able to
penetrate the xylem (Ducrey et al., 1996), trees showed clear signs
of water stress. An immediate decrease of sap flux density, lowered
stomatal conductance and increased daily trunk diameter variations
led the authors to speculate about possible heat-related hydraulic
limitations in the xylem.

IV. Hydraulic dysfunction


More recently, a growing number of studies increasingly have been
documenting fire impacts on xylem hydraulic function. Strong
support for a hydraulic mechanism affecting postfire tree physiol-
ogy was found by Balfour & Midgley (2006). By combining
manual girdling (bark, phloem and cambium were removed
without wounding the xylem) with heat girdling experiments, they
were able to analyze the effects of heat on the xylem. All heat-treated
trees started to shed leaves immediately and experienced stem death 100 200 400 800 1500 3000
within 4 wk, whereas manually girdled trees did not show any sign Electrical resistivity ( m)
of rapid leaf loss. Leaf-shedding and subsequent stem death of Fig. 4 Electrical resistivity tomograms of control and fire-exposed Picea
burnt plants were attributed to heat-induced reduction of sapwood abies stems. Electrical resistivity tomography can be used to estimate the
impaired sapwood area within living plants after forest fires. With this
area, which was detected using staining techniques. Also consistent
nondestructive method, the electrical resistivities of cross-sections can be
with these effects of a wildfire, was a study demonstrating that measured and functional sapwood can be differentiated from heartwood
branch death in fire-injured Protea repens specimens occurred much (Guyot et al., 2013) and impaired sapwood (Bieker et al., 2010). Water-
quicker than in manually girdled ones, and was linked to hydraulic conducting sapwood appears as a distinct ring (blue, low electrical
dysfunction and subsequent dehydration (Midgley et al., 2011). resistivity), surrounding the central area of heartwood (red, high electrical
resistivity) in intact control trees. The resistivity shifts in tomography patterns
Further indications for hydraulic dysfunction after wildfires were
of stems, which were exposed to a natural forest fire, indicate profound
provided by the visualization of nonfunctional sapwood via dye losses of functional sapwood areas. Note that the upper limit of the displayed
techniques (B€ar et al., 2018), electrical resistivity tomography resistivity range was set to 3000 Om. Resistivity values within fire-impaired
(Fig. 4) and wood-anatomical observations (Shigo & Marx, 1977; areas can by far exceed this limit.

Ó 2019 The Authors New Phytologist (2019) 223: 1728–1741


New Phytologist Ó 2019 New Phytologist Trust www.newphytologist.com
New
1734 Review Tansley review Phytologist

possible underlying mechanisms, which can be categorized as formation. Experimental confirmation of heat plume-induced
nonstructural (b1 in Fig. 3) and structural effects (b5 in Fig. 3). xylem embolism (b2 in Fig. 3) was later provided by West et al.
(2016), although the VPDs achieved experimentally are likely to be
substantially higher than those experienced in wildfires, where fuel
1. Nonstructural effects on xylem hydraulics
drying and combustion lead to relatively humid fire plumes
Perhaps the most important nonstructural impact of heat is reduced (O’Brien et al., 2018). Heat-plume simulations at 100°C caused
hydraulic efficiency (hydraulic conductivity of the xylem) due to an hydraulic conductivity losses of up to 80% in Eucalyptus cladocalyx
enhanced risk of embolism formation (Michaletz et al., 2012; and K. africana shoots (West et al., 2016). High VDP deficits may
Lodge et al., 2018). Leaf transpiration causes water to evaporate account for rapid tree mortality after fires as such high conductivity
from mesophyll cell walls into the surrounding atmosphere losses are suggested to be lethal (Adams et al., 2017). However,
(Venturas et al., 2017). As described by the cohesion-tension more tests are required to understand the relevance of this effect.
theory (Dixon & Joly, 1895), tensile forces at the evaporative
surface in cell walls pull water molecules through the hydraulic
2. Structural effects on xylem hydraulics
network of the xylem from the soil to leaves. This water transport
requires a continuous column of metastable water, which can be By contrast to nonstructural impacts, heating by fire also can cause
disrupted by cavitation on high tensions (i.e. low water potential) in structural effects (b5 Fig. 3, Fig. 5) related to direct, physical
the xylem sap (Oertli, 1971; Pickard, 1981). Xylem cavitation alterations of the xylem conduits. Xylem conduit walls are
occurs through the entry of air (air seeding) from adjacent air-filled composed of viscoelastic polymers (lignin, cellulose and hemicel-
compartments and ultimately leads to embolization of conduits (b2 luloses) that start to soften and behave like viscous liquids above
in Fig. 3; Cochard et al., 1992; Tyree et al., 1994a; Tyree & temperature thresholds (‘glass transition point’). Although the glass
Zimmermann, 2002). It is initiated when the pressure difference transition of lignin occurs between 60 and 90°C, it is thought that
across an air–water interface in pit membranes exceeds the the glass transition of hemicellulose can occur at c. 50°C (Irvine,
cavitation pressure required to displace the meniscus from the 1984; Hillis & Rozsa, 1985; Olsson & Salmen, 1997, 2004).
pore (Oertli, 1971; Pickard, 1981; Tyree & Zimmermann, 2002). However, additional research is needed to better quantify glass
In the case of conifers the valve like torus-margo architecture can transitions of cell wall polymers and to understand possible
fail and allow air entry, whereby surface tension forces also may play interactions with tensile sap water. When lignin and hemicellulose
a role (Tyree & Zimmermann, 2002; Hacke & Jansen, 2009; temperatures exceed their glass transition points, the overall xylem
Delzon et al., 2010; Losso et al., 2017). Embolized conduits are not rigidity decreases and alterations of the conduit cell walls can occur
able to transport water and, as a consequence, cause a reduction of in response to stresses imposed by tensile sap water (Hacke et al.,
the conductive xylem area and the hydraulic efficiency of the 2001). On cooling, cell wall polymers return to their glassy state
sapwood (b3 in Fig. 3). and heat-induced structural changes can become manifest in two
Increased temperatures during forest fires enhance cavitation important hydraulic aspects.
events as the surface tension decreases with heating (Vargaftik et al., First, the hydraulic efficiency may be permanently reduced by
1983). Lowered sap surface tensions have been shown to crucially heat-induced structural alterations (b3 in Figs 3, 5; Michaletz et al.,
impact the vulnerability to cavitation (Cochard et al., 2009; 2012; West et al., 2016; B€ar et al., 2018). During thermal
Michaletz et al., 2012; West et al., 2016; Losso et al., 2017; Lodge softening, destabilized conduit cell walls are prone to deformation
et al., 2018). If the hydraulic integrity of embolized conduits and rupture. The simultaneous occurrence of low water potentials
cannot be restored by refilling (Zwieniecki & Holbrook, 2009; in the xylem generates additional stress on cell walls (Hacke et al.,
Nardini et al., 2011; Brodersen & McElrone, 2013; Mayr et al., 2001; Cochard et al., 2004; Bouche et al., 2016) and may increase
2014), hydraulic limitations may persist and increase the risk of the likelihood of destabilization. Deformations can affect overall
postfire mortality, for example, by induction of run-away wall geometry as well as pit pore geometry, whereas ruptured parts
embolism (b4 in Fig. 3) under drought conditions. However, of the cell wall can potentially occlude pit membrane pores. In
additional work is required to understand the interplay of theory, both aspects can lead to reductions in hydraulic conduc-
instantaneous effects on water potential and embolism formation tivity (Tyree & Zimmermann, 2002; Sperry et al., 2006; Choat
during fire as well as long-term effects on xylem function. et al., 2007).
It also has been hypothesized that wildfires cause abrupt shifts of Michaletz et al. (2012) were the first to discover heat-related
atmospheric conditions as the hot and dry air of a heat plume reductions of xylem conductivity in laboratory experiments on
suddenly creates a high leaf-to-air vapor pressure deficit (VPD; Populus balsamifera. Heating xylem to 65°C caused permanent and
Kavanagh et al., 2010). Plants respond to VPD elevations with significant conductivity losses. Experimentally induced, long-
stomatal closure to prevent water loss and high tensions within the lasting conductivity limitations were further demonstrated for
xylem (Merilo et al., 2017). However, Kavanagh et al. (2010) K. africana (West et al., 2016) and Fagus sylvatica (B€ar et al., 2018).
assumed that the stomatal response is not fast enough to prevent The first evidence that hydraulic conductivity also can be reduced
extensive water loss, decreases of water potential (b6 in Fig. 3) and by natural forest fires was provided by B€ar et al. (2018), who found
xylem embolism. By modeling the xylem water potential for trees the xylem of F. sylvatica to be 39% less conductive in fire-damaged
exposed to high VPD during fires, they predicted that the branches than in undamaged control branches, a result not related
atmospheric conditions during a fire can cause embolism to embolism. In that study, two coniferous species (Picea abies and

