Next Article in Journal
Hydrocarbon-Degrading Bacteria Found Tightly Associated with the 50–70 μm Cell-Size Population of Eukaryotic Phytoplankton in Surface Waters of a Northeast Atlantic Region
Next Article in Special Issue
Quorum Sensing Regulation in Phytopathogenic Bacteria
Previous Article in Journal
The Limits and Avoidance of Biases in Metagenomic Analyses of Human Fecal Microbiota
Previous Article in Special Issue
The Evanescent GacS Signal
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Innovation and Application of the Type III Secretion System Inhibitors in Plant Pathogenic Bacteria

1
Department of Plant, Soil, and Microbial Sciences, Michigan State University, East Lansing, MI 48824, USA
2
Department of Biological Sciences, University of Wisconsin-Milwaukee, Milwaukee, WI 53211, USA
*
Authors to whom correspondence should be addressed.
Microorganisms 2020, 8(12), 1956; https://doi.org/10.3390/microorganisms8121956
Submission received: 10 November 2020 / Revised: 5 December 2020 / Accepted: 7 December 2020 / Published: 9 December 2020
(This article belongs to the Special Issue Global Regulatory Networks in Plant Pathogenic Bacteria)

Abstract

:
Many Gram-negative pathogenic bacteria rely on a functional type III secretion system (T3SS), which injects multiple effector proteins into eukaryotic host cells, for their pathogenicity. Genetic studies conducted in different host-microbe pathosystems often revealed a sophisticated regulatory mechanism of their T3SSs, suggesting that the expression of T3SS is tightly controlled and constantly monitored by bacteria in response to the ever-changing host environment. Therefore, it is critical to understand the regulation of T3SS in pathogenic bacteria for successful disease management. This review focuses on a model plant pathogen, Dickeya dadantii, and summarizes the current knowledge of its T3SS regulation. We highlight the roles of several T3SS regulators that were recently discovered, including the transcriptional regulators: FlhDC, RpoS, and SlyA; the post-transcriptional regulators: PNPase, Hfq with its dependent sRNA ArcZ, and the RsmA/B system; and the bacterial second messenger cyclic-di-GMP (c-di-GMP). Homologs of these regulatory components have also been characterized in almost all major bacterial plant pathogens like Erwinia amylovora, Pseudomonas syringae, Pectobacterium spp., Xanthomonas spp., and Ralstonia spp. The second half of this review shifts focus to an in-depth discussion of the innovation and development of T3SS inhibitors, small molecules that inhibit T3SSs, in the field of plant pathology. This includes T3SS inhibitors that are derived from plant phenolic compounds, plant coumarins, and salicylidene acylhydrazides. We also discuss their modes of action in bacteria and application for controlling plant diseases.

1. Introduction

Type III secretion systems (T3SSs) are well-studied protein secretion/translocation systems found in almost all Gram-negative bacterial pathogens of plants and animals [1,2]. Structurally, the T3SSs are highly conserved among bacteria. They are syringe-like nanomachines consisting of inner and outer membrane rings, known as a basal body, and an apparatus that enables bacteria to inject diverse effector proteins directly into the host cell cytoplasm, which participate in the regulation of host cell functions to benefit the bacterial survival and multiplication [3,4]. In the plant pathogenic bacteria, the T3SSs have attracted much attention due to their ability to elicit the hypersensitive response (HR), a plant defense mechanism characterized by rapid cell death, in resistant or non-host plants and induce disease symptoms in host plants [4,5]. In Erwinia amylovora, for instance, the T3SS is a major pathogenicity factor as the T3SS-deficient mutants are unable to cause fire blight disease in Rosaceous plant hosts [6,7]. Studies of the T3SS from Pseudomonas syringae pv. tomato DC3000, which causes bacterial speck of tomato and the model plant Arabidopsis thaliana, have provided valuable information for the understanding of host–microbe interactions by signifying the T3SS and its effectors as essential players involved in the repression of plant defense mechanisms during the bacterial infection [8,9,10].
As is the case for many T3SS-expressing pathogens, a growing number of studies have explored novel disease management strategies using the T3SS as a target for the inhibition of bacterial pathogenesis [11,12,13]. In the studies of plant pathogenic bacteria, small molecules that target and inhibit the T3SSs have been discovered. The soft rot plant pathogen Dickeya dadantii served as the model organism due to its well-characterized regulatory pathways of the T3SS. It has been well established that the T3SSs from plant pathogenic bacteria are encoded by the hypersensitive response and pathogenicity (hrp) operon [14]. Based on the similarities of operon organization and regulatory systems, plant pathogenic bacterial hrp gene clusters are divided into two groups [15,16]. The group I hrp gene cluster includes those of P. syringae, E. amylovora, and Dickeya spp., in which the alternative sigma factor HrpL is the master regulator that activates the expression of most hrp genes [17,18,19]. Ralstonia spp. and Xanthomonas spp. possess the group II hrp gene clusters, whose expression is heavily relied on an AraC-type transcriptional activator HrpX [20,21]. Genetic and molecular analyses have identified multiple important regulators that control the expression of T3SS in plant pathogenic bacteria, suggesting that the regulation of T3SS is a highly modulated process during the infection of plants by pathogens, and such regulation often occurs in a hierarchical manner.
In this review, we summarize our current understanding of the regulation of T3SS, focusing on bacterial second messengers, transcriptional regulators, and post-transcriptional regulators that control the T3SS in a model pathogen D. dadantii. Similar studies of T3SSs from other plant pathogens, such as P. syringae and E. amylovora, have been thoroughly reviewed elsewhere [22,23,24,25,26]. Furthermore, we discuss early innovations of the T3SS inhibitors and their modes of action in D. dadantii and E. amylovora, highlight the recent discovery of novel T3SS inhibitors in rice pathogen X. oryzae pv. oryzae, bacterial wilt pathogen R. solanacearum, and P. syringae pv. tomato, and provide remaining challenges and future perspectives on the application of T3SS inhibitors for managing plant diseases.

2. Regulation of T3SS in D. dadantii

The expression of T3SS in D. dadantii is governed by the master regulator HrpL. This alternative sigma factor activates T3SS gene expression by binding to the highly conserved hrp box (GGAACC-N15/16-CCACNNA) in their promoter regions. Genes such as hrpA, hrpN, and dspE, which encode the T3SS pilus protein, a harpin, and a virulence effector, respectively, are all transcribed in a HrpL-dependent manner [17,18,27,28]. The regulation of HrpL is achieved via two independent regulatory cascades in D. dadantii (Figure 1) [29,30,31,32]. Transcriptionally, HrpS is a σ54 (RpoN) enhancer binding protein, which binds the σ54 (RpoN)-containing RNA polymerase holoenzyme and initiates the expression of hrpL. A two-component signal transduction system (TCSTS) HrpX/HrpY is responsible for the activation of hrpS [29,33]. Post-transcriptionally, hrpL is regulated by RsmA, a small RNA-binding protein, which binds the 5′ untranslated region of hrpL mRNA and facilitates its degradation [34]. RsmB is a non-coding RNA containing multiple RsmA binding sites that bind RsmA proteins to antagonize their negative effect on hrpL mRNA, resulting in a positive effect on the downstream T3SS gene expression [33,35]. The GacS/GacA TCSTS has been shown to upregulate the production of RsmB RNA in D. dadantii [31].

2.1. Bacterial Second Messengers Regulate T3SS

D. dadantii can infect a wide range of host plants and has the ability to survive in soil and groundwater [36,37]. In the past decade, several regulators have been identified to be involved in the regulation of T3SS, implying a sophisticated regulatory network that allows D. dadantii to control its virulence gene expression to adapt to various environmental conditions. Bis-(3′-5′)-cyclic dimeric guanosine monophosphate (c-di-GMP), a ubiquitous bacterial second messenger in most major bacterial phyla, is one of the most critical and well-studied nodes of this network [38,39]. The metabolism of c-di-GMP is dependent on two kinds of enzymes: The GGDEF domain-containing diguanylate cyclase (DGC) enzymes that synthesize c-di-GMP from two molecules of guanosine-5′-triphosphate (GTP) and the EAL or HD-GYP domain-containing phosphodiesterase (PDE) enzymes that degrade c-di-GMP to 5′-phosphoguanylyl-(3′-5′)-guanosine or two molecules of guanosine monophosphate (GMP), respectively [40,41,42,43,44].
The presence of GGDEF and/or EAL domain protein-encoding genes is abundant in many Gram-negative bacteria. For example, Escherichia coli K-12 contains 29 genes, Vibrio cholerae contains 53, and E. amylovora contains 8 [45,46,47,48]. In D. dadantii, genomic analysis has identified twelve gcp (GGDEF domain-containing protein), four ecp (EAL domain-containing protein), and two egcp (EAL-GGDEF domain-containing protein) genes. Interestingly, genes encoding the HD-GYP domain-containing protein have not yet been annotated. Exploratory studies by Yi and colleagues demonstrated that two PDEs, EGcpB and EcpC, positively regulate the T3SS gene expression, swimming motility, and the production of pectate lyase (Pel), one of the major plant cell wall degrading enzymes [37,49], whereas negatively regulate biofilm formation, suggesting a pleiotropic effect of c-di-GMP in controlling D. dadantii pathogenesis [50,51]. These observations are also in agreement with reports from other bacteria, suggesting that bacterial second messenger c-di-GMP plays an essential role in promoting the transition from motile to sessile lifestyles [39,52]. Yi and colleagues further demonstrated that deletions of egcpB or ecpC cause elevated intracellular c-di-GMP levels, which in turn post-transcriptionally repress the expression of rpoN (Figure 1) [50]. Why D. dadantii utilizes two functionally redundant enzymes to modulate the same biological target remains unknown. To elucidate the function of DGCs in D. dadantii, Yuan and Tian et al. performed a phenotypical study on twelve gcp deletion mutants and found that a DGC GcpA synthesizes c-di-GMP to repress T3SS and Pel via negatively regulating rsmB at the post-transcriptional level and positively regulate RsmA protein levels [53]. H-NS, a nucleoid-associated protein known to activate Pel [54,55,56], is negatively regulated by GcpA at the level of post-transcription, which leads to the repression of rsmB [53]. Interestingly, H-NS is not involved in the regulation of T3SS by GcpA, possibly due to its role in modulating the DNA topology in D. dadantii [56]. Together, c-di-GMP has been shown to control the expression of T3SS in D. dadantii via two pathways: The RpoN-HrpL pathway and the RsmA/RsmB-HrpL pathway.
The expression of T3SS in plant pathogenic bacteria is known to be repressed in nutrient-rich media and induced in minimal media, which may correspond to the nutrient-deficient and low pH environment in plant apoplast [32,57,58]. To study the role of carbon source in T3SS regulation, the D. dadantii strain containing a transcriptional fusion of the hrpN gene to the E. coli lacZ gene was used as a reporter system [59]. The expression of the hrpN::lacZ fusion was under the control of the hrpN promoter and determined by the β-galactosidase assay. Nasser and colleagues reported that the expression of hrpN::lacZ fusion in D. dadantii was induced 5-fold in a minimal medium when supplemented with sucrose compared with glycerol [59], suggesting that the carbon source is involved in T3SS regulation. Yuan et al. recently showed that citrate, one of the major intermediates in the tricarboxylic acid (TCA) cycle, represses Pel production [60] and the T3SS gene expression (C.-H. Yang, unpublished data) in a c-di-GMP-dependent manner in D. dadantii. Deletions of TCA cycle enzymes reduce intracellular c-di-GMP levels in D. dadantii [60]; application of exogenous citrate induces intracellular c-di-GMP via promoting the expression of gcpA while repressing the expression of egcpB [60]. As the production of c-di-GMP has been reported to be induced by citrate in P. fluorescens [61], these data suggest that c-di-GMP might generally serve as an intracellular signal that regulates T3SSs in response to various environmental signals.
The nucleotide second messengers guanosine tetraphosphate (ppGpp) and guanosine pentaphosphate (pppGpp) are stringent response regulators that globally reprogram bacterial transcription via interacting with RNA polymerase and its binding protein DksA [62,63]. Several recent studies highlighted the role of (p)ppGpp in regulating T3SS in E. amylovora and P. syringae, showing that the (p)ppGpp-triggered T3SS gene expression might rely on the RpoN-HrpL cascade in E. amylovora [64,65,66,67]. However, whether this regulation occurs in D. dadantii has not yet been determined.

