You are on page 1of 32

PY55CH20-Wang ARI 13 June 2017 13:36

Review in Advance first posted online


V I E W
E on June 21, 2017. (Changes may
R

still occur before final publication

S
online and in print.)

C E The Candidatus
I N

A
D V A
Liberibacter–Host Interface:
Insights into Pathogenesis
Mechanisms and Disease
Control
Nian Wang,1 Elizabeth A. Pierson,2
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

João Carlos Setubal,3 Jin Xu,1 Julien G. Levy,2


Yunzeng Zhang,1 Jinyun Li,1 Luiz Thiberio Rangel,3
and Joaquim Martins Jr.3
1
Citrus Research and Education Center, Department of Microbiology and Cell Science,
Institute of Food and Agricultural Sciences, University of Florida, Lake Alfred, Florida 33850;
email: nianwang@ufl.edu
2
Department of Horticultural Sciences, Texas A&M University, College Station, Texas 77843
3
Department of Biochemistry, Institute of Chemistry, University of São Paulo, São Paulo,
SP, 05508-000, Brazil

Annu. Rev. Phytopathol. 2017. 55:20.1–20.32 Keywords


The Annual Review of Phytopathology is online at Huanglongbing, zebra chip, citrus, potato, pathogenicity, disease control
phyto.annualreviews.org

https://doi.org/10.1146/annurev-phyto-080516- Abstract
035513
“Candidatus Liberibacter” species are associated with economically devas-
Copyright  c 2017 by Annual Reviews. tating diseases of citrus, potato, and many other crops. The importance of
All rights reserved
these diseases as well as the proliferation of new diseases on a wider host
range is likely to increase as the insects vectoring the “Ca. Liberibacter”
species expand their territories worldwide. Here, we review the progress
on understanding pathogenesis mechanisms of “Ca. Liberibacter” species
and the control approaches for diseases they cause. We discuss the Liberib-
acter virulence traits, including secretion systems, putative effectors, and
lipopolysaccharides (LPSs), as well as other important traits likely to con-
tribute to disease development, e.g., flagella, prophages, and salicylic acid
hydroxylase. The pathogenesis mechanisms of Liberibacters have been dis-
cussed. Liberibacters secrete Sec-dependent effectors (SDEs) or other vir-
ulence factors into the phloem elements or companion cells to interfere
with host targets (e.g., proteins or genes), which cause cell death, necro-
sis, or other phenotypes of phloem elements or companion cells, leading
to localized cell responses and systemic malfunction of phloem. Receptors

20.1

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

on the remaining organelles in the phloem, such as plastid, vacuole, mitochondrion, or endoplas-
mic reticulum, interact with secreted SDEs and/or other virulence factors secreted or located on
the Liberibacter outer membrane to trigger cell responses. Some of the host genes or proteins
targeted by SDEs or other virulence factors of Liberibacters serve as susceptibility genes that
facilitate compatibility (e.g., promoting pathogen growth or suppressing immune responses) or
disease development. In addition, Liberibacters trigger plant immunity response via pathogen-
associated molecular patterns (PAMPs, such as lipopolysaccharides), which leads to premature cell
death, callose deposition, or phloem protein accumulation, causing a localized response and/or
systemic effect on phloem transportation. Physical presence of Liberibacters and their metabolic
activities may disturb the function of phloem, via disrupting osmotic gradients, or the integrity
of phloem conductivity. We also review disease management strategies, including promising new
technologies. Citrus production in the presence of Huanglongbing is possible if the most promis-
ing management approaches are integrated and approaches are discussed in the context of local,
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

area-wide, and regional Huanglongbing/Asian Citrus Psyllid epidemiological zone. For zebra chip
disease control, aggressive psyllid management enables potato production, although insecticide
resistance is becoming an issue. Meanwhile, new technologies such as clustered regularly inter-
spaced short palindromic repeat (CRISPR)-derived genome editing provide an unprecedented
opportunity to provide long-term solutions.

INTRODUCTION
Diseases caused by “Candidatus Liberibacter” species have gained recent worldwide notoriety.
Initially, this was due to the rapid and devastating proliferation of citrus Huanglongbing (HLB),
which nearly destroyed the citrus industry in Florida in the United States and heavily impacts
citrus production in most citrus-producing regions of the world. Outbreaks of zebra chip (ZC)
on potatoes in the Americas and New Zealand and of diseases on other economically impor-
tant vegetable crops further raised awareness of the importance of “Ca. Liberibacter” species as
pathogens. Currently, “Ca. Liberibacter” species are associated with diseases of crops in a limited
number of families (Table 1). Although many different isolates have been sequenced or otherwise
distinguished, e.g., by 16S rRNA sequencing, to date only one member of the genus, Liberibacter
crescens, has been cultured. Despite Koch’s postulates not having been satisfied because of the
inability to culture the putative pathogens, ample evidence has established Liberibacters as the
causative agents of HLB, ZC, and certain other plant diseases (Table 1). Hereafter, we refer to all
candidate members of the genus as Liberibacters. In all cases to date, specific psyllids have been
identified as the Liberibacter vectors (Table 1).
The dependence of Liberibacters on a highly mobile insect vector dramatically complicates
disease epidemiology and management. The development of diseases caused by Liberibacters
relies on more than the temporal and spatial convergence of a virulent pathogen, susceptible
host, and disease-conducive environment, which is referred to as the disease triangle (56). The
epidemiological outcomes for Liberibacter-associated plant diseases are determined by how vari-
ous environmental factors affect plant-Liberibacter-psyllid interactions and collectively are better
described by a disease pyramid, with the insect vector as an additional obligatory component
(Figure 1). It is now apparent that Liberibacters caused HLB in citrus well before their
twentieth century identification as the cryptic assassin. New diseases caused by Liberibacters,
such as ZC, are being discovered with alarming frequency. It is likely that deeper investigation

20.2 Wang et al.

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

Table 1 Diseases caused by Liberibacters, their hosts, and distributions


Pathogen Diseases Plant hosts Insect vector Distribution Reference
Candidatus HLB Citrus and citrus Asian citrus Widespread in most 20
Liberibacter relatives psyllid citrus-producing areas
asiaticus (Las) (Diaphorina of Asia, Africa, and the
citri ) Americas
Ca. L. africanus HLB Citrus and citrus African citrus South Africa 58, 82
(Laf ) relatives psyllid (Trioza
erytreae)
Ca. L. africanus Unknown Cape chestnut African citrus South Africa 59
subsp. capensis (Calodendrum capense) psyllid
(LafC) (T. erytreae)
Ca. L. africanus Unknown Horsewood (Clausena) African citrus South Africa 147
subsp. clausenae psyllid
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

(LafCl) (T. erytreae)


Ca. L. africanus Unknown Small forest knobwood African citrus South Africa 148
subsp. zanthoxyli (Zanthoxylum) psyllid
(LafZ) (T. erytreae)
Ca. L. africanus Unknown White ironwood (Vepris) African citrus South Africa 148
subsp. vepridis psyllid
(LafV) (T. erytreae)
Ca. L. africanus Unknown Flaky cherry-orange African citrus South Africa 148
subsp. tecleae (Teclea gerrardii ) psyllid
(LafT) (T. erytreae)
Ca. L. americanus HLB Citrus and citrus Asian citrus Brazil 167
(Lam) relatives psyllid (D. citri )
Ca. L. solanacearum ZC Solanaceous crops Bactericera Central America, 66, 100,
(Lso) cockerelli western Mexico, 128
Haplotype A western United States,
New Zealand
Lso haplotype B ZC Solanaceous crops B. cockerelli Eastern Mexico, central 128
United States
Lso haplotype C Yellows decline Carrot Trioza apicalis Finland, Sweden, 128
and vegetative France, Norway,
disorders Netherlands, Germany
Lso haplotype D Yellows decline Carrot Bactericera Spain, Morocco 129
and vegetative trigonica
disorders
Lso haplotype E Vegetative Celery and carrots B. trigonica Spain, France, Morocco 166, 169,
disorders (likely) 170
Ca. L. europaeus Asymptomatic Rosaceae plants, Cacopsylla pyri Italy, Hungary 143
(Leu) including apple (Malus (L.)
domestica), blackthorn
(Prunus spinosa),
hawthorn (Crataegus
monogyna), and pear
(Pyrus)
(Continued )

www.annualreviews.org • Liberibacter Pathogenesis Mechanisms and Disease Control 20.3

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

Table 1 (Continued )
Pathogen Diseases Plant hosts Insect vector Distribution Reference
Leu Stunted growth Scotch broom Broom psyllid New Zealand, Australia 173
of shoots, (Arytainilla (?), Europe (?)
shortened spartiophila)
internodes, leaf
dwarfing, and
leaf tip chlorosis
Liberibacter crescens Unknown Mountain papaya (Carica Unknown Puerto Rico 94
(Lcr) stipulata × C. pubescens)

Abbreviations: HLB, Huanglongbing; ZC, Zebra chip.

will reveal more Liberibacter species and that the inevitable introduction of Liberibacters into
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

more plant/psyllid vector combinations will continue to give rise to new diseases in other crops.
Expansion of Liberibacter distribution and associated diseases has been and will continue to be
facilitated by the introduction of the vector and/or susceptible plant species to new geographic
regions, such as occurred with the introduction of the potato/tomato psyllid into New Zealand
(172). Future disease expansions also may be aided by increases in the geographic and host ranges
of the psyllids, which are associated with the movement of plants and changes in psyllid behaviors,
as described below.
The rapid proliferation of plant diseases caused by Liberibacters and the devastating economic
consequences associated with them create urgency and present challenges for disease control.
The proliferation of Liberibacter diseases also presents unique opportunities to study plant-insect-
pathogen coevolution contributing to their origin and the plant response, and pathogenesis mech-
anisms underlying disease symptomology. Tremendous progress has been made in understanding
the ecological and evolutionary underpinnings of the Liberibacter disease pyramid. Many of these
observations have been summarized previously in several excellent reviews (20, 37, 61, 63, 65,
103, 181). In particular, Bové (20) has provided an insightful historical review of HLB. In this
review, we focus our efforts on the biology of Liberibacters within the plant host. We include a
genomic comparison of Liberibacters and summarize current findings on potential pathogenicity
mechanisms with an aim toward improvements in disease control.

Plant host
(e.g., citrus and potato)

Psyllid
Environment

Candidatus Liberibacter
Figure 1
The disease pyramid for Candidatus Liberibacter–associated plant diseases.

20.4 Wang et al.

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

DISEASES CAUSED BY LIBERIBACTERS AND WORLDWIDE


DISTRIBUTION
Multiple species of “Candidatus Liberibacter” have been identified from plants grown worldwide
(Table 1). These include “Ca. L. asiaticus” (Las), “Ca. L. africanus” (Laf ) (82), and “Ca. L.
americanus” (Lam) (167), which affect citrus and cause HLB (also called citrus greening); “Ca. L.
solanacearum” (Lso; a.k.a. “Ca. L. psyllaurous”), which causes ZC of potatoes and diseases on many
other solanaceous hosts (66, 100, 101) as well as on carrots in Finland, Sweden, Norway, Spain,
and Morocco and on celery in Spain (118, 120); “Ca. L. europaeus” (Leu), which causes disease
on scotch broom (173) and is also found in multiple Rosaceae plants, including apple, blackthorn,
hawthorn, and pear, without causing disease symptoms (143); and Liberibacter crescens (Lcr) found
in the tropical Babaco plant (also known as the hybrid mountain papaya) (94) (Table 1).
Among the diseases caused by “Ca. Liberibacter” spp., HLB and ZC spread worldwide and
caused dramatic economic losses, whereas other diseases may just be starting to gain such promi-
nence. We focus on HLB and ZC and the Liberibacters responsible for these diseases.
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

Huanglongbing
HLB is caused by Las, Lam, and Laf, with Las and Lam transmitted by the Asian citrus psyllid
(ACP; Diaphorina citri ) and Laf by the African citrus psyllid (Trioza erytreae). How Las estab-
lished its association with citrus remains unknown. It has been hypothesized that Laf was present
on the African continent before the introduction of citrus, possibly in an indigenous Rutaceous
species in Africa such as Calodendrum capense, Clausena anisate, Vepris lanceolate, or Zanthoxylum
capense (59, 135, 148). Liberibacters closely related to Laf, i.e., LafC, LafCl, LafV, and LafZ,
have been identified from the aforementioned native rutaceous hosts of T. erytreae in South Africa
(Table 1) (59, 148). Similarly, Lam might have evolved from a Liberibacter resembling Leu in-
troduced into Brazil from Europe via infected plants, considering ACP was first reported in Brazil
in 1942, whereas Lam was first found on citrus in Brazil in 2004. As shown in Figure 2, Lam
and Leu are closely related. The establishment of T. erytreae, which can transmit both Las and
Laf (20), in mainland Europe (30) raises the question of whether Leu or other Liberibacters will
ultimately evolve to infect citrus in the Mediterranean area, one of the major citrus-producing
areas yet without HLB.
Even though Las, Laf, and Lam have followed different evolutionary paths to establish their
associations with citrus and their genomic sequences are quite different (41, 105, 185), the symp-
toms caused by the three Liberibacters on citrus are similar. Las was reported to cause more severe
symptoms than Laf (31). Characteristic symptoms associated with HLB include yellow shoots;
blotchy mottling leaves; upright, hardened, and small leaves; leaves showing zinc deficiency and
corky vein; twig dieback; stunted growth, and tree decline. The canopy of HLB-diseased trees be-
comes thinner. Off-season flowering, smaller or lopsided fruit with aborted seeds, and preharvest
fruit drop are frequently observed with HLB-diseased trees. Root degradation is also commonly
observed (87).
HLB, D. citri, and T. erytreae have expanded rapidly into most citrus-producing areas world-
wide. Currently, HLB is confirmed in most citrus-producing areas in Asia, the Americas, and
Africa, although HLB has not been identified in the Mediterranean or Australian citrus-producing
areas. However, the presence of Las and ACP in Papua New Guinea (39) and Indonesia poses a
threat to the Australian citrus industry. Establishment of Las in Ethiopia, Mauritius, and Réunion
in Africa also threatens the African citrus industry. African citrus psyllids have been found in Spain
(30). T. erytreae also can transmit Las (115). In both Brazil and the United States, citrus psyllids
invaded the citrus groves and became established, followed by Las invasion.

www.annualreviews.org • Liberibacter Pathogenesis Mechanisms and Disease Control 20.5

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

Agrobacterium tumefaciens (D14500.1)


Rhizobium leguminosarum SEMIA 2083 (FJ025096.1) Outgroup
Rhizobium etli CFN 42 (NR074499.1)
Lcr BT0
Lcr
NR102476.1
FJ263693.1
Lam
13

FN678792.1
JX244258.1
JX244259.1 Leu
JX244260.1
27 Laf
GQ468843.1
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

100 Lso

GU991651.1
FJ871059.1
JN245974.1
DQ431997.1

DQ432004.1
6

23

2
KP063235.1
FJ811891.1
FJ811892.1 Las
FJ811893.1
FJ811894.1
JN245975.1
FJ811895.1
JN245978.1

276
0.01
Tree scale JQ867418.1
EU265646.1
EU371107.1
JN245979.1

Figure 2
The maximum-likelihood (ML) phylogenetic tree of the Candidatus Liberibacter genus based on 16S rDNA sequences. The high-
quality 16S rDNA sequences (>1,000 bp, totaling 478 sequences, including the three outgroups) extracted from the NCBI (National
Center for Biotechnology Information) nucleotide database were used for phylogenetic analysis. The sequences were aligned using
MUSCLE (multiple sequence comparison by log-expectation) (44), and the generated alignment was trimmed using Gblocks Version
0.91b (25). The ML tree was constructed using FastTree 2 (141) with gtr (general time reversible) and gamma options. The tree file was
midpoint rooted using iTOL (interactive tree of life) (95). The branches with lengths shorter than 0.01 were collapsed, with exceptions
[such as Lcr and Ca. L. europaeus (Leu clade), and the node number inside the collapsed branch was indicated. The bootstrap supports
for the branches were colored red to green (0–1). “Ca. Liberibacter psyllaurous” was located in the “Ca. L. solanacearum” (Lso) clade.