New Phytologist (2019) 223: 1728–1741 Ó 2019 The Authors


www.newphytologist.com New Phytologist Ó 2019 New Phytologist Trust
New
Phytologist Tansley review Review 1735

Pinus sylvestris) also were tested for their susceptibility to heat- midday or already under nonsevere drought. Therefore, heat-
induced conductivity losses. No heat-related effects were found in induced reductions in hydraulic conductivity can, on the one hand,
either species, and it was hypothesized that the high wall result in long-term decreases in stomatal conductance and
reinforcement (conduit thickness-to-span ratio; Hacke et al., respective reductions in assimilation and productivity (b8 in Fig. 3;
2001; Sperry et al., 2006) of conifer tracheids might provide Tyree et al., 1994b; Thompson et al., 2017). If the residual
mechanical protection against heat-induced destabilization. photosynthetic activity is not able to support the plant’s metabolic
Although heat-induced reductions of xylem conductivity were C demand, a heat-induced loss of hydraulic conductivity can
demonstrated in heat experiments and after a wildfire, available initiate the cascading effect of C starvation (a3, a4 in Fig. 3). On the
data are still limited, and species-specific responses and thermal other hand, with increased xylem tensions, cavitation and
thresholds for structural alterations still have to be evaluated. Also, embolism (b2 in Fig. 3) become more likely (Sperry et al., 1993).
no attention has yet been paid to the potential role of living cells of The formation of embolism can further increase hydraulic
the xylem, where, for example, heat-induced impairments of resistances and, unless stomata close, even lead to ‘run-away’
aquaporins may impair radial conductance or capacitance (Sevanto embolism (b4 in Fig. 3; Tyree & Sperry, 1988), which ends in fatal
et al., 2011). hydraulic failure, and consequently, to irreversible whole-plant
Reductions in hydraulic efficiency can have important impacts desiccation. In addition, the effect of deformed conduits on
on postfire plant physiology. For example, they can cause steep hydraulic conductivity might be enhanced by the occurrence of
water potential gradients within the xylem (Sperry et al., 1993), low traumatic resinosis as response to fire-induced wounding (Lom-
leaf water potentials and eventually turgor loss (b6 in Fig. 3). As bardero et al., 2006; Arbellay et al., 2014; Smith et al., 2016). Resin
stomatal regulation is closely connected to leaf water status gets mobilized to seal the fire wound and to build a chemical and
(Brodribb et al., 2003; Tombesi et al., 2015), low leaf water physical barrier against postfire insect attack and pathogen
potentials can trigger stomatal closure (b7 in Fig. 3), for instance at infection (Franceschi et al., 2005). This also can affect the water

Gymnosperms Angiosperms
(a) (b)

Functional
torus sealing
Water Air Air Water
Intact

Meniscus-stabilizing
sap surface tension

Pw Pa Pa Pw

ΔP ΔP
(c) (d)

Torus/porus
modifications
Heat-damaged

Reduced sap
surface tension

Pit membrane
microfibril
Cell wall reorientation
deformations
and ruptures

Fig. 5 Mechanisms of xylem dysfunction by heat for gymnosperm and angiosperm conduits. For comparison, intact (a, b) and heat-damaged (c, d) conduits are
illustrated, respectively. Aspiration of gaseous bubbles via the pits from adjacent, already embolized conduits (air seeding) is initiated when the pressure
difference (DP = Pw Pa; where Pw is the negative pressure in the water-filled and Pa is the atmospheric pressure in the air-filled conduit) between the
compartments exceeds the cavitation pressure required to cause a rupture of an air-water meniscus in the pit membrane pore or a failure of the torus sealing
mechanism. During a fire, air seed cavitation and subsequent embolism become more likely as xylem heating reduces the surface tension of sap water, affecting
the critical cavitation pressure threshold. Thermal softening of cell wall polymers can result in modifications of the pit porus and/or torus, and in reorientation of
pit membrane microfibrils, both of which affect the hydraulic safety of conduits by increasing the likelihood of air seeding. Limitations in hydraulic conductivity
can be caused by heat-induced cell wall deformations affecting overall wall geometry, by ruptured cell wall particles occluding pit membrane pores, or by
changes of the pit pore geometry. Low Pw in water-filled conduits and high pressure differences between water-filled and air-filled conduits potentially create
additional stresses on cell walls (hoop and bending stresses; Hacke et al., 2001), that favor deformations and ruptures during the thermal softening phase.

Ó 2019 The Authors New Phytologist (2019) 223: 1728–1741


New Phytologist Ó 2019 New Phytologist Trust www.newphytologist.com
New
1736 Review Tansley review Phytologist

