Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Heliyon 10 (2024) e34275

Contents lists available at ScienceDirect

Heliyon
journal homepage: www.cell.com/heliyon

Review article

An insight into pathogenicity and virulence gene content of


Xanthomonas spp. and its biocontrol strategies
Riddha Dey , Richa Raghuwanshi *
Department of Botany, Mahila Mahavidyalaya, Banaras Hindu University, Varanasi, 221005, Uttar Pradesh, India

A R T I C L E I N F O A B S T R A C T

Keywords: The genus Xanthomonas primarily serves as a plant pathogen, targeting a diverse range of
Xanthomonas economically significant crops on a global scale. Xanthomonas spp. utilizes a collection of toxins,
Pathogenicity adhesins, and protein effectors as part of their toolkit to thrive in their surroundings, and establish
Biocontrol
themselves within plant hosts. The bacterial secretion systems (Type 1 to Type 6) assist in
Virulence genes
delivering the effector proteins to their intended destinations. These secretion systems are
specialized multi-protein complexes responsible for transporting proteins into the extracellular
milieu or directly into host cells. The potent virulence and systematic infection system result in
rapid dissemination of the bacteria, posing significant challenges in management due to com­
plexities and substantial loss incurred. Consequently, there has been a notable increase in the
utilization of chemical pesticides, leading to bioaccumulation and raising concerns about adverse
health effects. Biological control mechanisms through beneficial microorganism (Bacillus, Pseu­
domonas, Trichoderma, Burkholderia, AMF, etc.) have proven to be an appropriate alternative in
integrative pest management system. This review details the pathogenicity and virulence factors
of Xanthomonas, as well as its control strategies. It also encourages the use of biological control
agents, which promotes sustainable and environmentally friendly agricultural practices.

1. Introduction

Xanthomonas is a Gram-negative, yellow-colored, plant-associated bacterium and is classified under the Gammaproteobacteria,
comprising about 30 species that cause severe diseases in around 400 plant hosts. This includes economically important crops, like
rice, tomato, citrus, pepper etc. This phytopathogen shows high specificity towards the host plant and tissue. It invades xylem elements
and intercellular space between mesophyll parenchyma tissue. Xanthomonas genomes encompass different Mobile Genetic Elements
(MGEs), which are major factors for virulence, genetic variations, and genome structure [1]. Examples of MGEs are transposons (Tn),
Genomic Islands (GI), Insertion Sequence (IS), and plasmids. Genomic islands in Xanthomonas spp. serve multifaceted functions that
are intertwined with bacterial pathogenicity, persistence, and evolutionary processes [1]. Recombination and horizontal gene transfer
play pivotal roles in shaping the population structure and fostering diversity within the pathosystems of Xanthomonas [2]. Xantho­
monas possess six main secretion systems among which the type III secretion system plays a substantial role in suppressing host de­
fenses and promoting the progression of the disease. A multitude of factors like type III effectors, adhesins, lipopolysaccharides, and
transcription factors, are identified to be critical elements regarding host specificity and bacterial pathogenicity [3]. Comprehending
the pathogenicity of Xanthomonas has consistently posed a challenge because of its extensive host range and intricate infection

* Corresponding author.
E-mail address: richabhu@yahoo.co.in (R. Raghuwanshi).

https://doi.org/10.1016/j.heliyon.2024.e34275
Received 12 October 2023; Received in revised form 24 June 2024; Accepted 7 July 2024
Available online 8 July 2024
2405-8440/© 2024 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/).
R. Dey and R. Raghuwanshi Heliyon 10 (2024) e34275

Glossary

AMF Arbuscular mycorrhizal fungi


BCA Biological control agents
EPS Exopolysaccharides
ETI Effector-Triggered Immunity
Hrp: Hypersensitive response and pathogenicity
LPS Lipopolysaccharides
M/DAMP-TI Microbe/Damage Associated Molecular Pattern-Triggered Immunity
MDCA 3,4-(methylenedioxy) cinnamic acid
NRPS-PKS Non-Ribosomal Peptide Synthetase-Polyketide Synthase
PTI Pathogen-Triggered Immunity
RTX Repeats-in-toxins
T1SS Type I Secretion System
T2SS Type II Secretion System
T3E T3 effector
T3SS Type III Secretion System
T4SS Type IV Secretion System
T5SS Type V Secretion System
T6SS Type VI Secretion System
TALE Transcription activator-like effectors
Xop: Xanthomonas outer proteins

pathways. Functional and comparative genomic studies have proven invaluable in uncovering how this bacterial group has evolved to
exploit an exceptionally wide array of plant hosts. Enhanced insights into the pathogenic adaptations exhibited by Xanthomonas spp.
and their infection pathways hold the potential to drive advancements in the prevention and management of bacterial diseases of
plants.
This review focuses on an elaborate study of the lifecycle of Xanthomonas spp. and consolidates the latest developments encom­
passing the macromolecular secretion systems, ranging from type I to type VI (T1SS to T6SS), as they pertain to Xanthomonas spp. This
also includes an overview of chemical and biological control methods for the agricultural management of Xanthomonas to mitigate the
occurrence of diseases.

1.1. Xanthomonas lifestyle

The genus Xanthomonas is known for its phytopathogenic diversity. It belongs to the family Xanthomonadaceae, subclass Gam­
maproteobacteria, and is a Gram-negative bacterial genus. It also has a contrasting phenotypic uniformity. The Xanthomonas colonies
are morphologically yellow due to membrane-bound xanthomonadin pigment, which protects them from photobiological damage
[4–6]. Initially, members of the Xanthomonas genus were classified into distinct species based on their infection capability to specific
hosts. However, following classical naming conventions, the majority of these species were subsequently consolidated into a singular
entity called Xanthomonas campestris (X. campestris) [7]. This collective group was further categorized into various pathovars.
Nonetheless, owing to multiple taxonomic revisions, there remains ongoing debate regarding the nomenclature of the roughly 19
species and 4140 pathovars within the Xanthomonas genus [8].
Different Xanthomonas strains cause bacterial disease in more than 400 hosts, including cabbage, tomato, cassava, wheat, rice,
pepper, citrus, and beans [9]. Although they show varied disease phenotypes and diverse host ranges, the maximum similarity is
observed at the genomic level. Most of the genes are conserved sequences. A single phylogenetic group was represented by the first two
sequenced X. perforans strains discovered in Florida in the 1990s [2,10]. This genus has been diversely studied due to its economic loss
and large genetic diversity. X. citri subsp. citri causal organism for citrus canker, which is broadly spread around the world, especially in
Brazil, USA, Mexico, India, and Japan [11]. X. campestris pv. campestris is the most destructive phytopathogen responsible for causing
black rot disease. The infection results in the darkening of vascular tissue and the development of foliar lesions, which can vary from
yellowing to tissue death (necrosis). This leads to vegetable decay and renders it unsuitable for consumption [12].
It causes significant economic losses to the Brassicaceae family including cabbage, broccoli, cauliflower, and kale [13]. Bacterial
leaf spot is a highly challenging disease observed in tomato and pepper resulting in a great economic loss. The causal agents include
X. vesicatoria, X. euvesicatoria, X. perforans, and X. gardneri [14,15]. Xanthomonas wilt, caused by X. campestris pv. musacearum, is a
significant disease that has spread globally. It was initially observed in Ensete in 1968 [16] and later in bananas in Ethiopia in 1974
[17]. Since 2001, it has been frequently reported in various countries, including Kenya [18], Tanzania [19], Uganda [20], the
Democratic Republic of the Congo [21], Rwanda, Burundi [18]. The disease has lately been recorded in the Democratic Republic of the
Congo in the South Kivu communities of Uvira and Fizi, in the northern Katanga territory of Kalemie, and in the Tshopo area of
Oriental Province.
Xanthomonas exhibit both epiphytic and endophytic life cycles, which involve distinct interactions with their host plants as

2
R. Dey and R. Raghuwanshi Heliyon 10 (2024) e34275

depicted in Fig. 1. The epiphytic life cycle of Xanthomonas refers to the phase of the bacteria’s lifecycle when it resides on the external
surfaces of plant tissues without causing an active infection. It starts with adhesion and colonization. They proceed to infiltrate the host
tissue through natural openings or injuries on the plant’s surface. Conventionally, they are exposed to harsh environmental conditions
on the surface of plants [22]. Despite these challenges, Xanthomonas is capable of persisting on leaf surfaces for prolonged durations
[23]. The survival time of Xanthomonas on leaf surfaces is not certain; it can vary from a few weeks to several months [23]. Kayaaslan
et al. [24], stated that Xanthomonas spp. have the ability to persist epiphytically on seeds, and it can be viable for up to 18 months. Once
attached to the surface, the bacteria start to aggregate and form microcolonies on the plant’s surface. These microcolonies provide
protection from environmental stresses, such as desiccation and UV radiation. These microcolonies can develop into more complex
structures known as biofilms to provide additional protection and help bacteria withstand harsh conditions [25]. Xanthomonas biofilm
is a dynamic structure, and both its formation and dispersal are regulated by the quorum sensing signal molecule known as the
diffusible signal factor (DSF). DSF, in conjunction with the internal secondary messenger cyclic di-GMP, plays a role in promoting
biofilm formation by inducing the production of exopolysaccharides (EPS) and the assembly of pili [26].
The primary constituent of the biofilm matrix in Xanthomonas spp. is typically Xanthan gum [27]. During the endophytic life cycle,
they gain access to the host tissue through natural openings (hydathodes, stomata, and wounds). Bacterial cells need to adjust their
physiology according to the particular conditions they confront during colonization, whether it is within the host mesophyll (for
non-vascular pathogens) or xylem vessel tissues (for vascular pathogens). Subsequently, they progress by migrating towards the central
veins for multiplication during which bacteria proliferate within the host tissue. As the bacterial population increases, they resurface
on the leaf’s exterior. Through dispersion via rain and wind, they find their way into a new host, initiating a fresh infection cycle. When
the bacterial population reaches a critical threshold, they invade the mesophyll tissues, giving rise to the characteristic disease
symptoms on the leaves [28].

1.2. Virulence factors

The severity of pathogen infection is virulence. The virulence and pathogenicity imparting genes and gene clusters in Xanthomonas
spp. consist of six types of protein secretion systems. Each protein type differs in function and composition [29] and plays an important
role in host-pathogen interaction (Fig. 2).

1.3. Type I secretion system (T1SS)

The main role of the T1SS is to secrete unfolded cognate substrates directly from the cytoplasm to the extracellular medium
involving an intermediary periplasmic state [30] which results in adhesion to the host cell. The T1SS is composed of a heterotrimeric
protein complex, which includes an ATP-binding cassette transporter in the inner leaflet, a protein channel in the outer leaflet, and a

Fig. 1. Life cycle of Xanthomonas spp.

3
R. Dey and R. Raghuwanshi Heliyon 10 (2024) e34275

Fig. 2. Mechanism of action of secretion systems of Xanthomonas spp.

membrane fusion protein that bridges the inner and outer leaflet. RaxST (tyrosine sulfotransferase), RaxA (membrane fusion protein),
RaxB (adenosine triphosphate–binding cassette transporter), RaxC (outer membrane protein) marked the existence of T1SS [31].
These proteins recognize effectors containing double glycine leader peptide. It was reported that in X. perforans, raxST, raxA, and raxB
are encoded in a same operon [32] whereas the raxC is located outside the raxSTAB operon [33]. The raxABC T1SS is regulated by raxH
and raxR which are located downstream of the raxSTAB operon [34]. It was stated that on mutating the raxX gene, the virulence in rice
was impaired, concluding that it promotes infection of X. oryzae pv. oryzae [35]. T1SS substrates belong to three different groups
including bacteriocin, proteins with non-peptide RTX motif (Repeats in Toxins), and large adhesin containing RTX at the C-terminal.
RTX proteins make up the majority of T1SS effectors. It includes proteases, adhesins, lipases, toxins, and may feature different
quantities of RTX motifs along with an uncleavable T1SS-specific C-terminal secretion signal. RTX proteins are conveyed in their
unfolded state, and once they are secreted, extracellular Ca2+ binds to the RTX motif, prompting their folding [36]. Fonseca et al. [37],
reported that in X. citri pv. citri 306 has hlyDB T1SS that secretes the two RTX-related effectors XAC2197 and XAC2198 which plays role
in its pathogenicity. However, it was also observed that X. citri pv. aurantifolii genome lacks the hlyDB T1SS, which makes it less
aggressive than X. citri pv. citri.

1.4. Type II secretion system (T2SS)

T2SS is crucial for secreting enzymes like proteases, lipases, cell wall lyase enzymes, and xylanases which results in host cell wall
degradation. It also transports protein from Xanthomonas periplasm to the extracellular milieu. The T2SS apparatus comprises a total of
12 proteins, which are responsible for various functions. These proteins include a cytoplasmic ATPase known as gspE, an inner
membrane platform complex composed of gspC, gspF, gspL, and gspM, an outer membrane channel represented by gspD, a prepilin
peptidase called gspO, and a pseudopilus formed by gspG, gspH, gspI, gspJ, and gspK. Together, these components work in concert to
eject the substrate. Recent transcriptome analysis indicates that the gspD mutant in the X. translucens pv. cerealis strain NXtc01 exhibits
a marked reduction in disease development in wheat, highlighting the pivotal role of the T2SS in enhancing virulence [38]. The
secretin protein family forms a complex transmembrane channel in the outer membrane. Substances are secreted through this channel
via a periplasmic pilus, which undergoes constant assembly and disassembly, effectively propelling T2SS substrates [39].
In the Xanthomonas genus, there are two distinct gene clusters associated with T2SS: xps, which is found consistently across the
entire genus, and xcs, which is present alongside xps in certain species like X. citri and X. campestris but is absent in others, including
X. oryzae strains [40,41]. In X. perforans, the T2SS is encoded by both the xps and xcs gene clusters [42]. While xps in T2SS is widely
considered the primary mechanism for secreting effectors and influencing virulence in various Xanthomonas species, the role of xcs

4
R. Dey and R. Raghuwanshi Heliyon 10 (2024) e34275

T2SS remains unclear as it has not been associated with any specific function or role to date [43,44]. Nonetheless, a recent RNA-seq
investigation carried out by Zhang et al., [45] showed that when the genes for stringent response regulators like the ppGpp synthases
SpoT and RelA and transcription factor DskA are mutated in X. citri pv. citri, it led to a significant reduction in xcs gene expression, while
xps genes remained unaffected. This finding raised curiosity because the stringent response typically played a role when the organism
faces stressors like limited nutrients. It suggested a potential, albeit unexplored, role for this enigmatic T2SS in aiding cell survival
during challenging environmental conditions. In X. euvesicatoria, the T2SS contributes to enhancing virulence, and its activity is
governed by master regulators HrpX and HrpG, which also control the T3SS [46]. While the detailed role of T2SS in pathogenicity is not
thoroughly elucidated in X. perforans, it is speculated to be comparable to its function in X. euvesicatoria.