20.6 Wang et al.

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

Zebra Chip and Other Diseases Caused by “Candidatus Liberibacter


solanacearum”
ZC is the best characterized disease caused by Lso, and currently all commercial potato (Solanum
tuberosum) varieties are susceptible. The name describes the major tuber symptom, which is a
brown discoloration of the phloem and medullary rays that is often apparent upon slicing fresh
tubers but intensifies with frying. This defect results in potato tubers that are unsuitable for
production of potato chips or French fries. ZC was reported first in Mexico in 1994, in Texas
in 2000, and in the Pacific Northwest region of the United States in 2010 (34, 36). ZC is also a
problem in Central America, including Guatemala and Honduras, and in New Zealand (1, 100).
The symptoms of ZC are visually apparent first in the aboveground parts of the plants and include
curling, purpling, and/or chlorosis of leaves. As the disease progresses, plants appear stunted,
develop swollen internodes, and may form aerial tubers. In addition to the tuber defects, severe
disease reduces tuber production and often causes plant death (118). The bacteria may be passed
to the next generation of plants via infection of tubers used as seed pieces (137). Lso was confirmed
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

as the causative agent of ZC based on graft transmission (35, 152).


Currently, five Lso haplotypes have been identified (Lso A, Lso B, Lso C, Lso D, and Lso E)
infecting different crops (8, 104, 120, 169) (Table 1). The Lso haplotypes were identified using
single nucleotide polymorphism genotyping of the 16S rRNA, 16S/23S ISR, and the 50S rplJ and
rplL ribosomal protein genes (128); multilocus sequence typing markers (60); and simple sequence
repeats (104). Lso A and Lso B are vectored by the tomato/potato psyllid Bactericera cockerelli and
cause ZC (101, 102). Both haplotypes are present in the Americas and may be found separately or
together in vectors and plant hosts. Lso A currently is the only haplotype found in New Zealand.
In addition to ZC, Lso A and B cause diseases of other economically important solanaceous hosts,
including tomato, pepper, eggplant, tamarillo, tobacco, and cape gooseberry (Physalis peruviana) in
North and Central America and New Zealand and are probably transmitted to these solanaceous
hosts by B. cockerelli. Lso may be seed transmitted in most of these crops, having been demonstrated
to do so in chili pepper (23). Other weedy solanaceous species are known to host Lso A and B,
showing reduced or no disease symptoms (101, 171, 182).
Lso C and Lso D cause disease in carrot in Europe and Africa, where they are vectored by the
carrot psyllids Trioza apicalis and Bactericera trigonica, respectively (8, 120). Haplotype E causes
disease in carrot and celery in Spain but no confirmed psyllid vector has been described (169).
Recent reports suggest that in Spain, B. trigonica, Bactericera tremblayi, and Bactericera nigricornis
as well as another Bactericera species carry Lso and could be vectors (170). Lso can also be seed
transmitted in both carrot and celery (16, 169).

LIBERIBACTERS IN PLANTA: CLUES TO UNDERSTANDING


LIBERIBACTER PATHOGENICITY
One of the hallmarks of diseases caused by Liberibacter is the relatively consistent symptomology
that occurs among different symptom-expressing hosts. Interestingly, this pattern of symptomol-
ogy is regularly observed in citrus and potato when infected by other phloem-limited pathogens.
For example, citrus stubborn disease caused by Spiroplasma citri (158) and citrus disease caused by
“Ca. Phytoplasma” (168) produce symptoms similar to HLB (20). Recognition of ZC in potato is
complicated by the similarities of the early symptoms to potato purple top caused by “Ca. Phyto-
plasma” (93) and potato leafroll virus (35). Heavy psyllid feeding itself causes a condition referred
to as psyllid yellows that also results in leaf chlorosis and purpling, shortened internodes, the for-
mation of aerial tubers, and reduced tuber yields (154). These observations support the hypothesis

www.annualreviews.org • Liberibacter Pathogenesis Mechanisms and Disease Control 20.7

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

that disruption of phloem function, regardless of the pathogen, mechanism, or host, may result in
the expression of consistent disease symptoms.
Another hallmark of diseases caused by Liberibacters is the cryptic nature of the pathogen
and the inconsistent relationship between pathogen titer in a particular tissue and symptomology.
Indeed much effort has been invested in the detection of this often obscure pathogen in planta.
Understanding the movement of Liberibacters in planta and the relationship between Liberib-
acter titer and disease symptomology is critical to unveiling how Liberibacters cause disease and
designing suitable disease control strategies. Evidence suggests that in plants, Liberibacters reside
solely within phloem tissues. For example, Las has been reported to be restricted to the inside of
sieve tubes (54, 73); however, one study also found Las in companion cells (57). Other Liberib-
acters such as Lso also live inside the sieve elements but have not been observed in associated
companion or mesophyll cells (152, 169).
Systemic infection of the host plant follows the direction of phloem sap translocation, moving
from sources (e.g., mature leaves) to sinks (e.g., roots, tubers, young leaves, flushes, and fruit),
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

although how this is accomplished and the speed of movement are subject to debate. Despite the
presence of structural genes for the production of pili and flagella in the genomes of all Liberibacter
(20, 41), these structures have yet to be observed on any Liberibacter in planta. This suggests that
passive movement occurs with the mass flow of phloem sap, which explains the temporal patterns
of downward and upward movement to roots and flushes and young leaves and fruit, respectively.
Microscopic analysis of Las in planta has shown that Las cells float freely in the phloem sap without
attaching to the sieve tube cell walls or each other (i.e., without forming biofilms) (73, 90, 91).
Similar observations have been made of Lso in planta (131). Circumstantial evidence suggests that
Liberibacter movement occurs primarily in the vertical direction along the sieve tubes through the
sieve pores rather than horizontally to adjacent sieve tubes. This hypothesis was first suggested by
observations of the uneven colonization of Las in dodder, where adjacent phloem vessel elements
were observed to be either completely full of Las or free of the pathogen (70). It is likely that sieve
plate pores with diameters ranging from <1 μm to ≈14 μm (46) do not prevent Liberibacters
from passing between sieve elements. For example, the diameter of Las has been reported to be
approximately 0.1 to 0.3 μm in citrus phloem (20, 73). Both Las and Lso have been observed
traversing the sieve plate pores (54, 131, 159).
The direction and rate of movement of Liberibacters within the host plant have been studied.
In growth chamber experiments with potato and tomato plants, Levy et al. (96) found that Lso
spread from the site of Lso infection upon psyllid feeding to the main stem within seven days.
Subsequent work demonstrated Lso dispersal out of the infected leaf occurred within 24 hours
(24). The movement of the bacteria within the plant followed the direction of translocation of
carbohydrates from the source leaves to the sinks, resulting in bacterial titers that were measurable
first in the newly developing leaves. Rapid accumulation of Lso following infection also occurred
in other sinks, including the roots and tuber stolons (182). After Lso inoculation of potato, Lso
was confirmed in the inner phloem above the export leaf and in the outer phloem below the
export leaf, consistent with upward movement of phloem sap in the inner phloem and downward
movement of phloem sap in the outer phloem (33) of the bicollateral vascular system of solanaceous
plants.
In citrus, it has been estimated that the speed at which sap moves in the phloem approximates
3.94 μm/s, or 0.34 m/day (47). However, Las movement has been estimated to be approximately
2–3 cm/day in citrus, i.e., at reduced speed compared to the mass flow of phloem sap (S. Lopes,
personal communication). The apparent difference in Las movement and mass flow of phloem sap
is a topic that needs to be addressed. Detection of Las in citrus foliar tissues has been problematic
because of the spatial and seasonal patterns of pathogen movement in the plant. Once Las arrives

20.8 Wang et al.

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

at the roots, it multiplies and serves as a source of inoculum to young flushes or fruit and is
a relatively reliable site of pathogen detection (87). The movement of Las within the plant is
limited by plant sectoriality, resulting from the subdivision of physically coherent vasculature.
Many plants exhibit sectoriality when physiological activities are integrated within some parts of
their structures, but the same activities display a strong degree of independence at other structural
scales because of vascular structure (179). Sectoriality leads to segmented patterns of infection and
symptom development typically observed in early citrus infections. When large populations of
ACP infect multiple branches simultaneously, the sectorial pattern of HLB symptoms disappears.
The expression of disease symptoms typically lags plant infection by several months to several
years, which is very problematic for disease management.
In field-grown potatoes, although there is typically a 2–3-week lag between plant infection and
symptom development, there is a strong correlation between the timing of plant infection and
the severity of tuber disease symptoms. Symptoms may be evident in the tubers within 3 weeks of
exposure of the plants to Lso-infected psyllids, which may proceed or coincide with the initiation
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

of foliar symptoms (22, 96). The cessation of tuber development typically coincides with the
development of disease symptoms, and tubers may not form if the plants are infected before tuber
initiation (22). However, even late-season infections are economically important for growers.
Rashed et al. (145) demonstrated that tubers of plants infected just four days before harvest may
become infected despite the plants and tubers being asymptomatic and testing negative for Lso at
harvest. Bacterial titers were found to increase in many of these tubers to detectable levels during
cold storage, and tubers after storage were as defective in sprouting as tubers showing symptoms
before storage (145). Of significance, recent findings demonstrated that plants infested with psyllids
only 48 hours before vine kill still developed symptoms in tubers left to mature in the ground for
one month (149). Collectively, these findings demonstrate the rapidity of systemic infection by Lso
and the susceptibility of plants to infection throughout the growing season. They also demonstrate
that disease symptoms in specific tissues such as tubers may develop independently from other
plant organs, as demonstrated by the progressive infection and symptom development in tubers
after vine kill and in cold storage.
The previous discussion demonstrates that Liberibacters probably live solely within phloem
tissues and are capable of rapid, systemic infection that is consistent with the sectorality of the plant
vasculature. Even though most symptoms caused by Liberibacters seem to result from disruption
of phloem function, organ-specific plant response to infection has been observed (e.g., symptom
development in detached tubers).

GENOMIC FEATURES OF LIBERIBACTERS


Genome analysis of Liberibacters has helped understand the biology, evolution, and virulence
mechanism of the pathogens. To date, 18 genomes of 5 Liberibacter species, i.e., Las (8 genomes),
Laf (1 genome), Lam (2 genomes), Lso (5 genomes), and Lcr (2 genomes) have been sequenced
(Supplemental Table 1). The genome sizes of Liberibacters range from 1.1 to 1.45 Mb with
a low GC content, approximately 35.3%. The average nucleotide identity (ANI) values between
different strains of the same species are 99.6% to 99.97%, demonstrating highly conserved genomic
sequence among these strains. However, among the different species, the ANI values were only
67.94% to 80.79%, indicating the highly diverse genomic traits among the different Liberibacter
species.
The pan-genome size of all Liberibacters is 2,394 genes, and the core-genome size is 305 genes.
The pan-genome plot does not reach a plateau, suggesting an open pan-genome and the need
for additional genome sequencing. Some Lso haplotypes infecting carrot or celery and the Leu

www.annualreviews.org • Liberibacter Pathogenesis Mechanisms and Disease Control 20.9

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

genome have yet to be sequenced. Sequencing of those aforementioned Liberibacter species will
help us gain a better understanding of the genomic features and their virulence mechanisms.
Among the 2,394 shared genes, only 619 have at least one representative in each of the
five species, which suggests a relatively low level of inter-species gene sharing. This is also sug-
gested by the high level of gene specificity: 1,506 genes of the pan genome are species specific,
meaning that almost two-thirds of all genes in the pan genome are not shared among different
species of the genus. Even though there is variation in gene specificity (the two extremes being
45.5% in Lcr and 16.0% in Laf ), we believe a general hypothesis for the high level of specificity
is the obligate intracellular lifestyle of all members of the genus, which creates a barrier for gene
exchange (138). The different hosts also represent different environments, and each one could
have specific selective pressures.
By using the clustered regularly interspaced short palindromic repeat (CRISPR) recognition
tool (17), possible CRISPR arrays have been found in most of the 18 sequenced Liberibacter
genomes (Supplemental Table 1) ( J. Xu & N. Wang, unpublished data) (194). The CRISPR
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

system is a prokaryotic immune system that confers resistance against phages and plasmids and
is present in most archaea and many bacteria (13). A complete CRISPR/Cas system consists of a
CRISPR array, which contains short direct repeats separated by short variable spacer sequences,
and diverse Cas genes located adjacent to this array (13). The CRISPR array and Cas genes have
been predicted in Las (194). Six of the 18 genomes contain a type I CRISPR/Cas system. Although
most CRISPR/Cas systems have been found in chromosomes, the CRISPR/Cas systems of many
Liberibacters are located on a prophage, especially in Las, thus suggesting that CRISPR/Cas might
contribute to the dominance of a single prophage in Las strains (194), as described in below. The
limited number of spacers in the CRISPR/Cas system is consistent with the limited exposure of
Liberibacters to phage and plasmids either in the phloem or in psyllids.

PATHOGENESIS MECHANISMS OF LIBERIBACTERS


With the exception of Lcr (94), Liberibacters have not been cultured in artificial media; therefore,
traditional molecular and genetic analyses have been difficult to perform. This difficulty has greatly
hampered efforts to understand the virulence mechanisms of Liberibacters. To date, most insights
into Liberibacter pathogenesis mechanisms have been derived from analyses of Liberibacter ge-
nomic sequences (41, 48, 94, 105, 185), studies using surrogates such as Sinorhizobium (176) and
E. coli (140), and expression in planta as well as indirectly from host transcriptomic, proteomic,
and metabolomic responses to Las, Lam, and Lso infection (50, 81, 112–114, 160, 190, 195, 196).
In this section, we summarize the virulence traits and pathogenicity mechanisms of Liberibacters.
The host responses to Liberibacter infection based on transcriptome, proteome, and metabolome
approaches were recently reviewed (37, 181) and are not reviewed in detail here.

Virulence Traits of Liberibacters


Owing to the lifestyle in the phloem and reductive evolution, Liberibacters contain very few
virulence traits compared to free-living plant-pathogenic bacteria such as Xanthomonas and Pseu-
domonas. We focus on the recent progress in the study of Liberibacter secretion systems and
effectors, lipopolysaccharides, flagella, prophage, and salicylic acid, which have been suggested to
play important roles in causing disease symptoms, contributing to bacterial survival in planta, and
suppressing plant defenses.