flow through the sapwood as conduit pits can occlude with resin future drought events and more prone to cavitation (b2 in Fig. 3).
release. However, the extent to which resin-caused blockages can This particularly affects species that already operate near their point
reduce xylem conductivity is currently unknown. of catastrophic hydraulic failure (Tyree & Sperry, 1988; Choat
Second, the hydraulic safety of the xylem may be impacted by et al., 2012).
heat-induced structural alterations (b9 in Figs 3, 5). There are Structural alterations of xylem conduits will affect the hydraulic
species-specific thresholds in water potential, at which air can enter integrity of trees until respective sapwood areas are completely
water-filled conduits (see section ‘IV 1. Nonstructural effects on replaced. Sapwood lifespans vary strongly with species and range
xylem hydraulics’) and block water transport. Vulnerability to from few years up to > 100 yr (Zweifel & Sterck, 2018). Species
drought-induced embolism is thus closely linked to pit architecture with few active tree rings, such as oaks, are able to replace large
and especially to the distribution of pit membrane pore sizes and pit proportions of impaired xylem within the first year postfire (e.g.
membrane thickness (Wheeler et al., 2005; Hacke et al., 2006; Li trees with sapwood lifespan of 5 yr renew c. 20% conductive xylem
et al., 2016). Any physical irregularities in pit structures are thought area each year). Different replacement times of impaired xylem may
to reduce the embolism-resistance (Plavcova et al., 2013). It is thus help to explain species-specific susceptibility to postfire tree
assumed that softened lignin and hemicellulose polymers allow decline and mortality.
movements of the (more heat-resistant) cellulose microfibrils in
conduit walls (Hillis & Rozsa, 1985). Microfibril movement can
V. Biotic attacks
widen pit membrane pore diameters, which increases the likelihood
of air aspiration from adjacent air-filled conduits in angiosperms (air Postfire mortality of surviving trees often is associated with insect
seeding; Oertli, 1971; Tyree et al., 1994a), and it can destabilize the attacks and microbial infections (c1 in Fig. 3; e.g. McHugh et al.,
connection between fibrils, hemicellulose and lignin within conduit 2003; Lombardero et al., 2006; Parker et al., 2006; Hood & Bentz,
walls, which favors pit membrane ruptures and torus sealing failures 2007; Breece et al., 2008; Conedera et al., 2010; Maringer et al.,
in gymnosperms (Sperry & Tyree, 1990; Cochard et al., 2009; 2016; Catry et al., 2017; Westlind & Kelsey, 2019). Trees respond
Delzon et al., 2010). In theory, the liquefaction of wall components to fire injuries with compartmentalization and wound closure
may further lead to alterations of the porus opening, deformations processes to avoid infestation and wood decay. It can take several
of the pit torus, as well as to detachments of the torus from the margo years until wound closure processes are completed, and barrier
membrane (Fengel, 1966; Kollmann & Sachs, 1967). These zones can adequately protect against insect, pathogen and air entry
changes may weaken the reliability of pit-sealing mechanisms (Smith & Sutherland, 2001; Smith et al., 2016). During this time,
during drought stress, which increases the risk of postfire hydraulic wounded trees are particularly vulnerable as their defense system
failure. Heat transfer into the xylem also may affect living can be additionally weakened by fire-induced changes in carbohy-
parenchyma cells that are connected to angiosperm vessels. They drate dynamics (McDowell, 2011; Wiley et al., 2016). Although
are thought to regulate the sap composition by releasing surfactants postfire insect attacks are often short-lived and recede within a few
into the vessel (Morris et al., 2018). According to Schenk et al. years after fire (Hood & Bentz, 2007; Davis et al., 2012), the
(2015, 2017), surfactants coat and stabilize nanobubbles and, defense system of trees can benefit in return from fires over the
therefore, reduce their likelihood to expand under negative pressure longer term. For example, low-severity fires stimulate resin duct-
and form embolism. If vessel-associated parenchyma cells are killed related defenses in ponderosa pine, increasing their resistance to
by heat, this regulating mechanisms could potentially break down, insect attacks. It has been shown that these defenses decline without
increasing the vulnerability to xylem embolism. fire disturbances (Hood et al., 2015).
The ability of heat to shift vulnerability thresholds was tested in Fire-injured trees release high amounts of ethanol and volatile
controlled heating experiments (P. balsamifera: Michaletz et al., terpenes, which attract bark beetles and other wood-boring insects
2012; P. abies, P. sylvestris, F. sylvatica: B€ar et al., 2018) and after (Wood, 1982; Kelsey & Westlind, 2017a,b; Valor et al., 2017).
prescribed (Pinus pinea: Battipaglia et al., 2016) and natural Bark and ambrosia beetles constitute one of the major biotic agents
wildfires (B€ar et al., 2018). Heat experiments and wildfires caused affecting postfire tree health (Parker et al., 2006). Within these
pronounced effects on the hydraulic safety of P. sylvestris and subfamilies of weevils (Curculionidae), two different feeding
F. sylvatica, whereas small or negligible effects were observed in the strategies can be found. Most of the true bark beetles (Scolytinae)
other species. However, it has to be mentioned that the stem surface construct their breeding galleries in the inner bark of trees, where
temperatures and residence times reported for P. pinea were larvae feed on phloem and cambium (phloeophagous). By contrast,
insufficient to cause neither cambium necrosis (Battipaglia et al., ambrosia beetles (Platypodinae) are predominantly xylophagous –
2016), nor cavitation or deformation, which require even higher they colonize the xylem for reproduction. They actively introduce
heat fluxes (Michaletz et al., 2012); thus, results of Battipaglia et al. fungal symbionts via their excavated tunnels to the sapwood, where
(2016) cannot be interpreted as evidence against the xylem the fungi predigest xylem for the larvae. This symbiosis habit can
dysfunction hypothesis. Further research is needed to understand cause severe damage to the xylem and also is known from
how heat can structurally alter the hydraulic safety in different phloeophagous bark beetles. Therefore, phloeophagous attacks (c2
species and to clearly demonstrate the underlying biophysical in Fig. 3) are often accompanied by xylophagous attacks (c3 in
processes. Fire-caused reductions of the hydraulic safety imply that Fig. 3; Hulcr et al., 2007; Six, 2012). Wood borers (insects
injured trees have to cope with reduced hydraulic ‘safety margins’ in belonging to the families of Buprestidae, Cerambycidae and
the field (b10 in Fig. 3), being less tolerant to water stress during Siricidae) also have been shown to affect the xylem of fire-weakened

New Phytologist (2019) 223: 1728–1741 Ó 2019 The Authors


www.newphytologist.com New Phytologist Ó 2019 New Phytologist Trust
New
Phytologist Tansley review Review 1737