1.5. Type III secretion system (T3SS)

T3SS plays the main role in pathogenesis. Effector protein translocation inside host cells and pilus structure assembly are the major
virulence factors enabling host specialization and infection [47]. This system alters the behavior of the host cell and directly delivers
the T3 effector (T3E) proteins into it through Hrp pilus. The core secretion apparatus is conserved among pathogenic bacteria. T3SS in
Xanthomonas spp is encoded by the hypersensitive response and pathogenicity (hrp) cluster, which in turn is regulated by hrpX and
hrpG master switches, consisting of an array of genes organized in transcriptional units [48]. The Hrp pilus spans the host plant cell
wall and connects the T3SS translocon. Individual translocon components mutate, leading to reduction or loss of pathogenicity. The
T3E protein translocation plays a vital role in both bacterial cell proliferation and the emergence of disease symptoms [49–51].
Colonization of the host milieu is promoted by T3E proteins, which suppress the immune response. The T3E Xanthomonas spp. is a
member of the AvrBs3/PthA family. The absence of either pthA or AvrBs3 from their respective strains led to a significant decrease in
the extent of infection of pathogen replication capacity within the host tissue leading to reduced disease symptoms. The existence of
the Xop effector gene represents a distinct characteristic specific to Xanthomonas spp. Yan et al. [52] discovered the Min system,
composed of MinC, MinD, and MinE proteins, constituted a regulatory mechanism that negatively influenced bacterial cell division.
This system is encoded by the genes minC, minD, and minE and is involved in the negative regulation of hrpB1 and hrpF, in X. oryzae pv.
oryzae. Advanced approaches like machine learning have been developed for detailed searching of T3E in Xanthomonas genomes based
on parameters like GC content, codon usage, gene homology, structural patterns, amino-terminal secretion signals, and HrpG/HrpX
signaling pathway [53]. Effectors play a dual role in Xanthomonas biology, being not only crucial for pathogenicity but also acting as
factors that determine host specificity and contribute to pathogen fitness.

1.6. T3SS effectors

TAL effectors: Transcription activator-like effectors (TALE) are a discrete family of T3E proteins present in most of the species of
the genus Xanthomonas, involved in enhancing the bacterial plasticity to host plants. The rearrangeable repetitive domains aid the
binding of promoters of host susceptibility genes sequence-specifically [54,55]. The rearranged central-repeat domains of TAL proteins
recognize EBE (Effector Binding Elements) situated in promoters [56]. Further, TALEs are not observed in X. gardneri, X. campestris, and
X. euvesicatoria [10,57]. However, TALEs are predominant in X. oryzae [58]. Additionally, X. translucens pv. undulosa deploys the TALE
Tal8 that modifies the abscisic acid pathway to increase disease susceptibility in wheat [59] and X. oryzae pv. oryzae. TALE targets the
sugar transporters (encoded by SWEET genes) in rice leading to bacterial blight disease. OsSWEET11/Os8N3/Xa13 was the first gene
discovered among the SWEET gene family involved in host-bacterial pathogen interaction in bacterial blight of rice caused by X. oryzae
pv. oryzae [60,61]. The rice plants with silenced OsSWEET11 showed resistance towards Xanthomonas oryzae pv. oryzae [61]. X. oryzae
pv. oryzae strain PXO99A released a distinct type III TAL effector PthXo1, which attached to the promoter of OsSWEET11 in rice,
thereby triggering its transcription [62]. Streubel et al. [63], proposed that PthXo3 (TalBH), TalC, and AvrXa7 targeted OsSWEE­
T14/Os11N3. This resulted in the accumulation of sugar in the apoplast and xylem that led to bacterial colonization [64,65]. In
X. oryzae pv. oryzae TALE (PthXo1 and PthXo2) target OsSWEET11 and OsSWEET13 [65]. The SWEET gene (GhSWEET10) belonging to
clade III in Gossypium sp. (cotton) similarly made the plant susceptible to X. citri subsp. malvacearum infection, which is targeted by the
TAL effector Avrb6 [66]. TAL20Xam66, secreted by X. axonopodis pv. manihoti, the causal agent of bacterial blight, induced the
expression of MeSWEET10a in Manihot esculentum (cassava), aiding the pathogen in obtaining hexose from the host plants [67]. X. citri
subsp. citri secretes TAL as a component of its infection strategy, stimulating the expression of CsSWEET1 in Citrus aurantiifolia (lime),
thereby leading to the development of bacterial canker disease [68]. To date, only clade III SWEET genes have been identified as targets
of TAL effectors produced by bacterial pathogens.

1.7. Xop (Xanthomonas outer proteins)

Xanthomonas is exposed to host immune responses after entering the plant tissues. Some examples of the responses are ETI
(Effector-Triggered Immunity) and M/DAMP-TI (Microbe/Damage Associated Molecular Pattern-Triggered Immunity [69]. T3SS ef­
fectors (T3E) are Xop proteins or Avr (avirulence), that aid Xanthomonas to cope with the immune responses. These Xop are delivered
inside the host cell using T3SS [69]. There are 53 Xop families (XopA- Xop BA) identified to date, which are elaborately discussed by
White et al. [70]. Further, the PTI pathways are targeted by Xanthomonas T3Es such as XopR, XopK, XopP, XopN, XopL, XopS, XopB,
and AvrAC, AvrBs2 [71,72] whereas DAMP-TI is inhibited by XopQ, XopX, XopZ, and XopN. ETI and PTI responses are interfered by
XopH and XopQ [31,73]. XopD and XopJ disrupt the host’s ubiquitin proteasomal system, leading to suppressing immunity targeting
the PTI factors and defense mediated by salicylic acid [31]. XopD consists of SUMO protease (Small Ubiquitin-like Modifier) (C48

5
R. Dey and R. Raghuwanshi Heliyon 10 (2024) e34275

protease family member). XopD is responsible for the de-sumoylation process from SUMO-modified plant proteins [74,75]. It is also
observed that XopD is nuclear localized and has nonspecific DNA binding activity [74,76]. Teper et al. [77] proposed that XopD
possesses a helix-loop-helix domain essential for DNA binding, along with two conserved ERF-associated amphiphilic motifs that
suppress transcription induced by salicylic acid and jasmonic acid in plants. Also, XopD inhibits the development of symptoms during
the latter phases of infection in tomato leaves. Enhanced population and genomic databases as well as machine-learning techniques,
have advanced the ability to identify and understand Xops and their interactions with hosts [78].

1.8. Type IV secretion system (T4SS)

The primary function of the T4SS is to facilitate the transfer of proteins and protein-DNA complexes into specific target cells. This
system is responsible for various processes, including bacterial conjugation, delivery of T-DNA into A. tumefaciens, the transmission of
protein-based effectors into host cells as well as the conveyance of harmful effectors into the cells of competing bacterial species. In
Xanthomonas spp. two classes of T4SS are present, the established T4ASS and T4BSS [42]. T4BSS is larger and more complex than
T4ASS. Xanthomonas spp. carries genes responsible for encoding the essential 12 subunits present in T4ASS. These subunits can be
grouped into three main categories: subunits responsible for the outer membrane pore formation, VirB7, VirB9, and VirB10; subunits
responsible for inner membrane pore formation, VirB3, VirB4, VirB6, VirB8, VirB11, VirD4, and the N-terminal domain of VirB10; and
subunits VirB2 and VirB5, are anticipated to create an extracellular pilus structure. Additionally, VirB1 is predicted to function as a
periplasmic lytic transglycosylase, contributing to peptidoglycan remodeling during the biogenesis of the T4SS [79]. Nonetheless,
recent research in X. citri has revealed that a T4SS located on a plasmid can transfer toxic effectors into neighboring bacteria during
competition, leading to the demise of those other bacterial cells [80]. VirD4, an ATPase located in the inner membrane, identifies and
transports harmful substances through a shared C-terminal domain known as ‘Xanthomonas VirD4 interacting protein conserved do­
mains’ (XVIPCDs), utilizing the T4SS pilus as a conduit [31]. The transport of phytotoxins ultimately disrupts the electric potential
across the plant membrane, culminating in the demise of plant cells [81,82]. X. oryzae pv. oryzae strain PXO99A lacks the entire T4SS.

1.9. Type V secretion system (T5SS)

The primary function of the T5SS is to transport effectors, including adhesins, enzymes, and toxins. It shows non-fimbrial adhesins.
These effectors have a significant role in processes such as adhesion to both host and non-host plants, the formation of biofilms, and the
aggregation of cells [31]. It is structurally simple consisting of two main domains: β-barrel present in the outer membrane and a
secreted passenger protein. T5SS have been categorized into subclasses (Va to Ve), with a newly described Vf subclass identified
specifically in H. pylori. This does not require sources of energy such as ATP or electrochemical gradients for transportation [83]. Va
plays the role of auto-transporter, whereas Vb is known as TPS (Two Partner Secretion System) constituted of a translocator protein
and a cognate passenger protein [84]. The transportation process to periplasm from the cytoplasm is carried out through Sec trans­
locase pathway [85]. Vc are known as TAA (Trimeric Autotransporter Adhesins) and remain in the outer membrane. The exported
effectors either get released to the extracellular environment or remain attached to the transporter. Current research strongly indicates
that labeling T5SS as “autotransporters” is inaccurate. These proteins, in fact, necessitate the involvement of numerous external factors
to facilitate their export to the bacterial cell surface, with distinct requirements at various stages of the process [86].

1.10. Type VI secretion system (T6SS)

The primary function of T6SS is to facilitate antagonistic behaviour by delivering toxic substances. This system is classified into five
clades (i1 to i5) and is present in all bacterial species that are associated with plants, whether it’s pathogenic or non-pathogenic. T6SS
is much more involved in bacterial competition rather than virulence. It confers advantages in terms of fitness by enhancing
competitiveness against other members of the bacterial community [87]. There are three main structures of T6SS in the cell envelope,
which accumulate together to function in the contractile system: a membrane complex, a baseplate, and an inner tube extension
enveloped by a contractile sheath. Recent studies proved that T6SS is subdivided into three main clusters on the basis of gene content;
T6SS-I, T6SS-II, and T6SS-III. T6SS-I were found in X. phaseoli, X. euvesicatoria, X. oryzae, X. axonopodis, and X. vasicola. Only three
species including X. oryzae, X. hortorum, and X. fragarie contain a full T6SS-II cluster. T6SS-III is present in X. citri pv. citri and
phylogenetically related species. X. maliensis is an exception as it has complete T6SS-I and partial T6SS-II and T6SS-III clusters [88].
Moreover, according to studies by Choi et al. [89] and Zhu et al. [90], the removal of hcp in T6SS-2 had a partial detrimental effect on
the virulence of X. oryzae pv. oryzae and X. oryzea pv. oryzicola. However, when hcp was deleted in T6SS-1, it did not lead to any
significant differences in virulence. However, the function of all the factors of T6SS in colonization and antagonism is yet to be
explored in detail.

1.11. Bacterial structures causing virulence

Structural elements responsible for virulence are crucial for the infection and the pathogenic acclimatization inside the host.
Bacterial capsular polysaccharides are important in virulence as they obstruct the actions of host phagocytes [91]. These virulence
factors like xanthan, toxins, and adhesion factors are highly accountable for the pathogenicity of Xanthomonas.
EPS Xanthan: The pathogenic factors in Xanthomonas include exopolysaccharides (EPS) that form biofilms. Xanthan, EPS, is a
distinctive feature found in all Xanthomonads, giving rise to the mucous or mucoid appearance of the bacterial colonies which is a

6
R. Dey and R. Raghuwanshi Heliyon 10 (2024) e34275

major component of its biofilm matrix [92]. Xanthan is made of cellulose-like (1 → 4)-β-D-glucose, containing a (1 → 3)-linked
trisaccharide side chain of β-d-mannose-(1 → 4)-βD-glucuronic acid-(1 → 2)-α-D-mannose. Xanthan’s capacity to shield bacteria from
environmental challenges like desiccation and harmful substances stems from its high-water content and anionic nature. However,
xanthan, in vascular pathogens, can lead to blocking of water flow in xylem vessels resulting in the wilting of host plants [93,94]. It also
suppresses callose deposition inducing plant susceptibility to Xanthomonas [95]. Biofilms formed by Xanthomonas secure it from host
defense response and antibiotics produced. Furthermore, it likely controls the bacteria’s survival on the plant’s surface before entering
the intercellular space, or it may affect the adherence of vascular bacteria to xylem vessels [96]. Several genetic loci such as the highly
conserved gum gene cluster (consisting of a total of 12 genes from gumB to gumM) direct the production of xanthan [97,98].
Toxins: Among Xanthomonas spp., the production of the toxin is presumed to be constrained to the X. albilineans. X. albilineans, the
causal organism of leaf scald disease in sugarcane, lacks T3SS and also does not produce xanthan gum [99]. However, it secretes
phytotoxin albicidin, which is a DNA gyrase inhibitor and is synthesized by nonribosomal peptide synthases [100]. It is synthesized by
NRPS-PKS (Non-Ribosomal Peptide Synthetase-Polyketide Synthase). Albicidin inhibits chloroplast formation, resulting in the whit­
ening of leaves. Nevertheless, it was not the predominant factor affecting the prevalence of X. albilineans since mutants lacking
albicidin retained their virulence [101]. Moreover, some Xanthomonas spp. like X. oryzae pv. oryzicola and X. axonopodis pv. citri
encode genes for the NRPS enzymes. These are much similar to phytotoxin syringomycin, which is encoded by syringomycin syn­
thetase genes [28]. There is ample opportunity for comprehensive research to clarify the correlation between the presence of NRPS
coding genes and toxin secretion in Xanthomonas species, as it remains unclear at present.
Xanthomonas LPS: LPS (lipopolysaccharides) are tripartite glycoconjugates in which the glycolipid part is known as lipid A. It
carries a core oligosaccharide and a polysaccharide (O-antigen), which constitutes the outer layer of the outer membrane. Addi­
tionally, rough-type LPS or lipooligosaccharides (LOS) are LPS lacking O-antigen components. LPS serves as a protective barrier
against antimicrobial substances that enhance the pathogen growth. It is also critical for the outer membrane’s structural properties.
LPSs also mediate adhesion and host recognition [102]. Defense-related responses are observed to be induced by LPS belonging to
different Xanthomonas spp. in plants [103,104]. The ability of lipid A in X. campestris to stimulate the innate immunity in the host plant
(Arabidopsis thaliana) is governed by a pattern of acylation and phosphorylation [105]. Lipid A is conserved throughout the
Gram-negative bacteria. Among core oligosaccharides, more variation is observed. Also, structures of the O-antigen of Xanthomonas
species are specific. However, to date, the structural details of only a few LPS have been determined.
Adhesins: Adhesion to the biotic surface is the initial step toward bacterial invasion in the host plant tissues. The adhesins
anchored in the bacterial outer membrane influence bacterial attachment. These adhesins are classified into fimbrial and nonfimbrial
adhesins. Non-fimbrial adhesins consist of XadA and XadB (homolog to adhesin YadA and the autotransporter YapH from Yersinia
spp.), FhaB (filamentous hemagglutinin-like proteins). These adhesins help in Xanthomonas attachment on leaves as well as entry
inside the host cell. Fimbrial adhesins consist of type IV pilus secretin PilQ. Gerlach and Hensel [29] reported that the predicted
structure of the periplasmic pilus of T2SS systems matches the filamentous proteinaceous structure of type IV pili [106,107]. These aids
in Xanthomonas replication and spread of infection in host tissue. Up to now, adhesins in X. oryzae pv. oryzae, X. fuscans ssp. fuscans and
X. axonopodis pv. citri, and have been demonstrated to play a role in bacterial virulence and in the attachment to biotic surfaces
[108–110]. It is well known that bacterial adhesin binds to host surface receptors, but the virulence function is intensively studied in
animal pathogenic bacteria, but it is lesser known in plant pathogenic bacteria.