20.10 Wang et al.

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

Extracellular space

Autotransporters β-barrel proteins OM vesicles

OM
β

Chaperones
N
Molecules
C other than
mature SDE
Periplasm
N proteins
Mature
SDE proteins C
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

SPase

IM YajC SecF SecD SecYEG YidC


SecA

ADP + Pi
SP
SecB ATP

Cytoplasm Preprotein

Figure 3
Candidatus Liberibacter spp. contain basic components of the Sec machinery. Putative Sec-dependent
effector (SDE) proteins (brown) have been predicted to be translocated via the Sec machinery. SecB is the
Sec-system-specific chaperone that channels the preprotein to the Sec translocation pathway, binding to
SecA. The preprotein-carrying SecA binds to the membrane. The SecYEG complex constitutes a channel
for polypeptide movement. Continued translocation requires ATP hydrolysis to provide energy force. Las
may utilize some noncanonical mechanisms to secrete proteins across the outer membrane. B-barrel proteins
in the outer membrane may have adopted the ability to transport proteins to the extracellular space. Some
SDEs proteins are autotransporters. In addition, some of these SDE proteins might be secreted in outer
membrane vesicles. Abbreviations: IM, inner membrane; OM, outer membrane; SP, signal peptide; SPase,
signal peptidases.

Secretion systems and putative effectors. Systems capable of secreting bacterial proteins called
effectors into host cells are among the most important virulence factors of bacterial pathogens.
Protein effectors often suppress plant defenses or manipulate developmental processes within the
host to benefit the pathogen (88). Interestingly, Liberibacters contain type I secretion systems
(T1SSs) and a complete general secretory pathway (Sec) (Figure 3) but lack other secretion
systems (41, 48, 105, 185). In addition to Sec, Lcr contains the twin arginine transport system,
which is not present in other Liberibacters.
Genes encoding T1SSs consisting of TolC, HlyD, and PrtD have been reported in Liberibac-
ters (41, 105). Based on analyses of 18 sequenced Liberibacter genomes (Supplemental Table 1),
a complete T1SS has been identified in Las, Laf, and Lso but not in Lam or Lcr. Putative T1SS

www.annualreviews.org • Liberibacter Pathogenesis Mechanisms and Disease Control 20.11

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

substrates, e.g., genes encoding serralysin and hemolysin, have been identified in Liberibacter
genomes (41, 105). A putative serralysin gene has been found in most sequenced Las, Lso, and
Laf genomes but not in Lam or Lcr, whereas a predicted hemolysin gene has been identified in
all sequenced Liberibacters. The functionality and contribution of T1SSs and their substrates to
the virulence of Liberibacters remain to be determined.
The Sec machinery facilitates the majority of protein transport across the cytoplasmic mem-
brane (Figure 3) and is essential for bacterial viability (153). The Sec apparatus also secretes
important virulence factors in some plant-pathogenic bacteria. For example, in phloem-dwelling
phytoplasmas, the Sec pathway is critical for the secretion of vital virulence factors, including
SAP54, SAP11, and Tengu, into plant cells (76, 111, 163). Importantly, phytoplasmas cause similar
symptoms as do Liberibacters (168). The genome sequences and virulence factors of phytoplasmas
that infect citrus remain to be determined but may reveal common effectors that are important
for the symptom development in diseases caused by both groups of pathogens.
It is unknown how Liberibacters translocate proteins from periplasm to the outside of the outer
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

membrane, where these proteins may be directly involved in host interactions. Interestingly, outer
membrane vesicles (OMVs) have been observed on the surface of Lso (131). OMVs have been
shown to secrete various virulence factors in other pathogens (19). It is probable that Liberibacters
use OMVs to secrete Sec-dependent effectors (SDEs) (Figure 3) and other putative virulence
factors.
Signal peptide (SP)-containing extracytoplasmic proteins (SPEPs), which are potential sub-
strates of the Sec apparatus, have been predicted for Las—i.e., Las_psy62 (166 SPEPs), Las_gxpsy
(168 SPEPs) and Las Ishi-1 (164 SPEPs) strains, Lam (133 SPEPs), Laf (141 SPEPs), Lso (171
SPEPs), and Lcr (214 SPEPs) (140). The three Las strains from the United States (Las_psy62),
China (Las_gxpsy), and Japan (Las Ishi-1) show high uniformity in their predicted Sec-dependent
extracytoplasmic proteins, with 151 proteins overlapping in all three strains. However, signifi-
cant differences have been observed among Las, Laf, and Lam, even though they all cause HLB.
Only 45 Sec-dependent extracytoplasmic proteins show homology between these species. To ex-
perimentally validate the predicted extracytoplasmic proteins, E. coli–based alkaline phosphatase
(PhoA) gene fusion assays have been conducted. A total of 86 of the 166 predicted Las proteins
were experimentally validated to contain an SP (140). In addition, 16 and 86 putative SDEs also
have been predicted by Pitino et al. (136) and Cong et al. (32), respectively. Among the genes
encoding SP-containing proteins, 36 have been shown to be highly expressed in planta compared
with in psyllids, whereas eight others are highly expressed in psyllids compared with in planta
(188), thus suggesting that those proteins may play critical roles in Liberibacter adaptation to
living in the plant and insect hosts.
Phenotypic changes have been observed via transient or transgenic expression of multiple
SDEs in planta, including in Nicotiana benthamiana and citrus (136) (N. Wang, unpublished data).
Specifically, Las5315 is localized in the chloroplast and induces cell death when it is transiently
expressed in N. benthamiana. It also induces callose deposition. How Las5315 contributes to the
virulence of Las in citrus remains unknown. Notably, homologs of Las5315 have not been found
in Lam (136) and other sequenced Liberibacters.
In addition to the SDEs, two type V autotransporters, LasAI and LasA, also rely on the Sec
machinery for transport and have been proposed to exist in Las. Neither LasAI nor LasA contains
a typical SP domain for Sec-dependent translocation (67). It remains to be determined whether
either protein is involved in the virulence of Las.

Lipopolysaccharides. Lipopolysaccharides (LPSs), which are composed of lipid A, an oligosac-


charide core, and an O-antigen, are critical components of the outer membranes of gram-negative

20.12 Wang et al.

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

bacteria. LPSs contribute to bacterial fitness and virulence by maintaining the cell structural in-
tegrity and protecting bacteria from certain chemical stresses, such as antimicrobial compounds.
Lipid A is somewhat conserved, whereas the oligosaccharide core and O-antigen are much more
variable. LPSs of bacterial pathogens often trigger plant immune responses, such as oxidative
bursts, salicylic acid (SA) accumulation, and callose deposition during pathogen-plant interactions
(197), whereas those of symbiotic bacteria may suppress plant defense responses as they invade
their partner host (7). Las induces callose deposition, which might be partially responsible for
phloem blockage in citrus (51, 90, 91). Whether the LPSs of Liberibacters trigger callose deposi-
tion remains to be determined. Interestingly, the Las genes encoding LPS synthesis and transport
are highly expressed in planta but not in the psyllid vector (188).
LPSs are synthesized on the inner membrane and typically require transporters, such as the
two ABC transporters MsbA and LptB2FG to be transported to the outer membrane (132). The
genes involved in the biosynthesis of the three LPS components are found in all the sequenced
Liberibacters except Lam (185). A majority of the LPS synthesis genes are missing in Lam, and
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

only two genes, lpxC and lpxD, are present (185). Loss of LPS has been observed previously in some
bacteria and may allow them to adapt to stressed niches (164). The different LPSs of Las, Lam,
and Laf might affect their induction of host defenses. LPSs have been reported to activate two of
the biosynthetic steps of the phenolic conjugates coumaroyltyramine (CT) and feruloyl-tyramine
(FT) (130). Lso causes an increase in the level of phenolics in infected potato tubers, which might
result from potato cells recognizing the LPSs (126). Las-induced phenolics in citrus also have
been reported (72). A lack of LPSs might allow Lam to avoid recognition by plant cells but may
also make Lam more susceptible to externally applied or plant-derived antimicrobial compounds.

Flagella. Bacterial flagella are responsible for rotational propulsion, promote host colonization
via adherence, and induce plant immunity. A nearly complete set of genes involved in the flagellar
synthesis pathway are present in the sequenced Liberibacter genomes (41, 48, 94, 105, 185).
Flagella have been observed on Lso cells in psyllids (29). Expression of all of these genes has
been documented in the transcriptome of Liberibacter-infected psyllids (53). For Las, flagella
have not been observed on the surface of bacterial cells in planta (20). The expression of genes
involved in flagellar assembly, including fliF, flgI, and flgD, and of the motB gene involved in motor
function is upregulated in planta. In contrast, flgL, flgK, and fliE are overexpressed in the psyllid
(188). The flagellar expression pattern suggests that Las may have a functional flagellum in some
environments. The Lso flagellin gene is also highly expressed in B. cockerelli (190). Unlike Las,
Lso, and other Liberibacters, the Lam genome does not contain flbT. FlbT functions as a key
regulatory checkpoint for flagellin gene expression, and flagellin is not expressed in flbT mutants
(52). The lack of flbT in Lam suggests that flagellar genes might not be expressed so as not to
activate plant defenses in planta (185).
Liberibacter flagella have been reported to stimulate plant defense. Las encodes a flagellin
that contains a conserved 22 amino acid domain (Flg22). The synthetic Flg22Las peptide does not
induce plant cell death in tobacco but has retained the ability to induce callose deposition at a
concentration of 20 μM or above (198). Transient expression of FlaLso in tobacco induces a delayed
increase in reactive oxygen species (ROS) production and slow necrotic cell death. The flg22Lso
peptide does not induce ROS production in tobacco, tomato, or potato but does induce callose
deposition in them (68). Flg22Lso and the flagellin FlaLso induce expression of genes associated
with PAMP-triggered immunity (PTI) in N. benthamiana (68).

Prophages. A prophage is a temperate bacteriophage genome that has integrated into the bac-
terial chromosome or exists as a plasmid. There are currently two known types of Las prophages,

www.annualreviews.org • Liberibacter Pathogenesis Mechanisms and Disease Control 20.13

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

Type 1 (SC1 or FP1-like) and Type 2 (SC2 or FP2-like) (41, 192). SC1 becomes replicative in
planta, and when lytic genes are expressed, bacteriophage particles are readily formed in periwin-
kle (192) and citrus (57). By contrast, SC2 is a replicative excision plasmid that lacks lytic genes.
An analysis of the 18 sequenced Liberibacter genomes (Supplemental Table 1) has revealed that
either Type 1 or 2 prophage is present in each strain. Interestingly, although most Las strains
contain a single prophage and a few strains contain both prophages, strain Ishi-1 from Japan (89)
does not contain any prophages. The pattern of a single prophage present in most Las strains also
has been observed in an extensive survey of Las isolates in Thailand and southern China (142,
194). Among strains observed in an extensive survey of Las isolates in China, it was typical for
Las to have a single prophage, with Guangdong isolates harboring mainly the type 2 prophage,
whereas isolates from Yunnan are dominated by the type 1 prophage (194). Notably, the Lso NZ1
contains three prophages (174).
Prophages have been suggested to assist Las in suppressing plant defenses (83). For exam-
ple, SC2-gp095 encodes a ROS-scavenging peroxidase, which is a predicted lysogenic conversion
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

factor. SC2-gp095 expression is suppressed in psyllids, but it is expressed at a high level in peri-
winkle. Expression of the peroxidase in Lcr, but not in E. coli, results in enhanced resistance to
H2 O2 . Nonclassical secretion has been predicted for SC2-gp095, and its secretion from Lcr has
been confirmed by enzymatic and western blot analyses. Transient expression of the peroxidase
in planta results in strong transcriptional downregulation of RbohB, the key gatekeeper of the
H2 O2 -mediated defense signaling in plants. It has been suggested that Las evades early detection
in the phloem by virtue of its secreted peroxidase activity, enabling it to build up a high titer
in citrus over several years before manifesting disease symptoms, thus resulting in the long in-
cubation period often associated with HLB progression (83). However, the lack of a prophage
in many Las strains does not appear to relate to the lack of HLB symptoms because Ishi-1 and
the Guangdong isolates, which do not contain any prophages, induce similar HLB symptoms as
isolates containing prophages (89, 194). Overall, this evidence suggests that prophages contribute
to bacterial virulence but are not required for Las pathogenicity.

Salicylic acid hydroxylase. Plant immunity responses involve the perception of the invading
pathogen and the activation of specific defense signal transduction pathways that lead to activation
of PTI, effector-triggered immunity (ETI), and systemic acquired resistance (SAR) (42, 88). SA
and its derivatives play a central role in plant defense by mediating responses against pathogens
in many plant species, and they are important for activating PTI, ETI, and SAR (42). SA is an
endogenous defense signal, and its derivative MeSA (methyl salicylate) is one of the signals for SAR
(134). Interestingly, all Liberibacters contain an SA hydroxylase protein encoded by sahA ( J. Li &
N. Wang, unpublished results). It has been suggested that Las uses SA hydroxylase to degrade SA
and suppress plant defenses. Purified SahA displays strong enzymatic activity to degrade SA and its
derivatives. In transgenic tobacco plants, SahA abolishes SA accumulation and the hypersensitive
response induced by incompatible pathogens. Foliar sprays of 2,1,3-benzothiadiazole (BTH) and
2,6-dichloroisonicotinic acid (INA), SA functional analogs not degraded by SahA, are moderately
effective in suppressing Las population growth and HLB disease progression in infected citrus
trees under field conditions ( J. Li & N. Wang, unpublished results).