trees by creating holes and extended tunnels in sap- and heartwood, periderm and the restoration of phloem and cambium continuity
which often provide gateways for fungal infections (Parker et al., (Franceschi et al., 2005). Although these mechanisms can poten-
2006; Costello et al., 2011; Negron et al., 2016). The feeding habit tially stop insects and fungi from causing fatal damage to the
and the possible introduction of fungi decide upon the type of hydraulic system, especially bark and ambrosia beetles are known to
infestation mechanism and thus how a tree is weakened or even the attack and kill trees with a weakened defense system (e.g. Negron
tree’s death caused. The creation of galleries, larval activity, spread et al., 2016; Catry et al., 2017). Therefore, the state of the defense
of the fungal symbiont and the tree’s defense response (exuded resin system, which is connected to the tree’s pre-fire condition, extents
also can affect phloem integrity) during a phloeophagous attack can of fire injuries and postfire physiology, is crucial for the vulnera-
cause serious damage to the carbohydrate translocation pathway. bility to biotic attacks and for postfire recovery.
For plants that are not injured by fire, it is assumed that the mere
phloem feeding activity is not a primary cause of mortality
VI. Conclusion
(Craighead, 1928; Paine et al., 1997; Hubbard et al., 2013; Wiley
et al., 2016). Rather, the accompanied introduction of fungi to the Postfire functional limitations and mortality of injured trees are
xylem can induce fatal impairments of the hydraulic system, as governed by complex physiological mechanisms that can act
fungal growth can substantially reduce xylem conductivity. Rapid independently or in combination. Accordingly, processes of the C
declines in sap flux and transpiration along with lowered pre-dawn starvation and/or hydraulic failure pathway may be initiated either
water potential were demonstrated after xylophagous attacks separately by cambial necrosis, hydraulic dysfunction and biotic
(Edburg et al., 2012; Hubbard et al., 2013; Frank et al., 2014), agents, or by a combination of these triggers. Even though the
all of which reflect fungal-caused hydraulic conductivity losses. knowledge of postfire physiological responses is growing, our
However, beetle-related phloem consumption in partially girdled understanding of how these processes are linked or coupled and
trees, which have been shown to exhibit an increased probability of how they can be mutually activated is incomplete. Evaluating the
infestation followed by mortality (Amman & Ryan, 1991; precise relationships between these physiological mechanisms is
Rasmussen et al., 1996), further reduces the already limited crucial for better understanding how fires affect the functional
downward translocation of C (a2 in Fig. 3). Even though the exact integrity of trees and for improving models of postfire tree mortality
contribution of insect activity to tree death cannot not yet be (Hood et al., 2018). Continued research is needed to enhance our
assessed, phloem feeding can theoretically lead to a complete knowledge on the underlying biophysical processes (e.g. cell wall
disruption of the C transport pathway within partially girdled kinetics during thermal softening) and to better understand how
stems and initiate the cascading processes of C starvation (a4 in different species respond to fire injuries under varying postfire
Fig. 3). environmental and ecological conditions. Holistic process-based
Maringer et al. (2016) and Conedera et al. (2010) highlighted approaches, considering heat transfer processes, first-order effects
the protective function of bark and the decay resistance of wood and second-order physiological responses, will improve predictions
against secondary fungal infections after fires. Both studies of postfire tree growth and mortality. Therefore, data on functional
attributed the high postfire mortality risk of F. sylvatica to fungal traits that control heat transfer (e.g. bark insulation capability), heat
activity, which was favored by cracks and open lesions in the thin resistance of tissues, wound closure behavior and hydraulics need to
bark and the slow wounding response after heating. Further, root- be linked with postfire physiological parameters to develop
inhabiting fungi (e.g. Poria spp., Leptographium spp. and improved management and risk assessment tools.
Heterobasidium annosum) were observed to be involved in postfire
tree decline and mortality (Littke & Gara, 1986; Otrosina et al.,
1999, 2002). Xylem decay and tissue occlusion by hyphal growth Acknowledgements
and resinosis during infestation may lead to considerable hydraulic This work was supported by grants of the ‘Doktoratsstipendium
conductivity losses in roots (Joseph et al., 1998). Any additional neu aus der Nachwuchsf€orderung der Leopold-Franzens-
loss in xylem conductivity (b3 in Fig. 3) caused by biotic agents may Universit€at Innsbruck’. The study was conducted in the frame of
amplify hydraulic limitations in fire-injured trees and thus be the research area ‘Mountain Regions’ of the University of
critical for tree survival. Innsbruck. STM was supported by SERDP project RC18-1346.
Vigorous trees may successfully defend themselves from biotic
attacks by killing or compartmentalizing the invading organism.
The highly evolved defense mechanisms of conifers (Franceschi ORCID
et al., 2005) are based on resin stored in axial and radial ducts.
Andreas B€ar https://orcid.org/0000-0002-0059-3964
When attacked, coniferous trees can alter the nature of the resin by
Stefan Mayr https://orcid.org/0000-0002-3319-4396
producing toxic chemicals and initiate resin flow to the wound site,
Sean T. Michaletz https://orcid.org/0000-0003-2158-6525
thereby compartmentalizing the intruding organism. Angiosperms
respond with accumulation of tannins, formation of tyloses, as well
as anatomical and chemical modifications to isolate the infected
area (Shigo, 1984; Salle et al., 2014). Subsequently, callus forma- References
tion in the cambial zone constitutes an important wound-closure Adams HD, Zeppel MJB, Anderegg WRL, Hartmann H, Landh€a usser SM, Tissue
and healing process that initiates the development of a wound DT, Huxman TE, Hudson PJ, Franz TE, Allen CD et al. 2017. A multi-species

Ó 2019 The Authors New Phytologist (2019) 223: 1728–1741


New Phytologist Ó 2019 New Phytologist Trust www.newphytologist.com
New
1738 Review Tansley review Phytologist