1.12. Ice-nucleation activity

Discovery of ice-nucleation activity in bacteria was done in phytopathogen Pseudomonas syrigae during the 1970s. Following this,
researchers identified several other ice-nucleating bacteria belonging to families Pseudomonadaceae, Enterobacteriaceae, Xantho­
monadaceae, and Lysinibacillus [111]. Xanthomonas spp. too possess ice-nucleation activity, enabling them to initiate ice formation
even at relatively high temperatures. Freezing injury tends to increase the likelihood or severity of disease and the freezing/thawing
process aids in the passive entry of pathogens as reported for Pseudomonas syringae into cherry leaves [112]. Azad and Schaad [113]
found that wheat, barley, beans and corn plants inoculated with X. campestris pv. translucens exhibit increased susceptibility to frost
damage compared to control plants irrigated solely with water. Frost-exposed plants (− 3 ◦ C) showed faster lesion development than
those unexposed. Moreover, the number and size of lesions increased with longer incubation periods between inoculation and freezing.
X. campestris pv. translucens strains isolated from diseased plants exhibited ice nucleation activity within the temperature range of − 2 to
− 8 ◦ C. This ability of bacteria to induce ice formation is attributed to the specialized proteins that are anchored to the outer membrane
of the bacterial cell [114]. As a plant pathogen, Xanthomonas spp. exacerbate frost injury to plant tissues by elevating the nucleation
temperature of water, thereby gaining access to nutrients within the plant. The ice nucleation activity of X. campestris pv. translucens
shows both quantitative and qualitative variations. A significant correlation was observed between the level of ice nucleation activity
and the extent of frost injury in inoculated maize seedlings at − 5 ◦ C [114].

1.13. Control mechanisms

Xanthomonas causes complex diseases and needs intense strategic management. Most of the crop plants are susceptible towards
phytopathogens. Disease control mechanisms can include cultural practices, chemical control, biological control, and host genetic
resistance. The development of disease-free seeds, resistant varieties, strict phyto-sanitation, and bactericidal applications can play key
management against Xanthomonas infection. Chemical, biological, and biotechnological tools can be of substantial use in the control
mechanism which are discussed in the following sections.

7
R. Dey and R. Raghuwanshi Heliyon 10 (2024) e34275

1.14. Chemical control

Professor Millardet in 1882, developed Bordeaux mixture that constituted of copper, lime, and sulfate significantly controlled grape
downy mildew [115]. The recommended dose of metallic copper was 0.5–0.7 g/L under field conditions [116]. Gradually copper
became a key component in pesticides for plant disease control, which effectively induced and killed persistent cells of Xanthomonas
[117]. Later research by Martins et al. [118] unveiled that proline functions as a signaling molecule that triggers the reawakening of
X. citri. However, contrary to this, isoleucine had an inhibitory impact, obstructing the revival of persisters even when proline was
present. Along with extensive copper use, in the 1950’s the dominant bactericide used was streptomycin to treat Xanthomonas
infection, but gradually streptomycin-resistant Xanthomonas strains developed [119]. Further prolonged use of copper resulted in
copper-tolerant Xanthomonas strains in the late 1960s [120]. The addition of ethylene-bis-dithiocarbamates along with copper bac­
tericides came into practice to check bacterial spots caused by Xanthomonas [121]. It was also reported that compared to basic copper
sulfate, applying a mixture of copper and mancozeb effectively controlled Xanthomonas [121]. Mancozeb increases copper solubility
hence increasing its effectiveness in infection control. Presently, commercial copper bactericides constitute micron-size metallic
insoluble copper in the form of copper (II) hydroxide, copper (II) oxide, and copper (II) oxychloride [122]. Kocide 3000 is a
commercially available bactericide marketed by DuPont, Wilmington, DE, which contains micron-size metallic copper (II) hydroxide
(5 μM) [119].
Further control of Xanthomonas biofilm formation can be effectively prevented by the treatment of D-leucine and 3-indolyl
acetonitrile (IAN) [123]. Carvacrol, extracted from essential oils of Zataria multiflora, is reported to effectively control phytopatho­
gens It leads to a notable reduction in the intensity of bacterial leaf spot disease in tomatoes and enhances the effectiveness of copper
treatment against copper-resistant strains of X. euvesicatoria pv. perforans [124]. Acibenzolar-S-methyl (ASM) acts as a plant activator
that induces ISR in pepper and tomato. Weekly foliar treatment showed alleviation in bacterial leaf spot in pepper [125]. Another
substitute to copper treatment is the use of zinc as it also has antimicrobial properties and is a safer product. A zinc-based formulation,
Zinkicide, has been reported to reduce 38–42 % of the canker lesions in pineapple leaves under greenhouse conditions [126]. The use
of zinc lowers the metal accumulation in soil, unlike copper. Ternary solution (TSOL) constituted of zinc, salicylic acid, and hydrogen
peroxide, is reported to be effective against X. citri [32]. Presently, a hybrid nanoparticle developed from copper-zinc is reported to
exhibit antimicrobial activities against phytopathogens. Foliar treatment with benzoic acid has also led to an increase in resistance
against X. campestris pv. campestris, which causes black rot disease in cauliflower [127]. X. campestris pv. campestris inoculated Brassica
rapa when sprayed with 3,4-(methylenedioxy) cinnamic acid (MDCA), effectively enhanced disease emergence [128] as MDCA acts as
the inhibitor of 4-coumarate–CoA ligase which is responsible for the formation of hydroxycinnamic acids and flavonoids.
Although chemical controls provide quick prevention against plant diseases, in the long run, it results in bio-accumulation, soil
nutrient loss, pathogenic resistance and many more undesirable effects. Biological control as an alternative promising disease man­
agement strategy as it is environmentally sustainable and economically viable.

Fig. 3. Biocontrol mechanisms of Xanthomonas infection.

8
R. Dey and R. Raghuwanshi Heliyon 10 (2024) e34275

1.15. Biological control

Biological control is one of the eco-friendly approaches under the integrated disease management techniques. Microorganisms
residing in plant rhizospheres and phyllosphere have a great role in disease management. They secrete antimicrobial or SAR-inducing
compounds that help to induce host plant immunity. Plant-associated microorganisms are attractive biocontrol agents and need to be
explored critically. Before implementing biocontrol, especially in situations of severe disease spread, it is crucial to gain a deeper
understanding of the relationship between plants, pathogens, and environmental factors. In addition to the microbial applications, the
plant extracts can be utilized in the form of biofertilizers and biopesticides as well as gene products also serve as effective biological
control agents (BCA). Bioactive compounds derived from bacteria and fungi play a significant role in regulating the growth and
infection of Xanthomonas in the host plant. They achieve this in pathogen by generating reactive oxygen species, causing DNA damage,
enhancing cell permeability, or by interfering with the nutrients transport to Xanthomonas from host as depicted in Fig. 3.

1.16. Bacterial control

Bacteria inhabiting the rhizospheric surface of plants are proficient in repressing phytopathogens as well as growth promotion of
the host plants. These beneficial bacteria contribute to the natural defense mechanism of host plants from phytopathogens by syn­
thesizing organic compounds having antimicrobial activity (direct mechanism) or by inducing and improving plant immunity (indirect
mechanism). A lot of work has been done to identify such rhizospheric and endophytic bacteria and exploit them in combating
Xanthomonas infections in the host plants. These bacteria enhance competition, parasitism, and antibiosis, as well as produce bioactive
compounds that stimulate SAR (Systemic Acquired Resistance) and ISR (Induced Systemic Resistance). Pathogen cell lysis is mediated
by hydrolytic enzymes like cellulase, chitinase, glucanase, and protease produced by the beneficial bacteria. Antibiosis occurs through
the production of various substances such as antibiotics like oomycin A, 2,4-diacetyl phloroglucinol (DAPG), pyoluteorin, phenazine-1-
carboxamide, phenazine-1-carboxylic acid, pyrrolnitrin, as well as organic compounds, HCN, toxins, and hydrolytic enzymes like
β-xylosidase, chitinase, pectin methylesterase, and β-1,3-glucanase. These substances break down glycosidic linkages in the bacterial
cell wall, aiding the host plant in its defense against pathogens [129,130]. Some of the bacterial bioactive compounds responsible for
the inhibition of Xanthomonas are listed in Table 1.
Bacillus spp. is widely known for its biocontrol activity [146]. The treatment of B. subtilis TKS1-1 and B. amyloliquefaciens WG6-14
resulted in reduced colonization and diminished biofilm formation by X. citri subsp. citri which in turn, led to a decrease in the
development of citrus canker disease [147]. The production of lipopeptides including iturin, bacilysin, surfactin, bacillomycin,

Table 1
Bioactive compounds isolated from bacterial extracts showing inhibitory activity against Xanthomonas spp.
S. Bacteria Bioactive Compound Pathogen Disease References
No.

1. Pseudomonas aeruginosa Octadecanoic acid 2-oxo methyl ester X. campestris Bacterial blight [131]
VRKK1 in cowpea
2. Bacillus velezensis ELS2 E− 4-Bromo-3-methyl-1-(tetrahydro-2-pyranyloxy)-2-butene X. citri pv. Bacterial blight [132]
malvacearum in cotton
3. Bacillus amyloliquefaciens Difficidin; Bacilysin X. oryzae pv. Bacterial blight [133]
FZB42 oryzae in rice
4. Bacillus velezensis 25 Bacillomycin-D X. citri subsp. citri Citrus canker [134]
5. Bacillus albus CWTS 10 2-deoxystreptamine; Miserotoxin; Fumitremorgin C; X. oryzae Bacterial leaf [135]
Pipercide; Pipernonaline; Gingerone A; Deoxyvasicinone streak in rice
6. Pseudomonas fluorescens 2,4-diacetylphloroglucinol X. campestris pv Black rot disease [136]
campestris in cabbage
7. Bacillus velezensis HN-2 C15 surfactin A X. oryzae pv. Bacterial rice [137]
oryzae leaf blight
8. Pseudomonas entomophila Diketopiperazines X. citri subsp. citri Citrus canker [138]
9. Pseudomonas mosselii strain Pseudoiodinine X. oryzae pv. Bacterial leaf [139]
923 oryzicola streak
10. Consortia of Priestia 2-thio-5(N-methylaminomethyl) uridine, Diprotin A (Isoleucyl X. vesicatoria Leaf spot in [140]
megaterium T3 and Bacillus Prolyl Isoleucine), Plipastatin, Gramicidin, Kurstakins, tomato
cereus T4 Fusaricidins, Surfactin, Polymyxin
11. Paenibacillus peoriae strain Polymyxin A X. perforans T4 Bacterial spot of [141]
To99 tomato
12. Pseudomonas aeruginosa 1-hydroxy-phenazine, pyocyanin, pyochellin, rhamnolipids, 4- X. oryzae pv. Bacterial Leaf [142]
strain BRp3 hydroxy-2-alkylquinolines, 2,3,4-trihydroxy-2-alkylquinolines oryzae Blight in Basmati
and 1,2,3,4-tetrahydroxy-2-alkylquinolines rice
13. Pseudomonas orientalis X2-1P phenazine-1-carboxylic acid X. campestris pv. Black rot in [143]
campestris oilseed rape
14. Streptomyces sp. AN090126 2-propyl furan, 3-methyl-1-butanol, 2-methyl-1-butanol, and X. euvesicatoria Leaf spot in red [144]
dimethyl disulfide pepper
15. Paenibacillus yonginensis hydroperoxy-dihydroxy-octadecenoic acid, tricin hexoside, X. oryzae pv. Bacterial blight [145]
DCY84T corchorifatty acid F, pinellic acid, dihydroxy-octadecadienoic oryzae in rice
acid and dihydroxy-octadecenoic acid

9
R. Dey and R. Raghuwanshi Heliyon 10 (2024) e34275

zwittermicin, and fengycin has been reported in Bacillus spp. which plays an active role in antibiosis [148]. It is well documented that
foliar treatment of naturally isolated strains of Bacilli and Streptomyces significantly decreases Xanthomonas infections and symptom
reduction in various crops when tested in field conditions [149,150]. Pajcin et al. [151] reported that several BCA, including Lacto­
bacillus MK3, QM 9414, and P. aeruginosa 1128, exhibit substantial antibiosis capabilities against X. euvesicatoria pv. euvesicatoria,
which is responsible for bacterial leaf spots in chili and tomato plants. Up-to 65 % reduction in symptoms was observed in inoculated
B. velezensis IP22 pepper leaves in X. euvesicatoria pv. euvesicatoria infected plants. This was due to the lipopeptides from fengycin and
locillomycin families produced by B. velezensis IP22 [151]. Mishra and Arora [136] have evaluated fifty-four rhizo-bacterial strains of
B. campestris for their antagonism against X. campestris pv. campestris. Among them, two isolates, P. aeruginosa KA19 and B. thuringiensis
SE, exhibited inhibition halos wider than 11 mm. Both isolates applied separately and in combinations with various techniques (foliar
spray/seed soak/soil drench) performed well in controlling black rot disease. When used as a mixture, the results were significantly
better. Seed priming with rhizobacterial strains can stimulate phenylpropanoid and isoflavonoid pathways by accumulating iso­
flavonoid phytoalexins and phenolic compounds. This leads to the improvement of the biocontrol mechanism as well as promotes plant
growth [152]. A significant reduction of about 50 % is observed in bacterial leaf spots severity when humic acids along with
H. seropedicae suspension (108 cfu/ml) was used in tomato [153]. Le et al. [154] reported that treating chili with Paenibacillus elgii
significantly decreased the bacterial leaf spot by 67 %. Tomato plants when treated with hrpG mutant strain of X. euvesicatoria pv.
euvesicatoria 75-3S showed a decrease of 57 % compared to control plants for leaf spot disease under both greenhouse and field
conditions [155]. Some of the bacteria like B. subtilis strain QST 713 are commercially available under the trademark of Serenade®
OPTI (BAYER) which are used against Xanthomonas infection. The Bacillus subtilis group also secretes nitrogen-containing VOCs that
can be categorized based on their cyclization rate. Among the cyclic compounds, azoles are the predominant group exhibiting anti­
microbial properties [156]. Azole groups are also used in the agronomy sector to design effective fertilizers. 1,3,4-oxadiazole is a
potential oxadiazole-based agrochemical that induces the starvation mechanisms in Xanthomonas spp. by manipulating the type III
secretion system. It may indirectly suppress the inducible activation of host SWEET genes (Fig. 3), thus mitigating sucrose availability
during the infection process to the phytopathogenic bacteria, thereby being used as a biocontrol agent in agriculture [157].
Phenazine-1-carboxylic acid which is a bioactive compound of Pseudomonas is formulated under the trade name of shenqinmycin and
is globally used as a biopesticide to control Xanthomonas infection. Bacillus subtilis strain QST 713 (1.34 %) is formulated under the
trade name Cease®, which is used to treat bacterial spot caused by Xanthomonas spp and gray mold caused by Botrytis cinerea. Bacillus
subtilis strain IAB/BS03 (0.08 %) is formulated under the trade name AVIV®, which is used to treat bacterial spot caused by Xan­
thomonas spp. These are used commercially as a crop protectant worldwide.