2.2. Transcriptional Regulators Control T3SS

The T3SS and flagellum are evolutionarily closely related due to their similarities in structure, function, and sequences of their main components [68,69,70,71]. In P. syringae, flagellin could be translocated into plant cells by the T3SS and induce immune responses [72]. Genetically, the master regulator of the flagellar assembly genes FlhDC has been shown to be essential for the activation of T3SS genes in D. dadantii and Pectobacterium carotovorum [51,73,74,75,76]. A detailed genetic study conducted by Yuan et al. demonstrated that D. dadantii FlhDC transcriptionally initiates the expression of ecpC, which regulates the T3SS through the c-di-GMP-mediated RpoN-HrpL pathway (Figure 1) [50,51]. Interestingly, the homolog of EcpC, named YhjH, in E. coli is also positively regulated by FlhDC. However, different from E. coli, the alternative sigma factor (σ28) FliA, which is under the control of FlhDC, is not required for the activation of ecpC via FlhDC in D. dadantii [51,77]. Additionally, since the consensus FlhDC binding sequence, AA(C/T)G(C/G)N2-3AAATA(A/G)CG [78,79], is not present in the promoter region of ecpC, the activation of ecpC by FlhDC might be indirect. FlhDC and YcgR, a PilZ domain protein, have also been shown to be involved in the regulation of T3SS through the RsmA/RsmB system (Figure 1) [51] and the function of YcgR has been shown to be activated by the binding to c-di-GMP [80,81]. However, in vitro studies demonstrated that their contributions to T3SS are not significant compared with the FlhDC-EcpC-RpoN-HrpL pathway [51].
The sigma factor RpoS is the master regulator for the stress response in bacteria [82]. In E. coli, RpoS is degraded by the ClpXP protease with the assistance of the recognition factor RssB [83,84,85]. RpoS from P. carotovorum has been reported to affect the production of Pel, T3SS gene hrpN expression, and virulence in plant through its regulation on RsmA [86]. Subsequent studies from D. dadantii and E. amylovora showed similar findings with a more detailed regulatory mechanism proposed in D. dadantii: The ClpXP-RssB-RpoS regulatory cascade controls the expression of rsmA, which post-transcriptionally affects HrpL; RpoS represses hrpL transcription without affecting HrpS or RpoN (Figure 1) [87,88], suggesting that a novel RpoS-HrpL pathway might exist in D. dadantii.
SlyA belongs to the member of the SlyA/MarR family transcriptional regulator [89]. In D. dadantii, SlyA was first characterized as a Pel regulator, and the deletion of slyA significantly reduced disease symptoms in planta [90]. Since SlyA has been reported to regulate the expression of T3SS in Salmonella enterica serovar Typhimurium [91], the role of SlyA homologue in controlling D. dadantii T3SS was investigated. Zou and colleagues reported that SlyA negatively regulates HrpL via two pathways: It upregulates the expression of rsmA and downregulates the transcription of hrpS that is independent of the TCSTS HrpX/HrpY (Figure 1). Interestingly, despite its negative impact on hrpL, SlyA positively regulates the expression of hrp regulon genes, such as hrpA and hrpN, in parallel with HrpL (Figure 1) [92], suggesting that multiple factors might be involved in the transcriptional regulation of T3SS genes in D. dadantii. Indeed, PecS, another MarR family transcriptional regulator [93,94], and PecT, a LysR family transcriptional regulator [37], have been shown to repress the transcription of hrpN (Figure 1) [59]. In vitro DNA-protein binding and DNase I footprinting analyses confirmed that PecS directly interacts with the promoter of hrpN to repress its transcription [59]. The mechanism of PecT-mediated regulation on hrpN remains unclear. Nevertheless, it will be interesting to know how and under what circumstances these transcriptional regulators modulate T3SS regulon gene expression in a cooperative or rather competitive manner.

2.3. Post-Transcriptional Regulators Control T3SS

Post-transcriptional regulation is an essential mechanism to control gene expression in bacteria [95,96]. Progress made in D. dadantii has identified several post-transcriptional regulators that control the expression of T3SS. Polynucleotide phosphorylase (PNPase) is one of the post-transcriptional regulators conserved in bacteria and eukaryotes [97,98,99]. Known as an exoribonuclease, PNPase is majorly involved in RNA decay [100]. The PNPase homologues in Yersinia spp. and Salmonella spp. have been reported to control T3SS [101,102]. The deletion of pnpase in D. dadantii significantly increased the transcriptional activities and mRNA levels of hrpA, hrpN, and DspE, suggesting that PNPase downregulates T3SS through HrpL in D. dadantii [103]. Further analyses confirmed that PNPase negatively regulates the stability of rpoN mRNA, which in turn affects the transcription of hrpL; PNPase also stimulates the decay of hrpL mRNA by reducing the amount of available RsmB transcripts (Figure 1) [103]. In Salmonella and E. coli, PNPase has been reported to be essential for the RNA decay of RsmB homologues CsrB [104,105].
The expression of RsmB was recently shown to be regulated by Hfq, an RNA chaperone [106,107], via a feed-forward signaling circuit in D. dadantii (Figure 1) [108]. Hfq and the Hfq-dependent small regulatory RNA (sRNA), ArcZ, repress the translation of pecT. PecT auto-inhibits its own transcription [109] and, more importantly, downregulates the transcription of rsmB that contributes to the Hfq-mediated regulation on the T3SS and Pel. As the PecT homologue HexA is also known to repress RsmB in P. carotovorum [110], it is reasonable to speculate that this regulation occurs at the level of transcription. RsmB is also post-transcriptionally regulated by Hfq since the deletion of hfq elevated the intracellular c-di-GMP levels owing to the increased productions of two DGCs, GcpA and GcpL. Both DGCs are required for the Hfq-mediated T3SS regulation. In E. amylovora, Zeng et al. found that Hfq and ArcZ were required for the virulence in host plants and the T3SS-dependent HR in non-host tobacco plants [111,112], but the mechanism has not yet been reported.

3. Discovery of T3SS Inhibitors in Plant Pathogens and Their Regulatory Mechanisms

Small molecules that could specifically inhibit the synthesis or functionality of the T3SS are referred to as T3SS inhibitors. Unlike traditional antibiotics that often target bacterial survival, T3SS inhibitors display negligible effects on bacterial growth, thus reducing the selective pressure for the development of resistance [12,113,114]. In plant pathogenic bacteria, extensive studies have led to the discovery of a group of plant-derived compounds and several chemically synthesized compounds that modulate the expression of T3SS in major plant pathogens (Table 1). Further studies detailed the mode of action of these compounds and investigated their potential for disease management.

3.1. Plant Phenolic Compounds as T3SS Inhibitors

Plant phenolic compounds are one of the most widespread secondary metabolites in plants. They range from low molecular weight and single aromatic ringed compounds to large and complex-polyphenols and are involved in diverse physiological activities in plants, such as pigmentation, growth, and defense mechanisms [123]. Plant phenolic compounds also function as signal molecules that either induce or repress the microbial gene expression during the interactions between plants and microbes [124]. As the expression of D. dadantii T3SS genes was induced in planta [125], Yang et al. discovered that two plant phenolic compounds, o-coumaric acid and t-cinnamic acid, positively regulated the expression of T3SS genes in D. dadantii [32]. o-coumaric acid upregulated hrpL at the post-transcriptional level via the RsmA/RsmB system and had no impact on the HrpX/HrpY-HrpS-HrpL pathway. Although o-coumaric acid and t-cinnamic acid are the biosynthetic precursors of salicylic acid, which plays an important role in plant defense responses [126,127], application of salicylic acid did not affect the expression of T3SS in D. dadantii [32].
To identify potential phenolic compounds as T3SS inhibitors, Li and colleagues evaluated the effects of 29 analogs and isomers of o-coumaric acid and t-cinnamic acid on the transcriptional activity of hrpA:gfp fusion in D. dadantii. One compound, p-coumaric acid, was shown to repress the expression of T3SS genes via the HrpS-HrpL pathway, as the addition of 100 µM p-coumaric acid significantly reduced the transcriptional activities and mRNA levels of hrpS and hrpL without affecting the bacterial growth [117]. p-coumaric acid is an intermediate in the phenylpropanoid biosynthesis pathway. Phenylpropanoids are plant secondary metabolites that act as defense molecules in response to microbial attack [128,129]. It is worth noting that the discovery of p-coumaric acid as a T3SS inhibitor has provided fundamental aspects for compound modification and greatly encouraged the innovation of chemically synthesized plant phenolic compound derivatives. A subsequent library screening of 50 p-coumaric acid derivatives identified trans-4-hydroxycinnamohydroxamic acid as a T3SS inhibitor in D. dadantii. The same study found that this synthetic compound increased the inhibitory potency towards T3SS by eightfold compared with the naturally occurring p-coumaric acid [118]. Furthermore, trans-4-hydroxycinnamohydroxamic acid was shown to repress hrpL transcriptionally through both the RpoN-HrpL pathway and the HrpX/HrpY-HrpS-HrpL pathway. Trans-4-hydroxycinnamohydroxamic acid repressed rsmB via unknown mechanisms, negatively contributing to the post-transcriptionally regulation of hrpL [118].
Using similar strategies as Li et al. [117,118], Khokhani and colleagues reported several plant phenolic compounds and derivatives that can modulate T3SS in E. amylovora [115]. o-coumaric acid and t-cinnamic acid, which induced the expression of T3SS in D. dadantii [32], repressed the E. amylovora T3SS. The T3SS inhibitors of E. amylovora also include benzoic acid, salicylic acid, and 4-methoxy-cinnamic acid. Benzoic acid negatively regulated hrpS expression, while 4-methoxy-cinnamic acid repressed HrpL, respectively. HR analysis demonstrated that co-infiltration of E. amylovora with benzoic acid or 4-methoxy-cinnamic acid reduced the HR development in Nicotiana tabacum cv. Xanthi leaves, suggesting that both phenolic compounds are functionally inhibiting T3SS in planta. Earlier studies on phenylpropanoid metabolism in Vanilla planifolia have reported that 4-methoxy-cinnamic acid is one of the intermediates in the biosynthetic conversion of cinnamic acid to benzoic acid [130]. Benzoic acid could be produced via the conversion of t-cinnamic acid to salicylic acid [131].
X. oryzae pv. oryzae is the causal agent of bacterial blight, one of the major rice diseases in the world. X. oryzae pv. oryzicola, which causes the bacterial leaf streak disease, is another rice pathogen [132]. Both pathogens process a T3SS encoded by the group II hrp gene clusters, in which HrpG and HrpX are two key regulators [16,21,30,133]. HrpG is a response regulator of the OmpR family of TCSTS, which positively regulates the expression of hrpX. HrpX is an AraC family transcriptional regulator responsible for activating the transcription of hrp genes [21,133]. By combining the transcriptional screening of the T3SS genes and HR assay in tobacco, Fan et al. identified four plant phenolic compounds, including o-coumaric acid, trans-2-methoxycinnamic acid, trans-2-phenylcyclopropane-1-carboxylic acid, and trans-2-methylcinnamic acid, actively inhibited the T3SS gene expression in vitro likely through the HrpG-HrpX regulatory cascade. Application of these T3SS inhibitors reduced the HR of X. oryzae pv. oryzae in non-host plants and the water soaking and disease symptoms of the pathogen in rice [116]. Since no impact on other virulence factors of X. oryzae pv. oryzae, such as the type II secretion system (T2SS), exopolysaccharide (EPS), and lipopolysaccharide (LPS) [134,135], was observed, the reduced virulence caused by the plant phenolic compounds in X. oryzae pv. oryzae is T3SS-dependent. A recent study was conducted to further understand the response of X. oryzae pv. oryzae to o-coumaric acid using transcriptomic analysis [136].
In R. solanacearum, the bacterial wilt pathogen of tomato, much progress has been made in identifying coumarins as T3SS inhibitors. Coumarins, consisting of fused benzene and α-pyrone rings, are a family of plant-derived secondary metabolites containing a large class of phenolic substances [137]. Members of the coumarins have been extensively studied mainly for their antimicrobial properties [138]. Umbelliferone is a 7-hydroxycoumarin that has recently been reported to repress the expression of hrpG and multiple T3SS regulon genes in R. solanacearum [119,120]. It reduced biofilm formation and suppressed the wilting disease process by reducing the colonization and proliferation of R. solanacearum in planta [120]. Besides, six plant phenolic compounds and derivatives, including p-coumaric acid, that repress the T3SS in D. dadantii, failed to modulate the expression of R. solanacearum T3SS genes [122].