synthesis of physiological mechanisms in drought-induced tree mortality. Nature Cochard H, Breda N, Granier A, Aussenac G. 1992. Vulnerability to air embolism
Ecology and Evolution 1: 1285–1291. of three European oak species (Quercus petraea (Matt) Liebl, Q. pubescens Willd,
Amman GD, Ryan KC. 1991. Insect infestation of fire-injured trees in the Greater Q. robur L). Annals of Forest Science 49: 225–233.
Yellowstone Area. Research Note INT-398. Odgen, UT, USA: USDA Forest Cochard H, Froux F, Mayr S, Coutand C. 2004. Xylem wall collapse in water-
Service, Intermountain Forest and Range Experiment Station. stressed pine needles. Plant Physiology 134: 401–408.
Anderegg WRL, Hicke JA, Fisher RA, Allen CD, Aukema J, Bentz B, Hood S, Cochard H, H€oltt€a T, Herbette S, Delzon S, Mencuccini M. 2009. New insights
Lichstein JW, Macalady K, McDowell N et al. 2015. Tree mortality from into the mechanisms of water-stress-induced cavitation in conifers. Plant
drought, insects, and their interactions in a changing climate. New Phytologist 208: Physiology 151: 949–954.
674–683. Conedera M, Lucini L, Valese E, Ascoli D, Pezzati GB. 2010. Fire resistance and
Arbellay E, Stoffel M, Sutherland EK, Smith KT, Falk DA. 2014. Resin duct size vegetative recruitment ability of different deciduous trees species after low- to
and density as ecophysiological traits in fire scars of Pseudotsuga menziesii and moderate-intensity surface fires in southern Switzerland. Proceedings of the VI
Larix occidentalis. Annals of Botany 114: 973–980. International Conference on Forest Fire Research, Coimbra, Portugal: 15–18.
Aubrey DP, Mortazavi B, O’Brien JJ, McGee JD, Hendricks JJ, Kuehn KA, Teskey Costello SL, Negron JF, Jacobi WR. 2011. Wood-boring insect abundance in fire-
RO, Mitchell RJ. 2012. Influence of repeated canopy scorching on soil CO2 injured ponderosa pine. Agricultural and Forest Entomology 13: 373–381.
efflux. Forest Ecology and Management 282: 142–148. Craighead FC. 1928. Interrelation of tree-killing barkbeetles (Dendroctonus) and
Balfour DA, Midgley JJ. 2006. Fire induced stem death in an African acacia is not blue stains. Journal of Forestry 26: 886–887.
caused by canopy scorching. Austral Ecology 31: 892–896. Daudet F-A, Ameglio T, Cochard H, Archilla O, Lacointe A. 2005. Experimental
B€ar A, Nardini A, Mayr S. 2018. Post-fire effects in xylem hydraulics of Picea abies, analysis of the role of water and carbon in tree stem diameter variations. Journal of
Pinus sylvestris and Fagus sylvatica. New Phytologist 217: 1484–1493. Experimental Botany 56: 135–144.
Battipaglia G, De Micco V, Fournier T, Aronne G, Carcaillet C. 2014a. Isotopic Davis RS, Hood S, Bentz BJ. 2012. Fire-injured ponderosa pine provide a pulsed
and anatomical signals for interpreting fire-related responses in Pinus halepensis. resource for bark beetles. Canadian Journal of Forest Research 42: 2022–2036.
Trees – Structure and Function 28: 1095–1104. De Micco V, Zalloni E, Balzano A, Battipaglia G. 2013. Fire influence on Pinus
Battipaglia G, Savi T, Ascoli D, Castagneri D, Esposito A, Mayr S, Nardini A. halepensis: wood responses close and far from the scars. IAWA Journal 34: 446–
2016. Effects of prescribed burning on ecophysiological, anatomical and stem 458.
hydraulic properties in Pinus pinea L. Tree Physiology 36: 1–13. De Schepper V, Steppe K, Van Labeke M-C, Lemeur R. 2010. Detailed analysis of
Battipaglia G, Strumia S, Esposito A, Giuditta E, Sirignano C, Altieri S, Rutigliano double girdling effects on stem diameter variations and sap flow in young oak
FA. 2014b. The effects of prescribed burning on Pinus halepensis Mill. as revealed trees. Environmental and Experimental Botany 68: 149–156.
by dendrochronological and isotopic analyses. Forest Ecology and Management Delzon S, Douthe C, Sala A, Cochard H. 2010. Mechanism of water-stress induced
334: 201–208. cavitation in conifers: bordered pit structure and function support the hypothesis
Bergman TL, Incropera FP. 2011. Fundamentals of heat and mass transfer. of seal capillary-seeding. Plant, Cell & Environment 33: 2101–2111.
Hoboken, NJ, USA: John Wiley & Sons. Dickinson MB, Johnson EA. 2004. Temperature-dependent rate models of vascular
Bieker D, Kehr R, Weber G, Rust S. 2010. Non-destructive monitoring of early cambium cell mortality. Canadian Journal of Forest Research 559: 546–559.
stages of white rot by Trametes versicolor in Fraxinus excelsior. Annals of Forest Dixon HH, Joly J. 1895. On the ascent of sap. Philosophical Transactions of the Royal
Science 67: 210. Society of London. B: Biological Sciences 186: 563–576.
Bond WJ, Keeley JE. 2005. Fire as a global ‘herbivore’: the ecology and evolution of Ducrey M, Duhoux F, Huc R, Rigolot E. 1996. The ecophysiological and growth
flammable ecosystems. Trends in Ecology and Evolution 20: 387–394. responses of Aleppo pine (Pinus halepensis) to controlled heating applied to the
Bouche PS, Delzon S, Choat B, Badel E, Brodribb TJ, Burlett R, Cochard H, base of the trunk. Canadian Journal of Forest Research 26: 1366–1374.
Charra-Vaskou K, Lavigne B, Li S et al. 2016. Are needles of Pinus pinaster more Edburg SL, Hicke JA, Brooks PD, Pendall EG, Ewers BE, Norton U, Gochis D,
vulnerable to xylem embolism than branches? New insights from X-ray computed Gutmann ED, Meddens AJH. 2012. Cascading impacts of bark beetle-caused
tomography. Plant, Cell & Environment 39: 860–870. tree mortality on coupled biogeophysical and biogeochemical processes. Frontiers
Brando PM, Nepstad DC, Balch JK. 2012. Fire-induced tree mortality in a in Ecology and the Environment 10: 416–424.
neotropical forest: the roles of bark traits, tree size, wood density and fire behavior. Fahnestock GR, Hare RC. 1964. Heating of tree trunks in surface fires. Journal of
Global Change Biology 18: 630–641. Forestry 62: 799–805.
Breece CR, Kolb TE, Dickson BG, McMillin JD, Clancy KM. 2008. Prescribed fire Fengel D. 1966. On the changes of the wood and its components within the
effects on bark beetle activity and tree mortality in southwestern ponderosa pine temperature range up to 200 C - Part III. Holz als Roh- und Werkstoff 11: 529–536.
forests. Forest Ecology and Management 255: 119–128. Franceschi VR, Krokene P, Christiansen E, Krekling T. 2005. Anatomical and
Brodersen CR, McElrone AJ. 2013. Maintenance of xylem network transport chemical defenses of conifer bark against bark beetles and other pests. New
capacity: a review of embolism repair in vascular plants. Frontiers in Plant Science Phytologist 167: 353–376.
4: 1–11. Frank JM, Massman WJ, Ewers BE, Huckaby LS, Negron JF. 2014. Ecosystem
Brodribb TJ, Holbrook NM, Edwards EJ, Gutierrez MV. 2003. Relations between CO2/H2O fluxes are explained by hydraulically limited gas exchange during tree
stomatal closure, leaf turgor and xylem vulnerability in eight tropical dry forest mortality from spruce bark beetles. Journal of Geophysical Research: Biogeosciences
trees. Plant, Cell & Environment 26: 443–450. 119: 1195–1215.
Catry FX, Branco M, Sousa E, Caetano J, Naves P, Nobrega F. 2017. Presence and Gessler A, Schaub M, McDowell NG. 2017. The role of nutrients in drought-
dynamics of ambrosia beetles and other xylophagous insects in a Mediterranean induced tree mortality and recovery. New Phytologist 214: 513–520.
cork oak forest following fire. Forest Ecology and Management 404: 45–54. Gill AM. 1974. Toward an understanding of fire-scar formation: field observation
Chatziefstratiou EK, Bohrer G, Bova AS, Subramanian R, Frasson RPM, Scherzer and laboratory simulation. Forest Science 20: 198–205.
A, Butler BW, Dickinson MB. 2013. FireStem2D – A two-dimensional heat Granzow-de la Cerda I, Lloret F, Ruiz JE, Vandermeer JH. 2012. Tree mortality
transfer model for simulating tree stem injury in fires. PLoS ONE 8: 1–14. following ENSO-associated fires and drought in lowland rain forests of Eastern
Choat B, Cobb AR, Jansen S. 2007. Structure and function of bordered pits: new Nicaragua. Forest Ecology and Management 265: 248–257.
discoveries and impact on whole-plant function. New Phytologist 177: 608–626. Groszmann M, Osborn HL, Evans JR. 2017. Carbon dioxide and water transport
Choat B, Jansen S, Brodribb TJ, Cochard H, Delzon S, Bhaskar R, Bucci SJ, Feild through plant aquaporins. Plant, Cell & Environment 40: 938–961.
TS, Gleason SM, Hacke UG et al. 2012. Global convergence in the vulnerability Gutsell SL, Johnson EA. 1996. How fire scars are formed: coupling a disturbance
of forests to drought. Nature 491: 752–755. process to its ecological effects. Canadian Journal of Forest Research 26: 166–174.
Clarke PJ, Lawes MJ, Midgley JJ, Lamont BB, Ojeda F, Burrows GE, Enright NJ, Guyot A, Ostergaard KT, Lenkopane M, Fan J, Lockington DA. 2013. Using
Knox KJE. 2013. Resprouting as a key functional trait: How buds, protection and electrical resistivity tomography to differentiate sapwood from heartwood:
resources drive persistence after fire. New Phytologist 197: 19–35. application to conifers. Tree Physiology 33: 187–194.

New Phytologist (2019) 223: 1728–1741 Ó 2019 The Authors


www.newphytologist.com New Phytologist Ó 2019 New Phytologist Trust
New
Phytologist Tansley review Review 1739