1.17. Fungal control

The plant-fungus symbiosis exhibiting great diversity and ubiquity is an important interaction in ecological functioning. Studies
have shown that some symbiotic fungi which are crucial in plants adapting towards stress can be exploited as fungal bio-control agents
[158]. These may include Trichoderma spp., arbuscular mycorrhizas (AMF), yeasts, ectomycorrhizas, and fungal endophytes. They
show significant antagonism against fungal pathogens including Alternaria, Candida, Aspergillus, Fusarium, Pichia, Penicillium, Pythium,
Verticillium and Talaromyces, which cause widespread diseases in host plants. Several mechanisms are applied by these fungal
biocontrol agents against different pathogens. T. harzianum exhibits parasitism against S. sclerotiorum by degrading the hyphal growth
of the pathogens [159]. Likewise, P. lilacinum performs antibiosis, by secreting various extracellular enzymes to inhibit white mold
[160]. P. lilacinum secrete extracellular enzymes which digests the eggs and parasitize the egg-laying female nematode, Meloidogyne
enterolobii [161]. Trichoderma spp. are filamentous fungi found in the soil, renowned for their effectiveness in a wide range of ap­
plications that promote plant health and benefits [162]. These strains utilize a sophisticated mechanism for controlling pathogens like
mycoparasitism [163,164], competing for physical space for root colonization, preventing the proliferation of phytopathogens by cell
lytic enzymes secretion, antagonistic secondary compounds production, plant defense responses stimulation, enhancing plant growth,
and increasing plant resilience against both biotic and abiotic stress factors [165].
T. harzianum and T. strigosum are reported to protect tomato against X. euvesicatoria causing leaf spot disease effectively [166].
Trichoderma produces terpenoid phytoalexins that is responsible for the antagonistic behavior towards the phytopathogens [167].
Trichoderma longibrachiatum, is efficient against X. campestris pv. campestris and X. arboricola, by secreting trichogin GA IV peptaibol,
which possesses antimicrobial activity [168]. Zhang et al. [169] reported in their study the antibacterial effects of Trichokonins A and
peptaibols produced by Trichoderma longibrachiatum SMF2 to control the infection of Xanthomonas oryzae pv. oryzae causing bacterial
leaf blight on rice (Fig. 3). Consortia of T. asperellum strain ICC 012 and T. gamsii strain ICC 080 (2 % + 0.08 %) formulated under the
trade name Bio-Tam®, is used to treat wilt caused by Xanthomonas spp. This is also used commercially against Phytophthora,
Rhizoctonia, Pythium and Verticillium.
Besides Trichoderma spp., Arbuscular mycorrhizal fungi (AMF) are the most prevalent, holding the largest biomass and having an
antagonistic effect on phytopathogens, making it a promising beneficial fungus [170]. A notable increase in defense-related tran­
scription factors in Medicago truncatula when infected with X. campestris indicates that AMF symbiosis is responsible for the induction
of ISR response [171]. In X. campestris pv. oryzae infected rice plants causing bacterial blight it is observed that AMF colonization at the
seedling stage is lowest and enhances significantly with maturity imparting a high degree of resistance towards the pathogen during
maturity re-establishing the fact that percent AMF colonization plays a direct role in biotic stress management [172].
Yeast is also an effective biocontrol agent thriving with minimal culture requirements and posing limited biosafety concerns.
A. pullulans, C. albidus, C. oleophila, S. cerevisiae, and M. fructicola are broadly used as BCA. The mechanisms used by them to combat
pathogens are competition, secretion of volatiles, mycoparasitism, enzymes and toxins production, and stimulation of phytoimmunity

10
R. Dey and R. Raghuwanshi Heliyon 10 (2024) e34275

in host. Black rot of crucifers caused by X. campestris pv. campestris have been reported to be biologically controlled using epiphytic
yeast isolated from cabbage leaves with 72–79 % reduction in disease severity [173]. It is also reported that edible strains of
S. cerevisiae NJ-1 isolated from the pomace of Ficus carica secrete 3-methyl-1-butanol, a volatile organic compound that inhibits the
production of aflatoxin [174]. Yeast can be used as an active biocontrol agent but the paucity of exploration and exploitation limits its
full utilization.

1.18. Phage-based biocontrol

An integrative disease management system has picked up interest in applications of phage-mediated Xanthomonas biocontrol in
agricultural fields. Bacteriophages are viruses that infect and replicate within bacterial cells. They are recognized as the most abundant
biological entities on the planet and are extensively distributed in the environment. The replication cycle consists of temperate and
lytic phases. The lytic pathway is a pivotal element in biocontrol activity. In this pathway, bacteriophages adhere to the bacterial
surface, introduce their genetic material, and replicate within the pathogen cell. This ultimately leads to the multiplication of phage,
which cause lysis and the demise of the pathogen [175]. In the 19th century, the biocontrol activity of phage drew attention of re­
searchers, when a filtrate derived from decomposing cabbage was used, it effectively halted the spread of cabbage-rot disease induced
by X. campestris pv. campestris [176]. Later, Civerolo and Kiel [177] reported a similar experiment to check bacterial spot caused by
X. campestris pv. pruni in peach. It is now widely progressed in field trials as well as explicit research is carried out on phage biocontrol
mechanisms. Bacteriophage alter the mechanism of DNA methylation, phosphorylation, and transcription in the host cell. In the case of
Phage Xp12 infecting X. oryzae pv. oryzae, this infection prompts the production of 5-methylcytosine, replacing all cytosine residues
only in the DNA of Xp12 resulting in change of the physical and chemical properties of DNA, like low buoyant density and high melting
temperature ultimately causing death of the cell. Phage biocontrol is widely used against bacterial spot of tomato caused by
X. campestris pv. vesicatoria [178], geranium bacterial blight caused by X. campestris pv. pelargonii [179], X. axonopodis pv. allii causing
leaf blight of onion [180], X. axonopodis pv. citri and X. axonopodis pv. citrumelo causing citrus canker and citrus bacterial spot [181],
X. axonopodis pv. citri causing Asiatic citrus canker [182], X. oryzae pv. oryzae, causal organism for bacterial leaf blight of rice [183]
X. axonopodis pv. allii, causal organism for bacterial leaf blight of Welsh onions [184]. On application of lytic phage cocktails like
Xp3-A and Xp3-I on peach seedlings, they resulted in a 51–54 % reduction in symptom emergence of bacterial spot caused by X. pruni
under environmental conditions [185]. To date, two commercial Xanthomonas phage products are available, manufactured by Agri­
phage™ which are reported to successfully regulate citrus canker disease and tomato and pepper spot disease [186]. Detailed research
can result in more commercial products leading to the exploitation of the complete potential of phage as a BCA.

2. Conclusion

The efforts invested in understanding Xanthomonas ecology, epidemiology, and phytopathology are ultimately aimed at controlling
and suppressing plant diseases caused by these bacteria. While there has been notable advancement in comprehending the functions of
Xanthomonas secretion systems, primarily focusing on T2SS and T3SS substrates, functionally characterizing the protein effectors
remains limited in scope. Consequently, ongoing efforts directed at identifying and characterizing effectors produced by the other
secretion systems, hold great promise as avenues for unveiling new and unique mechanisms and functions utilized by Xanthomonas to
manipulate target cells and flourish in their specific environmental niches. This knowledge has started to find practical applications in
numerous strategies, including chemical control of pathogens, utilizing biocontrol agents, and developing resistant plant varieties.
Furthermore, the increasing recognition of the plant microbiome’s role in enhancing plant growth promotion and phyto-immunity
raises the possibility of harnessing endophytic and epiphytic microbial communities as potential biocontrol agents, which could be
integrated into pest management systems. However, it is important to mention that research on the influence of the microbiome on
Xanthomonas colonization and strategical disease management is still in its nascent stages and needs to be emphasized minutely.

Data availability statement

The authors declare that no data associated with our study has been deposited into a publicly available repository since no data was
used for the research described in the article.

Ethics declarations

Review and/or approval by an ethics committee was not needed for this study because it donot include any human or animal
participation.

CRediT authorship contribution statement

Riddha Dey: Writing – original draft, Formal analysis. Richa Raghuwanshi: Writing – review & editing, Supervision, Resources,
Conceptualization.

11
R. Dey and R. Raghuwanshi Heliyon 10 (2024) e34275

Declaration of competing interest

The authors declare the following financial interests/personal relationships which may be considered as potential competing in­
terests:Richa Raghuwanshi reports administrative support and equipment, drugs, or supplies were provided by Banaras Hindu Uni­
versity. If there are other authors, they declare that they have no known competing financial interests or personal relationships that
could have appeared to influence the work reported in this paper.

Acknowledgements

Authors are thankful to the Banaras Hindu University, IoE, Faculty grant scheme for providing the necessary facilities and support.