Mechanistic View of Liberibacter Pathogenicity


Based on current knowledge of Liberibacters and plant interactions, we suggest that the pathogen-
esis mechanisms of Liberibacters can be summarized in the following model (Figure 4). Liberib-
acters actively secrete SDEs or other virulence factors into the phloem, causing disease symptoms,

20.14 Wang et al.

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

Xylem Source
SE
CC

Water RFO Symplastic


High- pathway
pressure
Sucrose Apoplastic
pathway

Organelles (e.g., ER
Sucrose or mitochondria)
Polyols and receptors on
RFO the membranes of
the organelles
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

Some putative effectors


might affect expression
of target genes directly
or indirectly.
Las and PAMPs
on cell surface

Transport phloem

Release phloem
Low-
pressure

Symplastic
pathway

Sucrose Apoplastic
pathway

Sink
Figure 4
Illustration of Liberibacter virulence mechanisms in plant. Liberibacters (brown ovals) live in the phloem
elements. Phloem is responsible for transportation of carbohydrates from source to sink. Nutrients are
transmitted into phloem via a symplastic pathway or an apoplastic pathway. Liberibacters might secrete
Sec-dependent effectors (SDEs) or other virulence factors (represented by a blue dot, maroon triangle, red dot, or
orange trapezoid ) into the phloem sieve elements (SEs) or companion cells (CCs) to interfere with host
targets (proteins or genes, shown as enlarged), which cause cell death, necrosis, or other phenotypes of
phloem elements or companion cells, leading to localized cell responses and systemic malfunction of
phloem. Receptors on the remaining organelles in the phloem, such as the plastid, vacuole, mitochondrion,
or endoplasmic reticulum (ER), might be able to interact with secreted SDEs and/or other virulence factors
secreted or located on Liberibacter outer membrane to trigger cell responses. Some putative SDEs might
affect the expression of target genes directly or indirectly. In addition, Liberibacter might trigger plant
immunity response via pathogen-associated molecular patterns (PAMPs), such as lipopolysaccharides and
outer membrane proteins, leading to premature cell death, callose deposition, or phloem protein
accumulation, which lead to a localized response and/or systemic effect on phloem transportation. The
physical presence of Liberibacters and their metabolic activities may disturb the function of phloem via
disrupting osmotic gradients, or the integrity of phloem conductivity. Abbreviations: RFO, raffinose family
oligosaccharide.
www.annualreviews.org • Liberibacter Pathogenesis Mechanisms and Disease Control 20.15

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

most likely by interfering with host targets either in the phloem or companion cells (e.g., proteins
or genes). Because Liberibacters live in the phloem elements, an intracellular environment, the
receptors on the remaining organelles in the phloem, such as the plastid, vacuole, mitochondrion,
or endoplasmic reticulum (71), might be able to interact with secreted SDEs and/or other viru-
lence factors secreted or located on the Liberibacter outer membrane. Some of the host genes or
proteins targeted by SDEs or other virulence factors of Liberibacters might serve as susceptibility
genes that facilitate compatibility (e.g., promoting pathogen growth or suppressing immune re-
sponses) or disease development. One of the potential effects of the effectors or virulence factors
is phloem malfunction, e.g., due to destructive effects on components of sieve tubes and elements
themselves or interacting with companion cells, which supply proteins and transcripts to the sieve
elements (109). Investigation of Liberibacter secretory systems, potential effectors, and suscepti-
bility genes might lead to improvements in understanding Liberibacter pathogenicity and disease
control. Genome editing of susceptibility genes or their promoter region has been previously
shown to successfully generate disease-resistant citrus varieties (84, 86).
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

In addition, Liberibacters cause disease symptoms by eliciting self-destructive plant responses


and/or affecting phloem function, which, in turn, contributes to the development of disease symp-
toms. Liberibacters cause malfunctions in phloem loading, unloading, and transport, which even-
tually lead to disease symptom development. The physical presence of Liberibacters and their
metabolic activities may disturb the function of phloem, via disrupting osmotic gradients, or
the integrity of phloem conductivity. In addition, plant recognition of PAMPs (e.g., flagella and
LPSs) of Liberibacters might cause self-destructive plant responses, including the accumulation
of phloem proteins and callose in citrus phloem in response to Las infection as well as the defense
responses in the tubers to Lso.
Interestingly, sieve tubes full of Lso cells have been observed in carrot (131). For Las, fully
occupied sieve tubes have been observed in sink tissues, including young leaves, seed coat, and
roots, but not in collection or transportation phloem. The fully occupied phloem is unlikely to
be functional, resembling the loss of function of Xylella-occupied xylem (26). The fully occupied
sieve tubes at sink tissues might affect the integrity of the phloem and disrupt the osmotically
generated pressure gradient that drives the phloem mass flow. Alternatively, as mentioned earlier,
Liberibacters contain many PAMPs (41, 105, 185). Liberibacter PAMPs have been known to
induce plant defense, although mechanistically how host plants respond to these PAMPs remains
to be determined. Plant immunity responses observed previously in Las-infected citrus include the
production of secondary metabolites, such as phenolics, and accumulation of phloem proteins and
callose (51, 72, 90, 91). Lso causes an increase in phenolic levels in infected potato tubers, which
may result from their recognition of PAMPs (126). Additionally, Liberibacters consumption of
photoassimilates may disrupt osmotic gradients, thus affecting phloem loading, unloading, and
transport. Liberibacters are unable to take up and metabolize sucrose but utilize other sugars such
as glucose in the phloem. This may cause an imbalance in carbohydrate partitioning, such as the
accumulation of sucrose in the source leaves, in turn giving rise to the inhibition of photosynthesis
and subsequent physiological disorders (49).
Currently, it is unclear which factor plays the most important role in disrupting the phloem
function, i.e., whether it is due to the virulence factors, the host responses to the PAMPs, the
impact of the metabolic activity of the pathogens, or the physical occupation of the phloem
by the pathogens. It is important to acknowledge that despite not knowing the exact causes,
malfunction of phloem loading, unloading, and transport can explain many initial phenotypical
changes in plants infected with Liberibacter. We postulate that many later symptoms are probably
the cumulative and indirect effects of root decline, reduced photosynthesis, and reduced nutrient
transportation associated with phloem malfunction rather than the direct outcome of Liberibacter

20.16 Wang et al.

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

infection. In HLB-affected trees, phloem malfunction may explain the blotchy mottle symptoms
and root decline; later symptoms such as zinc deficiency and twig dieback may be the end result
of the decline process. For ZC, malfunction of phloem loading, unloading, and transportation
likely explain the starch accumulation in leaves and decreased starch levels in the tubers (9, 22).
However, it is important to note that Liberibacter titers and ZC symptoms increased in tubers
following cold storage and in tubers on vine-killed plants, i.e., under conditions in which tubers
are no longer influenced by whole-plant translocation processes. These observations demonstrate
that the development of symptoms in specific organs may occur independently of the condition
of the rest of the plant, suggesting the local symptoms are due to either action of virulence factors
or plant responses to PAMPs.

DISEASE CONTROL
Knowledge of the biology of Liberibacters and the mechanisms underlying bacterial pathogenicity
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

and symptom development has led to the design of a variety of management strategies. For all
Liberibacter-caused diseases in citrus, potato, and other crops, disease control currently focuses
on psyllid management and inoculum removal, owing to the universal absence of disease resistance
in all crops. However, successful long-term disease management depends on designing effective
control strategies based on an understanding of the characteristics of and interactions among the
crops, insects, and Liberibacters.

Huanglongbing Disease Management


For various citrus-producing areas worldwide, different strategies are needed, depending on the
threat level of Las and ACP and the stage of HLB or ACP epidemiology (e.g., initial entry or
endemic). Quarantine measures have been adopted to keep HLB away from Australia and the
Mediterranean citrus-producing areas. For many newly infected citrus-producing areas such as
California, quarantine and eradication measures are being implemented. Whether those newly
infected areas can prevent HLB from becoming endemic not only depends on the strict implemen-
tation of quarantine and eradication measures but also relies on the control effort in neighboring
citrus-producing areas of the HLB/ACP epidemiological zone, as detailed below. For newly in-
fected areas, early and accurate HLB diagnosis and Las detection are critical for inoculum removal.
Readers are referred to the review by Arredondo Valdés et al. (10) on HLB diagnosis and Las de-
tection. For HLB-endemic regions, economic and profitable citrus production is still possible
by integrating the most promising HLB and ACP management approaches. Both chemical and
biological psyllid control are critical components of HLB management. Readers are referred to
several excellent reviews on psyllid control (18, 63). In addition, RNAi has been suggested to be
a promising approach to control ACP, and readers are referred to a review by Hunter et al. (80).

Huanglongbing/Asian citrus psyllid epidemiological zone. Based on the frequency of ACP


immigration and transmission of Las within the components of the citrus-producing areas, the
citrus-producing areas can be classified into three levels: local (no geographical separation or bar-
rier to prevent ACP from traveling between citrus groves); area wide (clear separation between the
citrus groves prevents psyllids from flying between groves, but the psyllids can still be introduced
via wind or other mild weather events); and regional (geographical separation or barriers prevent
ACP from flying between the citrus-producing areas, and ACP rarely moves among the different
citrus-producing areas unless there are major weather events such as hurricanes). The United
States, Mexico, and countries in the Caribbean and Central America belong to a single regional

www.annualreviews.org • Liberibacter Pathogenesis Mechanisms and Disease Control 20.17

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

HLB/ACP epidemiological zone, as evidenced by the timing of HLB occurrence and detection
of the same haplotype of Las in this region. HLB was found in Florida in 2005, Cuba in 2007, the
Dominican Republic in 2008, and Belize and Puerto Rico in 2010, and was found almost simulta-
neously in Costa Rica, Nicaragua, Honduras, Guatemala, Mexico, and Jamaica. Haplotype A Las
has been found in the United States, Mexico, and countries in the Caribbean and Central America
(116). In addition, de León et al. (40) discovered that ACP in Florida, Texas, and California in
the United States and in Mexico, Costa Rica, and Belize belongs to the same group, whereas ACP
in Brazil belongs to a different group, indicating Brazil belongs to a separate regional HLB/ACP
epidemiological zone from the United States, Mexico, and countries in the Caribbean and Cen-
tral America. Boykin et al. (21) reported that there is no relationship between the United States
and Caribbean introductions of the vector. However, studies of the Las population data support
a relationship among the introductions in the United States, Mexico, and countries in Central
America and the Caribbean. Both Mexico and the United States have haplotype B Las, thus sug-
gesting a separate introduction from haplotype A (116). Within local and area-wide HLB/ACP
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

epidemiology zones, coordinated HLB/ACP management is being implemented. For the different
citrus-producing areas in the regional HLB/ACP epidemiology zones, HLB/ACP management
is difficult to coordinate but is needed. Quarantine measures have been implemented to limit the
spread of HLB-positive plant materials and ACP from one location to another within the regional
HLB/ACP epidemiology zones.

Plant resistance to Huanglongbing and Asian citrus psyllid. No HLB-resistant commercial


citrus cultivars have been identified to date (55), although tolerance has been observed in trifoliate
orange (Poncirus trifoliata L. Raf.) and some of its hybrids (6). P. trifoliata is a trifoliate species
that is graft compatible with Citrus spp. and is used as a rootstock. In the presence of HLB, well-
defined HLB disease symptoms are absent in trifoliate orange and some of its hybrids (6). Albrecht
et al. (4) compared the metabolomics of whole leaves of Las-infected and healthy greenhouse-
grown seedlings of the tolerant cultivar US-897 Citrus reticulata Cleopatra × Poncirus trifoliata) and
susceptible cultivar Cleopatra mandarin (C. reticulata). The number of metabolites differentially
produced in Las-infected versus uninfected leaves was greatest in the susceptible cultivar Cleopatra
mandarin and lowest in the tolerant cultivar US-897, where few differences in the metabolite
profiles were found. Tolerance to HLB did not appear to be associated with the accumulation
of higher amounts of protective metabolites in response to Las infection (4). Ramadugu et al.
(144) conducted a six-year field trial of various Citrus varieties under natural disease challenge
conditions in an HLB-endemic region in Florida. This study included 65 Citrus accessions and
33 accessions from 20 other closely related genera. On the basis of Las titers and disease symptoms,
the authors have tentatively categorized 2 accessions as immune, 6 as resistant, and 14 as tolerant.
Resistance and tolerance may be attributed to multiple factors, including psyllid colonization
ability and the absence of Las multiplication. Specifically, this study used seedlings, thus avoiding
the possible influence of rootstocks on the HLB response. Many citrus relatives appear to support
the replication of Las for a short period of time before subsequently suppressing them completely.
Indian wood apple (Limonia acidissima) is a transient host of Las (79). Murraya species support
only very low titers of Las and are possible transient hosts of Liberibacters (180). Those citrus
relatives showing HLB resistance/tolerance, such as Eremocitrus and Microcitrus, which are
sexually compatible with citrus, may be useful in future breeding to impart HLB resistance to
cultivated citrus.
Some newly released citrus scion varieties, such as Sugar Belle, have been suggested to display
HLB tolerance (F. Gmitter, personal communication). Tolerance of scions to HLB is affected by
rootstocks. The foliar HLB symptoms of scions on US-897 are fewer than those on many other

20.18 Wang et al.

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

rootstocks. Scions grafted on Volkamer lemon also show less canopy damage (5). These results
suggest that optimization of rootstock and scion combination is a useful strategy for control of
HLB.
Plant resistance to ACP includes the prevention of spread and multiplication of ACP in a
plant, thus providing an effective, economical, environmentally safe, and sustainable method of
controlling ACP (146). Westbrook (183) surveyed 87 accessions, primarily in the Rutaceae orange
subfamily Aurantioideae, in the field for an abundance of ACP eggs, nymphs, and adults. Very low
abundances of all life stages of ACP were found on two accessions of P. trifoliata. Richardson &
Hall (146) have further tested the resistance of 46 accessions of P. trifoliata to ACP, and observed
no eggs on 36% of the P. trifoliata accessions. P. trifoliata appears to have antixenosis and antibiosis
resistance to ACP.

Enhanced nutrient program. Foliar applications of micronutrients have been widely used in
Florida to mitigate HLB-induced mineral deficiencies and counter the debilitating effects of the
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

disease. The nutrient program often includes micronutrients combined with SA and/or phos-
phite. The effectiveness of these treatments has been controversial (62, 157, 161). In some studies,
no differences have been observed in tree health, Las titer, and fruit quality and yield between
trees treated with an enhanced nutritional program and control trees (62). In contrast, Shen et al.
(157) reported that the enhanced nutrient program has reduced Las content and increased leaf
size and weight after at least three years of application. A positive effect on citrus fruit yields by
foliar nutrients was also observed (161). However, Xia et al. (187) surveyed nutritional approaches
to HLB management in China and found no consistent evidence to support the nutritional ap-
proach as an effective means of maintaining grove productivity or slowing disease progression in
HLB-diseased plants. The authors suggest, however, that the HLB-diseased trees may remain
productive for several years under a good nutritional program when combined with other cultural
practices such as good irrigation. SAR inducers and other plant defense inducers are regularly
included in such nutrient programs. Spraying of HLB-diseased trees with plant defense inducers,
including β-aminobutyric acid, BTH, and INA, individually or in combination slows Las pop-
ulation growth in planta and decreases HLB disease severity by approximately 15% to 30%, as
compared with the negative control (98). Notably, the application of plant defense inducers via
foliar spraying did not reduce Las titers (98). Overall, these studies have suggested that foliar nutri-
ents promote tree growth in asymptomatic trees or in those planted in citrus groves with poor soil
nutrients at an early stage of infection and that plant defense inducers can confer a moderate level
of disease control but cannot decrease Las titers in infected trees. However, the enhanced nutrient
program appears to have no significant effect on HLB-positive trees at advanced stages of infection.