3.2. Salicylidene Acylhydrazides and Their Derivatives as T3SS Inhibitors

Salicylidene acylhydrazides are one of the earliest T3SS inhibitors that were first identified through large-scale chemical screening approaches in Yersinia spp. [139,140,141]. Later studies have shown that salicylidene acylhydrazide family compounds actively against T3SSs from other human pathogenic bacteria, including S. enterica, Shigella, Chlamydia, P. aeruginosa, and enteropathogenic E. coli [142,143,144,145,146,147,148,149,150]. However, whether salicylidene acylhydrazides inhibit the plant pathogenic bacterial T3SSs remained unknown. To address this question, Yang and colleagues screened a small library of compounds. They found that several salicylidene acylhydrazides repressed the expression of T3SS genes, including the master regulator encoding gene hrpL, in E. amylovora [121]. Microarray analysis of E. amylovora treated with compound 3, also known as ME0054 [benzoic acid N′-(2,3,4-trihydroxy-benzylidene)-hydrazide] [141], confirmed its inhibitory impact on T3SS, as the majority of the 38 known T3SS genes was downregulated [121]. In addition, compound 3 suppressed the production of EPS amylovoran, one of the major pathogenicity factors of E. amylovora [7,151], and reduced the disease development of E. amylovora on crab apple flowers [121].
Puigvert and colleagues recently applied a similar approach to the T3SS of R. solanacearum to discover T3SS inhibitors [122]. Three synthetic salicylidene acylhydrazide derivatives, ME0054, ME0055 [4-nitrobenzoic acid N′-(2,4-dihydroxy-benzylidene)-hydrazide], and ME0177 [2-nitro-benzoic acid N′-(3,5-dichloro-2-hydroxy-benzylidene)-hydrazide] [141,152], were shown to inhibit the expression of T3SS genes through the inhibition of the regulator encoding gene hrpB [122]. All three salicylidene acylhydrazides could suppress the multiplication of R. solanacearum in planta and protect tomato from bacterial speck caused by P. syringae pv. tomato [122]. Since Yang et al. reported that ME0054 and ME0055 are T3SS inhibitors in E. amylovora [121], these findings suggest that salicylidene acylhydrazides and their derivatives are effective against different plant pathogenic bacteria.

4. Conclusions and Perspectives

Bacterial plant diseases, such as bacterial wilt, fire blight, soft rot, citrus greening, and bacterial leaf blight of rice, cause significant economic losses ($100M) annually on a global scale. Current control options are limited and involve applying chemicals, copper, and antibiotics, and biological control agents like bacteriophage [113,153]. In the field of fire blight management, for instance, three antibiotics have been used, including streptomycin, oxytetracycline, and a newly registered antibiotic kasugamycin. They are proven to be the most effective method in controlling fire blight in apple and pear orchards in the United States [154,155]. However, the widespread of resistance to antibiotics has placed major constraints on antibiotic usage. An increasing number of studies has implied the preexistence of antibiotic-resistance genes in environmental microbiome, such as soil, ground water, phyllosphere, and animal gut, before introducing antibiotics [114]. In addition, the extensive application of antibiotics in agriculture could greatly endanger human health as antibiotic-resistance genes are able to transfer between bacterial species via horizontal gene transfer [156]. Thus, the discovery and development of alternative control methods for controlling bacterial plant diseases are urgently needed.
T3SS is an essential virulence factor in many gram-negative bacterial pathogens [1,2]. Plant pathogenic bacteria use T3SSs to translocate various effector proteins into host cells to manipulate plant signaling behaviors and repress host immune responses [4]. The innovation of novel molecules that specifically target and inhibit primary virulence factors, such as T3SS, without lethal selective pressure serves as a compelling control option compared with conventional antibiotics. However, unlike antibiotics, often with known modes of action, the mechanisms of newly discovered virulence inhibitors are not well established in bacteria. This could be due to the lack of fundamental knowledge for the regulation of virulence factors and will be facilitated by analyzing the landscape of bacterial responses to virulence inhibitors at the whole-genome transcription level via next-generation sequencing techniques [136].
In the model microorganism D. dadantii, several regulatory components, including transcriptional and post-transcriptional regulators, sRNAs, and bacterial second messengers, have been reported to regulate the T3SS (Figure 1). The majority of these regulators, like c-di-GMP, are genetically and functionally conserved between different bacterial species, implicating that small molecules proven to be effective in one bacterium can also be applied in other bacteria. Several exemplary studies have already proved this concept via studying the impact of plant phenolic compounds and salicylidene acylhydrazides on T3SSs in various animal and plant pathogens [118,121,122]. Both small peptides and molecules (natural or synthetic) have been extensively studied in human pathogens P. aeruginosa and V. cholerae due to their impact on the c-di-GMP signaling. These molecules exert their function via binding to c-di-GMP directly, affecting the enzymatic activity of DGCs that synthesize c-di-GMP, or mimicking c-di-GMP as a competitor in bacteria [157,158,159,160]. However, similar studies have not been reported in plant pathogenic bacteria, and the idea of using small molecules that target c-di-GMP for the management of plant diseases needs to be further evaluated. On the other hand, controlling bacterial diseases using T3SS inhibitors may raise a concern whether they will interfere with the T3SS in some non-pathogenic or host beneficial bacteria that is also important for host interactions [161]. Ecologically, T3SS has been shown to contribute to the interaction between bacteria and fungi in soil and related habitats [162]. Thus, besides the capacity for disease management, the impact of T3SS inhibitors on the host microbiomes and the host plants should be monitored and evaluated.
Another challenge for the T3SS or any other virulence inhibitors is a successful implementation in the field. Unlike laboratory conditions, most commercially grown plants face fluctuating environmental conditions, such as sunlight, precipitation, and temperature. Therefore, it is important to evaluate the stability and efficacy of T3SS inhibitors under various environmental conditions. For example, a recent study showed that oxytetracycline and kasugamycin could be degraded by sunlight and early evening application is suggested to maximize the efficacy in controlling fire blight in the field [163]. In 2014 and 2015, we conducted field assays using a phenolic T3SS inhibitor, trans-4-phenylcinnamic acid, against fire blight on apple trees. Our results showed that it reduced blossom blight with an efficacy similar to kasugamycin at a concentration of 5 mM (Yang and Sundin, unpublished data), suggesting that application of T3SS inhibitor is a promising alternative method for controlling fire blight in the field. Meanwhile, trial application and evaluation of T3SS inhibitors on other bacterial plant diseases are planned. With more new discoveries, these antibiotic alternatives are expected to make agriculture more prepared for the upcoming challenges from continuously evolving pathogens.

Author Contributions

All authors contributed to drafting or revising the manuscript, gave final approval of the version to be published, and agreed to be accountable for all aspects of the work. All authors have read and agreed to the published version of the manuscript.

Funding

This project was supported by grants from the United States Department of Agriculture National Institute of Food and Agriculture (nifa.usda.gov) grant no. 2016-67030-24856, 2017-51106-27001, and 2020-70006-32999, the Research Growth Initiative of the University of Wisconsin-Milwaukee, the UWM Research Foundation (Catalyst Grant in Advanced Automation), and T3 Bioscience LLC.

Conflicts of Interest

This research is partially sponsored by T3 Bioscience LLC (T3 B) and may lead to the development of products which may be licensed to T3 B. In accordance with my ethical obligation as a researcher, I am reporting that I have financial and business interests in T3 B. I have disclosed those interests fully to the publisher and have in place an approved plan for managing any potential conflicts arising from this arrangement with Research Office of University of Wisconsin, Milwaukee and University of Wisconsin, Milwaukee Research Foundation.