Hacke UG, Jansen S. 2009. Embolism resistance of three boreal conifer species Losso A, Beikircher B, D€amon B, Kikuta S, Schmid P, Mayr S. 2017. Xylem sap
varies with pit structure. New Phytologist 182: 675–686. surface tension may be crucial for hydraulic safety. Plant Physiology 175: 1135–
Hacke UG, Sperry JS, Pockman WT, Davis SD, McCulloh KA. 2001. Trends in 1143.
wood density and structure are linked to prevention of xylem implosion by van Mantgem P, Schwartz M. 2003. Bark heat resistance of small trees in
negative pressure. Oecologia 126: 457–461. Californian mixed conifer forests: testing some model assumptions. Forest Ecology
Hacke UG, Sperry JS, Wheeler JK, Castro L. 2006. Scaling of angiosperm xylem and Management 178: 341–352.
structure with safety and efficiency. Tree Physiology 26: 689–701. van Mantgem PJ, Stephenson NL, Knapp E, Battles J, Keeley JE. 2011. Long-term
Hare RC. 1961. Heat effects on living plants. Occasional Paper 183. New Orleans, LA, effects of prescribed fire on mixed conifer forest structure in the Sierra Nevada,
USA: USDA Forest Service, Southern Forest Experiment Station. California. Forest Ecology and Management 261: 989–994.
Hare RC. 1965. Chemical test for fire damage. Journal of Forestry 63: 939. van Mantgem PJ, Stephenson NL, Mutch LS, Johnson VG, Esperanza AM,
Hartmann H. 2015. Carbon starvation during drought-induced tree mortality – are Parsons DJ. 2003. Growth rate predicts mortality of Abies concolor in both burned
we chasing a myth ? Journal of Plant Hydraulics 2: e-0005. and unburned stands. Canadian Journal of Forest Research 33: 1029–1038.
Hillis WE, Rozsa AN. 1985. High temperature and chemical effects on wood Maringer J, Ascoli D, K€ uffer N, Schmidtlein S, Conedera M. 2016. What drives
stability. Wood Science and Technology 19: 57–66. European beech (Fagus sylvatica L.) mortality after forest fires of varying severity?
Hood S, Bentz B. 2007. Predicting postfire Douglas-fir beetle attacks and tree Forest Ecology and Management 368: 81–93.
mortality in the northern Rocky Mountains. Canadian Journal of Forest Research Marshall JD, Waring RH. 1985. Predicting fine root production and turnover by
37: 1058–1069. monitoring root starch and soil temperature. Canadian Journal of Forest Research
Hood S, Sala A, Heyerdahl EK, Boutin M. 2015. Low-severity fire increases tree 15: 791–800.
defense against bark beetle attacks. Ecology 96: 1846–1855. Mayr S, Schmid P, Laur J, Rosner S, Charra-Vaskou K, Damon B, Hacke UG.
Hood SM, Varner JM, Van Mantgem P, Cansler AC. 2018. Fire and tree death: 2014. Uptake of water via branches helps timberline conifers refill embolized
understanding and improving modeling of fire-induced tree mortality. xylem in late winter. Plant Physiology 164: 1731–1740.
Environmental Research Letters 13: 113004. McDowell N, Pockman WT, Allen CD, Breshears DD, Cobb N, Kolb T, Plaut J,
Hubbard RM, Rhoades CC, Elder K, Negron J. 2013. Changes in Sperry J, West A, Williams DG et al. 2008. Mechanisms of plant survival and
transpiration and foliage growth in lodgepole pine trees following mountain mortality during drought: why do some plants survive while others succumb to
pine beetle attack and mechanical girdling. Forest Ecology and Management drought? New Phytologist 178: 719–739.
289: 312–317. McDowell NG. 2011. Mechanisms linking drought, hydraulics, carbon
Hulcr J, Mogia M, Isua B, Novotny V. 2007. Host specificity of ambrosia and bark metabolism, and vegetation mortality. Plant Physiology 155: 1051–1059.
beetles (Col., Curculionidae: Scolytinae and Platypodinae) in a New Guinea McDowell NG, Allen CD, Anderson-Teixeira K, Brando P, Brienen R, Chambers
rainforest. Ecological Entomology 32: 762–772. J, Christoffersen B, Davies S, Doughty C, Duque A et al. 2018. Drivers and
Irvine GM. 1984. The glass transitions of lignin and hemicellulose and their mechanisms of tree mortality in moist tropical forests. New Phytologist 219: 851–
measurement by differential thermal analysis. Tappi Journal 67: 118–121. 869.
Joseph G, Kelsey RG, Thies WG. 1998. Hydraulic conductivity in roots of McDowell NG, Beerling DJ, Breshears DD, Fisher RA, Raffa KF, Stitt M. 2011.
ponderosa pine infected with black-stain (Leptographium wageneri) or annosus The interdependence of mechanisms underlying climate-driven vegetation
(Heterobasidion annosum) root disease. Tree Physiology 18: 333–339. mortality. Trends in Ecology and Evolution 26: 523–532.
Kavanagh KL, Dickinson MB, Bova AS. 2010. A way forward for fire-caused tree McHugh CW, Kolb TE, Wilson JL. 2003. Bark beetle attacks on ponderosa pine
mortality prediction: modeling a physiological consequence of fire. Fire Ecology 6: following fire in northern Arizona. Environmental Entomology 32: 510–522.
80–94. Mencuccini M, Minunno F, Salmon Y, Martinez-Vilalta J, H€oltt€a T. 2015.
Keeley JE. 2009. Fire intensity, fire severity and burn severity: a brief review and Coordination of physiological traits involved in drought-induced mortality of
suggested usage. International Journal of Wildland Fire 18: 116–126. woody plants. New Phytologist 208: 396–409.
Keeley JE, Pausas JG, Rundel PW, Bond WJ, Bradstock RA. 2011. Fire as an Merilo E, Yarmolinsky D, Jalakas P, Parik H, Tulva I, Rasulov B, Kilk K, Kollist H.
evolutionary pressure shaping plant traits. Trends in Plant Science 16: 406–411. 2017. Stomatal VPD response: there is more to the story than ABA. Plant
Kelsey RG, Westlind DJ. 2017a. Physiological stress and ethanol accumulation in Physiology 176: 851–864.
tree stems and woody tissues at sublethal temperatures from fire. BioScience 67: Michaletz ST. 2018. Xylem dysfunction in fires: towards a hydraulic theory of plant
443–451. responses to multiple disturbance stressors. New Phytologist 217: 1391–1393.
Kelsey RG, Westlind DJ. 2017b. Ethanol and primary attraction of red turpentine Michaletz ST, Johnson EA. 2006. A heat transfer model of crown scorch in forest
beetle in fire stressed ponderosa pine. Forest Ecology and Management 396: 44–54. fires. Canadian Journal of Forest Research 36: 2839–2851.
Keyser TL, Smith FW, Shepperd WD. 2010. Growth response of Pinus ponderosa Michaletz ST, Johnson EA. 2007. How forest fires kill trees: a review of the
following a mixed-severity wildfire in the Black Hills, South Dakota. Western fundamental biophysical processes. Scandinavian Journal of Forest Research 22:
Journal of Applied Forestry 25: 49–54. 500–515.
Kollmann FFP, Sachs IB. 1967. The effects of elevated temperature on certain wood Michaletz ST, Johnson EA. 2008. A biophysical process model of tree mortality in
cells. Wood Science and Technology 1: 14–25. surface fires. Canadian Journal of Forest Research 38: 2013–2029.
Lambert BS, Stohlgren TJ. 1988. Giant sequoia mortality in burned and unburned Michaletz ST, Johnson EA, Mell WE, Greene DF. 2013. Timing of fire relative to
stands. Journal of Forestry 86: 44–46. seed development may enable non-serotinous species to recolonize from the aerial
Lawes MJ, Richards A, Dathe J, Midgley JJ. 2011. Bark thickness determines fire seed banks of fire-killed trees. Biogeosciences 10: 5061–5078.
resistance of selected tree species from fire-prone tropical savanna in north Michaletz ST, Johnson EA, Tyree MT. 2012. Moving beyond the cambium
Australia. Plant Ecology 212: 2057–2069. necrosis hypothesis of post-fire tree mortality: cavitation and deformation of
Li S, Lens F, Karimi Z, Klepsch MM, Schenk HJ, Schmitt M, Schuldt B, Jansen S. xylem in forest fires. New Phytologist 194: 254–263.
2016. Intervessel pit membrane thickness as a key determinant of embolism Midgley JJ, Kruger LM, Skelton R. 2011. How do fires kill plants? The hydraulic
resistance in angiosperm xylem. IAWA Journal 37: 152–171. death hypothesis and Cape Proteaceae ‘fire-resisters’. South African Journal of
Littke WR, Gara RI. 1986. Decay of fire-damaged lodgepole pine in south-central Botany 77: 381–386.
Oregon. Forest Ecology and Management 17: 279–287. Morris H, Plavcova L, Gorai M, Klepsch MM, Kotowska M, Schenk HJ, Jansen S.
Lodge AG, Dickinson MB, Kavanagh KL. 2018. Xylem heating increases 2018. Vessel-associated cells in angiosperm xylem: highly specialized living cells at
vulnerability to cavitation in longleaf pine. Environmental Research Letters 13: the symplast–apoplast boundary. American Journal of Botany 105: 151–160.
055007. Nardini A, Lo Gullo MA, Salleo S. 2011. Refilling embolized xylem conduits: Is it a
Lombardero MJ, Ayres MP, Ayres BD. 2006. Effects of fire and mechanical matter of phloem unloading? Plant Science 180: 604–611.
wounding on Pinus resinosa resin defenses, beetle attacks, and pathogens. Forest Negron JF, McMillin J, Sieg CH, Fowler JF, Allen KK, Wadleigh LL, Anhold JA,
Ecology and Management 225: 349–358. Gibson KE. 2016. Variables associated with the occurrence of Ips beetles, red