References

[1] A.C. Oliveira, R.M. Ferreira, M.I.T. Ferro, J.A. Ferro, M. Chandler, A.M. Varani, Transposons and pathogenicity in Xanthomonas: acquisition of murein lytic
transglycosylases by TnXax1 enhances Xanthomonas citri subsp. citri 306 virulence and fitness, PeerJ 6 (2018) e6111, https://doi.org/10.7717/peerj.6111.
[2] S. Timilsina, J.A. Pereira-Martin, G.V. Minsavage, F. Iruegas-Bocardo, P. Abrahamian, N. Potnis, B. Kolaczkowski, G.E. Vallad, E.M. Goss, J.B. Jones, Multiple
recombination events drive the current genetic structure of Xanthomonas perforans in Florida, Front. Microbiol. 10 (2019) 448, https://doi.org/10.3389/
fmicb.2019.00448.
[3] E.A. Newberry, R. Bhandari, G.V. Minsavage, S. Timilsina, M.O. Jibrin, J. Kemble, E.J. Sikora, J.B. Jones, N. Potnis, Independent evolution with the gene flux
originating from multiple Xanthomonas species explains genomic heterogeneity in Xanthomonas perforans, Appl. Environ. Microbiol. 85 (20) (2019) e00885,
https://doi.org/10.1128/AEM.00885-19, 19.
[4] M.P. Starr, W.L. Stephens, Pigmentation and taxonomy of the genus Xanthomonas, J. Bacteriol. 87 (2) (1964) 293–302, https://doi.org/10.1128/jb.87.2.293-
302.1964.
[5] C.L. Jenkins, M.P. Starr, The brominated aryl-polyene (xanthomonadin) pigments of Xanthomonas juglandis protect against photobiological damage, Curr.
Microbiol. 7 (1982) 323–326, https://doi.org/10.1007/BF01566872.
[6] L. Rajagopal, C.S. Sundari, D. Balasubramanian, R.V. Sonti, The bacterial pigment xanthomonadin offers protection against photodamage, FEBS Lett. 415 (2)
(1997) 125–128, https://doi.org/10.1016/S0014-5793(97)01109-5.
[7] D.W. Dye, R.A. Lelliott, Genus Xanthomonas, in: R.E. Buchanan (Ed.), Bergey’s Manual of Determinative Bacteriology, 1974.
[8] J.L.W. Rademaker, F.J. Louws, M.H. Schultz, U. Rossbach, L. Vauterin, J. Swings, F.J. De Bruijn, A comprehensive species to strain taxonomic framework for
Xanthomonas, Phytopathology 95 (9) (2005) 1098–1111, https://doi.org/10.1094/PHYTO-95-1098.
[9] M.A. Jacques, M. Arlat, A. Boulanger, T. Boureau, S. Carrère, S. Cesbron, N.W. Chen, S. Cociancich, A. Darrasse, N. Denancé, M. Fischer-Le Saux, Using
ecology, physiology, and genomics to understand host specificity in Xanthomonas, Annu. Rev. Phytopathol. 54 (2016) 163–187, https://doi.org/10.1146/
annurev-phyto-080615-100147.
[10] A.R. Schwartz, N. Potnis, S. Timilsina, M. Wilson, J. Patané, J. Martins Jr, G.V. Minsavage, D. Dahlbeck, A. Akhunova, N. Almeida, G.E. Vallad, Phylogenomics
of Xanthomonas field strains infecting pepper and tomato reveals diversity in effector repertoires and identifies determinants of host specificity, Front.
Microbiol. 6 (2015) 535, https://doi.org/10.3389/fmicb.2015.00535.
[11] R. Gade, R. Lad, Biological management of major citrus diseases in Central India-a review, Int. J. Curr. Microbiol. App. Sci 6 (2018) 296–308.
[12] A.M. Alvarez, Black rot of crucifers, in: Mechanisms of Resistance to Plant Diseases, Dordrecht: Springer, Netherlands, 2000, pp. 21–52.
[13] W. Qian, Y. Jia, S.X. Ren, Y.Q. He, J.X. Feng, L.F. Lu, Q. Sun, G. Ying, D.J. Tang, H. Tang, W. Wu, Comparative and functional genomic analyses of the
pathogenicity of phytopathogen Xanthomonas campestris pv. campestris, Genome Res. 15 (6) (2005) 757–767, https://doi.org/10.1101/gr.3378705.
[14] J.B. Jones, G.H. Lacy, H. Bouzar, R.E. Stall, N.W. Schaad, Reclassification of the xanthomonads associated with bacterial spot disease of tomato and pepper,
Syst. Appl. Microbiol. 27 (6) (2004) 755–762, https://doi.org/10.1078/0723202042369884.
[15] N. Potnis, S. Timilsina, A. Strayer, D. Shantharaj, J.D. Barak, M.L. Paret, G.E. Vallad, J.B. Jones, Bacterial spot of tomato and pepper: diverse Xanthomonas
species with a wide variety of virulence factors posing a worldwide challenge, Mol. Plant Pathol. 16 (9) (2015) 907–920, https://doi.org/10.1111/mpp.12244.
[16] D. Yirgou, J.F. Bradbury, Bacterial wilt of enset (Ensete ventricosum) incited by Xanthomonas musacearum sp. n, Phytopathology 58 (1968) 111–112.
[17] D. Yirgou, J.F. Bradbury, A note on wilt of banana caused by the enset wilt organism Xanthomonas musacearum, East Afr. Agric. For. J. 40 (1) (1974) 111–114,
https://doi.org/10.1080/00128325.1974.11662720.
[18] B.A. Carter, R. Reeder, S.R. Mgenzi, M. Mwangi, V. Aritua, J. Smith, F.D. Beed, Z.M. Kinyua, V. Nakato, S. Miller, J.N. Mbaka, Identification of Xanthomonas
vasicola (formerly X. campestris pv. musacearum), causative organism of banana Xanthomonas wilt, in: Tanzania, Kenya and Burundi, 2010, https://doi.org/
10.1111/j.1365-3059.2009.02124.x.
[19] S.R.B. Mgenzi, J. Muchunguzi, T. Mutagwaba, F. Mkondo, R. Mohamed, V. Aritua, An outbreak of banana bacterial wilt disease in Muleba district, Kagera
region, Tanzania, African Crops Net (2006). http://www.africancrops.net/news/april06.
[20] W. Tushemereirwe, A. Kangire, J. Smith, F. Ssekiwoko, M. Nakyanzi, D. Kataama, C. Musiitwa, R. Karyaija, An outbreak of bacterial wilt on banana in Uganda,
InfoMusa 12 (2) (2003) 6–8.
[21] V. Ndungo, K. Bakelana, S. Eden-Green, G. Blomme, An outbreak of banana Xanthomonas wilt (Xanthomonas campestris pv. musacearum) in the Democratic
Republic of Congo, InfoMusa 13 (2) (2004) 43–44.
[22] S.E. Lindow, M.T. Brandl, Microbiology of the phyllosphere, Appl. Environ. Microbiol. 69 (4) (2003) 1875–1883, https://doi.org/10.1128/AEM.69.4.1875-
1883.2003.
[23] S. Zarei, S.M. Taghavi, H. Hamzehzarghani, E. Osdaghi, J.R. Lamichhane, Epiphytic growth of Xanthomonas arboricola and Xanthomonas citri on non-host
plants, Plant Pathol. 67 (3) (2018) 660–670, https://doi.org/10.1111/ppa.12769.
[24] Z. Kayaaslan, S. Belgüzar, Y. Yanar, M. Mirik, Epidemiology of Xanthomonas euvesicatoria in Tokat Province, Agronomy 13 (3) (2023) 677, https://doi.org/
10.3390/agronomy13030677.
[25] X. Yu, S.P. Lund, R.A. Scott, J.W. Greenwald, A.H. Records, D. Nettleton, S.E. Lindow, D.C. Gross, G.A. Beattie, Transcriptional responses of Pseudomonas
syringae to growth in epiphytic versus apoplastic leaf sites, Proc. Natl. Acad. Sci. 110 (5) (2013) E425–E434, https://doi.org/10.1073/pnas.1221892110.
[26] R.P. Ryan, Cyclic di-GMP signalling and the regulation of bacterial virulence, Microbiol. 159 (7) (2013) 1286, https://doi.org/10.1099/mic.0.068189-0.
[27] L. Crossman, J.M. Dow, Biofilm formation and dispersal in Xanthomonas campestris, Microb. Infect. 6 (6) (2004) 623–629, https://doi.org/10.1016/j.
micinf.2004.01.013.
[28] R.P. Ryan, F.J. Vorhölter, N. Potnis, J.B. Jones, M.A. Van Sluys, A.J. Bogdanove, J.M. Dow, Pathogenomics of Xanthomonas: understanding bacterium–plant
interactions, Nat. Rev. Microbiol. 9 (5) (2011) 344–355, https://doi.org/10.1038/nrmicro2558.
[29] R.G. Gerlach, M. Hensel, Salmonella pathogenicity islands in host specificity, host pathogen-interactions and antibiotics resistance of Salmonella enterica, Berl.
Münchener Tierärztliche Wochenschr. 120 (7/8) (2007) 317, https://doi.org/10.2376/0005-9366-120-317.
[30] O. Spitz, I.N. Erenburg, T. Beer, K. Kanonenberg, I.B. Holland, L. Schmitt, Type I secretion systems—one mechanism for all? Microbiol. Spectr. 7 (2) (2019)
10–1128, https://doi.org/10.1128/microbiolspec.psib-0003-2018.
[31] C.E. Alvarez-Martinez, G.G. Sgro, G.G. Araujo, M.R. Paiva, B.Y. Matsuyama, C.R. Guzzo, M.O. Andrade, C.S. Farah, Secrete or perish: the role of secretion
systems in Xanthomonas biology, Comput. Struct. Biotechnol. J. 19 (2021) 279–302, https://doi.org/10.1016/j.csbj.2020.12.020.

12
R. Dey and R. Raghuwanshi Heliyon 10 (2024) e34275

[32] Z. Liu, A.M. Peasley, J. Yang, Y. Li, G. Guan, J. Luo, H. Yin, K.A. Brayton, The Anaplasma ovis genome reveals a high proportion of pseudogenes, BMC Genom.
20 (1) (2019) 1–14, https://doi.org/10.1186/s12864-018-5374-6.
[33] A.R. da Silva, J.A. Ferro, F. Reinach, C.S. Farah, L.R. Furlan, R.B. Quaggio, C.B. Monteiro-Vitorello, M.V. Sluys, N.A. Almeida, L.M.C. Alves, A.M. Do Amaral,
Comparison of the genomes of two Xanthomonas pathogens with differing host specificities, Nature 417 (6887) (2002) 459–463, https://doi.org/10.1038/
417459a.
[34] S. Burdman, Y. Shen, S.W. Lee, Q. Xue, P. Ronald, RaxH/RaxR: a two-component regulatory system in Xanthomonas oryzae pv. oryzae required for AvrXa21
activity, Mol. Plant Microbe Interact. 17 (6) (2004) 602–612, https://doi.org/10.1094/MPMI.2004.17.6.602.
[35] R.N. Pruitt, B. Schwessinger, A. Joe, N. Thomas, F. Liu, M. Albert, M.R. Robinson, L.J.G. Chan, D.D. Luu, H. Chen, O. Bahar, The rice immune receptor XA21
recognizes a tyrosine-sulfated protein from a Gram-negative bacterium, Sci. Adv. 1 (6) (2015) e1500245, https://doi.org/10.1126/sciadv.1500245.
[36] S. Guo, T.D. Vance, C.A. Stevens, I. Voets, P.L. Davies, RTX adhesins are key bacterial surface megaproteins in the formation of biofilms, Trends Microbiol. 27
(5) (2019) 453–467, https://doi.org/10.1016/j.tim.2018.12.003.
[37] N.P. Fonseca, J.S. Patané, A.M. Varani, É.B. Felestrino, W.L. Caneschi, A.B. Sanchez, I.F. Cordeiro, C.G.D.C. Lemes, R.D.A.B. Assis, C.C.M. Garcia, J. Belasque
Jr, Analyses of seven new genomes of Xanthomonas citri pv. aurantifolii strains, causative agents of citrus canker B and C, show a reduced repertoire of
pathogenicity-related genes, Front. Microbiol. (2019) 2361, https://doi.org/10.3389/fmicb.2019.02361.
[38] S.M.A. Shah, M. Khojasteh, Q. Wang, F. Haq, X. Xu, Y. Li, L. Zou, E. Osdaghi, G. Chen, Comparative transcriptome analysis of wheat cultivars in response to
Xanthomonas translucens pv. cerealis and its T2SS, T3SS and TALEs deficient strains, Phytopathology (ja) (2023), https://doi.org/10.1094/PHYTO-02-23-0049-
SA.
[39] T.L. Johnson, J. Abendroth, W.G. Hol, M. Sandkvist, Type II secretion: from structure to function, FEMS Microbiol. Lett. 255 (2) (2006) 175–186, https://doi.
org/10.1111/j.1574-6968.2006.00102.x.
[40] H. Lu, P. Patil, M.A. Van Sluys, F.F. White, R.P. Ryan, J.M. Dow, P. Rabinowicz, S.L. Salzberg, J.E. Leach, R. Sonti, V. Brendel, Acquisition and evolution of
plant pathogenesis–associated gene clusters and candidate determinants of tissue-specificity in Xanthomonas, PLoS One 3 (11) (2008) e3828, https://doi.org/
10.1371/journal.pone.0003828.
[41] L.M. Moreira, R.F. de Souza, N.F. Almeida Jr, J.C. Setubal, J.C.F. Oliveira, L.R. Furlan, J.A. Ferro, A.C. da Silva, Comparative genomics analyses of citrus-
associated bacteria, Annu. Rev. Phytopathol. 42 (2004) 163–184, https://doi.org/10.1146/annurev.phyto.42.040803.140310.
[42] N. Potnis, K. Krasileva, V. Chow, N.F. Almeida, P.B. Patil, R.P. Ryan, M. Sharlach, F. Behlau, J.M. Dow, M.T. Momol, F.F. White, Comparative genomics reveals
diversity among xanthomonads infecting tomato and pepper, BMC Genom. 12 (2011) 1–23, https://doi.org/10.1186/1471-2164-12-146.
[43] N.P. Cianciotto, R.C. White, Expanding role of type II secretion in bacterial pathogenesis and beyond, Infect. Immun. 85 (5) (2017) 10–1128, https://doi.org/
10.1128/iai.00014-17.
[44] S. Pfeilmeier, A. Werz, M. Ote, M. Bortfeld-Miller, P. Kirner, A. Keppler, L. Hemmerle, C.G. Gaebelein, C.M. Pestalozzi, J.A. Vorholt, Dysbiosis of a leaf
microbiome is caused by enzyme secretion of opportunistic Xanthomonas strains, bioRxiv (2023) 5, https://doi.org/10.1101/2023.05.09.539948.
[45] Y. Zhang, D. Teper, J. Xu, N. Wang, Stringent response regulators (p) ppGpp and DksA positively regulate virulence and host adaptation of Xanthomonas citri,
Mol. Plant Pathol. 20 (11) (2019) 1550–1565, https://doi.org/10.1111/mpp.12865.
[46] R. Szczesny, D. Büttner, L. Escolar, S. Schulze, A. Seiferth, U. Bonas, Suppression of the AvrBs1-specific hypersensitive response by the YopJ effector homolog
AvrBsT from Xanthomonas depends on a SNF1-related kinase, New Phytol. 187 (4) (2010) 1058–1074, https://doi.org/10.1111/j.1469-8137.2010.03346.x.
[47] P. Ghosh, Process of protein transport by the type III secretion system, Microbiol. Mol. Biol. Rev. 68 (4) (2004) 771–795, https://doi.org/10.1128/
mmbr.68.4.771-795.2004.
[48] D. Büttner, U. Bonas, Getting across—bacterial type III effector proteins on their way to the plant cell, EMBO J. 21 (20) (2002) 5313–5322, https://doi.org/
10.1093/emboj/cdf536.
[49] D. Büttner, D. Nennstiel, B. Klüsener, U. Bonas, Functional analysis of HrpF, a putative type III translocon protein from Xanthomonas campestris pv. vesicatoria,
J. Bacteriol. 184 (9) (2002) 2389–2398, https://doi.org/10.1128/jb.184.9.2389-2398.2002.
[50] J.G. Kim, B.K. Park, C.H. Yoo, E. Jeon, J. Oh, I. Hwang, Characterization of the Xanthomonas axonopodis pv. glycines Hrp pathogenicity island, J. Bacteriol. 185
(10) (2003) 3155–3166, https://doi.org/10.1128/jb.185.10.3155-3166.2003.
[51] A. Sugio, B. Yang, F.F. White, Characterization of the hrpF pathogenicity peninsula of Xanthomonas oryzae pv. oryzae, Mol. Plant Microbe Interact. 18 (6)
(2005) 546–554, https://doi.org/10.1094/MPMI-18-0546.
[52] Y. Yan, Y. Wang, X. Yang, Y. Fang, G. Cheng, L. Zou, G. Chen, The MinCDE cell Division system Participates in the regulation of type III secretion system (T3SS)
genes, bacterial virulence, and Motility in Xanthomonas oryzae pv. oryzae, Microorganisms 10 (8) (2022) 1549, https://doi.org/10.3390/
microorganisms10081549.
[53] D. Merda, M. Briand, E. Bosis, C. Rousseau, P. Portier, M. Barret, M.A. Jacques, M. Fischer-Le Saux, Ancestral acquisitions, gene flow and multiple evolutionary
trajectories of the type three secretion system and effectors in Xanthomonas plant pathogens, Mol. Ecol. 26 (21) (2017) 5939–5952, https://doi.org/10.1111/
mec.14343.
[54] J. Boch, H. Scholze, S. Schornack, A. Landgraf, S. Hahn, S. Kay, T. Lahaye, A. Nickstadt, U. Bonas, Breaking the code of DNA binding specificity of TAL-type III
effectors, Science 326 (5959) (2009) 1509–1512, https://doi.org/10.1126/science.1178811.
[55] M.J. Moscou, A.J. Bogdanove, A simple cipher governs DNA recognition by TAL effectors, Science 326 (5959) (2009) 1501, https://doi.org/10.1126/
science.1178817, 1501.
[56] J. Boch, U. Bonas, T. Lahaye, TAL effectors–pathogen strategies and plant resistance engineering, New Phytol. 204 (4) (2014) 823–832, https://doi.org/
10.1111/nph.13015.
[57] F. Thieme, R. Koebnik, T. Bekel, C. Berger, J. Boch, D. Büttner, C. Caldana, L. Gaigalat, A. Goesmann, S. Kay, O. Kirchner, Insights into genome plasticity and
pathogenicity of the plant pathogenic bacterium Xanthomonas campestris pv. vesicatoria revealed by the complete genome sequence, J. Bacteriol. 187 (21)
(2005) 7254–7266, https://doi.org/10.1128/jb.187.21.7254-7266.2005.
[58] A.J. Bogdanove, R. Koebnik, H. Lu, A. Furutani, S.V. Angiuoli, P.B. Patil, M.A. Van Sluys, R.P. Ryan, D.F. Meyer, S.W. Han, G. Aparna, Two new complete
genome sequences offer insight into host and tissue specificity of plant pathogenic Xanthomonas spp, J. Bacteriol. 193 (19) (2011) 5450–5464, https://doi.org/
10.1128/jb.05262-11.
[59] S. Sapkota, M. Mergoum, Z. Liu, The translucens group of Xanthomonas translucens: Complicated and important pathogens causing bacterial leaf streak on
cereals, Mol. Plant Pathol. 21 (3) (2020) 291–302, https://doi.org/10.1111/mpp.12909.
[60] T. Ogawa, L. Lin, R.E. Tabien, G.S. Kush, A new recessive gene for resistance to bacterial blight of rice, Rice Genet. NewsL. 4 (98) (1987) 100.
[61] B. Yang, A. Sugio, F.F. White, Os8N3 is a host disease susceptibility gene for bacterial blight of rice, Proc. Natl. Acad. Sci. U.S.A. 103 (2006) 10503–10508,
https://doi.org/10.1073/pnas.0604088103.
[62] T. Yuan, X. Li, J. Xiao, S. Wang, Characterization of Xanthomonas oryzae-responsive cis-acting element in the promoter of rice race-specific susceptibility gene
Xa13, Mol. Plant 4 (2011) 300–309, https://doi.org/10.1093/mp/ssq076.
[63] J. Streubel, C. Pesce, M. Hutin, R. Koebnik, J. Boch, B. Szurek, Five phylogenetically close rice SWEET genes confer TAL effector-mediated susceptibility to
Xanthomonas oryzae pv. oryzae, New Phytol. 200 (3) (2013) 808–819, https://doi.org/10.1111/nph.12411.
[64] S. Mücke, M. Reschke, A. Erkes, C.A. Schwietzer, S. Becker, J. Streubel, R.D. Morgan, G.G. Wilson, J. Grau, J. Boch, Transcriptional reprogramming of rice cells
by Xanthomonas oryzae TALEs, Front. Plant Sci. 10 (2019) 162, https://doi.org/10.3389/fpls.2019.00162.
[65] J. Tian, G. Xu, M. Yuan, Precise editing enables crop broad-spectrum resistance, Mol. Plant 12 (12) (2019) 1542–1544.
[66] K.L. Cox, F. Meng, K.E. Wilkins, F. Li, P. Wang, N.J. Booher, S.C.D. Carpenter, L.-Q. Chen, H. Zheng, X. Gao, Y. Zheng, Z. Fei, J.Z. Yu, T. Isakeit, T. Wheeler, W.
B. Frommer, P. He, A.J. Bogdanove, L. Shan, TAL effector driven induction of a SWEET gene confers susceptibility to bacterial blight of cotton, Nat. Commun. 8
(2017) 15588, https://doi.org/10.1038/ncomms15588.

13
R. Dey and R. Raghuwanshi Heliyon 10 (2024) e34275

[67] M. Cohn, R.S. Bart, M. Shybut, D. Dahlbeck, M. Gomez, R. Morbitzer, B.H. Hou, W.B. Frommer, T. Lahaye, B.J. Staskawicz, Xanthomonas axonopodis virulence
is promoted by a transcription activator-like effector-mediated induction of a SWEET sugar transporter in cassava, Mol. Plant Microbe Interact. 27 (2014)
1186–1198, https://doi.org/10.1094/MPMI-06-14-0161-R.
[68] Y. Hu, J. Zhang, H. Jia, D. Sosso, T. Li, W.B. Frommer, B. Yang, F.F. White, N. Wang, J.B. Jones, Lateral organ boundaries1 is a disease susceptibility gene for
citrus bacterial canker disease, Proc. Natl. Acad. Sci. U.S.A. 111 (2014) E521–E529, https://doi.org/10.1073/pnas.1313271111.
[69] D. Büttner, U. Bonas, Regulation and secretion of Xanthomonas virulence factors, FEMS Microbiol. Rev. 34 (2) (2010) 107–133, https://doi.org/10.1111/
j.1574-6976.2009.00192.x.
[70] F.F. White, N. Potnis, J.B. Jones, R. Koebnik, The type III effectors of Xanthomonas, Mol. Plant Pathol. 10 (6) (2009) 749–766, https://doi.org/10.1111/j.1364-
3703.2009.00590.x.
[71] J. Qin, X. Zhou, L. Sun, K. Wang, F. Yang, H. Liao, W. Rong, J. Yin, H. Chen, X. Chen, J. Zhang, The Xanthomonas effector XopK harbours E3 ubiquitin-ligase
activity that is required for virulence, New Phytol. 220 (1) (2018) 219–231, https://doi.org/10.1111/nph.15287.
[72] S. Deb, P. Ghosh, H.K. Patel, R.V. Sonti, Interaction of the Xanthomonas effectors XopQ and XopX results in induction of rice immune responses, Plant J. 104 (2)
(2020) 332–350, https://doi.org/10.1111/tpj.14924.
[73] D. Sinha, M.K. Gupta, H.K. Patel, A. Ranjan, R.V. Sonti, Cell wall degrading enzyme induced rice innate immune responses are suppressed by the type 3
secretion system effectors XopN, XopQ, XopX and XopZ of Xanthomonas oryzae pv. oryzae, PLoS One 8 (9) (2013) e75867, https://doi.org/10.1371/journal.
pone.0075867.
[74] A. Hotson, R. Chosed, H. Shu, K. Orth, M.B. Mudgett, Xanthomonas type III effector XopD targets SUMO-conjugated proteins in planta, Mol. Microbiol. 50 (2)
(2003) 377–389, https://doi.org/10.1046/j.1365-2958.2003.03730.x.
[75] R. Chosed, D.R. Tomchick, C.A. Brautigam, S. Mukherjee, V.S. Negi, M. Machius, K. Orth, Structural analysis of Xanthomonas XopD provides insights into
substrate specificity of ubiquitin-like protein proteases, J. Biol. Chem. 282 (9) (2007) 6773–6782, https://doi.org/10.1074/jbc.M608730200.
[76] J.G. Kim, K.W. Taylor, A. Hotson, M. Keegan, E.A. Schmelz, M.B. Mudgett, XopD SUMO protease affects host transcription, promotes pathogen growth, and
delays symptom development in Xanthomonas-infected tomato leaves, Plant Cell 20 (7) (2008) 1915–1929, https://doi.org/10.1105/tpc.108.058529.
[77] D. Teper, A.M. Girija, E. Bosis, G. Popov, A. Savidor, G. Sessa, The Xanthomonas euvesicatoria type III effector XopAU is an active protein kinase that
manipulates plant MAP kinase signaling, PLoS Pathog. 14 (1) (2018) e1006880, https://doi.org/10.1105/tpc.108.058529.
[78] D. Teper, D. Burstein, D. Salomon, M. Gershovitz, T. Pupko, G. Sessa, Identification of novel Xanthomonas euvesicatoria type III effector proteins by a machine-
learning approach, Mol. Plant Pathol. 17 (3) (2016) 398–411, https://doi.org/10.1111/mpp.12288.
[79] C.E. Alvarez-Martinez, P.J. Christie, Biological diversity of prokaryotic type IV secretion systems, Microbiol. Mol. Biol. Rev. 73 (4) (2009) 775–808, https://
doi.org/10.1128/mmbr.00023-09.
[80] G.G. Sgro, G.U. Oka, D.P. Souza, W. Cenens, E. Bayer-Santos, B.Y. Matsuyama, N.F. Bueno, T.R. Dos Santos, C.E. Alvarez-Martinez, R.K. Salinas, C.S. Farah,
Bacteria-killing type IV secretion systems, Front. Microbiol. 10 (2019) 1078, https://doi.org/10.3389/fmicb.2019.01078.
[81] B.K. Scholz-Schroeder, J.D. Soule, D.C. Gross, The sypA, sypB, and sypC synthetase genes encode twenty-two modules involved in the nonribosomal peptide
synthesis of syringopeptin by Pseudomonas syringae pv. syringae B301D, Mol. Plant Microbe Interact. 16 (4) (2003) 271–280, https://doi.org/10.1094/
MPMI.2003.16.4.271.
[82] G. Agner, Y.A. Kaulin, P.A. Gurnev, Z. Szabo, L.V. Schagina, J.Y. Takemoto, K. Blasko, Membrane-permeabilizing activities of cyclic lipodepsipeptides,
syringopeptin 22A and syringomycin E from Pseudomonas syringae pv. syringae in human red blood cells and in bilayer lipid membranes, Bioelectrochemistry
52 (2) (2000) 161–167, https://doi.org/10.1016/S0302-4598(00)00098-2.
[83] I. Meuskens, A. Saragliadis, J.C. Leo, D. Linke, Type V secretion systems: an overview of passenger domain functions, Front. Microbiol. 10 (2019) 1163,
https://doi.org/10.3389/fmicb.2019.01163.
[84] J. Guérin, S. Bigot, R. Schneider, S.K. Buchanan, F. Jacob-Dubuisson, Two-partner secretion: combining efficiency and simplicity in the secretion of large
proteins for bacteria-host and bacteria-bacteria interactions, Front. Cell. Infect. Microbiol. 7 (2017) 148, https://doi.org/10.3389/fcimb.2017.00148.
[85] S. Fan, Y. Chang, G. Liu, S. Shang, L. Tian, H. Shi, Molecular functional analysis of auxin/indole-3-acetic acid proteins (Aux/IAAs) in plant disease resistance in
cassava, Physiol. Plantarum 168 (1) (2020) 88–97, https://doi.org/10.1111/ppl.12970.
[86] J.C. Leo, D. Linke, A unified model for BAM function that takes into account type Vc secretion and species differences in BAM composition, AIMS Microbiol 4
(3) (2018) 455.
[87] P. Bernal, M.A. Llamas, A. Filloux, Type VI secretion systems in plant-associated bacteria, Environmen. Microbiol. 20 (1) (2018) 1–15, https://doi.org/
10.1111/1462-2920.13956.
[88] S. Timilsina, N. Potnis, E.A. Newberry, P. Liyanapathiranage, F. Iruegas-Bocardo, F.F. White, E.M. Goss, J.B. Jones, Xanthomonas diversity, virulence and
plant–pathogen interactions, Nat. Rev. Microbiol. 18 (8) (2020) 415–427, https://doi.org/10.1038/s41579-020-0361-8.
[89] Y. Choi, N. Kim, M. Mannaa, H. Kim, J. Park, H. Jung, G. Han, H.H. Lee, Y.S. Seo, Characterization of type VI secretion system in Xanthomonas oryzae pv. oryzae
and its role in virulence to rice, Plant Pathol. J. 36 (3) (2020) 289.
[90] P.C. Zhu, Y.M. Li, X. Yang, H.F. Zou, X.L. Zhu, X.N. Niu, L.H. Xu, W. Jiang, S. Huang, J.L. Tang, Y.Q. He, Type VI secretion system is not required for virulence
on rice but for inter-bacterial competition in Xanthomonas oryzae pv. oryzicola, Res. Microbiol. 171 (2) (2020) 64–73, https://doi.org/10.1016/j.
resmic.2019.10.004.
[91] J.C. Paton, C. Trappetti, Streptococcus pneumoniae capsular polysaccharide, Microbiol. Spect. 7 (2) (2019) 10–1128, https://doi.org/10.1128/microbiolspec.
gpp3-0019-2018.
[92] J.M. Dow, L. Crossman, K. Findlay, Y.Q. He, J.X. Feng, J.L. Tang, Biofilm dispersal in Xanthomonas campestris is controlled by cell–cell signaling and is required
for full virulence to plants, Proc. Nat. Acad. Sci. 100 (19) (2003) 10995–11000, https://doi.org/10.1073/pnas.1833360100.
[93] T.P. Denny, Involvement of bacterial polysaccharides in plant pathogenesis, Annu. Rev. Phytopathol. 33 (1) (1995) 173–197, https://doi.org/10.1146/
annurev.py.33.090195.001133.
[94] J.W. Chan, P.H. Goodwin, The molecular genetics of virulence of Xanthomonas campestris, Biotechnol. Adv. 17 (6) (1999) 489–508, https://doi.org/10.1016/
S0734-9750(99)00025-7.
[95] M.H. Yun, P.S. Torres, M.E. Oirdi, L.A. Rigano, R. Gonzalez-Lamothe, M.R. Marano, A.P. Castagnaro, M.A. Dankert, K. Bouarab, A.A. Vojnov, Xanthan induces
plant susceptibility by suppressing callose deposition, Plant Physiol. 141 (1) (2006) 178–187, https://doi.org/10.1104/pp.105.074542.
[96] P. Stoodley, K. Sauer, D.G. Davies, J.W. Costerton, Biofilms as complex differentiated communities, Annu. Rev. Microbiol. 56 (1) (2002) 187–209, https://doi.
org/10.1146/annurev.micro.56.012302.160705.
[97] F. Katzen, D.U. Ferreiro, C.G. Oddo, M.V. Ielmini, A. Becker, A. Pühler, L. Ielpi, Xanthomonas campestris pv. campestris gum mutants: effects on xanthan
biosynthesis and plant virulence, J. Bacteriol. 180 (7) (1998) 1607–1617, https://doi.org/10.1128/jb.180.7.1607-1617.1998.
[98] A.A. Vojnov, A. Zorreguieta, J.M. Dow, M.J. Daniels, M.A. Dankert, Evidence for a role for the gumB and gumC gene products in the formation of xanthan from
its pentasaccharide repeating unit by Xanthomonas campestris, Microbiol. 144 (6) (1998) 1487–1493, https://doi.org/10.1099/00221287-144-6-1487.
[99] B. Zhang, H. Zhang, F. Li, Y. Ouyang, M. Yuan, X. Li, J. Xiao, S. Wang, Multiple alleles encoding atypical NLRs with unique central tandem repeats in rice
confer resistance to Xanthomonas oryzae pv. oryzae, Plant Comm. 1 (4) (2020), https://doi.org/10.1016/j.xplc.2020.100088.
[100] I. Mensi, J.H. Daugrois, I. Pieretti, D. Gargani, L.A. Fleites, J. Noell, F. Bonnot, D.W. Gabriel, P. Rott, Surface polysaccharides and quorum sensing are involved
in the attachment and survival of Xanthomonas albilineans on sugarcane leaves, Mol. Plant Pathol. 17 (2) (2016) 236–246, https://doi.org/10.1111/
mpp.12276.
[101] P. Rott, L. Fleites, G. Marlow, M. Royer, D.W. Gabriel, Identification of new candidate pathogenicity factors in the xylem-invading pathogen Xanthomonas
albilineans by transposon mutagenesis, Mol. Plant Microbe Interact. 24 (5) (2011) 594–605, https://doi.org/10.1094/MPMI-07-10-0156.
[102] A. Silipo, G. Erbs, T. Shinya, J.M. Dow, M. Parrilli, R. Lanzetta, N. Shibuya, M.A. Newman, A. Molinaro, Glyco-conjugates as elicitors or suppressors of plant
innate immunity, Glycobiol 20 (4) (2010) 406–419, https://doi.org/10.1093/glycob/cwp201.

14
R. Dey and R. Raghuwanshi Heliyon 10 (2024) e34275

[103] A. Meyer, A. Pühler, K. Niehaus, The lipopolysaccharides of the phytopathogen Xanthomonas campestris pv. campestris induce an oxidative burst reaction in cell
cultures of Nicotiana tabacum, Planta 213 (2001) 214–222, https://doi.org/10.1007/s004250000493.
[104] M.A. Newman, E. Von Roepenack-Lahaye, A. Parr, M.J. Daniels, J.M. Dow, Prior exposure to lipopolysaccharide potentiates expression of plant defenses in
response to bacteria, Plant J. 29 (4) (2002) 487–495, https://doi.org/10.1046/j.0960-7412.2001.00233.x.
[105] A. Silipo, L. Sturiale, D. Garozzo, G. Erbs, T.T. Jensen, R. Lanzetta, J.M. Dow, M. Parrilli, M.A. Newman, A. Molinaro, The acylation and phosphorylation
pattern of lipid A from Xanthomonas campestris strongly influence its ability to trigger the innate immune response in Arabidopsis, Chembiochem 9 (6) (2008)
896–904, https://doi.org/10.1002/cbic.200700693.
[106] B.M. Lee, Y.J. Park, D.S. Park, H.W. Kang, J.G. Kim, E.S. Song, I.C. Park, U.H. Yoon, J.H. Hahn, B.S. Koo, G.B. Lee, The genome sequence of Xanthomonas oryzae
pathovar oryzae KACC10331, the bacterial blight pathogen of rice, Nucleic Acids Res. 33 (2) (2005) 577–586, https://doi.org/10.1093/nar/gki206.
[107] S.L. Salzberg, D.D. Sommer, M.C. Schatz, A.M. Phillippy, P.D. Rabinowicz, S. Tsuge, A. Furutani, H. Ochiai, A.L. Delcher, D. Kelley, R. Madupu, Genome
sequence and rapid evolution of the rice pathogen Xanthomonas oryzae pv. oryzae PXO99A, BMC Genom. 9 (2008) 1–16, https://doi.org/10.1186/1471-2164-
9-204.
[108] S.K. Ray, R. Rajeshwari, Y. Sharma, R.V. Sonti, A high-molecular-weight outer membrane protein of Xanthomonas oryzae pv. oryzae exhibits similarity to non-
fimbrial adhesins of animal pathogenic bacteria and is required for optimum virulence, Mol. Microbiol. 46 (3) (2002) 637–647, https://doi.org/10.1046/
j.1365-2958.2002.03188.x.
[109] A. Das, N. Rangaraj, R.V. Sonti, Multiple adhesin-like functions of Xanthomonas oryzae pv. oryzae are involved in promoting leaf attachment, entry, and
virulence on rice, Mol. Plant Microbe Interact. 22 (1) (2009) 73–85, https://doi.org/10.1094/MPMI-22-1-0073.
[110] N. Gottig, B.S. Garavaglia, C.G. Garofalo, E.G. Orellano, J. Ottado, A filamentous hemagglutinin-like protein of Xanthomonas axonopodis pv. citri, the
phytopathogen responsible for citrus canker, is involved in bacterial virulence, PLoS One 4 (2) (2009) e4358, https://doi.org/10.1371/journal.pone.0004358.
[111] H.N. Fones, Presence of ice-nucleating Pseudomonas on wheat leaves promotes Septoria tritici blotch disease (Zymoseptoria tritici) via a mutually beneficial
interaction, Sci. Rep. 10 (1) (2020) 17738, https://doi.org/10.1038/s41598-020-74615-7, 2020.
[112] S. Sule, E. Seemuller, The role of ice formation in the infection of sour cherry leaves by Pseudomonas syringae pv. Syringae, Phytopathology 77 (2) (1987)
173–177.
[113] H. Azad, N.W. Schaad, The relationship of Xanthomonas campestris pv. translucens to frost and the effect of frost on black chaff development in wheat,
Phytopathology 78 (1) (1988) 95–100.
[114] H.K. Kim, C. Orser, S.E. Lindow, D.C. Sands, Xanthomonas campestris pv. translucens strains active in ice nucleation, Plant Dis. 71 (11) (1987) 994–997, https://
doi.org/10.1094/PD-71-0994.
[115] G.F. Johnson, The early history of copper fungicides, Agric. Hist. 9 (2) (1935) 67–79.
[116] F. Behlau, N.L. Barelli, Jr J. Belasque, Lessons from a case of successful eradication of citrus canker in a citrus-producing farm in São Paulo state, Brazil, J. Plant
Pathol. 96 (3) (2014).
[117] J.R. Lamichhane, E. Osdaghi, F. Behlau, J. Köhl, J.B. Jones, J.N. Aubertot, Thirteen decades of antimicrobial copper compounds applied in agriculture,
A review. Agron. Sustain. Dev. 38 (2018) 1–18, https://doi.org/10.1007/s13593-018-0503-9.
[118] P.M. Martins, T.K. Wood, A.A. de Souza, Persister cells form in the plant pathogen Xanthomonas citri subsp. citri under different stress conditions,
Microorganisms 9 (2) (2021) 384, https://doi.org/10.3390/microorganisms9020384.
[119] P.L. Thayer, R.E. Stall, The survey of Xanthomonas vesicatoria resistance to streptomycin, Proc. Fla. State Hortic. Soc. 75 (1962) 163–165.
[120] G.M. Marco, R.E. Stall, Control of bacterial spot of pepper initiated by strains of Xanthomonas campestris pv. vesicatoria that differ in sensitivity to copper, Plant
Dis. 6797 (1983) 779–781.
[121] R.A. Conover, N.R. Gerhold, Mixtures of copper and maneb or mancozeb for control of bacterial spot of tomato and their compatibility for control of fungus
diseases, Flo. State Hort. Soc. Proc. 94 (1981) 154–156.
[122] M. Young, S. Santra, Copper (Cu)–Silica nanocomposite containing valence-engineered Cu: a new strategy for improving the antimicrobial efficacy of Cu
biocides, J. Agric. Food Chem. 62 (26) (2014) 6043–6052, https://doi.org/10.1021/jf502350w.
[123] J. Li, N. Wang, Foliar application of biofilm formation–inhibiting compounds enhances control of citrus canker caused by Xanthomonas citri subsp. citri,
Phytopathology 104 (2) (2014) 134–142, https://doi.org/10.1094/PHYTO-04-13-0100-R.
[124] K. Qiao, Q. Liu, Y. Huang, Y. Xia, S. Zhang, Management of bacterial spot of tomato caused by copper-resistant Xanthomonas perforans using a small molecule
compound carvacrol, Crop Protect. 132 (2020) 105114, https://doi.org/10.1016/j.cropro.2020.105114.
[125] P.A. Abbasi, J. Al-Dahmani, F. Sahin, H.A.J. Hoitink, S.A. Miller, Effect of compost amendments on disease severity and yield of tomato in conventional and
organic production systems, Plant Dis. 86 (2) (2002) 156–161, https://doi.org/10.1094/PDIS.2002.86.2.156.
[126] J.H. Graham, K.M. Gerberich, C.L. Davis, Injection–infiltration of attached grapefruit with Xanthomonas citri subsp. citri to evaluate seasonal population
dynamics in citrus canker lesions, J. Phytopathol. 164 (7–8) (2016) 528–533, https://doi.org/10.1111/jph.12478.
[127] H.J. Krauthausen, N. Laun, W. Wohanka, Methods to reduce the spread of the black rot pathogen, Xanthomonas campestris pv. campestris, in Brassica transplants,
J. Plant Dis. Protect. 118 (1) (2011) 7–16, https://doi.org/10.1007/BF03356375.
[128] M.T. Islam, B.R. Lee, P.R. Das, H.I. Jung, T.H. Kim, Characterization of p-Coumaric acid-induced soluble and cell wall-bound phenolic metabolites in relation to
disease resistance to Xanthomonas campestris pv. campestris in Chinese cabbage, Plant Physiol. Biochem. 125 (2018) 172–177, https://doi.org/10.1016/j.
plaphy.2018.02.012.
[129] H.P. Jadhav, S.S. Shaikh, R.Z. Sayyed, Role of hydrolytic enzymes of Rhizoflora in biocontrol of fungal phytopathogens: an overview, in: S. Mehnaz (Ed.),
Rhizotrophs: Plant Growth Promotion to Bioremediation, Microorganisms for Sustainability, vol. 2, Springer, Singapore, 2017, https://doi.org/10.1007/978-
981-10-4862-3_9.
[130] S.F.S. Ab Rahman, E. Singh, C.M. Pieterse, P.M. Schenk, Emerging microbial biocontrol strategies for plant pathogens, Plant Sci. 267 (2018) 102–111, https://
doi.org/10.1016/j.plantsci.2017.11.012.
[131] K. Kanthaiah, R.K. Velu, Characterization of the bioactive metabolite from a plant growth promoting rhizobacteria Pseudomonas aeruginosa VRKK1 and
exploitation of antibacterial behavior against Xanthomonas campestris a causative agent of bacterial blight disease in cowpea, Arch. Phytopathol. Plant Protect.
55 (7) (2022) 797–814, https://doi.org/10.1016/j.biocontrol.2007.01.008.
[132] A. Sampathkumar, K.E.A. Aiyanathan, S. Nakkeeran, S. Manickam, Multifaceted Bacillus spp. for the management of cotton bacterial blight caused by
Xanthomonas citri pv. malvacearum, Biol. Control 177 (2023) 105111, https://doi.org/10.1016/j.biocontrol.2022.105111.
[133] L. Wu, H. Wu, L. Chen, X. Yu, R. Borriss, X. Gao, Difficidin and bacilysin from Bacillus amyloliquefaciens FZB42 have antibacterial activity against Xanthomonas
oryzae rice pathogens, Sci. Rep. 5 (1) (2015) 12975, https://doi.org/10.1038/srep12975.
[134] M.F. Rabbee, K.H. Baek, Detection of antagonistic compounds synthesized by Bacillus velezensis against Xanthomonas citri subsp. citri by Metabolome and RNA
sequencing, Microorganisms 11 (6) (2023) 1523, https://doi.org/10.3390/microorganisms11061523.
[135] S. Chandwani, S. Dewala, S.M. Chavan, D. Paul, K. Kumar, N. Amaresan, Genomic, LC–MS, and FTIR analysis of plant Probiotic potential of Bacillus albus for
managing Xanthomonas oryzae via different Modes of application in rice (Oryza sativa L.), Probiotics Antimicrob. Proteins (2023) 1–12, https://doi.org/
10.1007/s12602-023-10120-3.
[136] S. Mishra, N.K. Arora, Evaluation of rhizospheric Pseudomonas and Bacillus as biocontrol tool for Xanthomonas campestris pv campestris, World J. Microbiol.
Biotechnol. 28 (2012) 693–702, https://doi.org/10.1007/s11274-011-0865-5.
[137] P. Jin, Y. Wang, Z. Tan, W. Liu, W. Miao, Antibacterial activity and rice-induced resistance, mediated by C15surfactin A, in controlling rice disease caused by
Xanthomonas oryzae pv. oryzae, Pesticide Biochem. Physiol. 169 (2020) 104669, https://doi.org/10.1016/j.pestbp.2020.104669.
[138] S. Villamizar, J.A. Ferro, J.C. Caicedo, L.M.C. Alves, Bactericidal effect of entomopathogenic bacterium Pseudomonas entomophila against Xanthomonas citri
reduces citrus canker disease severity, Front. Microbiol. 11 (2020) 1431, https://doi.org/10.3389/fmicb.2020.01431.
[139] R. Yang, Q. Shi, T. Huang, Y. Yan, S. Li, Y. Fang, Y. Li, L. Liu, L. Liu, X. Wang, Y. Peng, The natural pyrazolotriazine pseudoiodinine from Pseudomonas mosselii
923 inhibits plant bacterial and fungal pathogens, Nat. Comm. 14 (1) (2023) 734, https://doi.org/10.1038/s41467-023-36433-z.

15
R. Dey and R. Raghuwanshi Heliyon 10 (2024) e34275

[140] S. Gupta, S. Pandey, S.P. Nandi, M. Singh, Modulation of ethylene and ROS-scavenging enzymes by multifarious plant growth-promoting endophytes in tomato
(Solanum lycopersicum) plants to combat Xanthomonas-induced stress, Plant Physiol. Biochem. 202 (2023) 107982, https://doi.org/10.1016/j.
plaphy.2023.107982.
[141] S. Olishevska, A. Nickzad, C. Restieri, F. Dagher, Y. Luo, J. Zheng, E. Déziel, Bacillus velezensis and Paenibacillus peoriae strains effective as biocontrol agents
against Xanthomonas bacterial spot, Appl. Microbiol. 3 (3) (2023) 1101–1119, https://doi.org/10.3390/applmicrobiol3030076.
[142] S. Yasmin, F.Y. Hafeez, M.S. Mirza, M. Rasul, H.M. Arshad, M. Zubair, M. Iqbal, Biocontrol of bacterial leaf blight of rice and profiling of secondary metabolites
produced by rhizospheric Pseudomonas aeruginosa BRp3, Front. Microbiol. 8 (2017) 1895, https://doi.org/10.3389/fmicb.2017.01895.
[143] A. Jelušić, T. Popović, I. Dimkić, P. Mitrović, K. Peeters, A.M. Višnjevec, Č. Tavzes, S. Stanković, T. Berić, Changes in the winter oilseed rape microbiome
affected by Xanthomonas campestris pv. campestris and biocontrol potential of the indigenous Bacillus and Pseudomonas isolates, Biol. Control 160 (2021)
104695, https://doi.org/10.1016/j.biocontrol.2021.104695.
[144] K.D. Le, N.H. Yu, A.R. Park, D.J. Park, C.J. Kim, J.C. Kim, Streptomyces sp. AN090126 as a biocontrol agent against bacterial and fungal plant diseases,
Microorganisms 10 (4) (2022) 791, https://doi.org/10.3390/microorganisms10040791.
[145] Y.H. Yoo, M.Y. Lee, Y.J. Kim, E.K. Ahn, K.H. Jung, C.H. Lee, LC–MS analysis for Paenibacillus yonginensis DCY84T and Silicon Informs the significance of
suppressed Unsaturated Fatty acids, J. Plant Biol. 66 (1) (2023) 15–23, https://doi.org/10.1007/s12374-022-09371-4.
[146] D. Fira, I. Dimkić, T. Berić, J. Lozo, S. Stanković, Biological control of plant pathogens by Bacillus species, J. Biotechnol. 285 (2018) 44–55, https://doi.org/
10.1016/j.jbiotec.2018.07.044.
[147] T.P. Huang, D.D.S. Tzeng, A.C. Wong, C.H. Chen, K.M. Lu, Y.H. Lee, W.D. Huang, B.F. Hwang, K.C. Tzeng, DNA polymorphisms and biocontrol of Bacillus
antagonistic to citrus bacterial canker with indication of the interference of phyllosphere biofilms, PLoS One 7 (7) (2012) e42124, https://doi.org/10.1371/
journal.pone.0042124.
[148] S.E. Khabbaz, L. Zhang, L.A. Cáceres, M. Sumarah, A. Wang, P.A. Abbasi, Characterisation of antagonistic Bacillus and Pseudomonas strains for biocontrol
potential and suppression of damping-off and root rot diseases, Ann. Appl. Biol. 166 (3) (2015) 456–471, https://doi.org/10.1111/aab.12196.
[149] S. Thapa, M. Babadoost, Effectiveness of chemical compounds and biocontrol agents for management of bacterial spot of pumpkin Caused by Xanthomonas
cucurbitae, Plant Health Prog. 17 (2) (2016) 106–113, https://doi.org/10.1094/PHP-RS-15-0037.
[150] N. Daranas, G. Roselló, J. Cabrefiga, I. Donati, J. Francés, E. Badosa, F. Spinelli, E. Montesinos, A. Bonaterra, Biological control of bacterial plant diseases with
Lactobacillus plantarum strains selected for their broad-spectrum activity, Ann. Appl. Biol. 174 (1) (2019) 92–105, https://doi.org/10.1111/aab.12476.
[151] I. Pajčin, V. Vlajkov, M. Frohme, S. Grebinyk, M. Grahovac, M. Mojićević, J. Grahovac, Pepper bacterial spot control by Bacillus velezensis: Bioprocess solution,
Microorganisms 8 (10) (2020) 1463, https://doi.org/10.3390/microorganisms8101463.
[152] K. Das, R. Prasanna, A.K. Saxena, Rhizobia: a potential biocontrol agent for soilborne fungal pathogens, Folia microbiologica 62 (2017) 425–435, https://doi.
org/10.1007/s12223-017-0513-z.
[153] A.P.S. da Silva, F.L. Olivares, C.P. Sudré, L.E.P. Peres, N.A. Canellas, R.M. da Silva, V. Cozzolino, L.P. Canellas, Attenuations of bacterial spot disease
Xanthomonas euvesicatoria on tomato plants treated with biostimulants, Chem. Biol. Technol. Agric. 8 (2021) 1–9, https://doi.org/10.1186/s40538-021-00240-
9.
[154] K.D. Le, J. Kim, N.H. Yu, B. Kim, C.W. Lee, J.C. Kim, Biological control of tomato bacterial wilt, kimchi cabbage soft rot, and red pepper bacterial leaf spot
using Paenibacillus elgii JCK-5075, Front. Plant Sci. 11 (2020) 775, https://doi.org/10.3389/fpls.2020.00775.
[155] W.P. Moss, J.M. Byrne, H.L. Campbell, P. Ji, U. Bonas, J.B. Jones, M. Wilson, Biological control of bacterial spot of tomato using hrp mutants of Xanthomonas
campestris pv. vesicatoria, Biol. Control 41 (2) (2007) 199–206, https://doi.org/10.1016/j.biocontrol.2007.01.008.
[156] S. Caulier, C. Nannan, A. Gillis, F. Licciardi, C. Bragard, J. Mahillon, Overview of the antimicrobial compounds produced by members of the Bacillus subtilis
group, Front. Microbiol. 10 (2019) 302, https://doi.org/10.3389/fmicb.2019.00302.
[157] D. Zeng, S.S. Liu, W.B. Shao, T.H. Zhang, P.Y. Qi, H.W. Liu, X. Zhou, L.W. Liu, H. Zhang, S. Yang, New inspiration of 1, 3, 4-oxadiazole agrochemical
candidates: manipulation of a type III secretion system-induced bacterial starvation mechanism to prevent plant bacterial diseases, J. Agri. Food Chem. 71 (6)
(2023) 2804–2816, https://doi.org/10.1021/acs.jafc.2c07486.
[158] R. Rodriguez, R. Redman, More than 400 million years of evolution and some plants still can’t make it on their own: plant stress tolerance via fungal symbiosis,
J. Exp. Bot. 59 (5) (2008) 1109–1114, https://doi.org/10.1093/jxb/erm342.
[159] J. Inbar, A.N.A. Menendez, I. Chet, Hyphal interaction between Trichoderma harzianum and Sclerotinia sclerotiorum and its role in biological control, Soil Biol.
Biochem. 28 (6) (1996) 757–763, https://doi.org/10.1016/0038-0717(96)00010-7.
[160] E.A. Elsherbiny, M.A. Taher, M.F. Elsebai, Activity of Purpureocillium lilacinum filtrates on biochemical characteristics of Sclerotinia sclerotiorum and induction
of defense responses in common bean, Eur. J. Plant Pathol. 155 (2019) 39–52, https://doi.org/10.1007/s10658-019-01748-5.
[161] S.D. Silva, R.M. Carneiro, M. Faria, D.A. Souza, R.G. Monnerat, R.B. Lopes, Evaluation of Pochonia chlamydosporia and Purpureocillium lilacinum for suppression
of Meloidogyne enterolobii on tomato and banana, J. Nematol. 49 (1) (2017) 77.
[162] V. Kumar, S.K. Dwivedi, Bioremediation mechanism and potential of copper by actively growing fungus Trichoderma lixii CR700 isolated from electroplating
wastewater, J. Environ. Manage. 277 (2021) 111370, https://doi.org/10.1016/j.jenvman.2020.111370.
[163] P.K. Mukherjee, A. Mendoza-Mendoza, S. Zeilinger, B.A. Horwitz, Mycoparasitism as a mechanism of Trichoderma-mediated suppression of plant diseases,
Fungal Biol. Rev. 39 (2022) 15–33, https://doi.org/10.1016/j.fbr.2021.11.004.
[164] A. Saxena, S. Mishra, S. Ray, R. Raghuwanshi, H.B. Singh, Differential reprogramming of defense network in Capsicum annum L. plants against Colletotrichum
truncatum infection by phyllospheric and rhizospheric Trichoderma strains, J. Plant Growth Regul. 39 (2020) 751–763, https://doi.org/10.1007/s00344-019-
10017-y.
[165] N.A. Zin, N.A. Badaluddin, Biological functions of Trichoderma spp. for agriculture applications, Annals Agric. Sci. 65 (2) (2020) 168–178, https://doi.org/
10.1016/j.aoas.2020.09.003.
[166] A.D.B. Fontenelle, S.D. Guzzo, C.M.M. Lucon, R. Harakava, Growth promotion and induction of resistance in tomato plant against Xanthomonas euvesicatoria
and Alternaria solani by Trichoderma spp, Crop Protect. 30 (11) (2011) 1492–1500, https://doi.org/10.1016/j.cropro.2011.07.019.
[167] C.R. Howell, L.E. Hanson, R.D. Stipanovic, L.S. Puckhaber, Induction of terpenoid synthesis in cotton roots and control of Rhizoctonia solani by seed treatment
with Trichoderma virens, Phytopathology 90 (3) (2000) 248–252, https://doi.org/10.1094/PHYTO.2000.90.3.248.
[168] R. Caracciolo, L. Sella, M. De Zotti, A. Bolzonello, M. Armellin, L. Trainotti, F. Favaron, S. Tundo, Efficacy of Trichoderma longibrachiatum trichogin GA IV
peptaibol analogs against the black rot pathogen Xanthomonas campestris pv. campestris and other phytopathogenic bacteria, Microorganisms 11 (2) (2023)
480, https://doi.org/10.3390/microorganisms11020480.
[169] Y.Q. Zhang, S. Zhang, M.L. Sun, H.N. Su, H.Y. Li, Y.Z. Zhang, X.L. Chen, H.Y. Cao, X.Y. Song, Antibacterial activity of peptaibols from Trichoderma
longibrachiatum SMF2 against gram-negative Xanthomonas oryzae pv. oryzae, the causal agent of bacterial leaf blight on rice, Front. Microbiol. 13 (2022)
1034779, https://doi.org/10.3389/fmicb.2022.1034779.
[170] C.M. Allsup, R.A. Lankau, K.N. Paige, Herbivory and soil water availability induce changes in arbuscular mycorrhizal fungal abundance and composition,
Microbial Ecol. 84 (1) (2022) 141–152, https://doi.org/10.1007/s00248-021-01835-3.
[171] H. Liu, W. Yang, B. Hu, F. Liu, Virulence analysis and race classification of Xanthomonas oryzae pv. oryzae in China, J. Phytopathol. 155 (3) (2007) 129–135,
https://doi.org/10.1111/j.1439-0434.2007.01197.x.
[172] Z. Qi, T.W. Mew, Adult-plant resistance of rice cultivars to bacterial blight, Plant Dis. 69 (10) (1985) 896–898.
[173] S.M. Assis, R.L. Mariano, S.J. Michereff, G. Silva, E.A. Maranhão, Antagonism of yeasts to Xanthomonas campestris pv. campestris on cabbage phylloplane in
field, Rev. Microbiol. 30 (1999) 191–195, https://doi.org/10.1590/S0001-37141999000300002.
[174] T. Yang, C. Wang, C. Li, R. Sun, M. Yang, Antagonistic effects of volatile organic compounds of Saccharomyces cerevisiae NJ-1 on the growth and toxicity of
Aspergillus flavus, Biol. Control 177 (2023) 105093, https://doi.org/10.1016/j.biocontrol.2022.105093.
[175] F.L. Gordillo Altamirano, J.J. Barr, Phage therapy in the post antibiotic era, Clin. Microbiol. Rev. 32 (2) (2019) 10–1128, https://doi.org/10.1128/cmr.00066-
18.

16
R. Dey and R. Raghuwanshi Heliyon 10 (2024) e34275

[176] W.L. Mallmann, C.A.R.L. Hemstreet, Isolation of an inhibitory substance from plants, Agric. Res 28 (1924) 599–602.
[177] E.L. Civerolo, H.L. Kiel, Inhibition of bacterial spot of peach foliage by Xanthomonas pruni bacteriophage, Phytopathology 59 (1969) 1966–1967.
[178] J.E. Flaherty, G.C. Somodi, J.B. Jones, B.K. Harbaugh, L.E. Jackson, Control of bacterial spot on tomato in the greenhouse and field with H-mutant
bacteriophages, Hortscience 35 (5) (2000) 882–884, https://doi.org/10.21273/HORTSCI.35.5.882.
[179] J.E. Flaherty, B.K. Harbaugh, J.B. Jones, G.C. Somodi, L.E. Jackson, H-mutant bacteriophages as a potential biocontrol of bacterial blight of geranium,
Hortscience 36 (1) (2001) 98–100, https://doi.org/10.21273/HORTSCI.36.1.98.
[180] J.M. Lang, D.H. Gent, H.F. Schwartz, Management of Xanthomonas leaf blight of onion with bacteriophages and a plant activator, Plant Dis. 91 (7) (2007)
871–878, https://doi.org/10.1094/PDIS-91-7-0871.
[181] B. Balogh, N.T.T. Nga, J.B. Jones, Relative level of bacteriophage multiplication in vitro or in phyllosphere may not predict in planta efficacy for controlling
bacterial leaf spot on tomato caused by Xanthomonas perforans, Front. Microbiol. 9 (2018) 2176, https://doi.org/10.3389/fmicb.2018.02176.
[182] A.A. Ahmad, A. Askora, T. Kawasaki, M. Fujie, T. Yamada, The filamentous phage XacF1 causes loss of virulence in Xanthomonas axonopodis pv. citri, the
causative agent of citrus canker disease, Front. Microbiol. 5 (2014) 321, https://doi.org/10.3389/fmicb.2014.00321.
[183] P. Ranjani, Y. Gowthami, S.S. Gnanamanickam, P. Palani, Bacteriophages: a new weapon for the control of bacterial blight disease in rice caused by
Xanthomonas oryzae, Microbiol. Biotechnol. Lett. 46 (4) (2018) 346–359, https://doi.org/10.4014/mbl.1807.07009.
[184] N.T.T. Nga, T.N. Tran, D. Holtappels, N.L. Kim Ngan, N.P. Hao, M. Vallino, D.T.K. Tien, N.H. Khanh-Pham, R. Lavigne, K. Kamei, J. Wagemans, Phage
biocontrol of bacterial leaf blight disease on Welsh onion caused by Xanthomonas axonopodis pv. allii, Antibiotics 10 (5) (2021) 517, https://doi.org/10.3390/
antibiotics10050517.
[185] E.L. Civerolo, Relationship of Xanthomonas pruni bacteriophages to bacterial spot disease in prunus, Phytopathology 63 (10) (1973) 1279.
[186] R. Nakayinga, A. Makumi, V. Tumuhaise, W. Tinzaara, Xanthomonas bacteriophages: a review of their biology and biocontrol applications in agriculture, BMC
Microbiol. 21 (1) (2021) 1–20, https://doi.org/10.1186/s12866-021-02351-7.

17

You might also like