References

  1. Deng, W.; Marshall, N.C.; Rowland, J.L.; McCoy, J.M.; Worrall, L.J.; Santos, A.S.; Strynadka, N.C.; Finlay, B.B. Assembly, structure, function and regulation of type III secretion systems. Nat. Rev. Microbiol. 2017, 15, 323. [Google Scholar] [CrossRef] [PubMed]
  2. Hueck, C.J. Type III protein secretion systems in bacterial pathogens of animals and plants. Microbiol. Mol. Biol. Rev. 1998, 62, 379–433. [Google Scholar] [CrossRef] [Green Version]
  3. Coburn, B.; Sekirov, I.; Finlay, B.B. Type III secretion systems and disease. Clin. Microbiol. Rev. 2007, 20, 535–549. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Büttner, D.; He, S.Y. Type III protein secretion in plant pathogenic bacteria. Plant. Physiol. 2009, 150, 1656–1664. [Google Scholar] [CrossRef] [Green Version]
  5. Balint-Kurti, P. The plant hypersensitive response: Concepts, control and consequences. Mol. Plant Pathol. 2019, 20, 1163–1178. [Google Scholar] [CrossRef] [Green Version]
  6. Oh, C.S.; Kim, J.F.; Beer, S.V. The Hrp pathogenicity island of Erwinia amylovora and identification of three novel genes required for systemic infection. Mol. Plant. Pathol. 2005, 6, 125–138. [Google Scholar] [CrossRef]
  7. Oh, C.-S.; Beer, S.V. Molecular genetics of Erwinia amylovora involved in the development of fire blight. FEMS Microbiol. Lett. 2005, 253, 185–192. [Google Scholar] [CrossRef] [Green Version]
  8. Cunnac, S.; Lindeberg, M.; Collmer, A. Pseudomonassyringae type III secretion system effectors: Repertoires in search of functions. Curr. Opin. Microbiol. 2009, 12, 53–60. [Google Scholar] [CrossRef]
  9. Jin, L.; Ham, J.H.; Hage, R.; Zhao, W.; Soto-Hernandez, J.; Lee, S.Y.; Paek, S.-M.; Kim, M.G.; Boone, C.; Coplin, D.L. Direct and indirect targeting of PP2A by conserved bacterial type-III effector proteins. PLOS Pathog. 2016, 12, e1005609. [Google Scholar] [CrossRef] [PubMed]
  10. Xin, X.-F.; Nomura, K.; Ding, X.; Chen, X.; Wang, K.; Aung, K.; Uribe, F.; Rosa, B.; Yao, J.; Chen, J.; et al. Pseudomonassyringae effector avirulence protein E localizes to the host plasma membrane and down-regulates the expression of the nonrace-specific disease resistance1/harpin-induced1-like13 gene required for antibacterial immunity in Arabidopsis. Plant. Physiol. 2015, 169, 793–802. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Duncan, M.C.; Linington, R.G.; Auerbuch, V. Chemical inhibitors of the type three secretion system: Disarming bacterial pathogens. Antimicrob. Agents Chemother. 2012, 56, 5433–5441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Pendergrass, H.A.; May, A.E. Natural product type III secretion system inhibitors. Antibiotics 2019, 8, 162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Gu, L.; Zhou, S.; Zhu, L.; Liang, C.; Chen, X. Small-molecule inhibitors of the type III secretion system. Mol. Cells 2015, 20, 17659–17674. [Google Scholar] [CrossRef] [PubMed]
  14. Arnold, D.L.; Pitman, A.; Jackson, R.W. Pathogenicity and other genomic islands in plant pathogenic bacteria. Mol. Plant Pathol. 2003, 4, 407–420. [Google Scholar] [CrossRef] [PubMed]
  15. Troisfontaines, P.; Cornelis, G.R. Type III secretion: More systems than you think. Physiology 2005, 20, 326–339. [Google Scholar] [CrossRef] [PubMed]
  16. Alfano, J.R.; Collmer, A. The type III (Hrp) secretion pathway of plant pathogenic bacteria: Trafficking harpins, Avr proteins, and death. J. Bacteriol. 1997, 179, 5655. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Wei, Z.-M.; Beer, S.V. hrpL activates Erwinia amylovora hrp gene transcription and is a member of the ECF subfamily of sigma factors. J. Bacteriol. 1995, 177, 6201–6210. [Google Scholar] [CrossRef] [Green Version]
  18. Xiao, Y.; Hutcheson, S.W. A single promoter sequence recognized by a newly identified alternate sigma factor directs expression of pathogenicity and host range determinants in Pseudomonas syringae. J. Bacteriol. 1994, 176, 3089–3091. [Google Scholar] [CrossRef] [Green Version]
  19. Xiao, Y.; Heu, S.; Yi, J.; Lu, Y.; Hutcheson, S. Identification of a putative alternate sigma factor and characterization of a multicomponent regulatory cascade controlling the expression of Pseudomonas syringae pv. syringae Pss61 hrp and hrmA genes. J. Bacteriol. 1994, 176, 1025–1036. [Google Scholar]
  20. Brito, B.; Aldon, D.; Barberis, P.; Boucher, C.; Genin, S. A signal transfer system through three compartments transduces the plant cell contact-dependent signal controlling Ralstonia solanacearum hrp genes. Mol. Plant-Microbe Interact. 2002, 15, 109–119. [Google Scholar] [CrossRef] [Green Version]
  21. Wengelnik, K.; Bonas, U. HrpXv, an AraC-type regulator, activates expression of five of the six loci in the hrp cluster of Xanthomonas campestris pv. vesicatoria. J. Bacteriol. 1996, 178, 3462–3469. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. McNally, R.R.; Zhao, Y.; Sundin, G.W. Towards understanding fire blight: Virulence mechanisms and their regulation in Erwinia amylovora. In Bacteria-Plant Interactions: Advanced Research and Future Trends; Caister Academy Press: Norfolk, UK, 2015; pp. 61–82. [Google Scholar]
  23. Zhao, Y.; Sundin, G. Exploring linear and cyclic (di)-nucleotides as messengers for regulation of T3SS and biofilm formation in Erwinia amylovora. J. Plant. Pathol. 2017, 99, 25–35. [Google Scholar]
  24. Zhao, Y. Genomics of Erwinia amylovora and related Erwinia species associated with pome fruit trees. In Genomics of Plant-Associated Bacteria; Springer: Berlin/Heidelberg, Germany, 2014; pp. 1–36. [Google Scholar]
  25. Malnoy, M.; Martens, S.; Norelli, J.L.; Barny, M.-A.; Sundin, G.W.; Smits, T.H.; Duffy, B. Fire blight: Applied genomic insights of the pathogen and host. Annu. Rev. Phytopathol. 2012, 50, 475–494. [Google Scholar] [CrossRef] [Green Version]
  26. Xie, Y.; Shao, X.; Deng, X. Regulation of type III secretion system in Pseudomonas syringae. Environ. Microbiol. 2019, 21, 4465–4477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Lam, H.N.; Chakravarthy, S.; Wei, H.-L.; BuiNguyen, H.; Stodghill, P.V.; Collmer, A.; Swingle, B.M.; Cartinhour, S.W. Global analysis of the HrpL regulon in the plant pathogen Pseudomonas syringae pv. tomato DC3000 reveals new regulon members with diverse functions. PLoS ONE 2014, 9, e106115. [Google Scholar] [CrossRef] [Green Version]
  28. Yang, S.; Peng, Q.; Zhang, Q.; Zou, L.; Li, Y.; Robert, C.; Pritchard, L.; Liu, H.; Hovey, R.; Wang, Q.; et al. Genome-Wide Identification of HrpL-Regulated Genes in the Necrotrophic Phytopathogen Dickeya dadantii 3937. PLoS ONE 2010, 5, e13472. [Google Scholar] [CrossRef]
  29. Yap, M.-N.; Yang, C.-H.; Barak, J.D.; Jahn, C.E.; Charkowski, A.O. The Erwinia chrysanthemi type III secretion system is required for multicellular behavior. J. Bacteriol. 2005, 187, 639–648. [Google Scholar] [CrossRef] [Green Version]
  30. Tang, X.; Xiao, Y.; Zhou, J.-M. Regulation of the type III secretion system in phytopathogenic bacteria. Mol. Plant.-Microbe Interact. 2006, 19, 1159–1166. [Google Scholar] [CrossRef] [Green Version]
  31. Yang, S.; Peng, Q.; Zhang, Q.; Yi, X.; Choi, C.J.; Reedy, R.M.; Charkowski, A.O.; Yang, C.-H. Dynamic regulation of GacA in type III secretion, pectinase gene expression, pellicle formation, and pathogenicity of Dickeya dadantii (Erwinia chrysanthemi 3937). Mol. Plant.-Microbe Interact. 2008, 21, 133–142. [Google Scholar] [CrossRef] [Green Version]
  32. Yang, S.; Peng, Q.; San Francisco, M.; Wang, Y.; Zeng, Q.; Yang, C.-H. Type III secretion system genes of Dickeya dadantii 3937 are induced by plant phenolic acids. PLoS ONE 2008, 3, e2973. [Google Scholar] [CrossRef]
  33. Chatterjee, A.; Cui, Y.; Chatterjee, A.K. Regulation of Erwinia carotovora hrpLEcc (sigma-LEcc), Which Encodes an Extracytoplasmic Function Subfamily of Sigma Factor Required for Expression of the HRP Regulon. Mol. Plant.-Microbe Interact. 2002, 15, 971–980. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Chatterjee, A.; Cui, Y.; Liu, Y.; Dumenyo, C.K.; Chatterjee, A.K. Inactivation of rsmA leads to overproduction of extracellular pectinases, cellulases, and proteases in Erwinia carotovora subsp. carotovora in the absence of the starvation/cell density-sensing signal, N-(3-oxohexanoyl)-L-homoserine lactone. Appl. Environ. Microbiol. 1995, 61, 1959–1967. [Google Scholar] [CrossRef] [Green Version]
  35. Liu, M.Y.; Gui, G.; Wei, B.; Preston, J.F.; Oakford, L.; Yüksel, Ü.; Giedroc, D.P.; Romeo, T. The RNA molecule CsrB binds to the global regulatory protein CsrA and antagonizes its activity in Escherichia coli. J. Biol. Chem. 1997, 272, 17502–17510. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Czajkowski, R.; Perombelon, M.C.; van Veen, J.A.; van der Wolf, J.M. Control of blackleg and tuber soft rot of potato caused by Pectobacterium and Dickeya species: A review. Plant. Pathol. 2011, 60, 999–1013. [Google Scholar] [CrossRef]
  37. Reverchon, S.; Nasser, W. Dickeya ecology, environment sensing and regulation of virulence programme. Environ. Microbiol. Rep. 2013, 5, 622–636. [Google Scholar] [PubMed]
  38. Römling, U.; Galperin, M.Y.; Gomelsky, M. Cyclic di-GMP: The first 25 years of a universal bacterial second messenger. Microbiol. Mol. Biol. Rev. 2013, 77, 1–52. [Google Scholar] [CrossRef] [Green Version]
  39. Hengge, R. Principles of c-di-GMP signalling in bacteria. Nat. Rev. Microbiol. 2009, 7, 263–273. [Google Scholar] [CrossRef]