Thermotherapy. Thermal therapy treatments have been used for decades to prevent or eliminate
diseases in plants (92). Continuous thermal exposure to 40◦ C to 42◦ C for a minimum of 48 h has
been reported to be sufficient to significantly decrease Las titers or eliminate Las entirely in
HLB-affected citrus seedlings (74). Hoffman et al. (74) indicated that heat treatment enhances
the vigor of Las-infected citrus and promotes new root growth. Ehsani and colleagues (45) also
reported a decrease in HLB symptoms after heat treatment of citrus trees in groves. The current
challenge with thermotherapy is that although sufficiently high temperatures may be reached
in the aboveground parts of the plant, killing temperatures are unlikely to be reached in roots
where the heat treatment is buffered by the soil (108). Thus, thermotherapy is unlikely to redue
Las populations in the roots, which then serve as reservoirs for reinfection of the canopy during
flushes. The effectiveness of thermotherapy on reducing Las populations in roots needs to be
improved for it to become a viable component of integrated HLB management.

www.annualreviews.org • Liberibacter Pathogenesis Mechanisms and Disease Control 20.19

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

Replanting. Owing to the loss of trees to HLB in HLB-endemic citrus-producing areas, re-
planting is critical to sustaining citrus production. Certified HLB-free nursery stock is required
for replanting. Younger orchards express symptoms much more quickly than older trees, normally
within 6 to 12 months after planting (61). Aggressive psyllid control on young trees can prevent
psyllid feeding and transfer of Las to the young seedlings (155). In addition, instead of replanting
one-year-old seedlings, planting two- or three-year-old seedlings allows profitable production
sooner after replantation, and the older trees are more tolerant of HLB compared to younger
replantation (G. Zhong, personal communication). Growing HLB-free trees in nurseries now
typically requires the production of citrus within totally enclosed structures to prevent contact
with ACP. This practice has been expanded to include the growth of mature trees. Citrus un-
dercover production systems (CUPSs), which utilize an enclosed structure to grow mature trees
for citrus production (151), are now being used commercially on a small scale in both the United
States and China.
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

Bactericides. Bactericides, including ampicillin, carbenicillin, cephalexin, oxytetracycline


(OTC), penicillin, rifampicin, streptomycin sulfate, and sulfadimethoxine, have been shown to
be highly effective in eliminating or suppressing Las in infected trees (191). Streptomycin sulfate,
OTC hydrochloride, and OTC calcium complex have been approved for use in controlling HLB
via foliar spray in Florida since 2016. The benefits and efficacy of these bactericides in the Florida
citrus industry via foliar spray remain to be determined.
Trunk injection of different bactericides, including OTC, has some efficacy for HLB control
(78). However, the challenges of trunk injection include potential damage to trees and the lack of a
suitable device for injecting trees at a commercial scale, thus leading to low efficacy and high labor
costs. Optimization of trunk injections and development of an efficient trunk injector are required
before the application of trunk injection of bactericides to control HLB can be considered a viable
commercial disease management tool.
It remains to be determined whether Las will develop resistance against bactericides. The
emergence of bactericide resistance typically results from horizontal gene transfer (HGT) of
mobile genetic elements that carry the resistance genes among co-occurring bacteria (178) or
by mutation of bacterial targets. HGT is widespread among bacteria. For example, tetracycline
resistance genes encoding either efflux or ribosomal protection proteins in many bacteria are
suggested to have been acquired by HGT from the tetracycline-producing streptomycetes (28).
HGT is thought to be less common in intracellular bacteria, including Liberibacters, because
they are sequestered from external microbial populations (117, 175). Several bacterial symbionts,
including Wolbachia spp., Carsonella DC, and “Candidatus Profftella armatura” (124), have been
discovered in the ACP gut and may potentially serve as donors of resistance genes. For Las, only
lysE has been proposed to have been acquired from “Ca. Profftella armatura” via HGT (124).
It remains to be determined whether any of the gut microbes contain genes for resistance to
bactericides such as OTC or streptomycin. Notably, Las contains the ABC transporters Msb1
and MsbA2, which share high identity with E. coli MsbA and have been predicted to confer
multidrug resistance to Las (99, 150). It remains to be determined whether MsbA-like genes may
contribute to the development of resistance to OTC, streptomycin, or other bactericides and
whether long-term use of bactericides will lead to antimicrobial resistance of Las in the future.
Efforts are being made to improve the application of bactericides and identify novel bactericides
to control HLB. Nanoemulsion formulations of ampicillin have been found to eliminate Las
in HLB-diseased citrus in planta more efficiently than non-nanoformulation controls (189). A
structure-based design approach has been implemented to identify small antimicrobial molecules
targeting SecA, an essential component of the Sec machinery (2, 3, 77). In addition, it has also been

20.20 Wang et al.

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

reported that benzbromarone inhibits LdtR, a Las homolog of the multidrug resistance regulator
MarR (133).

Area-wide Huanglongbing/Asian citrus psyllid management. One successful HLB manage-


ment approach in the United States, China, and Brazil is area-wide HLB/ACP management. As
reported by Bassanezi et al. (14), area-wide management decreases ACP populations compared
with those in groves not under such management, and relatively few immigrant psyllids that car-
ried Las were found in the study. The beginning of the HLB epidemic for a new plantation of
sweet orange under area-wide management was delayed by 10 months. Approximately five years
after planting, the disease incidence for the new plantation under area-wide management was
decreased by 90% compared with controls (14). Successful area-wide HLB/ACP management has
also been reported by Belasque et al. (15).
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

Genetic improvement of citrus against Huanglongbing. Breeding disease-resistant varieties


is the most efficient and sustainable approach to controlling most plant diseases. Traditional
breeding to generate HLB-resistant citrus varieties has been hindered by a lack of HLB-resistant
germplasm, polyembryony, pollen-ovule sterility, sexual and graft incompatibilities, and extended
juvenility (38). Transgenic approaches have been used to improve citrus resistance against HLB
by overexpressing NPR1 (nonexpressor of pathogenesis related genes 1), a key regulator in the
signal transduction pathway that leads to SAR (43), and overexpression of antimicrobial proteins
(e.g., thionins and plant defensins) (69) or antimicrobial peptides (e.g., cecropin B) (162). Citrus
varieties expressing spinach defensins developed by Erik Mirkov (Texas A&M University) have
been granted an experimental use permit by the US Environmental Protection Agency for field
testing of resistance against HLB. However, the regulatory hurdles to register transgenic citrus and
consumer acceptance of transgenic citrus may present many challenges to the commercialization
of such varieties.
Recent developments in targeted genome-editing technologies, however, have the potential
to facilitate the process to produce genetically modified citrus cultivars that lack transgenes (86).
Targeted genome-editing technologies based on zinc finger nucleases, transcription activator-like
effector nucleases, or CRISPR/Cas9–single guide RNA (sgRNA) (hereafter, Cas9-sgRNA) provide
a promising path to improve crops, including generating disease-resistant plants. Cas9-sgRNA
technology has been successfully used for genome editing in many plants (127, 156). The citrus
genome has been successfully modified by using Cas9-sgRNA (84, 85). Importantly, genome
editing via CRISPR technology has been used to generate disease-resistant citrus. Specifically,
the promoter and coding regions of the disease susceptibility gene CsLOB1, which favors citrus
bacterial canker disease, caused by Xanthomonas citri, have been modified to generate canker-
resistant plants (84, 86).
Nontransgenic approaches that generate plants without introducing foreign genes into the
genome are likely to have wider acceptance and an easier regulatory pathway. Nontransgenic,
foreign DNA–free, genome-modified plants have been generated by transfecting pre-assembled
complexes of purified Cas9 protein and sgRNA into plant protoplasts of Arabidopsis thaliana,
tobacco, lettuce, and rice to achieve targeted mutagenesis in regenerated plants at frequencies of
up to 46% (184). Nontransgenic wheat plants have also been generated with CRISPR technology
by regeneration of plants from callus cells transiently expressing CRISPR/Cas9 introduced as DNA
or RNA (193). Nontransgenic, genome-edited plants made using Cas9-sgRNA may provide an
unprecedented opportunity to generate HLB-resistant/tolerant citrus varieties.

www.annualreviews.org • Liberibacter Pathogenesis Mechanisms and Disease Control 20.21

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

Zebra Chip Disease Management


The control strategy of choice for managing ZC remains application of a rotating collection of
chemical insecticides, including imidacloprid, abamectin, spiromesifen, and spinosad. However,
insecticides can have difficulties in reaching their target because psyllids spend most of their time
within the plant canopy on the underside of leaves. Moreover, the efficacy of application may
vary depending on agronomical practices and growing conditions (139). The heavy usage and
dependence on insecticides have led to the development of resistance by psyllids to imidacloprid
(106), and, more recently, resistance to abamectin was discovered in the San Luis Potosı́ vicinity in
Mexico (27). Insect resistance to pesticides can lead to larger insect populations, with consequences
for greater disease incidence and the spread of disease to new crops. Although management sys-
tems, as currently practiced in commercial potato production fields, have been successful in man-
aging psyllid populations, effective rotation of insecticides with different chemistries, regionally
coordinated pesticide applications, and use of materials and practices that are friendlier to bene-
ficial arthropods are needed to extend the useful life of existing pesticides. Future strategies also
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

must take into account the growing evidence for the role of persistent, resident psyllid popula-
tions in areas where they were not previously thought to overwinter (75, 119, 121, 177) as well
as differences in the relative abundance, physiology, and behavior of insect haplotypes (107, 122,
123, 165). Importantly, field studies clearly demonstrate that plants remain susceptible to ZC until
harvest and that vine kill following infection does not protect tubers (149).
An additional complicating factor in the long-term management of ZC is the observation that
potato psyllids may carry and transmit one or both of Lso A and B to the host. Recent studies
reveal life-stage-specific effects of different Lso haplotypes on the fitness of the vector (190). These
observations have implications for inoculum dosage (24) as well as field-scale epidemiology. Both
Lso haplotypes individually kill potato plants, although the relationship between the severity of ZC
symptoms and Lso haplotype is unclear. Symptoms in Lso A–infected tomato are typically milder
(reduced growth) than symptoms in Lso B–infected tomato, which frequently displays dramatic
symptoms leading to plant death (C. Tamborindeguy, personal communication; E. Tienebo,
J. Levy, and E. Pierson, unpublished results).
Breeding for resistance to Lso, the psyllid vector, or both remains the most promising long-
term strategy for managing Lso-associated diseases. However, to date no resistance to ZC in any
commercial breeding lines has been identified. Several wild solanaceous species have been identi-
fied that show resistance to insect feeding. For example, Solanum habrochaites, a wild solanaceous
plant that is both antixenosis (repellent) and toxic to the whitefly Bemisia tabaci, was evaluated
for its potential to similarly affect B. cockerelli (97). Despite these effects, in choice studies with
Solanum lycopersicum at least 50% of the S. habrochaites plants tested positive for Lso, indicating
this resistance is not sufficient to protect solanaceous plants against Lso infection. Additionally,
unique germplasm has been derived from wild potato species with Solanum tuberosum, Solanum
berthaultii, Solanum raphanifolium, Solanum chacoense, and Solanum guerreroense in their pedigree
(110). Some of these exhibited antixenosis to B. cockerelli, and some displayed no Lso infection in
no-choice transmission assays. Further testing and selection for agronomical traits acceptable to
the industry are required.
In addition, phage-based therapies targeting the pathogen are promising approaches (12).
Strategies for genetic modifications incorporating disease resistance in potato may eventually
benefit from research in the more commercially valuable citrus crops. Finally, gene-silencing
strategies such as RNA interference and CRISPR-based gene editing may provide ways to control
insect populations or modulate host response to Lso infection. For example, Wuriyanghan et al.
(186) recently demonstrated that double-stranded RNAs and siRNAs, when expressed in the gut of

20.22 Wang et al.

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

B. cockerelli via injection or oral acquisition, were able to induce mortality in the recipient psyllids.
Ongoing investigations to identify genes that are important for maintaining endosymbionts (81)
may facilitate efforts to reduce Lso infection in psyllids.

CONCLUDING REMARKS
Liberibacters and their insect vectors are expanding their distribution worldwide, which represents
an intensifying global threat to the production of citrus and many other crops. Importantly,
weather-driven physiologically based demographic models predict that Egypt, Israel, Cyprus,
Sicily, Southern Spain, and Portugal are at great risk for ACP establishment and the spread of
HLB (64). Spatial models also predict that southern Australia is at risk of HLB introduction in
the future (11). Both models predict that climates in large areas of Africa, Latin America, and
North Australia are well suited to ACP and Las expansion (125). Additionally, recent findings
confirm the suspicion that psyllids may overwinter in colder conditions than initially predicted
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

and are using weeds as green bridges between seasons, possibly explaining recent expansions in
the geographic range of ZC in the United States and New Zealand (75, 119, 121, 177). Europe is
one of the major potato-producing areas in the world currently free of ZC. There is concern that
Lso haplotypes found in Europe, but not currently infecting potato, can or will jump to potato
via a psyllid vector utilizing potato as a host. Significant progress has been made in understanding
the pathogenesis mechanisms of Liberibacters, which will continue to be useful in the design of
effective disease management strategies. Currently, potato production is managed by aggressive
psyllid control. Citrus production in the presence of HLB is also possible if the most promising
management approaches are integrated. Meanwhile, new technologies such as CRISPR/Cas9 may
provide an unprecedented opportunity to provide long-term management solutions. There are
many similarities in the pathogenicity mechanisms of different Liberibacters. Thus, we believe
that there is much that can be learned by working together across commodities and disciplines to
tackle these scientifically intriguing yet economically destructive crop diseases.

DISCLOSURE STATEMENT
The authors are not aware of any affiliations, memberships, funding, or financial holdings that
might be perceived as affecting the objectivity of this review.

ACKNOWLEDGMENTS
We thank Drs. Dennis Gross, Leland Pierson, Charlie Rush, and Cecilia Tamborindeguy for
valuable discussions. This work has been supported by Florida Citrus Research and Development
Foundation, USDA-APHIS MAC program, USDA-SCRI, USDA-Specialty Block Grant, and
Florida Citrus Initiative program.