Ó 2019 The Authors New Phytologist (2019) 223: 1728–1741


New Phytologist Ó 2019 New Phytologist Trust www.newphytologist.com
New
1740 Review Tansley review Phytologist

turpentine beetle and wood borers in live and dead ponderosa pines with post-fire Sala A, Piper F, Hoch G. 2010. Physiological mechanisms of drought-induced tree
injury. Agricultural and Forest Entomology 18: 313–326. mortality are far from being resolved. New Phytologist 186: 274–281.
Nesmith JC, Das AJ, O’Hara KL, van Mantgem PJ. 2015. The influence of prefire Salle A, Nageleisen L, Lieutier F. 2014. Bark and wood boring insects involved in
tree growth and crown condition on postfire mortality of sugar pine following oak declines in Europe: current knowledge and future prospects in a context of
prescribed fire in Sequoia National Park. Canadian Journal of Forest Research 45: climate change. Forest Ecology and Management 328: 79–93.
910–919. Schenk HJ, Espino S, Romo DM, Nima N, Do AYT, Michaud JM,
Noel ARA. 1970. The girdled tree. Botanical Review 36: 162–195. Papahadjopoulos-Sternberg B, Yang J, Zuo YY, Steppe K et al. 2017. Xylem
Oberhuber W, Gruber A, Lethaus G, Winkler A. 2017. Stem girdling indicates surfactants introduce a new element to the cohesion-tension theory. Plant
prioritized carbon allocation to the root system at the expense of radial stem Physiology 173: 1177–1196.
growth in Norway spruce under drought conditions. Environmental and Schenk HJ, Steppe K, Jansen S. 2015. Nanobubbles: a new paradigm for air-seeding
Experimental Botany 138: 109–118. in xylem. Trends in Plant Science 20: 199–205.
O’Brien JJ, Hiers JK, Varner JM, Hoffman CM, Dickinson MB, Michaletz ST, Schoonenberg T, Pinard M, Woodward S. 2003. Responses to mechanical
Loudermilk EL, Butler BW. 2018. Advances in mechanistic approaches to wounding and fire in tree species characteristic of seasonally dry tropical forest of
quantifying biophysical fire effects. Current Forestry Reports 4: 161–177. Bolivia. Canadian Journal of Forest Research 33: 330–338.
Oertli JJ. 1971. Stability of water under tension in the xylem. Zeitschrift fur Sevanto S, H€oltt€a T, Holbrook NM. 2011. Effects of the hydraulic coupling
Pflanzenphysiologie 65: 195–209. between xylem and phloem on diurnal phloem diameter variation. Plant, Cell &
Olsson A-M, Salmen L. 1997. The effect of lignin composition on the viscoelastic Environment 34: 690–703.
properties of wood. Nordic Pulp and Paper Research Journal 12: 140–144. Sevanto S, McDowell NG, Dickman LT, Pangle R, Pockman WT. 2014. How do
Olsson A-M, Salmen L. 2004. The softening behavior of hemicelluloses related to trees die? A test of the hydraulic failure and carbon starvation hypotheses. Plant,
moisture. ACS Symposium Series 864: 184–197. Cell & Environment 37: 153–161.
Otrosina WJ, Bannwart D, Roncadori RW. 1999. Root-infecting fungi associated Shigo AL. 1984. Compartmentalization: a conceptual framework for understanding
with a decline of longleaf pine in the southeastern United States. Plant and Soil how trees grow and defend themselves. Annual Review of Phytopathology 22: 189–
217: 145–150. 214.
Otrosina WJ, Walkinshaw CH, Zarnoch SJ, Sung S, Sullivan BT. 2002. Root Shigo AL, Marx HG. 1977. Compartmentalization of decay in trees. Agriculture
disease, longleaf pine mortality, and prescribed burning. Proceedings of the eleventh Information Bulletin No. 405. Washington, DC, USA: USDA Forest Service.
biennial southern silvicultural research conference. Asheville, NC, USA: USDA Six DL. 2012. Ecological and evolutionary determinants of bark beetle - fungus
Forest Service, Southern Research Station,551–557. symbioses. Insects 3: 339–366.
Paine TD, Raffa KF, Harrington TC. 1997. Interactions among scolytid bark Smirnova E, Bergeron Y, Brais S, Granstr€om A. 2008. Postfire root distribution of
beetles, their associated fungi, and live host conifers. Annual Review of Entomology Scots pine in relation to fire behaviour. Canadian Journal of Forest Research 38:
42: 179–206. 353–362.
Parker TJ, Clancy KM, Mathiasen RL. 2006. Interactions among fire, insects and Smith KT, Arbellay E, Falk DA, Sutherland EK. 2016. Macroanatomy and
pathogens in coniferous forests of the interior western United States and Canada. compartmentalization of recent fire scars in three North American conifers.
Agricultural and Forest Entomology 8: 167–189. Canadian Journal of Forest Research 46: 535–542.
Pausas JG. 2015. Bark thickness and fire regime. Functional Ecology 29: 315–327. Smith KT, Sutherland EK. 1999. Fire-scar formation and compartmentalization in
Pausas JG, Keeley JE. 2017. Epicormic resprouting in fire-prone ecosystems. Trends oak. Canadian Journal of Forest Research 29: 166–171.
in Plant Science 22: 1008–1015. Smith KT, Sutherland EK. 2001. Terminology and biology of fire scars in selected
Pearson HA, Davis JR, Schubert GH. 1972. Effects of wildfire on timber and forage central hardwoods. Tree-Ring Research 57: 141–147.
production in Arizona. Journal of Range Management 25: 250–253. Sperry JS, Alder NN, Eastlack SE. 1993. The effect of reduced hydraulic
Pickard WF. 1981. The ascent of sap in plants. Progress in Biophysics and Molecular conductance on stomatal conductance and xylem cavitation. Journal of
Biology 37: 181–229. Experimental Botany 44: 1075–1082.
Plavcova L, Jansen S, Klepsch M, Hacke UG. 2013. Nobody’s perfect: can Sperry JS, Hacke UG, Pittermann J. 2006. Size and function in conifer tracheids
irregularities in pit structure influence vulnerability to cavitation? Frontiers in and angiosperm vessels. American Journal of Botany 93: 1490–1500.
Plant Science 4: 1–6. Sperry JS, Tyree MT. 1990. Water-stress-induced xylem embolism in three species
Pounden E, Greene DF, Michaletz ST. 2014. Non-serotinous woody plants behave of conifers. Plant, Cell & Environment 13: 427–436.
as aerial seed bank species when a late-summer wildfire coincides with a mast year. Stambaugh MC, Smith KT, Dey DC. 2017. Fire scar growth and closure rates in
Ecology and Evolution 4: 3830–3840. white oak (Quercus alba) and the implications for prescribed burning. Forest
Rasmussen LA, Amman GD, Vandygriff JC, Oakes RD, Munson AS, Gibson KE. Ecology and Management 391: 396–403.
1996. Bark beetle and wood borer infestation in the greater Yellowstone area during Swezy DM, Agee JK. 1991. Prescribed-fire effects on fine-root and tree mortality in
four postfire years. Research Paper INT-RP-487. Odgen, UT, USA: USDA Forest old-growth ponderosa pine. Canadian Journal of Forest Research 21: 626–634.
Service, Intermountain Forest and Range Experiment Station. Sword MA, Haywood JD. 1999. Effects of crown scorch on longleaf pine fine roots.
Reich PB, Abrams MD, Elllsworth DS, Kruger EL, Tabone TJ. 1990. Fire affects Tenth Biennial Southern Silvicultural Research Conference, Shreveport, LA,
ecophysiology and community dynamics of central Wisconsin oak forest USA: 223–227.
regeneration. Ecology 71: 2179–2190. Thompson MTC, Koyama A, Kavanagh KL. 2017. Wildfire effects on
Rosenberg B, Kemeny G, Switzer RC, Hamilton TC. 1971. Quantitative evidence physiological properties in conifers of central Idaho forests, USA. Trees 31: 545–
for protein denaturation as the cause of thermal death. Nature 232: 471–473. 555.
Ryan KC. 1993. Effects of fire-caused defoliation and basal girdling on water relations Tombesi S, Nardini A, Frioni T, Soccolini M, Zadra C, Farinelli D, Poni S,
and growth of ponderosa pine. Dissertation, University of Montana, Missoula, MT, Palliotti A. 2015. Stomatal closure is induced by hydraulic signals and maintained
USA. by ABA in drought-stressed grapevine. Scientific Reports 5: 12449.
Ryan KC. 2000. Effects of fire injury on water relations of ponderosa pine. Fire and Tunstall BR, Walker J, Gill MA. 1976. Temperature distribution around synthetic
forest ecology: innovative silviculture and vegetation management. Tall Timbers Fire trees during grass fires. Forest Science 22: 269–276.
Ecology Conference Proceedings, no. 21. Tallahassee, FL, USA: Tall Timbers Tyree MT, Davis SD, Cochard H. 1994a. Biophysical perspectives of xylem
Research Station, 58–66. evolution: is there a tradeoff of hydraulic efficiency for vulnerability to
Ryan KC, Frandsen WH. 1991. Basal injury from smoldering fires in mature Pinus dysfunction. IAWA Journal 15: 335–360.
ponderosa Laws. International Journal of Wildland Fire 1: 107–118. Tyree MT, Kolb KJ, Rood SB, Pati~ no S. 1994b. Vulnerability to drought-induced
Ryan KC, Peterson DL, Reinhardt ED. 1988. Modeling long-term fire-caused cavitation of riparian cottonwoods in Alberta: a possible factor in the decline of the
mortality of Douglas-fir. Forest Science 34: 190–199. ecosystem? Tree Physiology 14: 455–466.

New Phytologist (2019) 223: 1728–1741 Ó 2019 The Authors


www.newphytologist.com New Phytologist Ó 2019 New Phytologist Trust
New
Phytologist Tansley review Review 1741

Tyree MT, Sperry JS. 1988. Do woody-plants operate near the point of catastrophic Wallin KF, Kolb TE, Skov KR, Wagner MR. 2003. Effects of crown scorch on
xylem dysfunction caused by dynamic water-stress? Answers from a model. Plant ponderosa pine resistance to bark beetles in northern Arizona. Environmental
Physiology 88: 574–580. Entomology 32: 652–661.
Tyree MT, Zimmermann MH. 2002. Xylem structure and the ascent of sap. Berlin, West AG, Nel JA, Bond WJ, Midgley JJ. 2016. Experimental evidence for heat
Germany: Springer. plume-induced cavitation and xylem deformation as a mechanism of rapid post-
Valor T, Casals P, Altieri S, Gonza lez-Olabarria JR, Pique M, Battipaglia G. 2018. fire tree mortality. New Phytologist 211: 828–838.
Disentangling the effects of crown scorch and competition release on the Westlind DJ, Kelsey RG. 2019. Predicting post-fire attack of red turpentine or
physiological and growth response of Pinus halepensis Mill. using d13C and d18O western pine beetle on ponderosa pine and its impact on mortality probability in
isotopes. Forest Ecology and Management 424: 276–287. Pacific Northwest forests. Forest Ecology and Management 434: 181–192.
Valor T, Gonza lez-Olabarria JR, Pique M. 2015. Assessing the impact of prescribed Wheeler JK, Sperry JS, Hacke UWEG, Hoang N. 2005. Inter-vessel pitting and
burning on the growth of European pines. Forest Ecology and Management 343: cavitation in woody Rosaceae and other vesselled plants. Plant, Cell &
101–109. Environment 28: 800–812.
Valor T, Orme~ no E, Casals P, Niinemets U. € 2017. Temporal effects of prescribed Wiley E, Rogers BJ, Hodgkinson R, Landh€ausser SM. 2016. Nonstructural
burning on terpene production in Mediterranean pines. Tree Physiology 37: 1622– carbohydrate dynamics of lodgepole pine dying from mountain pine beetle attack.
1636. New Phytologist 209: 550–562.
Van Wagner CE. 1973. Height of crown scorch in forest fires. Canadian Journal of Wood DL. 1982. The role of pheromones, kairomones and allomones in the host
Forest Research 3: 373–378. selection and colonization behavior of bark beetles. Annual Review of Entomology
Vargaftik NB, Volkov BN, Voljak LD. 1983. International tables of the surface 27: 411–446.
tension of water. Journal of Physical and Chemical Reference Data 12: 817–820. Woolley T, Shaw DC, Ganio LM, Fitzgerald S. 2012. A review of logistic regression
Varner JM, Putz FE, Brien JJO, Hiers JK, Robert J, Gordon DR. 2009. Post-fire models used to predict post-fire tree mortality of western North American
tree stress and growth following smoldering duff fires. Forest Ecology and conifers. International Journal of Wildland Fire 21: 1–35.
Management 258: 2467–2474. Zweifel R, Sterck F. 2018. A conceptual tree model explaining legacy effects on stem
Venturas MD, Sperry JS, Hacke UG. 2017. Plant xylem hydraulics: what we growth. Frontiers in Forests and Global Change 1: 1–9.
understand, current research, and future challenges. Journal of Integrative Plant Zwieniecki MA, Holbrook NM. 2009. Confronting Maxwell’s demon: biophysics
Biology 59: 356–389. of xylem embolism repair. Trends in Plant Science 14: 530–534.
Wagener WW. 1961. Guidelines for estimating the survival of fire-damages trees in
California. Miscellaneous Paper 60. Berkeley, CA, USA: USDA Forest Service,
Pacific Southwest Forest and Range Experiment Station.

New Phytologist is an electronic (online-only) journal owned by the New Phytologist Trust, a not-for-profit organization dedicated
to the promotion of plant science, facilitating projects from symposia to free access for our Tansley reviews and Tansley insights.

Regular papers, Letters, Research reviews, Rapid reports and both Modelling/Theory and Methods papers are encouraged.
We are committed to rapid processing, from online submission through to publication ‘as ready’ via Early View – our average time
to decision is <26 days. There are no page or colour charges and a PDF version will be provided for each article.

The journal is available online at Wiley Online Library. Visit www.newphytologist.com to search the articles and register for table
of contents email alerts.

If you have any questions, do get in touch with Central Office (np-centraloffice@lancaster.ac.uk) or, if it is more convenient,
our USA Office (np-usaoffice@lancaster.ac.uk)

For submission instructions, subscription and all the latest information visit www.newphytologist.com

Ó 2019 The Authors New Phytologist (2019) 223: 1728–1741


New Phytologist Ó 2019 New Phytologist Trust www.newphytologist.com

View publication stats

You might also like