  40. Paul, R.; Weiser, S.; Amiot, N.C.; Chan, C.; Schirmer, T.; Giese, B.; Jenal, U. Cell cycle-dependent dynamic localization of a bacterial response regulator with a novel di-guanylate cyclase output domain. Genes Dev. 2004, 18, 715–727. [Google Scholar] [CrossRef] [Green Version]
  41. Whiteley, C.G.; Lee, D.-J. Bacterial diguanylate cyclases: Structure, function and mechanism in exopolysaccharide biofilm development. Biotechnol. Adv. 2015, 33, 124–141. [Google Scholar] [CrossRef]
  42. Ryan, R.P.; Fouhy, Y.; Lucey, J.F.; Crossman, L.C.; Spiro, S.; He, Y.-W.; Zhang, L.-H.; Heeb, S.; Cámara, M.; Williams, P.; et al. Cell-cell signaling in Xanthomonas campestris involves an HD-GYP domain protein that functions in cyclic di-GMP turnover. Proc. Natl. Acad. Sci. USA 2006, 103, 6712–6717. [Google Scholar] [CrossRef] [Green Version]
  43. Schmidt, A.J.; Ryjenkov, D.A.; Gomelsky, M. The ubiquitous protein domain EAL is a cyclic diguanylate-specific phosphodiesterase: Enzymatically active and inactive EAL domains. J. Bacteriol. 2005, 187, 4774–4781. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Tamayo, R.; Tischler, A.D.; Camilli, A. The EAL domain protein VieA is a cyclic diguanylate phosphodiesterase. J. Biol. Chem. 2005, 280, 33324–33330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Povolotsky, T.L.; Hengge, R. ‘Life-style’control networks in Escherichia coli: Signaling by the second messenger c-di-GMP. J. Biotechnol. 2012, 160, 10–16. [Google Scholar] [CrossRef] [PubMed]
  46. Waters, C.M.; Lu, W.; Rabinowitz, J.D.; Bassler, B.L. Quorum sensing controls biofilm formation in Vibrio cholerae through modulation of cyclic di-GMP levels and repression of vpsT. J. Bacteriol. 2008, 190, 2527–2536. [Google Scholar] [CrossRef] [Green Version]
  47. Edmunds, A.C.; Castiblanco, L.F.; Sundin, G.W.; Waters, C.M. Cyclic Di-GMP modulates the disease progression of Erwinia amylovora. J. Bacteriol. 2013, 195, 2155–2165. [Google Scholar] [CrossRef] [Green Version]
  48. Kharadi, R.R.; Castiblanco, L.F.; Waters, C.M.; Sundin, G.W. Phosphodiesterase Genes Regulate Amylovoran Production, Biofilm Formation, and Virulence in Erwinia amylovora. Appl. Environ. Microbiol. 2019, 85. [Google Scholar] [CrossRef] [Green Version]
  49. Hugouvieux-Cotte-Pattat, N.; Condemine, G.; Shevchik, V.E. Bacterial pectate lyases, structural and functional diversity. Environ. Microbiol. Rep. 2014, 6, 427–440. [Google Scholar] [CrossRef]
  50. Yi, X.; Yamazaki, A.; Biddle, E.; Zeng, Q.; Yang, C.H. Genetic analysis of two phosphodiesterases reveals cyclic diguanylate regulation of virulence factors in Dickeya dadantii. Mol. Microbiol. 2010, 77, 787–800. [Google Scholar] [CrossRef]
  51. Yuan, X.; Khokhani, D.; Wu, X.; Yang, F.; Biener, G.; Koestler, B.J.; Raicu, V.; He, C.; Waters, C.M.; Sundin, G.W.; et al. Cross-talk between a regulatory small RNA, cyclic-di-GMP signalling and flagellar regulator FlhDC for virulence and bacterial behaviours. Environ. Microbiol. 2015, 17, 4745–4763. [Google Scholar] [CrossRef]
  52. Jenal, U.; Reinders, A.; Lori, C. Cyclic di-GMP: Second messenger extraordinaire. Nat. Rev. Microbiol. 2017, 15, 271–284. [Google Scholar] [CrossRef] [Green Version]
  53. Yuan, X.; Tian, F.; He, C.; Severin, G.B.; Waters, C.M.; Zeng, Q.; Liu, F.; Yang, C.H. The diguanylate cyclase GcpA inhibits the production of pectate lyases via the H-NS protein and RsmB regulatory RNA in Dickeya dadantii. Mol. Plant. Pathol. 2018, 19, 1873–1886. [Google Scholar] [CrossRef] [Green Version]
  54. Nasser, W.; Faelen, M.; Hugouvieux-Cotte-Pattat, N.; Reverchon, S. Role of the nucleoid-associated protein H-NS in the synthesis of virulence factors in the phytopathogenic bacterium Erwinia chrysanthemi. Mol. Plant.-Microbe Interact. 2001, 14, 10–20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Nasser, W.; Reverchon, S. H-NS-dependent activation of pectate lyases synthesis in the phytopathogenic bacterium Erwinia chrysanthemi is mediated by the PecT repressor. Mol. Microbiol. 2002, 43, 733–748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Ouafa, Z.-A.; Reverchon, S.; Lautier, T.; Muskhelishvili, G.; Nasser, W. The nucleoid-associated proteins H-NS and FIS modulate the DNA supercoiling response of the pel genes, the major virulence factors in the plant pathogen bacterium Dickeya dadantii. Nucleic Acids Res. 2012, 40, 4306–4319. [Google Scholar] [CrossRef] [PubMed]
  57. Stauber, J.L.; Loginicheva, E.; Schechter, L.M. Carbon source and cell density-dependent regulation of type III secretion system gene expression in Pseudomonas syringae pathovar tomato DC3000. Res. Microbiol. 2012, 163, 531–539. [Google Scholar] [CrossRef] [PubMed]
  58. Alfano, J.R.; Kim, H.-S.; Delaney, T.P.; Collmer, A. Evidence that the Pseudomonas syringae pv. syringae hrp-linked hrmA gene encodes an Avr-like protein that acts in an hrp-dependent manner within tobacco cells. Mol. Plant-Microbe Interact. 1997, 10, 580–588. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Nasser, W.; Reverchon, S.; Vedel, R.; Boccara, M. PecS and PecT coregulate the synthesis of HrpN and pectate lyases, two virulence determinants in Erwinia chrysanthemi 3937. Mol. Plant.-Microbe Interact. 2005, 18, 1205–1214. [Google Scholar] [CrossRef]
  60. Yuan, X.; Zeng, Q.; Xu, J.; Severin, G.B.; Zhou, X.; Waters, C.M.; Sundin, G.W.; Ibekwe, A.M.; Liu, F.; Yang, C.-H. Tricarboxylic Acid (TCA) Cycle Enzymes and Intermediates Modulate Intracellular Cyclic di-GMP Levels and the Production of Plant Cell Wall–Degrading Enzymes in Soft Rot Pathogen Dickeya dadantii. Mol. Plant-Microbe Interact. Mpmi 2020, 33, 296–307. [Google Scholar] [CrossRef]
  61. Giacalone, D.; Smith, T.J.; Collins, A.J.; Sondermann, H.; Koziol, L.J.; O’Toole, G.A. Ligand-Mediated Biofilm Formation via Enhanced Physical Interaction between a Diguanylate Cyclase and Its Receptor. mBio 2018, 9, e01254-18. [Google Scholar] [CrossRef] [Green Version]
  62. Ross, W.; Vrentas, C.E.; Sanchez-Vazquez, P.; Gaal, T.; Gourse, R.L. The magic spot: A ppGpp binding site on E. coli RNA polymerase responsible for regulation of transcription initiation. Mol. Cell 2013, 50, 420–429. [Google Scholar] [CrossRef] [Green Version]
  63. Hauryliuk, V.; Atkinson, G.C.; Murakami, K.S.; Tenson, T.; Gerdes, K. Recent functional insights into the role of (p)ppGpp in bacterial physiology. Nat. Rev. Microbiol. 2015, 13, 298–309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Yang, H.-w.; Yu, M.; Lee, J.H.; Chatnaparat, T.; Zhao, Y. The stringent response regulator (p)ppGpp mediates virulence gene expression and survival in Erwinia amylovora. BMC Genom. 2020, 21, 1–19. [Google Scholar] [CrossRef] [PubMed]
  65. Ancona, V.; Lee, J.H.; Chatnaparat, T.; Oh, J.; Hong, J.-I.; Zhao, Y. The bacterial alarmone (p)ppGpp activates the type III secretion system in Erwinia amylovora. J. Bacteriol. 2015, 197, 1433–1443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Chatnaparat, T.; Li, Z.; Korban, S.S.; Zhao, Y. The bacterial alarmone (p)ppGpp is required for virulence and controls cell size and survival of Pseudomonas syringae on plants. Environ. Microbiol. 2015, 17, 4253–4270. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Chatnaparat, T.; Li, Z.; Korban, S.S.; Zhao, Y. The stringent response mediated by (p)ppGpp is required for virulence of Pseudomonas syringae pv. tomato and its survival on tomato. Mol. Plant.-Microbe Interact. 2015, 28, 776–789. [Google Scholar] [CrossRef] [Green Version]
  68. Young, G.M.; Schmiel, D.H.; Miller, V.L. A new pathway for the secretion of virulence factors by bacteria: The flagellar export apparatus functions as a protein-secretion system. Proc. Natl. Acad. Sci. USA 1999, 96, 6456–6461. [Google Scholar] [CrossRef] [Green Version]
  69. Lee, S.H.; Galán, J.E. Salmonella type III secretion-associated chaperones confer secretion-pathway specificity. Mol. Microbiol. 2004, 51, 483–495. [Google Scholar] [CrossRef] [Green Version]
  70. Pallen, M.J.; Beatson, S.A.; Bailey, C.M. Bioinformatics, genomics and evolution of non-flagellar type-III secretion systems: A Darwinian perpective. FEMS Microbiol. Rev. 2005, 29, 201–229. [Google Scholar] [CrossRef] [Green Version]
  71. Erhardt, M.; Namba, K.; Hughes, K.T. Bacterial nanomachines: The flagellum and type III injectisome. Cold Spring Harb. Perspect. Biol. 2010, 2, a000299. [Google Scholar] [CrossRef] [Green Version]
  72. Wei, H.L.; Chakravarthy, S.; Worley, J.N.; Collmer, A. Consequences of flagellin export through the type III secretion system of Pseudomonas syringae reveal a major difference in the innate immune systems of mammals and the model plant Nicotiana benthamiana. Cell. Microbiol. 2013, 15, 601–618. [Google Scholar] [CrossRef]
  73. Wang, S.; Fleming, R.T.; Westbrook, E.M.; Matsumura, P.; McKay, D.B. Structure of the Escherichia coli FlhDC complex, a prokaryotic heteromeric regulator of transcription. J. Mol. Biol. 2006, 355, 798–808. [Google Scholar] [CrossRef] [PubMed]
  74. Aldridge, P.D.; Karlinsey, J.E.; Aldridge, C.; Birchall, C.; Thompson, D.; Yagasaki, J.; Hughes, K.T. The flagellar-specific transcription factor, σ28, is the type III secretion chaperone for the flagellar-specific anti-σ28 factor FlgM. Genes Dev. 2006, 20, 2315–2326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Chilcott, G.S.; Hughes, K.T. Coupling of Flagellar Gene Expression to Flagellar Assembly in Salmonella enterica Serovar Typhimurium and Escherichia coli. Microbiol. Mol. Biol. Rev. 2000, 64, 694–708. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Cui, Y.; Chatterjee, A.; Yang, H.; Chatterjee, A.K. Regulatory network controlling extracellular proteins in Erwinia carotovora subsp. carotovora: FlhDC, the master regulator of flagellar genes, activates rsmB regulatory RNA production by affecting gacA and hexA (lrhA) expression. J. Bacteriol. 2008, 190, 4610–4623. [Google Scholar] [CrossRef] [Green Version]