LITERATURE CITED
1. Aguilar E, Sengoda VG, Bextine B, McCue KF, Munyaneza JE. 2013. First report of “Candidatus Liberib-
acter solanacearum” on tomato in Honduras. Plant Dis. 97(10):1375
2. Akula N, Trivedi P, Han FQ, Wang N. 2012. Identification of small molecule inhibitors against SecA
of Candidatus Liberibacter asiaticus by structure based design. Eur. J. Med. Chem. 54:919–24
3. Akula N, Zheng H, Han FQ, Wang N. 2011. Discovery of novel SecA inhibitors of Candidatus Liberib-
acter asiaticus by structure based design. Bioorg. Med. Chem. Lett. 21(14):4183–88

www.annualreviews.org • Liberibacter Pathogenesis Mechanisms and Disease Control 20.23

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

4. Albrecht U, Fiehn O, Bowman KD. 2016. Metabolic variations in different citrus rootstock cultivars
associated with different responses to Huanglongbing. Plant Physiol. Biochem. 107:33–44
5. Albrecht U, McCollum G, Bowman KD. 2012. Influence of rootstock variety on Huanglongbing disease
development in field-grown sweet orange (Citrus sinensis [l.] Osbeck) trees. Sci. Hortic. 138:210–20
6. Albrecht U, Bowman KD. 2012. Transcriptional response of susceptible and tolerant citrus to infection
with Candidatus Liberibacter asiaticus. Plant Sci. 185–186:118–30
7. Albus U, Baier R, Holst O, Pühler A, Niehaus K. 2001. Suppression of an elicitor-induced oxidative
burst reaction in Medicago sativa cell cultures by Sinorhizobium meliloti lipopolysaccharides. New Phytol.
151(3):597–606
8. Alfaro-Fernández A, Cebrián MC, Villaescusa FJ, de Mendoza AH, Ferrándiz JC, et al. 2012. First report
of “Candidatus Liberibacter solanacearum” in carrot in mainland Spain. Plant Dis. 96(4):582
9. Alvarado VY, Odokonyero D, Duncan O, Mirkov TE, Scholthof HB. 2012. Molecular and physiological
properties associated with zebra complex disease in potatoes and its relation with Candidatus Liberibacter
contents in psyllid vectors. PLOS ONE 7(5):e37345
10. Arredondo Valdés R, Delgado Ortiz JC, Beltrán Beache M, Anguiano Cabello J, Cerna Chávez E,
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

et al. 2016. A review of techniques for detecting Huanglongbing (greening) in citrus. Can. J. Microbiol.
62(10):803–11
11. Aurambout JP, Finlay KJ, Luck J, Beattie GAC. 2009. A concept model to estimate the potential dis-
tribution of the Asiatic citrus psyllid (Diaphorina citri Kuwayama) in Australia under climate change—a
means for assessing biosecurity risk. Ecol. Model. 220:2512–24
12. Balogh B, Jones JB, Iriarte FB, Momol MT. 2010. Phage therapy for plant disease control. Curr. Pharm.
Biotechnol. 11(1):48–57
13. Barrangou R, Fremaux C, Deveau H, Richards M, Boyaval P, et al. 2007. CRISPR provides acquired
resistance against viruses in prokaryotes. Science 315(5819):1709–12
14. Bassanezi RB, Montesino LH, Gimenes-Fernandes N, Yamamoto PT, Gottwald TR, et al. 2013. Efficacy
of area-wide inoculum reduction and vector control on temporal progress of Huanglongbing in young
sweet orange plantings. Plant Dis. 97(6):789–96
15. Belasque J, Bassanezi RB, Yamamoto PT, Ayres AJ, Tachibana A et al. 2010. Lessons from Huanglong-
bing management in São Paulo state, Brazil. J. Plant Path. 82:285–302
16. Bertolini E, Teresani GR, Loiseau M, Tanaka FAO, Barbé S, et al. 2015. Transmission of “Candidatus
Liberibacter solanacearum” in carrot seeds. Plant Pathol. 64(2):276–85
17. Bland C, Ramsey TL, Sabree F, Lowe M, Brown K, et al. 2007. CRISPR recognition tool (CRT): a tool
for automatic detection of clustered regularly interspaced palindromic repeats. BMC Bioinform. 8(1):209
18. Boina DR, Bloomquist JR. 2015. Chemical control of the Asian citrus psyllid and of Huanglongbing
disease in citrus. Pest Manag. Sci. 71(6):808–23
19. Bonnington KE, Kuehn MJ. 2014. Protein selection and export via outer membrane vesicles. Biochim.
Biophys. Acta 1843(8):1612–19
20. Bové JM. 2006. Huanglongbing: a destructive, newly-emerging, century-old disease of citrus. J. Plant
Pathol. 88(1):7–37
21. Boykin LM, De Barro P, Hall DG, Hunter WB, McKenzie CL, et al. 2012. Overview of worldwide
diversity of Diaphorina citri Kuwayama mitochondrial cytochrome oxidase 1 haplotypes: two old world
lineages and a new world invasion. Bull. Entomol. Res. 102(5):573–82
22. Buchman JL, Fisher TW, Sengoda VG, Munyaneza JE. 2012. Zebra chip progression: from inoculation
of potato plants with Liberibacter to development of disease symptoms in tubers. Am. J. Potato Res.
89(2):159–68
23. Camacho-Tapia M, Rojas-Martı́nez RI, Zavaleta-Mejı́a E, Hernández-Deheza MG, Carrillo-Salazar JA,
et al. 2011. Aetiology of chili pepper variegation from Yurécuaro, México. J. Plant Pathol. 93(2):331–35
24. Casteel CL, Hansen AK, Walling LL, Paine TD. 2012. Manipulation of plant defense responses by the
tomato psyllid (Bactericerca cockerelli ) and its associated endosymbiont Candidatus Liberibacter psyllau-
rous. PLOS ONE 7(4):e35191
25. Castresana J. 2000. Selection of conserved blocks from multiple alignments for their use in phylogenetic
analysis. Mol. Biol. Evol. 17:540–52

20.24 Wang et al.

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

26. Chatterjee S, Almeida RPP, Lindow S. 2008. Living in two worlds: the plant and insect lifestyles of
Xylella fastidiosa. Annu. Rev. Phytopathol. 46:243–71
27. Chávez EC, Bautista OH, Flores JL, Uribe LA, Fuentes YMO. 2015. Insecticide-resistance ratios of three
populations of Bactericera cockerelli (Hemiptera: Psylloidea: Triozidae) in regions of northern Mexico.
Fla. Entomol. 98(3):950–53
28. Chopra I, Roberts M. 2001. Tetracycline antibiotics: mode of action, applications, molecular biology,
and epidemiology of bacterial resistance. Microbiol. Mol. Biol. Rev. 65:232–60
29. Cicero JM, Fisher TW, Brown JK. 2016. Localization of “Candidatus Liberibacter solanacearum” and
evidence for surface appendages in the potato psyllid vector. Phytopathology 106(2):142–54
30. Cocuzza GEM, Alberto U, Hernández-Suárez E, Siverio F, Di Silvestro S, et al. 2016. A review on Trioza
erytreae (African citrus psyllid), now in mainland Europe, and its potential risk as vector of Huanglongbing
(HLB) in citrus. J. Pest Sci. 2004:1–17
31. Coletta-Filho HD, Targon MLPN, Takita MA, De Negri JD, Pompeu J, et al. 2004. First report of the
causal agent of Huanglongbing (“Candidatus Liberibacter asiaticus”) in Brazil. Plant Dis. 88(12):1382–82
32. Cong Q, Kinch LN, Kim B-H, Grishin NV. 2012. Predictive sequence analysis of the Candidatus Liberib-
acter asiaticus proteome. PLOS ONE 7(7):e41071
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

33. Cooper WR, Alcala PE, Barcenas NM. 2015. Relationship between plant vascular architecture and
within-plant distribution of “Candidatus Liberibacter solanacearum” in potato. Am. J. Potato Res.
92(1):91–99
34. Crosslin JM, Hamm PB, Eggers JE, Rondon SI, Sengoda VG, Munyaneza JE. 2012. First report of zebra
chip disease and “Candidatus Liberibacter solanacearum” on potatoes in Oregon and Washington State.
Plant Dis. 96(3):452
35. Crosslin JM, Munyaneza JE. 2009. Evidence that the zebra chip disease and the putative causal agent
can be maintained in potatoes by grafting and in vitro. Am. J. Potato Res. 86(3):183–87
36. Crosslin JM, Munyaneza JE, Brown JK, Liefting LW. 2010. Potato zebra chip disease: a phytopatho-
logical tale. Plant Health Prog. https://doi.org/10.1094/PHP-2010-0317-01-RV
37. da Graça JV, Douhan GW, Halbert SE, Keremane ML, Lee RF, et al. 2016. Huanglongbing: an overview
of a complex pathosystem ravaging the world’s citrus. J. Integr. Plant Biol. 58(4):373–87
38. Davey MR, Anthony P, Power JB, Lowe KC. 2005. Plant protoplasts: status and biotechnological per-
spectives. Biotechnol. Adv. 23(2):131–71
39. Davis RI, Gunua TG, Kame MF, Tenakanai D, Ruabete TK. 2005. Spread of citrus Huanglongbing
(greening disease) following incursion into Papua New Guinea. Australas. Plant Pathol. 34(4):517
40. de León JH, Sétamou M, Gastaminza GA, Buenahora J, Cáceres S, et al. 2011. Two separate introductions
of Asian citrus psyllid populations found in the American continents. Ann. Entomol. Soc. Am. 104(6):1392–
98
41. Duan Y, Zhou L, Hall DG, Li W, Doddapaneni H, et al. 2009. Complete genome sequence of citrus
Huanglongbing bacterium, “Candidatus Liberibacter asiaticus” obtained through metagenomics. Mol.
Plant-Microbe Interact. 22(8):1011–20
42. Durrant WE, Dong X. 2004. Systemic acquired resistance. Annu. Rev. Phytopathol. 42:185–209
43. Dutt M, Barthe G, Irey M, Grosser J, Duan Y, et al. 2015. Transgenic citrus expressing an Arabidopsis
NPR1 gene exhibit enhanced resistance against Huanglongbing (HLB; citrus greening). PLOS ONE
10(9):e0137134
44. Edgar RC. 2004. MUSCLE: multiple sequence alignment with high accuracy and high throughput.
Nucleic Acids Res. 32(5):1792–97
45. Ehsani R, Reyes De Corcuera JI, Khot L. 2013. The potential of thermotherapy in combating HLB.
Resour. Mag. 20(6):18–19
46. Esau K, Cheadle VI. 1959. Size of pores and their contents in sieve elements of dicotyledons. PNAS
45(2):156–62
47. Etxeberria E, Gonzalez P, Borges AF, Brodersen C. 2016. The use of laser light to enhance the uptake
of foliar-applied substances into citrus (Citrus sinensis) leaves. Appl. Plant Sci. 4(1):1500106
48. Fagen JR, Leonard MT, McCullough CM, Edirisinghe JN, Henry CS, et al. 2014. Comparative genomics
of cultured and uncultured strains suggests genes essential for free-living growth of Liberibacter. PLOS
ONE 9(1):e84469

www.annualreviews.org • Liberibacter Pathogenesis Mechanisms and Disease Control 20.25

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

49. Fan J, Chen C, Brlansky RH, Gmitter FG Jr., Li Z-G. 2010. Changes in carbohydrate metabolism in
Citrus sinensis infected with “Candidatus Liberibacter asiaticus.” Plant Pathol. 59(6):1037–43
50. Fan J, Chen C, Yu Q, Brlansky RH, Li Z-G, Gmitter FG. 2011. Comparative iTRAQ proteome and
transcriptome analyses of sweet orange infected by “Candidatus Liberibacter asiaticus.” Physiol. Plant.
143(3):235–45
51. Fan J, Chen C, Yu Q, Khalaf A, Achor DS, et al. 2012. Comparative transcriptional and anatomical
analyses of tolerant rough lemon and susceptible sweet orange in response to “Candidatus Liberibacter
asiaticus” infection. Mol. Plant-Microbe Interact. 25(11):1396–407
52. Ferooz J, Lemaire J, Letesson J-J. 2011. Role of FlbT in flagellin production in Brucella melitensis.
Microbiology 157(5):1253–62
53. Fisher TW, Vyas M, He R, Nelson W, Cicero JM, et al. 2014. Comparison of potato and Asian cit-
rus psyllid adult and nymph transcriptomes identified vector transcripts with potential involvement in
circulative, propagative Liberibacter transmission. Pathogens 3(4):875–907
54. Folimonova SY, Achor DS. 2010. Early events of citrus greening (Huanglongbing) disease development
at the ultrastructural level. Phytopathology 100(9):949–58
55. Folimonova SY, Robertson CJ, Garnsey SM, Gowda S, Dawson WO. 2009. Examination of the re-
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

sponses of different genotypes of citrus to Huanglongbing (citrus greening) under different conditions.
Phytopathology 99(12):1346–54
56. Francl LJ. 2001. The disease triangle: a plant pathological paradigm revisited. Plant Health Instr.
https://doi.org/10.1094/PHI-T-2001-0517-01
57. Fu SM, Hartung J, Zhou CY, Su HN, Tan J, Li ZA. 2015. Ultrastructural changes and putative phage
particles observed in sweet orange leaves infected with “Candidatus Liberibacter asiaticus.” Plant Dis.
99(3):320–24
58. Garnier M, Bové JM, Cronje CPR, Sanders GM, Korsten L, Le Roux HF. 2000. Presence of “Candidatus
Liberibacter africanus” in the Western Cape province of South Africa.” Proc. Conf. Int. Organ. Citrus
Virol., 14th, Campinas, São Paulo, Sept. 13–18, 1998, pp. 369–72. Riverside, CA: IOCV
59. Garnier M, Jagoueix-Eveillard S, Cronje PR, Le Roux HF, Bove JM. 2000. Genomic characterization
of a Liberibacter present in an ornamental rutaceous tree, Calodendrum capense, in the Western Cape
province of South Africa. Proposal of “Candidatus Liberibacter africanus subsp. capensis.” Int. J. Syst.
Evol. Microbiol. 50(6):2119–25
60. Glynn JM, Islam MS, Bai Y, Lan S, Wen A, et al. 2012. Multilocus sequence typing of “Candidatus
Liberibacter solanacearum” isolates from North America and New Zealand. J. Plant Pathol. 94(1):223–
28
61. Gottwald TR. 2010. Current epidemiological understanding of citrus Huanglongbing. Annu. Rev. Phy-
topathol. 48:119–39
62. Gottwald TR, Graham JH, Irey MS, McCollum TG, Wood BW. 2012. Inconsequential effect of nu-
tritional treatments on Huanglongbing control, fruit quality, bacterial titer and disease progress. Crop
Prot. 36:73–82
63. Grafton-Cardwell EE, Stelinski LL, Stansly PA. 2013. Biology and management of Asian citrus psyllid,
vector of the Huanglongbing pathogens. Annu. Rev. Entomol. 58:413–32
64. Gutierrez AP, Ponti L. 2013. Prospective analysis of the geographic distribution and relative abundance
of Asian citrus psyllid (Hemiptera: Liviidae) and citrus greening disease in North America and the
Mediterranean basin. Fla. Entomol. 96(4):1375–91
65. Haapalainen M. 2014. Biology and epidemics of Candidatus Liberibacter species, psyllid-transmitted
plant-pathogenic bacteria. Ann. Appl. Biol. 165(2):172–98
66. Hansen AK, Trumble JT, Stouthamer R, Paine TD. 2008. A new Huanglongbing species, “Candidatus
Liberibacter psyllaurous,” found to infect tomato and potato, is vectored by the psyllid Bactericera cockerelli
(Sulc). Appl. Environ. Microbiol. 74(18):5862–65
67. Hao G, Boyle M, Zhou L, Duan Y. 2013. The intracellular citrus Huanglongbing bacterium, “Candidatus
Liberibacter asiaticus” encodes two novel autotransporters. PLOS ONE 8(7):e68921
68. Hao G, Pitino M, Ding F, Lin H, Stover E, Duan Y. 2014. Induction of innate immune responses
by flagellin from the intracellular bacterium, “Candidatus Liberibacter solanacearum.” BMC Plant Biol.
14(1):211