  77. Pesavento, C.; Becker, G.; Sommerfeldt, N.; Possling, A.; Tschowri, N.; Mehlis, A.; Hengge, R. Inverse regulatory coordination of motility and curli-mediated adhesion in Escherichia coli. Genes Dev. 2008, 22, 2434–2446. [Google Scholar] [CrossRef] [Green Version]
  78. Claret, L.; Hughes, C. Interaction of the atypical prokaryotic transcription activator FlhD2C2 with early promoters of the flagellar gene hierarchy. J. Mol. Biol. 2002, 321, 185–199. [Google Scholar] [CrossRef]
  79. Stafford, G.P.; Ogi, T.; Hughes, C. Binding and transcriptional activation of non-flagellar genes by the Escherichia coli flagellar master regulator FlhD2C2. Microbiology 2005, 151, 1779. [Google Scholar] [CrossRef] [Green Version]
  80. Fang, X.; Gomelsky, M. A post-translational, c-di-GMP-dependent mechanism regulating flagellar motility. Mol. Microbiol. 2010, 76, 1295–1305. [Google Scholar] [CrossRef]
  81. Paul, K.; Nieto, V.; Carlquist, W.C.; Blair, D.F.; Harshey, R.M. The c-di-GMP binding protein YcgR controls flagellar motor direction and speed to affect chemotaxis by a “backstop brake” mechanism. Mol. Cell 2010, 38, 128–139. [Google Scholar] [CrossRef] [Green Version]
  82. Battesti, A.; Majdalani, N.; Gottesman, S. The RpoS-mediated general stress response in Escherichia coli. Annu. Rev. Microbiol. 2011, 65, 189–213. [Google Scholar] [CrossRef] [Green Version]
  83. Lange, R.; Hengge-Aronis, R. The cellular concentration of the sigma S subunit of RNA polymerase in Escherichia coli is controlled at the levels of transcription, translation, and protein stability. Genes Dev. 1994, 8, 1600–1612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Zhou, Y.; Gottesman, S.; Hoskins, J.R.; Maurizi, M.R.; Wickner, S. The RssB response regulator directly targets ςS for degradation by ClpXP. Genes Dev. 2001, 15, 627–637. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Schweder, T.; Lee, K.-H.; Lomovskaya, O.; Matin, A. Regulation of Escherichia coli starvation sigma factor (sigma s) by ClpXP protease. J. Bacteriol. 1996, 178, 470–476. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Mukherjee, A.; Cui, Y.; Ma, W.; Liu, Y.; Ishihama, A.; Eisenstark, A.; Chatterjee, A.K. RpoS (sigma-S) controls expression of rsmA, a global regulator of secondary metabolites, harpin, and extracellular proteins in Erwinia carotovora. J. Bacteriol. 1998, 180, 3629–3634. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Li, Y.; Yamazaki, A.; Zou, L.; Biddle, E.; Zeng, Q.; Wang, Y.; Lin, H.; Wang, Q.; Yang, C.-H. ClpXP protease regulates the type III secretion system of Dickeya dadantii 3937 and is essential for the bacterial virulence. Mol. Plant.-Microbe Interact. 2010, 23, 871–878. [Google Scholar] [CrossRef] [Green Version]
  88. Lee, J.H.; Zhao, Y. ClpXP-Dependent RpoS Degradation Enables Full Activation of Type III Secretion System, Amylovoran Production, and Motility in Erwinia amylovora. Phytopathology 2017, 107, 1346–1352. [Google Scholar] [CrossRef] [Green Version]
  89. Ellison, D.W.; Miller, V.L. Regulation of virulence by members of the MarR/SlyA family. Curr. Opin. Microbiol. 2006, 9, 153–159. [Google Scholar] [CrossRef]
  90. Haque, M.M.; Kabir, M.S.; Aini, L.Q.; Hirata, H.; Tsuyumu, S. SlyA, a MarR family transcriptional regulator, is essential for virulence in Dickeya dadantii 3937. J. Bacteriol. 2009, 191, 5409–5418. [Google Scholar] [CrossRef] [Green Version]
  91. Linehan, S.A.; Rytkönen, A.; Yu, X.-J.; Liu, M.; Holden, D.W. SlyA regulates function of Salmonella pathogenicity island 2 (SPI-2) and expression of SPI-2-associated genes. Infect. Immun. 2005, 73, 4354–4362. [Google Scholar] [CrossRef] [Green Version]
  92. Zou, L.; Zeng, Q.; Lin, H.; Gyaneshwar, P.; Chen, G.; Yang, C.-H. SlyA regulates type III secretion system (T3SS) genes in parallel with the T3SS master regulator HrpL in Dickeya dadantii 3937. Appl. Environ. Microbiol. 2012, 78, 2888–2895. [Google Scholar] [CrossRef] [Green Version]
  93. Hommais, F.; Oger-Desfeux, C.; Van Gijsegem, F.; Castang, S.; Ligori, S.; Expert, D.; Nasser, W.; Reverchon, S. PecS is a global regulator of the symptomatic phase in the phytopathogenic bacterium Erwinia chrysanthemi 3937. J. Bacteriol. 2008, 190, 7508–7522. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Reverchon, S.; Rouanet, C.; Expert, D.; Nasser, W. Characterization of indigoidine biosynthetic genes in Erwinia chrysanthemi and role of this blue pigment in pathogenicity. J. Bacteriol. 2002, 184, 654–665. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Romeo, T.; Vakulskas, C.A.; Babitzke, P. Post-transcriptional regulation on a global scale: Form and function of Csr/Rsm systems. Environ. Microbiol. 2013, 15, 313–324. [Google Scholar] [CrossRef] [PubMed]
  96. Van Assche, E.; Van Puyvelde, S.; Vanderleyden, J.; Steenackers, H.P. RNA-binding proteins involved in post-transcriptional regulation in bacteria. Front. Microbiol. 2015, 6, 141. [Google Scholar] [CrossRef] [Green Version]
  97. Kudla, J.; Hayes, R.; Gruissem, W. Polyadenylation accelerates degradation of chloroplast mRNA. Embo J. 1996, 15, 7137–7146. [Google Scholar] [CrossRef]
  98. Leszczyniecka, M.; Kang, D.-c.; Sarkar, D.; Su, Z.-z.; Holmes, M.; Valerie, K.; Fisher, P.B. Identification and cloning of human polynucleotide phosphorylase, hPNPaseold-35, in the context of terminal differentiation and cellular senescence. Proc. Natl. Acad. Sci. USA 2002, 99, 16636–16641. [Google Scholar] [CrossRef] [Green Version]
  99. Kinscherf, T.; Apirion, D. Polynucleotide phosphorylase can participate in decay of mRNA in Escherichia coli in the absence of ribonuclease II. Mol. Gen. Genet. MGG. 1975, 139, 357–362. [Google Scholar] [CrossRef]
  100. Cameron, T.A.; Matz, L.M.; De Lay, N.R. Polynucleotide phosphorylase: Not merely an RNase but a pivotal post-transcriptional regulator. PLoS Genet. 2018, 14, e1007654. [Google Scholar] [CrossRef]
  101. Rosenzweig, J.A.; Chromy, B.; Echeverry, A.; Yang, J.; Adkins, B.; Plano, G.V.; McCutchen-Maloney, S.; Schesser, K. Polynucleotide phosphorylase independently controls virulence factor expression levels and export in Yersinia spp. FEMS Microbiol. Lett. 2007, 270, 255–264. [Google Scholar] [CrossRef] [Green Version]
  102. Ygberg, S.E.; Clements, M.O.; Rytkönen, A.; Thompson, A.; Holden, D.W.; Hinton, J.C.; Rhen, M. Polynucleotide phosphorylase negatively controls spv virulence gene expression in Salmonella enterica. Infect. Immun. 2006, 74, 1243–1254. [Google Scholar] [CrossRef] [Green Version]
  103. Zeng, Q.; Ibekwe, A.M.; Biddle, E.; Yang, C.-H. Regulatory mechanisms of exoribonuclease PNPase and regulatory small RNA on T3SS of Dickeya dadantii. Mol. Plant.-Microbe Interact. 2010, 23, 1345–1355. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Viegas, S.C.; Pfeiffer, V.; Sittka, A.; Silva, I.s.J.; Vogel, J.; Arraiano, C.M. Characterization of the role of ribonucleases in Salmonella small RNA decay. Nucleic Acids Res. 2007, 35, 7651–7664. [Google Scholar] [CrossRef] [Green Version]
  105. Dressaire, C.; Pobre, V.; Laguerre, S.; Girbal, L.; Arraiano, C.M.; Cocaign-Bousquet, M. PNPase is involved in the coordination of mRNA degradation and expression in stationary phase cells of Escherichia coli. BMC Genom. 2018, 19, 848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Chao, Y.; Vogel, J. The role of Hfq in bacterial pathogens. Curr. Opin. Microbiol. 2010, 13, 24–33. [Google Scholar] [CrossRef] [PubMed]
  107. Vogel, J.; Luisi, B.F. Hfq and its constellation of RNA. Nat. Rev. Microbiol. 2011, 9, 578–589. [Google Scholar] [CrossRef] [Green Version]
  108. Yuan, X.; Zeng, Q.; Khokhani, D.; Tian, F.; Severin, G.B.; Waters, C.M.; Xu, J.; Zhou, X.; Sundin, G.W.; Ibekwe, A.M.; et al. A Feed-forward signaling circuit controls bacterial virulence through linking cyclic di-GMP and two mechanistically distinct sRNAs; ArcZ and RsmB. Environ. Microbiol. 2019, 21, 2755–2771. [Google Scholar] [CrossRef]
  109. Castillo, A.; Reverchon, S. Characterization of the pecT control region from Erwinia chrysanthemi 3937. J. Bacteriol. 1997, 179, 4909–4918. [Google Scholar] [CrossRef] [Green Version]
  110. Mukherjee, A.; Cui, Y.; Ma, W.; Liu, Y.; Chatterjee, A.K. hexA of Erwinia carotovora ssp. carotovora strain Ecc71 negatively regulates production of RpoS and rsmB RNA, a global regulator of extracellular proteins, plant virulence and the quorum-sensing signal, N-(3-oxohexanoyl)-L-homoserine lactone. Environ. Microbiol. 2000, 2, 203–215. [Google Scholar] [CrossRef]
  111. Zeng, Q.; Sundin, G.W. Genome-wide identification of Hfq-regulated small RNAs in the fire blight pathogen Erwinia amylovora discovered small RNAs with virulence regulatory function. BMC Genom. 2014, 15, 414. [Google Scholar] [CrossRef] [Green Version]
  112. Zeng, Q.; McNally, R.R.; Sundin, G.W. Global small RNA chaperone Hfq and regulatory small RNAs are important virulence regulators in Erwinia amylovora. J. Bacteriol. 2013, 195, 1706–1717. [Google Scholar] [CrossRef] [Green Version]
  113. Sundin, G.W.; Castiblanco, L.F.; Yuan, X.; Zeng, Q.; Yang, C.H. Bacterial disease management: Challenges, experience, innovation and future prospects: Challenges in bacterial molecular plant pathology. Mol. Plant. Pathol. 2016, 17, 1506–1518. [Google Scholar] [CrossRef]
  114. Sundin, G.W.; Wang, N. Antibiotic Resistance in Plant-Pathogenic Bacteria. Annu. Rev. Phytopathol. 2018, 56, 161–180. [Google Scholar] [CrossRef] [PubMed]
  115. Khokhani, D.; Zhang, C.; Li, Y.; Wang, Q.; Zeng, Q.; Yamazaki, A.; Hutchins, W.; Zhou, S.-S.; Chen, X.; Yang, C.-H. Discovery of plant phenolic compounds that act as type III secretion system inhibitors or inducers of the fire blight pathogen, Erwinia amylovora. Appl. Environ. Microbiol. 2013, 79, 5424–5436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Fan, S.; Tian, F.; Li, J.; Hutchins, W.; Chen, H.; Yang, F.; Yuan, X.; Cui, Z.; Yang, C.H.; He, C. Identification of phenolic compounds that suppress the virulence of Xanthomonas oryzae on rice via the type III secretion system. Mol. Plant Pathol. 2017, 18, 555–568. [Google Scholar] [CrossRef] [PubMed]
  117. Li, Y.; Peng, Q.; Selimi, D.; Wang, Q.; Charkowski, A.O.; Chen, X.; Yang, C.-H. The plant phenolic compound p-coumaric acid represses gene expression in the Dickeya dadantii type III secretion system. Appl. Environ. Microbiol. 2009, 75, 1223–1228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Li, Y.; Hutchins, W.; Wu, X.; Liang, C.; Zhang, C.; Yuan, X.; Khokhani, D.; Chen, X.; Che, Y.; Wang, Q. Derivative of plant phenolic compound inhibits the type III secretion system of Dickeya dadantii via HrpX/HrpY two-component signal transduction and Rsm systems. Mol. Plant. Pathol. 2015, 16, 150–163. [Google Scholar] [CrossRef]
  119. Yang, L.; Ding, W.; Xu, Y.; Wu, D.; Li, S.; Chen, J.; Guo, B. New insights into the antibacterial activity of hydroxycoumarins against Ralstonia solanacearum. Mol. Cells 2016, 21, 468. [Google Scholar] [CrossRef]
  120. Yang, L.; Li, S.; Qin, X.; Jiang, G.; Chen, J.; Li, B.; Yao, X.; Liang, P.; Zhang, Y.; Ding, W. Exposure to umbelliferone reduces Ralstonia solanacearum biofilm formation, transcription of type III secretion system regulators and effectors and virulence on tobacco. Front. Microbiol. 2017, 8, 1234. [Google Scholar] [CrossRef]
  121. Yang, F.; Korban, S.S.; Pusey, P.L.; Elofsson, M.; Sundin, G.W.; Zhao, Y. Small-molecule inhibitors suppress the expression of both type III secretion and amylovoran biosynthesis genes in Erwinia amylovora. Mol. Plant Pathol. 2014, 15, 44–57. [Google Scholar] [CrossRef]
  122. Puigvert, M.; Solé, M.; López-Garcia, B.; Coll, N.S.; Beattie, K.D.; Davis, R.A.; Elofsson, M.; Valls, M. Type III secretion inhibitors for the management of bacterial plant diseases. Mol. Plant Pathol. 2019, 20, 20–32. [Google Scholar] [CrossRef] [Green Version]
  123. Velderrain-Rodríguez, G.; Palafox-Carlos, H.; Wall-Medrano, A.; Ayala-Zavala, J.; Chen, C.O.; Robles-Sánchez, M.; Astiazaran-García, H.; Alvarez-Parrilla, E.; González-Aguilar, G. Phenolic compounds: Their journey after intake. Food Funct. 2014, 5, 189–197. [Google Scholar] [CrossRef] [PubMed]
  124. Siqueira, J.O.; Nair, M.G.; Hammerschmidt, R.; Safir, G.R.; Putnam, A.R. Significance of phenolic compounds in plant-soil-microbial systems. Crit. Rev. Plant. Sci. 1991, 10, 63–121. [Google Scholar] [CrossRef]
  125. Yang, S.; Perna, N.T.; Cooksey, D.A.; Okinaka, Y.; Lindow, S.E.; Ibekwe, A.M.; Keen, N.T.; Yang, C.-H. Genome-wide identification of plant-upregulated genes of Erwinia chrysanthemi 3937 using a GFP-based IVET leaf array. Mol. Plant-Microbe Interact. 2004, 17, 999–1008. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Montesano, M.; Brader, G.; Ponce de León, I.; Palva, E.T. Multiple defence signals induced by Erwinia carotovora ssp. carotovora elicitors in potato. Mol. Plant Pathol. 2005, 6, 541–549. [Google Scholar] [CrossRef]
  127. Vidal, S.; de León, I.P.; Denecke, J.; Palva, E.T. Salicylic acid and the plant pathogen Erwinia carotovora induce defense genes via antagonistic pathways. Plant. J. 1997, 11, 115–123. [Google Scholar] [CrossRef]
  128. Dixon, R.A.; Paiva, N.L. Stress-induced phenylpropanoid metabolism. Plant Cell 1995, 7, 1085. [Google Scholar] [CrossRef]
  129. Feys, B.J.; Parker, J.E. Interplay of signaling pathways in plant disease resistance. Trends Genet. 2000, 16, 449–455. [Google Scholar] [CrossRef]
  130. Funk, C.; Brodelius, P.E. Phenylpropanoid metabolism in suspension cultures of Vanilla planifolia Andr.: III. Conversion of 4-methoxycinnamic acids into 4-hydroxybenzoic acids. Plant. Physiol. 1990, 94, 102–108. [Google Scholar] [CrossRef] [Green Version]
  131. Yalpani, N.; Raskin, I. Salicylic acid: A systemic signal in induced plant disease resistance. Trends Microbiol. 1993, 1, 88–92. [Google Scholar] [CrossRef]
  132. Niño-Liu, D.O.; Ronald, P.C.; Bogdanove, A.J. Xanthomonas oryzae pathovars: Model pathogens of a model crop. Mol. Plant Pathol. 2006, 7, 303–324. [Google Scholar] [CrossRef]
  133. Wengelnik, K.; Van den Ackerveken, G.; Bonas, U. HrpG, a key hrp regulatory protein of Xanthomonas campestris pv. vesicatoria ls homologous to two-component response regulators. Mol. Plant-Microbe Interact. 1996, 9, 704–712. [Google Scholar] [CrossRef] [PubMed]
  134. Das, A.; Rangaraj, N.; Sonti, R.V. Multiple adhesin-like functions of Xanthomonas oryzae pv. oryzae are involved in promoting leaf attachment, entry, and virulence on rice. Mol. Plant-Microbe Interact. 2009, 22, 73–85. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Büttner, D.; Bonas, U. Regulation and secretion of Xanthomonas virulence factors. Fems Microbiol. Rev. 2010, 34, 107–133. [Google Scholar] [CrossRef] [Green Version]
  136. Fan, S.; Tian, F.; Fang, L.; Yang, C.-H.; He, C. Transcriptional responses of Xanthomonas oryzae pv. oryzae to type III secretion system inhibitor ortho-coumaric acid. BMC Microbiol. 2019, 19, 163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Stringlis, I.A.; De Jonge, R.; Pieterse, C.M. The age of coumarins in plant–microbe interactions. Plant. Cell Physiol. 2019, 60, 1405–1419. [Google Scholar] [CrossRef] [Green Version]
  138. Venugopala, K.N.; Rashmi, V.; Odhav, B. Review on natural coumarin lead compounds for their pharmacological activity. BioMed Res. Int. 2013, 2013, 963248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Kauppi, A.M.; Nordfelth, R.; Uvell, H.; Wolf-Watz, H.; Elofsson, M. Targeting bacterial virulence: Inhibitors of type III secretion in Yersinia. Chem. Biol. 2003, 10, 241–249. [Google Scholar] [CrossRef] [Green Version]
  140. Kauppi, A.M.; Nordfelth, R.; Hägglund, U.; Wolf-Watz, H.; Elofsson, M. Salicylanilides are potent inhibitors of type III secretion in Yersinia. In The Genus Yersinia; Springer: Boston, MA, USA, 2004; pp. 97–100. [Google Scholar]
  141. Nordfelth, R.; Kauppi, A.M.; Norberg, H.; Wolf-Watz, H.; Elofsson, M. Small-molecule inhibitors specifically targeting type III secretion. Infect. Immun. 2005, 73, 3104–3114. [Google Scholar] [CrossRef] [Green Version]
  142. Muschiol, S.; Bailey, L.; Gylfe, Å.; Sundin, C.; Hultenby, K.; Bergström, S.; Elofsson, M.; Wolf-Watz, H.; Normark, S.; Henriques-Normark, B. A small-molecule inhibitor of type III secretion inhibits different stages of the infectious cycle of Chlamydia trachomatis. Proc. Natl. Acad. Sci. USA 2006, 103, 14566–14571. [Google Scholar] [CrossRef] [Green Version]
  143. Muschiol, S.; Normark, S.; Henriques-Normark, B.; Subtil, A. Small molecule inhibitors of the Yersinia type III secretion system impair the development of Chlamydia after entry into host cells. BMC Microbiol. 2009, 9, 75. [Google Scholar] [CrossRef] [Green Version]
  144. Veenendaal, A.K.; Sundin, C.; Blocker, A.J. Small-molecule type III secretion system inhibitors block assembly of the Shigella type III secreton. J. Bacteriol. 2009, 191, 563–570. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Wolf, K.; Betts, H.; Chellas-Gery, B.; Hower, S.; Linton, C.; Fields, K. Treatment of Chlamydia trachomatis with a small molecule inhibitor of the Yersinia type III secretion system disrupts progression of the chlamydial developmental cycle. Mol. Microbiol. 2006, 61, 1543–1555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Bailey, L.; Gylfe, Å.; Sundin, C.; Muschiol, S.; Elofsson, M.; Nordström, P.; Henriques-Normark, B.; Lugert, R.; Waldenström, A.; Wolf-Watz, H. Small molecule inhibitors of type III secretion in Yersinia block the Chlamydia pneumoniae infection cycle. FEBS Lett. 2007, 581, 587–595. [Google Scholar] [CrossRef] [Green Version]
  147. Tree, J.J.; Wang, D.; McInally, C.; Mahajan, A.; Layton, A.; Houghton, I.; Elofsson, M.; Stevens, M.P.; Gally, D.L.; Roe, A.J. Characterization of the effects of salicylidene acylhydrazide compounds on type III secretion in Escherichia coli O157: H7. Infect. Immun. 2009, 77, 4209–4220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Hudson, D.L.; Layton, A.N.; Field, T.R.; Bowen, A.J.; Wolf-Watz, H.; Elofsson, M.; Stevens, M.P.; Galyov, E.E. Inhibition of type III secretion in Salmonella enterica serovar Typhimurium by small-molecule inhibitors. Antimicrob. Agents Chemother. 2007, 51, 2631–2635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Negrea, A.; Bjur, E.; Ygberg, S.E.; Elofsson, M.; Wolf-Watz, H.; Rhen, M. Salicylidene acylhydrazides that affect type III protein secretion in Salmonella enterica serovar typhimurium. Antimicrob. Agents Chemother. 2007, 51, 2867–2876. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Anantharajah, A.; Buyck, J.M.; Sundin, C.; Tulkens, P.M.; Mingeot-Leclercq, M.-P.; Van Bambeke, F. Salicylidene acylhydrazides and hydroxyquinolines act as inhibitors of type three secretion systems in Pseudomonas aeruginosa by distinct mechanisms. Antimicrob. Agents Chemother. 2017, 61, e02566-16. [Google Scholar] [CrossRef] [Green Version]