20.26 Wang et al.

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

69. Hao G, Stover E, Gupta G. 2016. Overexpression of a modified plant thionin enhances disease resistance
to citrus canker and Huanglongbing (HLB). Front. Plant Sci. 7:1078
70. Hartung JS, Paul C, Achor D, Brlansky RH. 2010. Colonization of dodder, Cuscuta indecora, by “Candi-
datus Liberibacter asiaticus” and “Ca. L. americanus.” Phytopathology 100(8):756–62
71. Heo J, Blob B, Helariutta Y. 2017. Differentiation of conductive cells: a matter of life and death. Curr.
Opin. Plant Biol. 35:23–29
72. Hijaz FM, Manthey JA, Folimonova SY, Davis CL, Jones SE, et al. 2013. An HPLC-MS characterization
of the changes in sweet orange leaf metabolite profile following infection by the bacterial pathogen
Candidatus Liberibacter asiaticus. PLOS ONE 8(11):e79485
73. Hilf ME. 2011. Colonization of citrus seed coats by “Candidatus Liberibacter asiaticus”: implications for
seed transmission of the bacterium. Phytopathology 101(10):1242–50
74. Hoffman MT, Doud MS, Williams L, Zhang M-Q, Ding F, et al. 2013. Heat treatment eliminates
“Candidatus Liberibacter asiaticus” from infected citrus trees under controlled conditions. Phytopathology
103(1):15–22
75. Horton DR, Cooper WR, Munyaneza JE, Swisher KD, Echegaray ER, et al. 2015. A new problem and
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

old questions: potato psyllid in the Pacific Northwest. Am. Entomol. 61(4):234–44
76. Hoshi A, Oshima K, Kakizawa S, Ishii Y, Ozeki J, et al. 2009. A unique virulence factor for proliferation
and dwarfism in plants identified from a phytopathogenic bacterium. PNAS 106(15):6416–21
77. Hu J, Akula N, Wang N, Gottwald T, Wang N, et al. 2016. Development of a microemulsion formulation
for antimicrobial SecA inhibitors. PLOS ONE 11(3):e0150433
78. Hu J, Wang N. 2016. Evaluation of the spatiotemporal dynamics of oxytetracycline and its control effect
against citrus Huanglongbing via trunk injection. Phytopathology 106(12):1495–503
79. Hung TH, Wu ML, Su HJ. 2000. Identification of alternative hosts of the fastidious bacterium causing
citrus greening disease. J. Phytopathol. 148(6):321–26
80. Hunter WB, Glick E, Paldi N, Bextine BR. 2012. Advances in RNA interference: dsRNA treatment in
trees and grapevines for insect pest suppression. Southwest. Entomol. 37(1):85–87
81. Ibanez F, Levy J, Tamborindeguy C. 2014. Transcriptome analysis of “Candidatus Liberibacter
solanacearum” in its psyllid vector, Bactericera cockerelli. PLOS ONE 9(7):e100955
82. Jagoueix S, Bove JM, Garnier M. 1994. The phloem-limited bacterium of greening disease of citrus is a
member of the alpha subdivision of the proteobacteria. Int. J. Syst. Bacteriol. 44(3):379–86
83. Jain M, Fleites LA, Gabriel DW. 2015. Prophage-encoded peroxidase in “Candidatus Liberibacter asi-
aticus” is a secreted effector that suppresses plant defenses. Mol. Plant-Microbe Interact. 28(12):1330–37
84. Jia H, Orbovic V, Jones JB, Wang N. 2016. Modification of the PthA4 effector binding elements
in type I CsLOB1 promoter using Cas9/sgRNA to produce transgenic Duncan grapefruit alleviating
XccpthA4:dCsLOB1.3 infection. Plant Biotechnol. J. 14(5):1291–301
85. Jia H, Wang N. 2014. Targeted genome editing of sweet orange using Cas9/sgRNA. PLOS ONE
9(4):e93806
86. Jia H, Zhang Y, Orbović V, Xu J, White FF, et al. 2016. Genome editing of the disease suscepti-
bility gene csLOB1 in citrus confers resistance to citrus canker. Plant Biotechnol. J. https://doi.org/
10.1111/pbi.12677
87. Johnson EG, Wu J, Bright DB, Graham JH. 2014. Root loss on presymptomatic Huanglongbing affected
trees is preceded by Candidatus Liberibacter asiaticus root infection but not phloem plugging. Plant Pathol.
63:290–98
88. Jones JDG, Dangl JL. 2006. The plant immune system. Nature 444(7117):323–29
89. Katoh H, Miyata S, Inoue H, Iwanami T. 2014. Unique features of a Japanese “Candidatus Liberibacter
asiaticus” strain revealed by whole genome sequencing. PLOS ONE 9(9):e106109
90. Kim J-S, Sagaram US, Burns JK, Li J-L, Wang N. 2009. Response of sweet orange (Citrus sinen-
sis) to “Candidatus Liberibacter asiaticus” infection: microscopy and microarray analyses. Phytopathology
99(1):50–57
91. Koh E-J, Zhou L, Williams DS, Park J, Ding N, et al. 2012. Callose deposition in the phloem plas-
modesmata and inhibition of phloem transport in citrus leaves infected with “Candidatus Liberibacter
asiaticus.” Protoplasma 249(3):687–97

www.annualreviews.org • Liberibacter Pathogenesis Mechanisms and Disease Control 20.27

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

92. Kunkel LO. 1936. Heat treatment for the control of yellows and other virus diseases of peach.
Phytopathology 26:809–30
93. Lee I-M, Bottner KD, Sun M. 2009. An emerging potato purple top disease associated with a new 16SrIII
group phytoplasma in Montana. Plant Dis. 93(9):970
94. Leonard MT, Fagen JR, Davis-Richardson AG, Davis MJ, Triplett EW. 2012. Complete genome se-
quence of Liberibacter crescens BT-1. Stand. Genom. Sci. 7(2):271–83
95. Letunic I, Bork P. 2007. Interactive Tree Of Life (iTOL): An online tool for phylogenetic tree display
and annotation. Bioinformatics 23:127–28
96. Levy J, Ravindran A, Gross D, Tamborindeguy C, Pierson E. 2011. Translocation of “Candidatus Liberib-
acter solanacearum”, the zebra chip pathogen, in potato and tomato. Phytopathology 101(11):1285–91
97. Levy J, Tamborindeguy C. 2014. Solanum habrochaites, a potential source of resistance against Bac-
tericera cockerelli (Hemiptera: Triozidae) and “Candidatus Liberibacter solanacearum”. J. Econ. Entomol.
107(3):1187–93
98. Li J, Trivedi P, Wang N. 2016. Field evaluation of plant defense inducers for the control of citrus
Huanglongbing. Phytopathology 106(1):37–46
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

99. Li W, Cong Q, Pei J, Kinch LN, Grishin N V. 2012. The ABC transporters in Candidatus Liberibacter
asiaticus. Proteins Struct. Funct. Bioinform. 80(11):2614–28
100. Liefting LW, Perez-Egusquiza ZC, Clover GRG, Anderson JAD. 2008. A new “Candidatus Liberibacter”
species in Solanum tuberosum in New Zealand. Plant Dis. 92(10):1474
101. Liefting LW, Sutherland PW, Ward LI, Paice KL, Weir BS, Clover GRG. 2009. A new “Candidatus
Liberibacter” species associated with diseases of solanaceous crops. Plant Dis. 93(3):208–14
102. Liefting LW, Ward LI, Shiller JB, Clover GRG. 2008. A new “Candidatus Liberibacter” species in
Solanum betaceum (tamarillo) and Physalis peruviana (cape gooseberry) in New Zealand. Plant Dis.
92(11):1588
103. Lin H, Gudmestad NC. 2013. Aspects of pathogen genomics, diversity, epidemiology, vector dynamics,
and disease management for a newly emerged disease of potato: zebra chip. Phytopathology 103(6):524–37
104. Lin H, Islam MS, Bai Y, Wen A, Lan S, et al. 2012. Genetic diversity of “Candidatus Liberibacter
solanacearum” strains in the United States and Mexico revealed by simple sequence repeat markers. Eur.
J. Plant Pathol. 132(2):297–308
105. Lin H, Lou B, Glynn JM, Doddapaneni H, Civerolo EL, et al. 2011. The complete genome sequence of
“Candidatus Liberibacter solanacearum”, the bacterium associated with potato zebra chip disease. PLOS
ONE 6(4):e19135
106. Liu D, Trumble JT. 2007. Comparative fitness of invasive and native populations of the potato psyllid
(Bactericera cockerelli ). Entomol. Exp. Appl. 123(1):35–42
107. Liu D, Trumble JT, Stouthamer R. 2006. Genetic differentiation between eastern populations and recent
introductions of potato psyllid (Bactericera cockerelli ) into western North America. Entomol. Exp. Appl.
118(3):177–83
108. Lopes SA, Luiz FQBF, Martins EC, Fassini CG, Sousa MC, et al. 2013. “Candidatus Liberibacter asiati-
cus” titers in citrus and acquisition rates by Diaphorina citri are decreased by higher temperature. Plant
Dis. 97(12):1563–70
109. Lucas WJ, Groover A, Lichtenberger R, Furuta K, Yadav S-R, et al. 2013. The plant vascular system:
evolution, development and functions. J. Integr. Plant Biol. 55(4):294–388
110. Machida-Hirano R. 2015. Diversity of potato genetic resources. Breed. Sci. 65(1):26–40
111. MacLean AM, Sugio A, Makarova O V, Findlay KC, Grieve VM, et al. 2011. Phytoplasma effector SAP54
induces indeterminate leaf-like flower development in Arabidopsis plants. Plant Physiol. 157(2):831–41
112. Mafra V, Martins PK, Francisco CS, Ribeiro-Alves M, Freitas-Astúa J, Machado MA. 2013. Candidatus
Liberibacter americanus induces significant reprogramming of the transcriptome of the susceptible citrus
genotype. BMC Genom. 14(1):247
113. Martinelli F, Reagan RL, Uratsu SL, Phu ML, Albrecht U, et al. 2013. Gene regulatory networks
elucidating Huanglongbing disease mechanisms. PLOS ONE 8(9):e74256
114. Martinelli F, Uratsu SL, Albrecht U, Reagan RL, Phu ML, et al. 2012. Transcriptome profiling of citrus
fruit response to Huanglongbing disease. PLOS ONE 7(5):e38039

20.28 Wang et al.

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

115. Massonié G, Garnier M, Bove JM. 1976. Transmission of Indian citrus decline by Trioza erytreae (Del
Guercio), the vector of South African greening. Proc. Conf. Int. Organ. Citrus Virol., 7th, Athens, pp. 18–20.
Riverside, CA: IOCV
116. Matos LA, Hilf ME, Chen J, Folimonova SY, da Graça J, et al. 2013. Validation of “variable number of
tandem repeat”-based approach for examination of “Candidatus Liberibacter asiaticus” diversity and its
applications for the analysis of the pathogen populations in the areas of recent introduction. PLOS ONE
8(11):e78994
117. McCutcheon JP, Moran NA. 2011. Extreme genome reduction in symbiotic bacteria. Nat. Rev. Microbiol.
10(1):13
118. Munyaneza JE. 2012. Zebra chip disease of potato: biology, epidemiology, and management. Am. J.
Potato Res. 89(5):329–50
119. Munyaneza JE, Crosslin JM, Buchman JL. 2009. Seasonal occurrence and abundance of the potato
psyllid, Bactericera cockerelli, in south central Washington. Am. J. Potato Res. 86(6):513–18
120. Munyaneza JE, Fisher TW, Sengoda VG, Garczynski SF, Nissinen A, Lemmetty A. 2010. Association
of “Candidatus Liberibacter solanacearum” with the psyllid, Trioza apicalis (Hemiptera: Triozidae) in
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

Europe. J. Econ. Entomol. 103(4):1060–70


121. Murphy AF, Rondon SI, Jensen AS. 2013. First repot of potato psyllids, Bactericera cockerelli, overwin-
tering in the Pacific Northwest. Am. J. Potato Res. 90:294–96
122. Mustafa T, Alvarez JM, Munyaneza JE. 2015. Effect of cyantraniliprole on probing behavior of the
potato psyllid (Hemiptera: Triozidae) as measured by the electrical penetration graph technique. J. Econ.
Entomol. 108(6):2529–35
123. Mustafa T, Horton DR, Cooper WR, Swisher KD, Zack RS, et al. 2015. Use of electrical penetration
graph technology to examine transmission of “Candidatus Liberibacter solanacearum” to potato by three
haplotypes of potato psyllid (Bactericera cockerelli; Hemiptera: Triozidae). PLOS ONE 10(9):e0138946
124. Nakabachi A, Nikoh N, Oshima K, Inoue H, Ohkuma M, et al. 2013. Horizontal gene acquisition of
Liberibacter plant pathogens from a bacteriome-confined endosymbiont of their psyllid vector. PLOS
ONE 8(12):e82612
125. Narouei-Khandan HA, Halbert SE, Worner SP, van Bruggen AHC. 2016. Global climate suitability of
citrus Huanglongbing and its vector, the Asian citrus psyllid, using two correlative species distribution
modeling approaches, with emphasis on the USA. Eur. J. Plant Pathol. 144(3):655–70
126. Navarre DA, Shakya R, Holden J, Crosslin JM. 2009. LC-MS analysis of phenolic compounds in tubers
showing zebra chip symptoms. Am. J. Potato Res. 86(2):88–95
127. Nekrasov V, Staskawicz B, Weigel D, Jones JDG, Kamoun S. 2013. Targeted mutagenesis in the model
plant Nicotiana benthamiana using Cas9 RNA-guided endonuclease. Nat. Biotechnol. 31(8):691–93
128. Nelson WR, Fisher TW, Munyaneza JE. 2011. Haplotypes of “Candidatus Liberibacter solanacearum”
suggest long-standing separation. Eur. J. Plant Pathol. 130(1):5–12
129. Nelson WR, Sengoda VG, Alfaro-Fernandez AO, Font MI, Crosslin JM, Munyaneza JE. 2013. A new
haplotype of “Candidatus Liberibacter solanacearum” identified in the Mediterranean region. Eur. J.
Plant Pathol. 135(4):633–39
130. Newman M-A, von Roepenack-Lahaye E, Parr A, Daniels MJ, Dow JM. 2002. Prior exposure to
lipopolysaccharide potentiates expression of plant defenses in response to bacteria. Plant J. 29(4):487–95
131. Nissinen AI, Haapalainen M, Jauhiainen L, Lindman M, Pirhonen M. 2014. Different symptoms in
carrots caused by male and female carrot psyllid feeding and infection by “Candidatus Liberibacter
solanacearum.” Plant Pathol. 63(4):812–20
132. Okuda S, Sherman DJ, Silhavy TJ, Ruiz N, Kahne D. 2016. Lipopolysaccharide transport and assembly
at the outer membrane: the PEZ model. Nat. Rev. Microbiol. 14(6):337–45
133. Pagliai FA, Gonzalez CF, Lorca GL. 2015. Identification of a ligand binding pocket in LdtR from
Liberibacter asiaticus. Front. Microbiol. 6:1314
134. Park S-W, Kaimoyo E, Kumar D, Mosher S, Klessig DF. 2007. Methyl salicylate is a critical mobile
signal for plant systemic acquired resistance. Science 318(5847):113–16
135. Phahladira MNB, Viljoen R, Pietersen G. 2012. Widespread occurrence of “Candidatus Liberibacter
africanus subspecies capensis” in Calodendrum capense in South Africa. Eur. J. Plant Pathol. 134(1):39–47

www.annualreviews.org • Liberibacter Pathogenesis Mechanisms and Disease Control 20.29

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

136. Pitino M, Armstrong CM, Cano LM, Duan Y. 2016. Transient expression of Candidatus Liberibacter
asiaticus effector induces cell death in Nicotiana benthamiana. Front. Plant Sci. 7:982
137. Pitman AR, Drayton GM, Kraberger SJ, Genet RA, Scott IAW. 2011. Tuber transmission of “Candidatus
Liberibacter solanacearum” and its association with zebra chip on potato in New Zealand. Eur. J. Plant
Pathol. 129(3):389–98
138. Popa O, Dagan T. 2011. Trends and barriers to lateral gene transfer in prokaryotes. Curr. Opin. Microbiol.
14(5):615–23
139. Prager SM, Vindiola B, Kund GS, Byrne FJ, Trumble JT. 2013. Considerations for the use of neon-
icotinoid pesticides in management of Bactericera cockerelli (Šulk) (Hemiptera: Triozidae). Crop Prot.
54:84–91
140. Prasad S, Xu J, Zhang Y, Wang N. 2016. SEC-translocon dependent extracytoplasmic proteins of
Candidatus Liberibacter asiaticus. Front. Microbiol. 7:1–9
141. Price MN, Dehal PS, Arkin AP. 2010. FastTree 2: approximately maximum-likelihood trees for large
alignments. PLOS ONE 5:e9490
142. Puttamuk T, Zhou L, Thaveechai N, Zhang S, Armstrong CM, Duan Y. 2014. Genetic diversity of
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

Candidatus Liberibacter asiaticus based on two hypervariable effector genes in Thailand. PLOS ONE
9(12):e112968
143. Raddadi N, Gonella E, Camerota C, Pizzinat A, Tedeschi R, et al. 2011. “Candidatus Liberibacter
europaeus” sp. nov. that is associated with and transmitted by the psyllid Cacopsylla pyri apparently
behaves as an endophyte rather than a pathogen. Environ. Microbiol. 13(2):414–26
144. Ramadugu C, Keremane ML, Halbert SE, Duan YP, Roose ML, et al. 2016. Long-term field evaluation
reveals Huanglongbing resistance in Citrus relatives. Plant Dis. 100(9):1858–69
145. Rashed A, Wallis CM, Paetzold L, Workneh F, Rush CM. 2013. Zebra chip disease and potato bio-
chemistry: tuber physiological changes in response to “Candidatus Liberibacter solanacearum” infection
over time. Phytopathology 103(5):419–26
146. Richardson ML, Hall DG. 2013. Resistance of Poncirus and Citrus × Poncirus germplasm to the Asian
citrus psyllid. Crop Sci. 53:183–88
147. Roberts R, Pietersen G. 2016. A novel subspecies of “Candidatus Liberibacter africanus” found on native
Teclea gerrardii (Family: Rutaceae) from South Africa. Antonie Van Leeuwenhoek 110:437
148. Roberts R, Steenkamp ET, Pietersen G. 2015. Three novel lineages of “Candidatus Liberibacter
africanus” associated with native rutaceous hosts of Trioza erytreae in South Africa. Int. J. Syst. Evol.
Microbiol. 65(Pt. 2):723–31
149. Rush CM, Workneh F, Rashed A. 2015. Significance and epidemiological aspects of late-season infections
in the management of potato zebra chip. Phytopathology 105(7):929–36
150. Schultz KM, Merten JA, Klug CS. 2011. Effects of the L511P and D512G mutations on the Escherichia
coli ABC transporter MsbA. Biochemistry 50(13):2594–602
151. Schumann A, Singerman A. 2016. The economics of citrus undercover production systems and whole
tree thermotherapy. Citrus Ind. 2016:14–18
152. Secor GA, Rivera VV, Abad JA, Lee I-M, Clover GRG, et al. 2009. Association of “Candidatus Liberibac-
ter solanacearum” with zebra chip disease of potato established by graft and psyllid transmission, electron
microscopy, and PCR. Plant Dis. 93(6):574–83
153. Segers K, Anné J. 2011. Traffic jam at the bacterial Sec translocase: Targeting the SecA nanomotor by
small-molecule inhibitors. Chem. Biol. 18(6):685–98
154. Sengoda VG, Munyaneza JE, Crosslin JM, Buchman JL, Pappu HR. 2010. Phenotypic and etiological
differences between psyllid yellows and zebra chip diseases of potato. Am. J. Potato Res. 87(1):41–49
155. Serikawa RH, Backus EA, Rogers ME. 2012. Effects of soil-applied imidacloprid on Asian citrus psyllid
(Hemiptera: Psyllidae) feeding behavior. J. Econ. Entomol. 105(5):1492–502
156. Shan Q, Wang Y, Li J, Zhang Y, Chen K, et al. 2013. Targeted genome modification of crop plants
using a CRISPR-Cas system. Nat. Biotechnol. 31(8):686–88
157. Shen W, Cevallos-Cevallos JM, Nunes da Rocha U, Arevalo HA, Stansly PA, et al. 2013. Relation between
plant nutrition, hormones, insecticide applications, bacterial endophytes, and Candidatus Liberibacter
Ct values in citrus trees infected with Huanglongbing. Eur. J. Plant Pathol. 137(4):727–42

20.30 Wang et al.

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

158. Shi J, Pagliaccia D, Morgan R, Qiao Y, Pan S, et al. 2014. Novel diagnosis for citrus stubborn disease
by detection of a Spiroplasma citri–secreted protein. Phytopathology 104(2):188–95
159. Shokrollah H, Abdullah TL, Sijam K, Abdullah SNA. Ultrastructures of Candidatus Liberibacter asiaticus
and its damage in Huanglongbing (HLB) infected citrus. Afr. J. Biotechnol. 9(36):5897–901
160. Slisz AM, Breksa AP, Mishchuk DO, McCollum G, Slupsky CM. 2012. Metabolomic analysis of citrus
infection by “Candidatus Liberibacter” reveals insight into pathogenicity. J. Proteome Res. 11(8):4223–30
161. Stansly PA, Arevalo HA, Qureshi JA, Jones MM, Hendricks K, et al. 2014. Vector control and foliar nu-
trition to maintain economic sustainability of bearing citrus in Florida groves affected by Huanglongbing.
Pest Manag. Sci. 70(3):415–26
162. Stover E, Stange RR Jr., McCollum TG, Jaynes J, Irey M, Mirkov E. 2013. Screening antimicrobial
peptides in vitro for use in developing transgenic citrus resistant to Huanglongbing and citrus canker.
J. Am. Soc. Hortic. Sci. 138:142–48
163. Sugio A, Kingdom HN, MacLean AM, Grieve VM, Hogenhout SA. 2011. Phytoplasma protein effector
SAP11 enhances insect vector reproduction by manipulating plant development and defense hormone
biosynthesis. PNAS 108(48):E1254–63
164. Sutcliffe IC. 2010. A phylum level perspective on bacterial cell envelope architecture. Trends Microbiol.
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

18(10):464–70
165. Swisher KD, Henne DC, Crosslin JM. 2014. Identification of a fourth haplotype of Bactericera cockerelli
(Hemiptera: Triozidae) in the United States. J. Insect Sci. 14(1):161
166. Tahzima R, Maes M, Achbani EH, Swisher KD, Munyaneza JE, De Jonghe K. 2014. First report of
“Candidatus Liberibacter solanacearum” on carrot in Africa. Plant Dis. 98(10):1426
167. Teixeira DC, Saillard C, Eveillard S, Danet JL, da Costa PI, et al. 2005. “Candidatus Liberibacter
americanus”, associated with citrus Huanglongbing (greening disease) in São Paulo state, Brazil. Int. J.
Syst. Evol. Microbiol. 55:1857–62
168. Teixeira DC, Wulff NA, Martins EC, Kitajima EW, Bassanezi R, et al. 2008. A phytoplasma closely
related to the pigeon pea witches’-broom phytoplasma (16Sr IX) is associated with citrus Huanglongbing
symptoms in the state of São Paulo, Brazil. Phytopathology 98(9):977–84
169. Teresani G, Bertolini E, Alfaro-Fernandez A, Martı́nez C, Tanaka FAO, et al. 2014. Association of
“Candidatus Liberibacter solanacearum” with a vegetative disorder of celery in Spain and development
of a real-time PCR method for its detection. Phytopathology 104(8):804–11
170. Teresani G, Hernández E, Bertolini E, Siverio F, Marroquı́n C, et al. 2015. Search for potential vec-
tors of “Candidatus Liberibacter solanacearum”: population dynamics in host crops. Span. J. Agric. Res.
13(1):e1002
171. Thinakaran J, Pierson E, Kunta M, Munyaneza JE, Rush CM, Henne DC. 2015. Silverleaf nightshade
(Solanum elaeagnifolium), a reservoir host for “Candidatus Liberibacter solanacearum”, the putative causal
agent of zebra chip disease of potato. Plant Dis. 99(7):910–15
172. Thomas KL, Jones DC, Kumarasinghe LB, Richmond JE, Gill GSC, Bullians MS. 2011. Investigation
into the entry pathway for tomato potato psyllid Bactericera cockerelli. N. Z. Plant Prot. 64:259–68
173. Thompson S, Fletcher JD, Ziebell H, Beard S, Panda P, et al. 2013. First report of “Candidatus Liberib-
acter europaeus” associated with psyllid infested scotch broom. New Dis. Rep. 27:6
174. Thompson SM, Johnson CP, Lu AY, Frampton RA, Sullivan KL, et al. 2015. Genomes of “Candidatus
Liberibacter solanacearum” haplotype a from New Zealand and the United States suggest significant
genome plasticity in the species. Phytopathology 105(7):863–71
175. Tyler HL, Roesch LFW, Gowda S, Dawson WO, Triplett EW. 2009. Confirmation of the sequence
of “Candidatus Liberibacter asiaticus” and assessment of microbial diversity in Huanglongbing-infected
citrus phloem using a metagenomic approach. Mol. Plant-Microbe Interact. 22(12):1624–34
176. Vahling-Armstrong CM, Zhou H, Benyon L, Morgan JK, Duan Y. 2012. Two plant bacteria, S. meliloti
and Ca. Liberibacter asiaticus, share functional znuABC homologues that encode for a high affinity zinc
uptake system. PLOS ONE 7(5):e37340
177. Vereijssen J SI. 2013. Psyllid can overwinter on non-crop host plants. N. Z. Grow. 68(1):14–15
178. von Wintersdorff CJH, Penders J, van Niekerk JM, Mills ND, Majumder S, et al. 2016. Dissemination
of antimicrobial resistance in microbial ecosystems through horizontal gene transfer. Front. Microbiol.
7:173

www.annualreviews.org • Liberibacter Pathogenesis Mechanisms and Disease Control 20.31

Changes may still occur before final publication online and in print
PY55CH20-Wang ARI 13 June 2017 13:36

179. Vuorisalo T, Hutchings MJ. 1996. On plant sectoriality, or how to combine the benefits of autonomy
and integration. Vegetatio 127:3
180. Walter AJ, Duan YP, Hall DG. 2012. Titers of “Ca. Liberibacter asiaticus” in Murraya paniculata and
Murraya-reared Diaphorina citri are much lower than in citrus and citrus-reared psyllids. HortScience
47:1449–52
181. Wang N, Trivedi P. 2013. Citrus Huanglongbing: a newly relevant disease presents unprecedented
challenges. Phytopathology 103(7):652–65
182. Wen A, Mallik I, Alvarado VY, Pasche JS, Wang X, et al. 2009. Detection, distribution, and genetic
variability of “Candidatus Liberibacter” species associated with zebra complex disease of potato in North
America. Plant Dis. 93(11):1102–15
183. Westbrook CJ, Hall DG, Stover E, Duan YP, Lee RF. 2011. Colonization of citrus and citrus-related
germplasm by Diaphorina citri (Hemiptera: Psyllidae). HortScience 46:997–1005
184. Woo JW, Kim J, Kwon SI, Corvalán C, Cho SW, et al. 2015. DNA-free genome editing in plants with
preassembled CRISPR-Cas9 ribonucleoproteins. Nat. Biotechnol. 33(11):1162–64
185. Wulff NA, Zhang S, Setubal JC, Almeida NF, Martins EC, et al. 2014. The complete genome sequence of
Annu. Rev. Phytopathol. 2017.55. Downloaded from www.annualreviews.org
Access provided by Columbia University on 06/22/17. For personal use only.

“Candidatus Liberibacter americanus”, associated with citrus Huanglongbing. Mol. Plant-Microbe Interact.
27(2):163–76
186. Wuriyanghan H, Rosa C, Falk BW. 2011. Oral delivery of double-stranded RNAs and siRNAs induces
RNAi effects in the potato/tomato psyllid, Bactericerca cockerelli. PLOS ONE 6(11):e27736
187. Xia Y, Ouyang G, Sequeira RA, Sequeira R, Sequeira Y, et al. 2011. A review of Huanglongbing
(citrus greening) management in citrus using nutritional approaches in China. Plant Health Prog.
https://doi.org/10.1094/PHP-2010-1003-01-RV
188. Yan Q, Sreedharan A, Wei S, Wang J, Pelz-Stelinski K, et al. 2013. Global gene expression changes in
Candidatus Liberibacter asiaticus during the transmission in distinct hosts between plant and insect. Mol.
Plant Pathol. 14(4):391–404
189. Yang C, Powell CA, Duan Y, Shatters R, Zhang M, et al. 2015. Antimicrobial nanoemulsion formulation
with improved penetration of foliar spray through citrus leaf cuticles to control citrus Huanglongbing.
PLOS ONE 10(7):e0133826
190. Yao J, Saenkham P, Levy J, Ibanez F, Noroy C, et al. 2016. Interactions “Candidatus Liberibacter
solanacearum”–Bactericera cockerelli: haplotype effect on vector fitness and gene expression analyses. Front.
Cell. Infect. Microbiol. 6:62
191. Zhang M, Guo Y, Powell CA, Doud MS, Yang C, Duan Y. 2014. Effective antibiotics against “Candidatus
Liberibacter asiaticus” in HLB-affected citrus plants identified via the graft-based evaluation. PLOS ONE
9(11):e111032
192. Zhang S, Flores-Cruz Z, Zhou L, Kang B-H, Fleites LA, et al. 2011. “Ca. Liberibacter asiaticus” carries
an excision plasmid prophage and a chromosomally integrated prophage that becomes lytic in plant
infections. Mol. Plant-Microbe Interact. 24(4):458–68
193. Zhang Y, Liang Z, Zong Y, Wang Y, Liu J, et al. 2016. Efficient and transgene-free genome editing in
wheat through transient expression of CRISPR/Cas9 DNA or RNA. Nat. Commun. 7:12617
194. Zheng Z, Bao M, Wu F, Chen J, Deng X. 2016. Predominance of single prophage carrying a CRISPR/Cas
system in “Candidatus Liberibacter asiaticus” strains in southern China. PLOS ONE 11(1):e0146422
195. Zheng Z-L, Zhao Y. 2013. Transcriptome comparison and gene coexpression network analysis provide
a systems view of citrus response to “Candidatus Liberibacter asiaticus” infection. BMC Genom. 14(1):27
196. Zhong Y, Cheng C-Z, Jiang N-H, Jiang B, Zhang Y-Y, et al. 2015. Comparative transcriptome and
iTRAQ proteome analyses of citrus root responses to Candidatus Liberibacter asiaticus infection. PLOS
ONE 10(6):e0126973
197. Zipfel C, Robatzek S. 2010. Pathogen-associated molecular pattern-triggered immunity: Veni, vidi...?
Plant Physiol. 154:551–54
198. Zou H, Gowda S, Zhou L, Hajeri S, Chen G, Duan Y. 2012. The destructive citrus pathogen, “Candidatus
Liberibacter asiaticus” encodes a functional flagellin characteristic of a pathogen-associated molecular
pattern. PLOS ONE 7(9):e46447

20.32 Wang et al.

Changes may still occur before final publication online and in print

You might also like