  151. Bellemann, P.; Geider, K. Localization of transposon insertions in pathogenicity mutants of Erwinia amylovora and their biochemical characterization. Microbiology 1992, 138, 931–940. [Google Scholar] [CrossRef] [Green Version]
  152. Dahlgren, M.K.; Zetterström, C.E.; Gylfe, Å.; Linusson, A.; Elofsson, M. Statistical molecular design of a focused salicylidene acylhydrazide library and multivariate QSAR of inhibition of type III secretion in the Gram-negative bacterium Yersinia. Bioorganic Med. Chem. 2010, 18, 2686–2703. [Google Scholar] [CrossRef]
  153. Buttimer, C.; McAuliffe, O.; Ross, R.P.; Hill, C.; O’Mahony, J.; Coffey, A. Bacteriophages and bacterial plant diseases. Front. Microbiol. 2017, 8, 34. [Google Scholar] [CrossRef] [Green Version]
  154. Stockwell, V.; Duffy, B. Use of antibiotics in plant agriculture. Rev. Sci. Et Tech. Off. Int. Des. Epizoot. 2012, 31, 199–210. [Google Scholar] [CrossRef] [PubMed]
  155. McGhee, G.C.; Sundin, G.W. Evaluation of kasugamycin for fire blight management, effect on nontarget bacteria, and assessment of kasugamycin resistance potential in Erwinia amylovora. Phytopathology 2011, 101, 192–204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Benveniste, R.; Davies, J. Aminoglycoside antibiotic-inactivating enzymes in actinomycetes similar to those present in clinical isolates of antibiotic-resistant bacteria. Proc. Natl. Acad. Sci. USA 1973, 70, 2276–2280. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Sambanthamoorthy, K.; Sloup, R.E.; Parashar, V.; Smith, J.M.; Kim, E.E.; Semmelhack, M.F.; Neiditch, M.B.; Waters, C.M. Identification of small molecules that antagonize diguanylate cyclase enzymes to inhibit biofilm formation. Antimicrob. Agents Chemother. 2012, 56, 5202–5211. [Google Scholar] [CrossRef] [Green Version]
  158. Kim, B.; Park, J.-S.; Choi, H.-Y.; Yoon, S.S.; Kim, W.-G. Terrein is an inhibitor of quorum sensing and c-di-GMP in Pseudomonas aeruginosa: A connection between quorum sensing and c-di-GMP. Sci. Rep. 2018, 8, 1–13. [Google Scholar] [CrossRef]
  159. Hee, C.-S.; Habazettl, J.; Schmutz, C.; Schirmer, T.; Jenal, U.; Grzesiek, S. Intercepting second-messenger signaling by rationally designed peptides sequestering c-di-GMP. Proc. Natl. Acad. Sci. USA 2020, 117, 17211–17220. [Google Scholar] [CrossRef]
  160. Cho, K.H.; Tryon, R.G.; Kim, J.-H. Screening for Diguanylate Cyclase (DGC) Inhibitors Mitigating Bacterial Biofilm Formation. Front. Chem. 2020, 8, 264. [Google Scholar] [CrossRef]
  161. Zamioudis, C.; Pieterse, C.M. Modulation of host immunity by beneficial microbes. Mol. Plant.-Microbe Interact. 2012, 25, 139–150. [Google Scholar] [CrossRef] [Green Version]
  162. Nazir, R.; Mazurier, S.; Yang, P.; Lemanceau, P.; Van Elsas, J.D. The ecological role of type three secretion systems in the interaction of bacteria with fungi in soil and related habitats is diverse and context-dependent. Front. Microbiol. 2017, 8, 38. [Google Scholar] [CrossRef] [Green Version]
  163. Slack, S.; Walters, K.J.; Outwater, C.; Sundin, G.W. Effect of kasugamycin, oxytetracycline, and streptomycin on in-orchard population dynamics of Erwinia amylovora on apple flower stigmas. Plant. Dis. 2020. [Google Scholar] [CrossRef]
Figure 1. Model of the type III secretion system (T3SS) regulation and modes of action of T3SS inhibitors in Dickeya dadantii. The expression of T3SS master regulator HrpL is transcriptionally regulated by the HrpX/HrpY-HrpS-RpoN pathway and post-transcriptionally regulated by the GacS/GacA-RsmB-RsmA pathway. Several transcriptional regulators, including FlhDC, SlyA, PecS, PecT, and RpoS, regulate the expression of T3SS genes via targeting multiply key components in the T3SS regulatory pathways. Hfq and its dependent sRNA ArcZ form a feed-forward signaling cascade that positively control the expression of RsmB. PNPase degrades RsmB sRNA and is also required for the stability of rpoN mRNA. c-di-GMP signaling is involved in the FlhDC-mediated RpoN regulation and the Hfq-mediated RsmB regulation. T3SS injects bacterial effectors into the plant cells to suppress plant immune responses. On the other hand, plants secrete a variety of phenolic compounds that are able to regulate the T3SS in bacteria. For D. dadantii, two plant phenolic compounds, o-coumaric acid (OCA) and t-cinnamic acid (TCA), upregulate T3SS gene expression via the GacS/GacA-RsmB-RsmA-HrpL pathway. Trans-4-hydroxycinnamohdroxamic acid (TS103), a plant phenolic compound derivative, represses the expression of T3SS via the HrpX/HrpY-HrpS-RpoN-HrpL pathway. ⊥ represents negative control; → represents positive control.
Figure 1. Model of the type III secretion system (T3SS) regulation and modes of action of T3SS inhibitors in Dickeya dadantii. The expression of T3SS master regulator HrpL is transcriptionally regulated by the HrpX/HrpY-HrpS-RpoN pathway and post-transcriptionally regulated by the GacS/GacA-RsmB-RsmA pathway. Several transcriptional regulators, including FlhDC, SlyA, PecS, PecT, and RpoS, regulate the expression of T3SS genes via targeting multiply key components in the T3SS regulatory pathways. Hfq and its dependent sRNA ArcZ form a feed-forward signaling cascade that positively control the expression of RsmB. PNPase degrades RsmB sRNA and is also required for the stability of rpoN mRNA. c-di-GMP signaling is involved in the FlhDC-mediated RpoN regulation and the Hfq-mediated RsmB regulation. T3SS injects bacterial effectors into the plant cells to suppress plant immune responses. On the other hand, plants secrete a variety of phenolic compounds that are able to regulate the T3SS in bacteria. For D. dadantii, two plant phenolic compounds, o-coumaric acid (OCA) and t-cinnamic acid (TCA), upregulate T3SS gene expression via the GacS/GacA-RsmB-RsmA-HrpL pathway. Trans-4-hydroxycinnamohdroxamic acid (TS103), a plant phenolic compound derivative, represses the expression of T3SS via the HrpX/HrpY-HrpS-RpoN-HrpL pathway. ⊥ represents negative control; → represents positive control.
Microorganisms 08 01956 g001
Table 1. Compounds that regulate T3SS in some major plant pathogens.
Table 1. Compounds that regulate T3SS in some major plant pathogens.
CompoundStructureKnown Modes of Action on T3SS in Plant PathogensReferences
o-coumaric acid Microorganisms 08 01956 i001Induce T3SS via the RsmA/RsmB-HrpL pathway in D. dadantii; inhibit T3SS in E. amylovora and T3SS in X. oryzae pv. oryzae via the HrpG-HrpX regulatory cascade.[32,115,116]
t-cinnamic acid Microorganisms 08 01956 i002Induce T3SS in D. dadantii; inhibit T3SS in E. amylovora.[32,115]
p-coumaric acid Microorganisms 08 01956 i003Inhibit T3SS via the HrpS-HrpL pathway in D. dadantii.[117]
trans-4-hydroxycinnamodydroxamic acid Microorganisms 08 01956 i004Inhibit T3SS via the RpoN-HrpL pathway, the HrpX/HrpY-HrpS-HrpL pathway, and the RsmB-HrpL pathway in D. dadantii.[118]
Benzoic acid Microorganisms 08 01956 i005Inhibit T3SS via targeting HrpS in E. amylovora.[115]
Salicylic acid Microorganisms 08 01956 i006Inhibit T3SS in E. amylovora.[115]
4-methoxy-cinnamic acid Microorganisms 08 01956 i007Inhibit T3SS via targeting HrpL in E. amylovora.[115]
trans-2-methoxycinnamic acid Microorganisms 08 01956 i008Inhibit T3SS via the HrpG-HrpX regulatory cascade in X. oryzae pv. oryzae.[116]
trans-2-phenylcyclopropane-1-carboxylic-acid Microorganisms 08 01956 i009Inhibit T3SS via the HrpG-HrpX regulatory cascade in X. oryzae pv. oryzae.[116]
trans-2-methylcinnamic acid Microorganisms 08 01956 i010Inhibit T3SS via the HrpG-HrpX regulatory cascade in X. oryzae pv. oryzae.[116]
Umbelliferone Microorganisms 08 01956 i011Inhibit T3SS via targeting HrpG in R. solanacearum.[119,120]
Benzoic acid N′-(2,3,4-trihydroxy-benzylidene)-hydrazide Microorganisms 08 01956 i012Inhibit multiple T3SS regulon genes in E. amylovora; inhibit T3SS via targeting HrpB in R. solanacearum.[121,122]
4-nitrobenzoic acid N′-(2,4-dihydroxy-benzylidene)-hydrazide Microorganisms 08 01956 i013Inhibit T3SS via targeting HrpB in R. solanacearum; inhibit T3SS via targeting HrpL in E. amylovora.[121,122]
2-nitro-benzoic acid N′-(3,5-dichloro-2-hydroxy-benzylidene)-hydrazide Microorganisms 08 01956 i014Inhibit T3SS via targeting HrpB in R. solanacearum.[122]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yuan, X.; Yu, M.; Yang, C.-H. Innovation and Application of the Type III Secretion System Inhibitors in Plant Pathogenic Bacteria. Microorganisms 2020, 8, 1956. https://doi.org/10.3390/microorganisms8121956

AMA Style

Yuan X, Yu M, Yang C-H. Innovation and Application of the Type III Secretion System Inhibitors in Plant Pathogenic Bacteria. Microorganisms. 2020; 8(12):1956. https://doi.org/10.3390/microorganisms8121956

Chicago/Turabian Style

Yuan, Xiaochen, Manda Yu, and Ching-Hong Yang. 2020. "Innovation and Application of the Type III Secretion System Inhibitors in Plant Pathogenic Bacteria" Microorganisms 8, no. 12: 1956. https://doi.org/10.3390/microorganisms8121956

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop