Pathogenesis Related (PR) Proteins in Plant Defense Mechanism

advertisement
Science against microbial pathogens: communicating current research and technological advances
_______________________________________________________________________________
A. Méndez-Vilas (Ed.)
Pathogenesis Related (PR) Proteins in Plant Defense Mechanism
Saboki Ebrahim1, K.Usha1 and Bhupinder Singh2
1
Division of Fruits and Horticultural Technology, 2 Division of Nuclear Research Laboratory, Indian Agricultural
Research Institute, New Delhi, India
When resistant plants recognize cognate or matching elicitors, intracellular signal transduction pathways are activated that
ultimately result in the derepression of a battery of genes called defense response genes. These latter genes encode
producing various pathogenesis related (PR) toxic proteins such as chitinases, glucanases, lysozyme-active proteins, or cell
wall strengthening proteins such as hydroxyproline rich glycoproteins. Response proteins may also be enzymes in
biosynthetic pathways for lignification of cell walls or the production of phytoalexins, low molecular weight toxic
chemicals that antagonize the invader. In the following portion, biochemical response of plant defence mechanism related
to PR-protein including chitinase and glucanase, in addition to plant lignin content will be explained.
Keywords : Pathogenesis related (PR) protein, Chitinase, β-1,3-glucanase, Lignin.
Pathogenesis-Related (PR) Proteins
Higher plants have a broad range of mechanisms to protect themselves against various threats including physical,
chemical and biological stresses, such as wounding, exposures to salinity, drought, cold, heavy metals, air pollutants
and ultraviolet rays and pathogen attacks, like fungi, bacteria and viruses [1]. Plant reactions to these factors are very
complex, and involve the activation of set of genes, encoding different proteins. These stresses can induce biochemical
and physiological changes in plants, such as physical strengthening of the cell wall through lignification, suberization,
and callose deposition; by producing phenolic compounds, phytoalexins and pathogenesis-related (PR) proteins which
subsequently prevent various pathogen invasion [2]. Among these, production and accumulation of pathogenesis related
proteins in plants in response to invading pathogen and/or stress situation is very important. Phytoalexins are mainly
produced by healthy cells adjacent to localized damaged and necrotic cells, but PR proteins accumulate locally in the
infected and surrounding tissues, and also in remote uninfected tissues. Production of PR proteins in the uninfected
parts of plants can prevent the affected plants from further infection [3, 4]. PR protein in the plants was first discovered
and reported in tobacco plants infected by tobacco mosaic virus [5]. Later, these proteins were found in many plants.
Most PR proteins in the plant species are acid-soluble, low molecular weight, and protease-resistant proteins [6, 7]. PR
proteins depending on their isoelectric points may be acidic or basic proteins but they have similar functions. Most
acidic PR proteins are located in the intercellular spaces, whereas, basic PR proteins are predominantly located in the
vacuole [8, 9, 10]. The PR proteins have been classically divided initially into 5 families based on molecular mass,
isoelectric point, and localization and biological activity [11]. Currently PR-proteins were categorized into 17 families
according to their properties and functions (Table 1), including β-1,3-glucanases, chitinases, thaumatin-like proteins,
peroxidases, ribosome-inactivating proteins, defenses, thionins, nonspecific lipid transfer proteins, oxalate oxidase, and
oxalate-oxidase-like proteins [10]. Among these PR proteins chitinases and β-1,3-glucanases are two important
hydrolytic enzymes that are abundant in many plant species after infection by different type of pathogens. The amount
of them significantly increase and play main role of defense reaction against fungal pathogen by degrading cell wall,
because chitin and β-1,3-glucan is also a major structural component of the cell walls of many pathogenic fungi. β-1,3glucanases appear to be coordinately expressed along with chitinases after fungal infection. This co-induction of the two
hydrolytic enzymes has been described in many plant species, including pea, bean, tomato, tobacco, maize, soybean,
potato, and wheat [12, 13, 14, 15, 16, 17, 18, 19, 20].
©FORMATEX 2011
1043
Science against microbial pathogens: communicating current research and technological advances
______________________________________________________________________________
A. Méndez-Vilas (Ed.)
Table 1: Classification of pathogenesis related proteins [10].
Families
Type member
Properties
PR-1
Tobacco PR-1a
Antifungal
PR-2
Tobacco PR-2
β-1,3-glucanase
PR-3
Tobacco P, Q
Chitinase type I,II, IV,V,VI,VII
PR-4
Tobacco ‘R’
Chitinase type I,II
PR-5
Tobacco S
Thaumatin- like
PR-6
Tomato Inhibitor I
Proteinase- inhibitor
PR-7
Tomato P69
Endoproteinase
PR-8
Cucumber chitinase
Chitinase type III
PR-9
Tobacco ‘lignin forming peroxidase’
Peroxidase
PR-10
Parsley ‘PR1’
Ribonuclease like
PR-11
Tobacco ‘class V’ chitinase
Chitinase, type I
PR-12
Radish Rs- AFP3
Defensin
PR-13
Arabidopsis THI2.1
Thionin
PR-14
Barley LTP4
Lipid- transfer protein
PR-15
Barley OxOa (germin)
Oxalate oxidase
PR-16
Barley OxOLP
Oxalate oxidase-like
PR-17
Tobacco PRp27
Unknown
Synergistic Effect of Chitinase and β-1, 3-Glucanase in Transgenic Plants
Chitinase genes, alone or together with β-1,3-glucanase genes, have been transferred to a number of plant species and
expressed and those effects were studied. In most cases, the resulting transgenic plants exhibit enhanced levels of fungal
disease resistance or delayed symptom development as compared to the control plants [15, 21, 22]. But several studies
have showed that plants transformed with chitinase or β-1, 3-glucanase gene alone did not exhibit resistance to certain
pathogens or showed less resistance compared to plants which were transformed with both the β-1, 3-glucanase and
chitinase genes. Like β-1, 3-glucanase, chitinases inhibit only a limited number of fungal species therefore these two
enzyme have synergistic effect. Plant chitinases alone usually affect only the hyphal tip and are unable to effectively
degrade harder chitin structures of fungi. But whenever these two enzymes are combined, a synergic effect can usually
be observed. For example, tomato plants expressing tobacco class I β-1,3-glucanase and chitinase transgenes showed
increased tolerance to infection by Fusarium oxysporum f.sp. lycopersici [22]. Also in tobacco plants transformed with
a barley class II basic β-1, 3-glucanase along with a barley class II basic chitinase gene showed enhanced levels of
protection against Rhizoctonia solani as compared to plants transformed with a single gene [15].
Age-Related Pathogen Resistance
Many studies reported a direct relationship between plant age and pathogen resistance or susceptibility
[23,24,25,26,27,28,30,31,32,33,34,35]. In most plants, resistance to pathogen usually increases with increasing age; but
in few cases susceptibility of plants to pathogens increases with development of plant. This difference in resistance to
1044
©FORMATEX 2011
Science against microbial pathogens: communicating current research and technological advances
_______________________________________________________________________________
A. Méndez-Vilas (Ed.)
pathogens depending on plant age is called age-related pathogen resistance. For example, the older leaves from bottoms
of the shoots of grapevines showed more β-1, 3-glucanase and consequently they exhibited a higher resistance to
Plasmopara viticola than younger leaves from upper shoots [31]. In addition, both young and mature leaves in older
tobacco and wheat plants showed higher resistance than the leaves in younger plants [25, 26]. Another study in tobacco
showed that tobacco plants naturally become resistant to blue mold caused by Peronospora tabacina as its age
increases. β-1,3-glucanase, chitinase and peroxidase activities increased in tobacco with age [35]. Enzyme activities
were observed higher in leaf tissue from the main stalk (resistant to blue mold) as compared to leaf tissue from
suckering stems (susceptible to blue mold) on the same plant [35].
Chitinase in Plant Defense
Chitinases (E.C. 3.2.1.14) are enzymes catalyzing the cleavage of a bond between C1 and C4 of two consecutive Nacetyl-D-glucosamine monomers of chitin. They are widely distributed in nature, occurring in bacteria, fungi, animals,
and plants. Chitin is a common component of fungal cell walls and of the exoskeleton of arthropods [36]. Plant
chitinases are usually endo-chitinases capable of degrading chitin, a major constituent of certain fungal cell walls as
well as inhibit fungal growth [37, 38]. Some class I chitinases are localized in the vacuole, whereas other chitinases,
including class III chitinases, are located outside the cell [39]. Extracellular chitinases may directly block the growth of
the hyphae invading intercellular spaces and possibly release fungal elicitors, which then induce additional chitinase
biosynthesis and further defense reactions in the host [40, 41,]. There are strong indications that chitinase, together with
β-1, 3 glucanase participate in the plant defense system against fungal pathogens. Chitinase are found in leaves of many
plants, whereas has no known function in growth and development of plant. Higher plants do not contain chitin and any
other known endogenous substrate for chitinase. However chitin and β-1, 3-glucanase are major components in the cell
wall of many fungi and there is possibility of plant chitinase and β-1, 3- glucanase target fungi cell wall components as
substrate and has anti fungal function [42.43].
Classification of Plant Chitinase
Chitinase are classified into two categories, endochitinases and exochitinases. Endochitinases cleave chitin randomly at
internal points within the polymer, producing soluble, low molecular weight multimers of N-acetylglucosamine, such as
chitotrose, chitotetraoseand and the dimer, di-acetylchitobios. Exochitinases are divided into two subcategories:
chitobioses, catalyzing release of di-acetylchitobioses from nondeoxydizing end of the chitin microfibril, and β-1, 4-Nacetylglucoseaminidase, cleaving oligomer products of endochitinases and chitobiosidases with the production of
monomer N-acetylglucosamine [29].
Based on their primary structures, plant chitinases have been classified into seven classes, class I through VII.
Different chitinase classes have no apparent correlation to being present in a particular plant species and plant organ or
tissue. However, certain chitinase isoforms are sometimes induced by a particular elicitor. For example, in potato, both
acidic and basic chitinase, can be active. Also only particular isoforms have antifungal activities and some isoforms
have shown another role like antifreeze activity [44, 45].
Class I chitinases have a cysteine-rich N-terminal chitin-binding domain (CBD) that is homologous to havein, a
chitin-binding lectin from the rubber tree. CBD is separated from the catalytic domain by a proline and glycine-rich
hinge or spacer region, variable both in size and composition. Class I chitinase, in tobacco due to deletion of CBD and
the spacer region singly or in combination reduces the hydrolytic activity by 50%, also antifungal function is reduced by
80% [46]. Class I is divided into subclasses Ia and Ib that include acidic and alkaline chitinases, respectively. Class I
chitinases occur only in plants, whereas chitinases of class II, also in fungi and bacteria. They are induced by the
pathogen attack and lack both N-terminal domains as the C-terminal signal peptide ensuring vacuolar localization. Class
I chitinase, in tobacco due to deletion of CBD and the spacer region singly or in combination reduces the hydrolytic
activity by 50%, also antifungal function is reduced by 80% [46]. Class II chitinases are similar to class I but they lack
the N-terminal CBD and the hinge region. Also, they have acidic properties.
Class III chitinases are unique in structure and have no relationship to any other class of plant chitinases. These
chitinases belong to the PR-8 family and family 18 of glycosyl-hydrolases. Class III chitinases generally have lysozyme
activity and appear to be more closely related to the bacterial chitinases. Class III chitinase enzyme have high amino
acid sequence homology to the bifunctional chitinase lysozyme from Hevea brasiliensis and include the chitinase from
Azukin bean and cucumber. This is unusual for a class III chitinase since they do not have a chitin-binding domain.
Class III chitinases show a wide range of isoelectric points, activity over a wide range of pH, and temperature stability
at 60-70oC. The B. hispida chitinase has a pH optimum of 2 and retains approximately 50% activity at pH 8 [47]. Some
class III chitinases, such as a yam enzyme, show two pH optima and heat stability at 80oC [48]. There is no posttranslational modification reported for class III enzymes analyzed to date.
Class IV, V, VI, and VII chitinases belong to the PR-3 family of pathogenesis related proteins. The structure of class
IV chitinases is similar to class I chitinases containing cysteine rich domain and a conserved main structure, except that
©FORMATEX 2011
1045
Science against microbial pathogens: communicating current research and technological advances
______________________________________________________________________________
A. Méndez-Vilas (Ed.)
they are shorter due to four deletions. Class IV chitinases represent a group of extracellular chitinases, Like class II
chitinases, they are found mainly in dicotyledon plants. Class V was not similar to any known plant chitinase but have
significant sequence similarity to bacterial exochitinases. Class VI chitinase are homologous, to sugar beet chitinase I
possessing only half of chitin binding domain and a long prolin rich sapacer.
I-VII classes of chitinase [49] have been reclassified into 4 families (Corresponding to 4 PR-protein families) namely
Chia, Chib, Chic and Chid and having total of nine classes [39].This classification is based on the presence or absence
of an N-terminal havein domain and on sequence similarity to an archetypal catalytic domain. Chia1 chitinase have an
N-terminal havein domain and a catalytic domain that is at least 50% identical to tobacco Chia1chitinase. The Chia2
chitinase lack an N-terminal havein domain, but contain catalytic domain that are at least 50% identical to the catalytic
domain of tobacco Chia1 chitinase. The Chia4 chitinase have N-terminal havein domain but shorter than Chia1
chitinase. They are at least 50% identical to Phaseolus vulgaris PR-4 chitinase. Chib1 chitinase sequence are different
than Chia1, Chia2 and Chia4 chitinases, but are at least 30% identical to tobacco Chib1 chitinases which also have
lysozyme activity. Chic1 chitinases have no similarity whit Chia1, 2, 4 or Chib1 but their amino acid sequences are at
least 50% identical to a group of tobacco chitinase that are similar to bacterial exochitinases. Chia5 chitinase has two
chitin binding domains and Chia6 chitinase has half chitin binding domain and are related to Chia1.Chia chitinase
belong to PR-3 protein family, Chib1 belong to PR-8 and Chic1 belong to PR-11 protein family. PR-4 has been found
to have chitinase activity, and makes the family Chid.
Chia1 enzymes in general have more specific activity than chia2. Compared to chia2, Chia1 has 3 fold more activity
in barley [50] and 6 to 15 fold more specific activity in tobacco [8]. Chia1 chitinase account for the majority of the
chitinolytic activity in plant. For example, two chia1 isoforms account for 68% of total activity, yet constitute only 25%
of the chitinolytic protein [8].
Purified chitinases show varying degrees of antifungal activities in vitro. In general, class I chitinases have the
highest antifungal activity, perhaps due to the presence of a chitin-binding domain [44]. All other chitinase classes have
lower to no antifungal activity as compared to class I chitinases.
Induction and Functions of Chitinases in the Plant
Chitinase have been known to be induced in the plant by fungal infection or other biotic and abiotic factors. Inhibitory
effect of plant chitinase about fungal growth was demonstrated by in vitro studies on the growth of fungi that contain
chitin as a component of their cell wall [12, 13, 38]. The induction of chitinases was initially shown in pea plants
infected with Fusarium solani [12]. Inhibitory effect of protein extracted from pea plant was infected by Fusarium
solani and showed to inhibit growth of 15 of the 18 fungal species tested in vitro.
Purified chitinase inhibited growth of only one fungal species whereas a combination of chitinase and another PRprotein, β-1, 3-glucanase, inhibited the growth of all fungi tested showing a synergism in activities [13]. The number of
studies verified these results in tobacco [51], grapes [52], chickpea [53], rice [54] and other plants. These studies
demonstrated that specific isoforms are induced in response to a particular pathogen and only certain isoforms are able
to inhibit specific fungi [44, 55]. For example, a class I chitinase from tobacco showed antifungal activity against
Fusarium solani [15].
Various studies have shown that chitinase expression against phyto-pathogen systems is higher and induction is
stronger in the resistant varieties in comparison to susceptible varieties in the sugar beet [56], wheat [57] and tomato
varieties [58]. Also another report showed no difference in the induction timing or amounts of PR-protein in resistant
and susceptible cultivars of cotton [59].
However, quick response in the resistant cultivars might affect the cell wall of germinating fungal spores, releasing
elicitors leading to the expression of PR-genes and disease resistance. It was shown for Alternaria solani that a basic
chitinase was only active on the germinating spores and not on the mature fungal cell wall for generation of elicitor
molecules to induce disease resistance [58].
Kragh [60] demonstrated chitinase activity in susceptible primary leaves of barley (Hordeum vulgare L.) fivefold and
threefold, respectively 7 days after inoculation with powdery mildew fungus (Erysiphe gramini f. sp. hordei). Metraux
and Boller [61] reported in cucumber, chitinase is induced up to 600-fold locally and up to 100-fold systemically in
response to infection by various pathogens. They showed systemic induction of chitinase in response to a localized
treatment and was correlated with the systemic induction of resistance against C. lagenarium. O'Garro and
Charlemange [62] studied response of the pepper leaves and flowers to infection with Xanthomonase campestris Pv.
Vesicaroria and showed chitinase and β-1, 3-glucanase were coordinately induced in infected leaf and flower tissue. A
grape class III chitinase was also shown to be induced in infected and non-infected leaves upon fungal infection [63].
The induction showed two maxima at 2 d and 6 d in the susceptible Vitis vinifera whereas the level was steeply induced
up to 4 d and declined to the basal level by day 7 in the resistant V. rupestris.
Induction of chitinase was elevated in transgenic plants. In bean (Phaseolus vulgaris) chitinase gene under the
control of CaMV 35S promoter was introduced into tobacco plants through Agrobacterium mediated transformation.
Level of chitinase activity was 20-40 folds as compared to control in the transgenic plants. Transgenic plants showed
increased resistance to infection by pathogenic fungi Rhizoctonia solani and delayed development of disease symptoms.
1046
©FORMATEX 2011
Science against microbial pathogens: communicating current research and technological advances
_______________________________________________________________________________
A. Méndez-Vilas (Ed.)
The transgenic plants showed no resistance to the non-chitin containing fungus Pythium aphanidermatum [64]. In
another study, tomato chitinase gene was transferred to oilseed rape (Brassica napus) and was challenged with three
different fungal pathogens at two field locations [65]. Over a period of 52 days, the protection level against three fungi
was 23% to 79% with both delayed appearance of symptoms and reduced lesion numbers in transgenic plants. The
highest protection was against Cylindrosporium concentricum which contains chitin as a major component in its cell
wall and at the tip of growing hyphae. The protection was lower against two other fungi whose cell walls have different
distributions and amounts of chitin.
Transformation of chitinase genes were performed in grapevine, rice and peanut and enhanced disease resistance has
been achieved. In grapevine, rice chitinase gene enhanced resistance of these plants against powdery mildew caused by
Uncinula necator [66]. . Chitinase gene transformed into rice showed enhanced resistance to sheath blight caused by
Rhizoctonia solani [67]; a tobacco chitinase gene transformed into peanut increased resistance to leaf spot disease
caused by Cercospora arachidicola [68].
Non-plant chitinases have also been transferred to model plants which enhanced resistance in susceptible varieties. A
chitinase gene from the mycoparasitic fungus Trichoderma harzianum was transformed into tobacco and potato plants.
The transgenic plants increased resistance to various fungal pathogens including Alternaria alternata, A. solani, Botrytis
cinerea and Rhizoctonia solani [69]. Another fungal chitinase gene from Rhizopus oligosporus showed suppression of
disease symptoms in transgenic tobacco plants when challenged with pathogenic fungi [70].
Plant β-1, 3-Glucanases in Plant Defense
Plant β-1, 3-glucanases are pathogenesis-related (PR) proteins, which belong to the PR-2 family of pathogenesis-related
proteins and are believed to play an important role in plant defense responses to pathogen infection. In addition to plant,
β-1, 3-glucanases have been found in yeasts, actinomycetes, bacteria, fungi, insects, and fish [71, 72].β-1,3-glucanases
are able to catalyze the cleavage of the β-1,3-glucosidic bonds in β-1,3-glucan [73].β-1,3-glucan is a another major
structural component of the cell walls of many pathogenic fungi [74, 75]. For example, Phytophthora infestans is an
oomycete pathogen that causes late blight of potato and tomato. Oomycetes have a cell wall that is comprised of 8090% β-1,3-glucan. Syntheses of these enzymes can be induced by pathogens or other stimuli. β-1,3-glucan (called
callose in plants) is unlike chitinases, the substrate for β-1,3-glucanases is widespread in plants and therefore these
enzymes may have other physiological functions as well as in plant defense. For example, callose acts as a permeability
barrier in pollen mother cell walls [76] and muskmelon endosperm envelopes [77]. β-1,3-glucanases in the plant also
have function, such as cell division and cell elongation [78, 79, 80], fruit ripening [81], pollen germination and tube
growth [ 82, 83], fertilization [84], somatic embryogenesis [85], seed germination [86, 87] and flower formation [88,
89].Plant β-1,3-glucanases are pathogenesis-related proteins and have been suggested as an important component of
plant defense mechanisms against pathogens [90, 41, 91, 92]. It has been suspected that β-1,3-glucanases have direct
effect in defending against fungi by hydrolyzing fungal cell walls, which consequently causes the lysis of fungal cells.
In addition, β-1,3-glucanases was showed to have an indirect effect on plant defense by causing the formation of
oligosaccharide elicitors, which elicit the production of other PR proteins or low molecular weight antifungal
compounds, such as phytoalexins [93, 94, 95].
Classification of Plant β-1, 3-Glucanases
β-1, 3-glucanase genes have been reported in a number of plants, including tobacco [96], soybean [97], rubber tree [98],
banana [99], and rice [100]. There are different β-1, 3-glucanase genes in different plant species and a single plant
species may have various copies of β-1, 3-glucanase genes. For instance, more than 14 β-1,3-glucanase genes have been
reported in tobacco plants, [6], A variety of β-1,3-glucanase genes have been identified in a wide range of plant species.
Plant β-1,3-glucanases having size from 30-40 kDa, with both acidic and basic isoforms. Based on their amino acid
sequence, structural properties and cellular localizations, β-1, 3-glucanases were classified into two major classes I and
II and two minor classes [84, 6, 101,102,103, 104, 105]. Most of these hydrolytic enzymes identified so far are class I
and class II enzymes. Most class I β-1,3-glucanases are usually basic proteins which are localized in the cell vacuole,
while most class II, III, and IV β-1,3-glucanases are acidic proteins which are secreted into the extracellular space.
Class I β-1, 3-glucanases contain a N-terminal signal peptide and a short C-terminal signal sequence that is
glycosylated. The C-terminal signal sequence commonly referred to as the C-terminal extension, is believed to be the
signal sequence responsible for vacuolar localization of these enzymes [84]. Classes II, III and IV β-1, 3- glucanases are
usually acidic and lack this C-terminal signal and therefore are secreted to extracellular spaces. Some β-1, 3-glucanases
are solely developmentally regulated and do not show a stress-related induction. Two examples are tobacco styler β-1,3glucanase [106] and a tobacco anther β-1,3-glucanase [107]. Several β-1, 3-glucanases in class I like enzymes from
tobacco and tomato have inhibitory activity on the growth of certain pathogenic fungi in vitro. Combinations of class I
β-1, 3-glucanases and class I chitinases showed synergistic effect. In contrast, class II β-1,3-glucanases from tobacco
and tomato do not have the in vitro fungal growth inhibitory effect [44, 108,109]. Class I β-1, 3-glucanase accumulated
©FORMATEX 2011
1047
Science against microbial pathogens: communicating current research and technological advances
______________________________________________________________________________
A. Méndez-Vilas (Ed.)
only at the site of tobacco mosaic virus (TMV) infection in tobacco plants. In contrast class II and III β-1, 3-glucanases
accumulated both at the site of infection and systemically [110, 111]. Similar to chitinases, β-1, 3-glucanases can
degrade the fungal cell wall by disrupting hyphal tips, especially in combination with a chitinase [12, 13].Glucanases
can also play roles in plant defense by generating pathogen-derived elicitors. Like class II enzymes β-1, 3-glucanases,
digest fungal cell walls, leading to the release of oligosaccharide elicitors which stimulate the production of PR proteins
and other defense-related molecules [112].
Many studies have shown that the synthesis of β-1, 3-glucanases is stimulated by pathogen infections. For example,
The induction of the β-1,3-glucanase gene transcript on tobacco usually occurs within 24 h to 48 h and β-1,3-glucanase
amount increasing up to 21-fold by the bacterium Pseudomonas syringae pv syringae [113, 114]. In barley, two β-1,3glucanase isozymes, GI and GIII, were observed in healthy barley leaves, while three isozymes, GI, GII, and GIII, were
observed in infected leaves at different expression levels [115], the β-1,3-glucanase activity also can change during
plant development. This change corresponds to the age-related pathogen resistance in plants [31].
Induction of β-1,3-Glucanases in the Plant
β-1,3-glucanases usually expressed at low concentration in plants, but when plants are infected by fungal, bacterial, or
viral pathogens, β-1,3-glucanases enzyme concentration increases dramatically, . For example, Van Kan et al [103]
showed that mRNA for a tomato acidic β-1, 3-glucanase accumulated to a higher level in leaves infected by the fungal
pathogen Cladosporium fulvum. In tobacco by infection with the bacterium Pseudomonas syringae pv syringae level of
β-1,3-glucanase was transcriptionally induced up to 21-fold [113, 114]. These increasing levels showed locally or/ and
systemically [116, 117, 118, 119]. Induction of β-1, 3-glucanases by pathogens can vary in different clones of same
plant species. For example, when the production of a β-1, 3-glucanase upon infection with Corynespora cassiicola was
compared in different clones of Hevea brasiliensis, variability of the enzyme’s activity was observed significantly
among different clones during pathogenesis. Enzyme activity increased in the tolerant clone, while in the susceptible
clone enzyme decreased [120].
Several studies showed that the expression levels of these enzymes increased after infected with pathogens, such as
barley infected by powdery mildew [121], maize infected with Aspergillus flavus [122]. pepper infected with
Xanthomonas campestris pv. vesicatoria and Phytophthora capsici soybean infected with Pseudomonas syringae [123],
wheat infected with Fusarium graminearum [20], chickpea infected with Ascochyta rabiei (Pass.) Labr. [124] and
peach infected with Monilinia fructicola [125].
Induction of β-1, 3-glucanases and other PR proteins in the plant can also occur due to some components of
pathogens or degraded components of pathogens. These elicitors may be components of the cell surface of the pathogen
that are released by host enzymes, including fungal β-glucan, chitin, chitosan, glycoproteins and Nacetylchitooligosaccharides [126, 127, 128]. They may also be synthesized and released by the pathogen after it enters
the host in response to host signals.
Plant β-1, 3-glucanases are induced not only by pathogen infection, but also by other factors. For example, salicylic
acid induced accumulation of mRNAs of classes II and III β-1, 3-glucanases in wild-type tobacco plants [9,102].Similar
effect was reported for abscisic acid (ABA) in tobacco [129, 130, 131] and methyl jasmonate, ethylene and gibberellin
A3 in tomato seeds and leaves [132]. Stress factors like wounding, drought, exposure to heavy metals, air pollutant
ozone, and ultraviolet radiation can stimulate synthesis of β-1,3-glucanases in some plants [89, 125, 133, 134, 135].
These various factors often appear to interact, resulting in a dynamic response to biotic, as well as abiotic stimuli.
Lignin
Lignin is principally derived from phenylpropanoid hydroxycinnamyl alcohols and it has an important role in host
defense against pathogen invasion. Lignification is a mechanism for resistance in plants. Lignin deposition during
pathogen attack is well documented as a plant defense response [136, 137]. After pathogen invades the plant, during
plant defense response, lignin or lignin-like phenolic compound accumulation was shown to occur in a variety of plantmicrobe interactions during the plant defense responses [138]. Plants assemble cell wall apposition at the sites of
attempted penetration of biotrophic fungi such as powdery mildew [139]. Lignin has critical role in cell wall apposition
(CWA)-mediated defense against pathogen fungus penetration into plant. Lignin, a major component of cell walls of
vascular plants, was shown to accumulate in cell wall apposition and surrounding halo areas [139, 140]. Hence it is
considered as first defense barrier against successful penetration of invasive pathogens. Lignification renders the cell
wall more resistant to mechanical pressure applied during penetration by fungal appressoria and water resistant and thus
less accessible to cell wall-degrading enzymes. Lignification is necessary for the structural integrity of plant cell walls
and that is crucial for plant development [141], but the monomeric composition of lignin can vary depending on the
developmental process: thus, defense lignin accumulated by an elicitor treatment was shown to be significantly different
from lignin in vascular tissues. Lange et al. [142], suggestied that lignin biosynthesis is differentially regulated.
1048
©FORMATEX 2011
Science against microbial pathogens: communicating current research and technological advances
_______________________________________________________________________________
A. Méndez-Vilas (Ed.)
Lignification may have several roles in defense as pointed out by Ride [143]. He suggested five ways that
lignification might hinder fungal growth through plant tissue. 1) Lignin may make walls more resistant to mechanical
penetration. It is generally held that lignin increases resistance of walls to compressive forces such as would be
expected at the growing point of a penetration peg. 2) Lignification of the wall at the point of attack may render it
resistant to dissolution by fungal enzymes. Induced lignification of wheat leaves makes them resistant to enzyme
degradation [144]. 3) Lignification of walls may restrict diffusion of enzymes and toxins from the fungus to host, and
water and nutrients from the host to fungus, in essence starving the fungus [145]. 4) Low molecular weight phenolic
precursors of lignin and free radicals produced during polymerization, may inactivate fungal membranes, enzymes,
toxins, and elicitors. Many phenolic compounds have antifungal activity. 5) The hyphal tip may become lignified and
lose plasticity necessary for growth [130]. Each of these roles can be tested and therefore used to evaluate lignification
as a resistance mechanism. Thus, if lignification is to play a primary role in restricting pathogen development, it must
occur early in the host-pathogen interaction and it must be localized near the invading pathogen [146].
Recently, Bhuiyan et al.,[147] by using an RNAi gene-silencing assay, they showed that monolignol biosynthesis
plays a critical role in cell wall apposition mediated defense against powdery mildew fungus penetration into diploid
wheat. Silencing monolignol genes led to super-susceptibility of wheat leaf tissues to an appropriate pathogen, Blumeria
graminis f. sp. tritici (Bgt), and compromised penetration resistance to a non-appropriate pathogen, B. graminis f. sp.
hordei. Autofluorescence of cell wall apposition regions was reduced significantly at the fungal penetration sites in
silenced cells. Their work indicates an important role for monolignol biosynthetic genes in effective cell wall apposition
formation against pathogen penetration. In this addendum, they show that silencing of monolignol genes also
compromised penetration resistant to Bgt in a resistant wheat line.
Conclusion
Higher plants protect themselves against fungal infection or other biotic and abiotic factors in different ways. Plants
defend themselves against such factors by physical strengthening of the cell wall through lignification, suberization, and
producing various pathogenesis-related (PR) proteins such as chitinases, β-1, 3-glucanases.
Pathogenesis-related proteins, including hydrolytic enzymes chitinases and β- 1, 3-glucanases, in plants in response
to invading pathogen are very important. Plant pathogenesis-related proteins are implicated in plant defense responses
against pathogen infection. Production of PR proteins in the remote uninfected parts of plants can lead to the occurrence
of systemic acquired resistance, protecting the affected plants from further infection.
It has been suspected that PR proteins have direct effect in defense mechanism against fungi by hydrolyzing fungal
cell walls, which consequently causes the lysis of fungal cells. In addition, PR proteins were showed to have an indirect
effect on plant defense by causing the formation of oligosaccharide elicitors, which elicit the production of other PR
proteins or low molecular weight antifungal compounds, such as phytoalexins.
Lignin is has an important role in plant defense against pathogen invasion. Lignification is a mechanism for
resistance in plants. After pathogen invades the plant, lignin or lignin-like phenolic compound accumulation was shown
to occur in a variety of plant-microbe interactions during the plant defense responses.
Endogenous enzymes chitinase, β-1, 3-glucanases and lignin content in plant leaves can be used as biochemical
markers for identifying plant varieties resistant to fungal infection or other biotic and abiotic factors. Also by
transferring pathogenesis-related (PR) proteins such as chitinases, β-1, 3-glucanases genes can induce resistance in
plants to various photogenes.
References
[1]. Agrios GN. Plant Pathology, 4th ed. San Diego: Academic Press, p: 93-114; 1997.
[2]. Bowles DJ. Defense-related proteins in higher plants. Annu. Rev. Biochem. 1990; 59:873-907.
[3]. Ryals JA, Neuenschwander UH, Willits MG, Molina A, Steiner HY, Hunt MD. Systemic acquired resistance. Plant Cell. 1996;
8: 1809-1819.
[4]. Delaney TP. Genetic dissection of acquired resistance to disease. Plant Physiol. 1997; 113: 5-12.
[5]. Van Loon LC, Van Kammen A. Polyacrylamide disc electrophoresis of the soluble leaf proteins from Nicotiana tabacum var
‘Samsun’ and Samsun NN. II. Changes in protein constitution after infection with tobacco mosaic virus. Virology. 1970; 40:
199-211.
[6]. Leubner-Metzger G, Meins FJ. Functions and regulation of plant β-1,3-glucanases (PR-2). In: Datta, S. K., Muthukrishnan, S.
eds. Pathogenesis-Related Proteins in Plants. CRC Press, Boca Raton, Florida, p. 77-105; 1999.
[7]. Neuhaus JM. Plant chitinases (PR-3, PR-4, PR-8, PR-11). In “Pathogenesis related proteins in plants”, eds. Datta, S.K. and
Mathukrishnan, S. CRC Press, Boca Raton, pp. 77-105; 1999.
[8]. Legrand M, Kauffmann S, Geoffroy P, Fritig B. Biological function of pathogenesis-related proteins: Four tobacco pathogenesisrelated proteins are chitinases. Proceeding of National Academy of Sciences. 1987; 84: 6750-6754.
[9].Niki T, Mitsuhara I, Seo S, Ohtsubo N, Ohashi Y. Antagonistic effect of salicylic acid and jasmonic acid on the expression of
pathogenesis-related (PR) protein genes in wounded mature tobacco leaves. Plant Cell Physiol. 1998; 39: 500-507.
©FORMATEX 2011
1049
Science against microbial pathogens: communicating current research and technological advances
______________________________________________________________________________
A. Méndez-Vilas (Ed.)
[10]. Van Loon LC, Van Strien EA. The families of pathogenesis-related proteins, their activities, and comparative analysis of PR-1
type proteins. Physiological and Molecular Plant Pathology. 1999; 55: 85-97.
[11]. Van Loon L C. Pathogenesis-related proteins. Plant Mol. Biol. 1985; 4: 111-116.
[12]. Mauch F, Hadwiger LA, Boller T. Antifungal hydrolases in pea tissue I. Purification and characterization of two chitinases and
two beta-1,3-glucanases differentially regulated during development and in response to fungal infection. Plant Physiol. 1988;
87: 325-333.
[13]. Mauch F, Mauch-Mani B, BollerT. Antifungal hydrolases in pea tissue. II. Inhibition of fungal growth by combinations of
chitinase and β-1,3-glucanase. Plant Physiol. 1988; 88: 936-942.
[14]. Vogelsang R, Barz W. Purification, characterization and differential hormonal regulation of a β-1,3-glucanase and two
chitinases from chickpea (Cicer arietinum L.). Planta. 1993; 189: 60-69.
[15]. Jach G, Gornhardt B, Mundy J, Logemann J, Pinsdorf E, Leah R, Schell J, Maas C. Enhanced quantitative resistance against
fungal disease by combinatorial expression of different barley antifungal proteins in transgenic tobacco. Plant J. 1995; 8: 97109.
[16]. Bettini P, Cosi E, Pellegrini M G, Turbanti L, Vendramin GG, Buiatti M. Modification of competence for in vitro response to
Fusarium oxysporum in tomato cells, III. PR-Protein gene expression and ethylene evolution in tomato cell lines transgenic for
phytohormone-related bacterial genes. Theor. Appl. Genet. 1998; 98: 575-583.
[17]. Lambais MR; Mehdy MC. Spatial distribution of chitinases and β-1,3-glucanase transcripts in bean arbbuscular mycorrhizal
roots under low and high soil phosphate conditions. New Phytol. 1998; 140: 33-42.
[18]. Petruzzelli. L, Kunz C, Waldvogel R, Meins F, Leubner-Metzger G. Distinct ethylene and tissue-specific regulation of β-1,3glucanases and chitinases during pea seed germination. Planta. 1999; 209:195-201.
[19]. Cheong YH, Kim C Y, Chun H J, Moon B C, Park HC, Kim JK, Lee SH, Han C D, Lee S Y, Cho M J. Molecular cloning of a
soybean class III β-1,3-glucanase gene that is regulated both developmentally and in response to pathogen infection. Plant Sci.
2000; 154: 71-81.
[20]. Li WL, Faris JD, Muthukrishnan S, Liu DJ, Chen PD; Gill BS. Isolation and characterization of novel cDNA clones of acidic
chitinases and β-1,3-glucanases from wheat spikes infected by Fusarium graminearum. Theor. Appl. Genet. 2001; 102: 353362.
[21] Zhu Q, Maher EA, Masoud S, Dixon RA, Lamb C J. Enhanced protection against fungal attack by constitutive co-expression of
chitinase and glucanase genes in transgenic tobacco. Bio Technol. 1994; 12: 807-811.
[22]. Jongedijk E, Tigelaar H, Van Roekel J S C, Bres-Vloemans S A, Dekker I, Van den Elzen P J M, Cornelissen B J C, Melchers
L S. Synergistic activity of chitinases and β-1,3-glucanases enhances fungal resistance in transgenic tomato plants. Euphytica.
1995, 85: 173-180.
[23]. Allen SJ, Brown JF, Kochman JK. Effects of leaf age, host growth stage, leaf injury and pollen on the infection of sunflower by
Alternaria helianthi. Phytopathology. 1983; 73: 896-898.
[24]. Miller ME. Relationship between onion leaf age and susceptibility to Alternaria porri. Plant Dis. 1983; 67: 284-286.
[25]. Reuveni M, Tuzun S, Cole J S, Siegel MR, Kuc J. The effects of plant age and leaf position on the susceptibility of tobacco to
blue mold caused by Peronospora tabacina. Phytopathology. 1986; 76: 455-458.
[26]. Pretorius ZA, Rijkenberg FHJ, Wilcoxson RD. Effects of growth stage, leaf position, and temperature on adult-plant resistance
of wheat infected by Puccinia recondita f.sp. tritici. Plant Pathol. 1988; 37: 36-44.
[27]. Memelink J, Linthorst JMH, Schilperoort AR, Hoge CH. Tobacco gene encoding acidic and basic isoforms of pathogenesisrelated proteins display different expression patterns. Plant Mol. Biol. 1990; 14:119-126.
[28]. Koch MF, Mew T W. Effects of plant age and leaf maturity on the quantitative resistance of rice cultivars to Xanthomonas
campestris pv. oryzae. Plant Dis. 1991; 75: 901-904.
[29]. Cohen-Kupiec R, Chet I. The molecular biology of chitin digestion. Curr. Opinion Biotechnol. 1998; 9: 270-277.
[30]. Rupe JC, Gbur Jr EE. Effects of plant age, maturity group, and environment on disease progress of sudden death syndrome of
soybean. Plant Dis. 1995; 79: 139-143.
[31]. Reuveni M. Relationships between leaf age, peroxidase and β-1,3-glucanase activity, and resistance to downy mildew in
grapevines. J. Phytopathol. 1998; 146: 525-530.
[32]. Kus JV, Zaton K, Sarkar R, Cameron R K. Age-related resistance in Arabidopsis is a developmentally regulated defense
response to Pseudomonas syringae. Plant Cell. 2002; 14: 479-490.
[33]. Panter SN, Jones DA. Age-related resistance to plant pathogens. Advances in Botanical Research incorporating Advance in
Plant Pathology. 2002;38: 251-280.
[34]. Gaudet DA, Laroche A, Frick M, Huel R, Puchalski B. Plant development affects the cold-induced expression of plant defenserelated transcripts in winter wheat. Physiol. Mol. Plant Pathol. 2003; 62: 175-184.
[35].Wyatt S, Pan S, Kuc J. β-1,3-Glucanase, chitinase and peroxidase activities in tobacco tissues resistant and susceptible to blue
mould as related to flowering, age and sucker development. Physiol. Mol. Plant Pathol. 1991; 39: 433-440.
[36]. Bartnicki-Garcia S. Cell wall chemistry, morphogeneseis, and taxonomy of fungi. Ann. Rev. Microbiol. 1968; 22: 87-108.
[37]. Broekaert WF, Van Parijs J, Allen A K, Peumans WJ. Comparison of some molecular, enzymatic and antifungal properties of
chitinases from thorn-apple, tobacco and wheat. Physiol. Mol. Plant Pathol. 1988; 33: 319-331.
[38]. Schlumbaum A, Mauch F, Vogeli U, Boller T. Plant chitinases are potent inhibitors of fungal growth. Nature. 1986; 324: 365367.
[39]. Neuhaus JM, Fritig B, Linthorst H J M, Meins F, Jr F M, Mikkelsen J D, Ryals J. A revised nomenclature for chitinase genes.
Plant Mol. Biol. Reporter. 1996; 14: 102-104.
[40]. Barber MS, Bertram RE, Ride JP. Chitin oligosaccharides elicit lignifications in wounded wheat leaves. Physiol. Mol. Plant
Pathol. 1989; 34: 3-12.
[41]. Mauch F, Staehelin L A. Functional implication of the subcellular localization of ethylene-induced chitinase and β-1,3glucanase in bean leaves. Plant Cell. 1989; 1: 447-457.
1050
©FORMATEX 2011
Science against microbial pathogens: communicating current research and technological advances
_______________________________________________________________________________
A. Méndez-Vilas (Ed.)
[42]. Abeles FB, Bosshart RT, Forrense LE, Habig WH. Preparation and purification of glucanase and chitinase from bean leaves.
Plant Physiol. 1970;.47: 129-134.
[43]. Pegg GF. Chitinase from Verticillium alba-atrum. Methods in Enzymology. 1988; 161: 474-479.
[44]. Sela-Buurlage MB, Ponstein AS, Bres-Vloemans SA, Melchers LS, Van Den Elzen PJ M, Cornelissen BJC. Only specific
tobacco (Nicotiana tabacum) chitinases and β-1,3-glucanases exhibit antifungal activity. Plant Physiol. 1993; 101: 857-863.
[45]. Yeh S, Moffatt BA, Griffith M, Xiong F, Yang D S C, Wiseman S B, Sarhan F, Danyluk J, Xue Y Q, Hew C L, Doherty-Kirby
A, Lajoie G. Chitinase genes responsive to cold encode antifreeze proteins in winter cereals. Plant Physiol. 2000; 124:12511263.
[46]. Suarez V, Staehelin C, Arango R, Holtorf H, Hofsteenge J, Meins JF. Substrate specificity and antifungal activity of
recombinant tobacco class I chitinases. Plant Mol. Biol. 2001; 45: 609-618.
[47]. Shih CT, Khan AA, Jia S, Wu J, Shih DS. Purification, characterization, and molecular cloning of a chitinase from the seeds of
Benincasa hispida. Biosci. Biotechnol. Biochem. 2001; 65: 501-509.
[48]. Tsukamoto T, Koga D, Ide A, Ishibashi T, Horino-Matsushige M, Yagshita K, Imoto T. Purification and some properties of
chitinases from yam, Dioscorea posita THUMB. Agric. Biol. Chem. 1984; 48: 931-939.
[49]. Meins F, Fritig B, Linthorst HJ M, Mikkelsen JD, Nauhaus JM,. Ryals J. Plant chitinase genes. Plant, Mol. Biol. Repoter. 1994;
12: 522-528.
[50]. Jacobsen S, Mikkelsen J D, Hejgaard. J. Characterization of two antifungal proteins from barley grain. Physiol. Plant. 1990; 79:
554-562.
[51].Yun D, D’Urzo M P, Abad L, Takeda S, Salzman R, Chen Z, Lee H, Hasegawa P M, Bressan RA. Novel osmotically induced
antifungal chitinases and bacterial expression of an active recombinant isoform. Plant Physiol. 1996; 111: 1219-1225.
[52]. Derckel J, Audran J, Haye B, Lambert B, Legendre L. Characterization, induction by wounding and salicylic acid, and activity
against Botrytis cinerea of chitinases and β-1,3-glucanases of ripening grape berries. Physiol. Plant. 1998; 104: 56-64.
[53]. Giri A P, Harsulkar A M, Patankar A G, Gupta VS, Sainani M N, Deshpande V V, Ranjekar PK. Association of induction of
protease and chitinase in chickpea roots with resistance to Fusarium oxysporum f. sp. ciceri. Plant Pathol. 1998; 47: 693-699.
[54]. Velazhahan R, Samiyappan R, Vidhyasekaran P. Purification of an elicitor-inducible antifungal chitinase from suspensioncultured rice cells. Phytoparasitica. 2000; 28: 131-139.
[55]. Ji C, Norton R A, Wicklow D T, Dowd PF. Isoform patterns of chitinase and β-1,3-glucanase in maturing corn kernels (Zea
mays L.) associated with Aspergillus flavus milk stage induction. J. Agric. Food Chem. 2000; 48: 507-511.
[56]. Nielsen KK, Mikkelsen JD, Dragh KM, Bojsen K. An acidic class III chitinase in sugar beet: induction by Cercospora beticola,
characterization, and expression in transgenic tobacco plants. Mol. Plant Microbe Interact. 1993; 6: 495-506.
[57]. Anguelova M V, Westhuizen VD, Pretorius ZA. Beta-1,3-glucanase and chitinase activities and the resistance response of
wheat to leaf rust. J. Phytopathol. 2001; 149: 381-384.
[58]. Lawrence CB, Singh NP, Qui J, Gardner RG, Tuzun S. Constitutive hydrolytic enzymes are associated with polygenic
resistance of tomato to Alternaria solani and may function as an elicitor release mechanism. Physiol. Mol. Plant Pathol. 2000;
57: 211-220.
[59]. McFadden HG, Chapple R, de Feyter R, Dennis E. Expression of pathogenesis-related genes in cotton stems in response to
infection by Verticillium dahliae. Physiol. Mol. Plant Pathol. 2001; 58: 119-131.
[60]. Kragh KM, Jacobsen S, Mikkelsen J D, Nielsen K A. Tissue specificity and induction of class I, II, and III chitinases in barley
(Hordeum vulgare). Physiologia Plantarum. 1993; 89: 490-498.
[61]. Metraux JP, Boller T. Local and systemic induction of chitinase in cucumber plants in response to viral, bacterial and fungal
infections. Physiol. Mol. Plant Pathol. 1986; 28: 161-169.
[62]. O'Garro LW, Charlemange E. Comparison of bacterial growth and activity glucanase and chitinase in pepper leaf and floral
infected with Xanthomonas campestris pv. Vesicatoria. Phsiol. and Mol. Plant Pathology. 1994; 45: 181-188.
[63]. Busam G, Kassemeyer H, Matern U. Differential expression of chitinases in Vitis vinifera L. responding to systemic acquired
resistance activators or fungal challenge. Plant Physiol. 1997; 115: 1029-1038.
[64]. Broglie K, Chet I, Holliday M, Cressman R, Biddle P, Knowlton S, Mauvais C J, Broglie R. Transgenic plants with enhanced
resistance to the fungal pathogen Rhizoctonia solani. Science. 1991; 254: 1194-1197.
[65]. Grison R, Grezes-Besset B, Schneider M, Lucante N, Olsen L, Leguay J, Toppan A. Field tolerance to fungal pathogens of
Brassica napus constitutively expressing a chimeric chitinase gene. Nature Biotechnol. 1996; 114: 643-646.
[66]. Yamamoto T, Iketani H, Leki H, Nishizawa Y, Notsuka K, Hibi T, Hayashi T, Matsuta N. Transgenic grapevine plants
expressing a rice chitinase with enhanced resistance to fungal pathogens. Plant Cell Reports. 2000; 19: 639-646.
[67]. Datta K, Tu J, Oliva N, Ona I, Velazhahan R, Mew T W, Muthukrishnan S, Datta SK. Enhanced resistance to sheath blight by
constitutive expression of infection-related rice chitinase in transgenic elite indica rice cultivars. Plant Sci. 2001; 160: 405-414.
[68]. Rohini VK, Rao S. Transformation of peanut (Arachi hypogaea L.) with tobacco chitinase gene: variable response of
transformants to leaf spot disease. Plant Sci. 2001; 160: 889-898.
[69]. Lorito M, Woo S L, Fernandez IG, Colucci G, Harman GE, Pintor-Toro J A, Filippone E, Muccifora S, Lawrence CB, Zoina A,
Tuzun S, Scala F. Genes from mycoparasitic fungi as a source for improving plant resistance to fungal pathogens. Proc. Natl.
Acad. Sci. 1998; 95: 7860-7865.
[70].Terakawa T, Takaya N, Horiuchi H, Koike M, Takagi M. A fungal chitinase gene from Rhizopus oligosporus confers antifungal
activity to transgenic tobacco. Plant Cell Rep. 1997; 16: 439-443.
[71]. Boller T. Induction of hydrolases as a defense reaction against pathogens. In: Key, J. L., and Kosuge, T., Eds. Cellular and
molecular biology of plant stress. Liss, New York, p. 247-262; 1985.
[72]. Pan SQ, Ye XS, Kuc J. Direct detection of β-1,3-glucanase isozymes on polyacrylamide electrophoresis and isoelectrofocusing
gels. Anal. Biochem. 1989; 182: 136-140.
[73]. Simmons CR. The physiology and molecular biology of plant 1,3-β-D-glucanases and 1,3;1,4-β-D-glucanases. Crit. Rev. Plant
Sci. 1994; 13: 325-387.
©FORMATEX 2011
1051
Science against microbial pathogens: communicating current research and technological advances
______________________________________________________________________________
A. Méndez-Vilas (Ed.)
[74].Wessels JGH, Sietsma JH. Fungal cell wall: a survey, in: W. Tanner, F. A. Loewus (Eds.), Encyclopedia of Plant Physiology,
New Series, Plant Carbohydrates II, vol. 13B, Springer-Verlag, p. 352-394; 1981.
[75].Adams DJ. Fungal cell wall chitinases and glucanases. Microbiology. 2004; 150: 2029-2035.
[76].Heslop-Harrison J. Cell walls, cell membranes and protoplasmic connections during meiosis and pollen development. In:
Linskens, H. ed. Pollen physiology and fertilization. North-Holland, Amsterdam, p. 29-47; 1964.
[77].Yim KO, Bradford KJ. Callose deposition is responsible for apoplastic semipermeability of the endosperm envelope of
muskmelon seeds. Plant Physiol. 1998; 118: 83-90.
[78].Masuda Y, Wada S. Effects of β-1,3-glucanase on the elongation growth of oat coleoptile. Bot. Mag. 1967; 80: 100-102.
[79].Heyn A N J. Glucanase activity in coleoptile of Avena. Arch. Biochem. Biophys. 1969; 132: 442-449.
[80].Fulcher RG, Mc Cully ME, Setterfield G, Sutherland J. β-1,3-Glucans may be associated with cell plate formation during
cytokinesis. Can. J. Bot. 1976; 54: 459-542.
[81].Hinton D M, Pressey R. Glucanase in fruits and vegetables. J. Amer. Soc. Hort. Sci. 1980; 105: 499-502.
[82].Roggen HPJR, Stanley RG. Cell-wall-hydrolysing enzymes in wall formation as measured by pollen-tube extension. Planta.
1969; 84: 295-303.
[83]. Meikle PJ, Bonig I, Hoogenraad NJ, Clarke AE, Stone BA. The location of (1-3)-β-glucans in the walls of pollen tubes of
Nicotiana alata using a (1-3)-β-glucan specific monoclonal antibody. Planta. 1991; 185: 1-8.
[84]. Ori N, Sessa G, Lotan T, Himmelhoch S, Fluhr R. A major stylar matrix polypeptide (Sp41) is a member of the pathogenesisrelated proteins superclass. EMBO J. 1990; 9: 3429-3436.
[85].Helleboid S, Chapman A, Hendricks T, Inze D, Vasseur J, Hilbert J L. Cloning of β-1,3-glucanases expressed during Cichorium
somatic embryogenesis. Plant Mol. Biol. 2000; 42: 377-386.
[86].Morohashi Y, Matsushima H. Development of β-1,3-glucanase activity in germinated tomato seeds. J. Exp. Bot. 2000; 51: 13811387.
[87].Buchner P, Rochat C, Wuilleme S, Boutin JP. Characterization of a tissue specific and developmentally regulated β-1,3glucanase gene in pea (Pisum sativum). Plant Mol. Biol. 2002; 49: 171-186.
[88].Neale AD, Wahleithner JA, Lund M, Bonnett H T, Kelly A, Meeks-Wagner DR, Peacock W J, Denis E S. Chitinase, β-1,3glucanase, osmotin, and extensin are expressed in tobacco explants during flower formation. Plant Cell. 1990; 2: 673-684.
[89].Akiyama T, Pillai M A, Sentoku N. Cloning, characterization and expression of OsGLN2, a rice endo-1,3-β-glucanase gene
regulated developmentally in flowers and hormonally in germinating seeds. Planta. 2004; 220: 129-139.
[90].Kauffmann S, Legrand M, Geoffroy P, Fritig B. Biological function of pathogenesis-related proteins: four PR proteins of
tobacco have 1,3-β-glucanase activity. EMBO J. 1987; 6: 3209-3212.
[91].Linthorst HJM. Pathogenesis-related proteins of plants. Crit. Rev. Plant Sci. 1991; 10: 123-150.
[92].Cordero MJ, Raventos D, San Segundo B. Differential expression and induction of chitinases and β-1,3-glucanases in response
to fungal infection during germination of maize seeds. Mol. Plant Microbe Interact.1994; 7: 23-31.
[93].Keen NT, Yoshikawa M. β-1,3-Endoglucanase from soybean releases elicitoractive carbohydrates from fungal cell walls. Plant
Physiol. 1983; 7: 460-465.
[94].Ham KS, Kauffmann S, Albersheim P, Darvill A G Host-pathogen interactions; A soybean pathogenesis-related protein with β1,3-glucanase activity releases phytoalexin elicitor-active heat stable fragments from fungal walls. Mol Plant Microbe Interact.
1991; 4: 545-552.
[95].Klarzynski O, Plesse B, Joubert JM, Yvin JC, Kopp M, Kloareg B, Fritig B. Linear β-1,3-glucans are elicitors of defense
responses in tobacco. Plant Physiol. 2000; 124: 1027-1037.
[96].De Loose M, Alliotte T, Gheysen G, Genetello C, Gielen J, Soetaert P, Van Montagu M, Inze D. Primary structure of a
hormonally regulated β-glucanase of Nicotiana plumbaginifolia. Gene. 1988; 70: 13-23.
[97].Takeuchi Y, Yoshikawa M, Takeba G, Tanaka K, Shibata D, Horino O. Molecular cloning and ethylene induction of mRNA
encoding a phytoalexin elicitorreleasing factor, β-1,3-endoglucanase, in soybean. Plant Physiol. 1990; 93: 673-682.
[98].Chye ML, Cheung KY. β-1, 3-Glucanase is highly-expressed in laticifers of Hevea brasiliensis. Plant Mol. Biol. 1995; 29: 397402.
[99].Peumans WJ, Barre A, Derycke V, Rouge P, Zhang W, May GD, Delcour JA, Van Leuven F, Van Damme EJ. Purification,
characterization and structural analysis of an abundant β-1,3-glucanase from banana fruit. Eur. J. Biochem. 2000; 267: 11881195.
[100].Yamaguchi T, Nakayama K, Hayashi T, Tanaka Y, Koike S. Molecular cloning and characterization of a novel β-1,3-glucanase
gene from rice. Biosci. Biotechnol. Biochem. 2002; 66: 1403-1406.
[101].Payne G, Ward E, Gaffney T, Goy PA, Moyer M, Harper A, Meins FJ, Ryals J. Evidence for a third structural class of β-1, 3glucanase in tobacco. Plant Mol. Biol. 1990; 15: 797-808.
[102].Ward ER, Uknes SJ, Williams S C, Dincher SS, Wiederhold DL, Alexander DC, Ahl-Goy P, Metraux J P, Ryals JA.
Coordinate gene activity in response to agents that induce systemic acquired resistance. Plant Cell. 1991; 3: 1085-1094.
[103].Van Kan JAL, Joosten MHAJ, Wagemakers CAM. Differential accumulation of mRNAs encoding extracellular and
intracellular PR proteins in tomato induced by virulent and avirulent races of Cladosporium fulvum. Plant Mol. Biol. 1992; 20:
513-527.
[103].Beerhues L, Kombrink E. Primary structure and expression of mRNAs encoding basic chitinase and 1,3-β-glucanase in potato.
Plant Mol. Biol. 1994; 24: 353-367.
[103].Domingo C, Conejero V, Vera P. Genes encoding acidic and basis class III β-1,3-glucanases are expressed in tomato plants
upon viroid infection. Plant Mol. Biol. 1994; 24: 725-732.
[104].Oh H Y, Yang M S. Nucleotide sequence of genomic DNA encoding the potato β-1,3-glucanase. Plant Physiol. 1995; 107:
1453.
[105].Buchel AS, Linthorst HJM. PR-1: A group of plant proteins induced upon pathogen infection. In: Pathogenesis-related proteins
in plants. Eds. S.K. Datta, S. Muthukrishnan, CRC Press LLC, Boca Raton, p. 21-47; 1999
1052
©FORMATEX 2011
Science against microbial pathogens: communicating current research and technological advances
_______________________________________________________________________________
A. Méndez-Vilas (Ed.)
[106].Shinshi H, Wenzler H, Neuhaus JM, Felix G, Hofsteenge J, Meins FJ. Evidence of amino and carboxyl-terminal processing of
a plant defense-related enzyme primary structure of tobacco β-1,3-glucanase. Proc. Natl. Acad. Sci. 1988; 85: 5541-5546.
[107].Bucciaglia.PA, Smith. A G. Cloning and characterization of Tag1, a tobacco anther β-1,3-glucanase expressed during tetrad
dissolution. Plant Mol. Biol. 1994; 24: 903-914.
[108].Joosten M H A J, Verbakel H M, Nettekoven M E, Vanleeuwen J, Vandervossen R T M, Dewit P J G M. The phytopathogenic
fungus Cladosporium fulvum is not sensitive to the chitinase and β-1,3-glucanase defense proteins of its host, tomato. Physiol.
Mol. Plant Pathol. 1995; 46: 45-59.
[109].Anfoka G, Buchenauer H. Systemic acquired resistance in tomato against Phytophthora infestans by pre-inoculation with
tobacco necrosis virus. Physiol. Mol. Plant Pathol. 1997; 50: 85-102.
[110].Vogeli-lange R, Frybdt C, Gart CM, Nagt FM, Meins FJ. Developmental, hormonal, and pathogenesis-related regulation of the
tobacco class I β-glucanase B promoter. Plant Mol. Biol. 1994; 25: 299-311.
[111].Livne B, Faktor O, Zeitoune S, Edelbaum O, Sela I. TMV-induced expression of tobacco β-glucanase promoter activity is
mediated by a single, inverted,GCC motif. Plant Sci. 1997; 130: 159-169.
[112].Ryan CA, Farmer EE. Oligosaccharide signals in plants: a current assessment. Annu. Rev. Plant Physiol. Mol. Biol. 1991; 42:
651-674.
[113].Castresana C, De Carvalho F Gheysen G, Habets M, Inze D, Van Montagu M. Tissue-specific and pathogen-induced regulation
of a Nicotiana plumbaginifolia β-1,3-glucanase gene. Plant Cell. 1990; 2: 1131-1144.
[114].Alonso E, De Carvalho Niebel F, Obregon P, Gheysen G, Inze D, Van Montagu M, Castresana C. Differential in vitro DNA
binding activity to a promoter element of the gn1 β-1,3-glucanase gene in hyper sensitively reacting tobacco plants. Plant J.
1995; 7: 309-320.
[115].Roulin S, Xu P, Brown AHD, Fincher GB. Expression of specific (1-3)-β-glucanase genes in leaves of near-isogenic resistant
and susceptible barley lines infected with the leaf scald fungus (Rhynchosporium secalis). Physiol. Mol. Plant Pathol. 1997; 50:
245-261.
[116].Bol JF, Buchel AS, Knoester M, Baladin T, Van Loon LC, Linthorst HJM. Regulation of the expression of plant defense genes.
Plant Growth Regul. 1996; 18: 87-91.
[117].Lusso M, Kuc J. Evidence for transcriptional regulation of β-1,3-glucanase as it relates to induced systemic resistance of
tobacco to blue mold. Mol. Plant Microbe Interact. 1995; 8: 473-475.
[118].Tuzun SS, Rao MN, Vogelie U, Schardl CL, Kuc J. Induced systemic resistance to blue mold: Early induction and
accumulation of β-1,3-glucanases, chitinases, and other pathogenesis-related proteins (β-proteins) in immunized tobacco.
Phytopathology. 1989; 79: 979-983.
[119].Ward ER, Payne GB, Moyer M B, Williams SC, Dincher S S, Sharkey K C, Beck J J, Taylor HT, Ahl-Goy P, Meins F J, Ryals
J A. Differential regulation of β-1,3-glucanase messenger RNAs in response to pathogen infection. Plant Physiol. 1991; 96:
390-397.
[120].Philip S, Joseph A, Kumar A, Jacob C, Kothandaraman R. Detection of β-1,3-glucanase isoforms against Corynespora leaf
disease of rubber (Hevea brasiliensis). Indian J. Nat. Rubber Res. 2001; 14: 1-6.
[121].Ignatius SMJ, Chopra R K. Effects of fungal infection and wounding on the expression of chitinase and β-1,3-glucanases in
near-isogenic lines of barley. Physiol. Plant. 1994; 90: 584-592.
[122].Lozovaya VV, Waranyuwat A, Widholm JM. β-1,3-glucanase and resistance to Aspergillus flavus infection in maize. Crop Sci.
1998. 38: 1255-1260.
[123].Jung HW, Hwang BK. Pepper gene encoding a basic β-1,3-glucanase is differentially expressed in pepper tissues upon
pathogen infection and ethephon or methyl jasmonate treatment. Plant Sci. 2000; 159: 97-106.
[124].Hanselle T, Barz W. Purification and characterization of the extracellular PR-2b β-1,3-glucanase accumulating in different
Ascochyta rabiei-infected chickpea (Cicer arietinum L.) cultivars. Plant Sci. 2001; 161: 773-781.
[125].Zemanek AB, Ko TS, Thimmapuram J, Hammerschlag FA, Korban S S. Changes in β-1,3-glucanase mRNA levels in peach in
response to treatment with pathogen culture filtrates, wounding, and other elicitors. J. Plant Physiol. 2002; 159: 877-889.
[126].Chang MM, Hadwiger LA, Horovitz D. Molecular characterization of a pea β-1, 3-glucanase induced by Fusarium solani and
chitosan challenge. Plant Mol. Biol. 1992; 20: 609-618.
[127].Kaku H, Shibuya N, Xu PL, Aryan AP, Fincher GB. N-acetylchitooligosaccharides elicit expression of a single (1-3)-βglucanase gene in suspension-cultured cells from barley (Hordeum vulgare). Physiol. Plant. 1997; 100: 111-118.
[128].Muench-Garthoff S, Neuhaus JM, Boller T, Kemmerling B, Kogel KH. Expression of β-1,3-glucanase and chitinase in healthy,
stem-rust-affected and licitortreated near-isogenic wheat lines showing Sr5- or Sr24-specified race-specific rust resistance.
Planta. 1997; 201: 235-244.
[129].Rezzonico E, Flury N, Meins F, Beffa R. Transcriptional down-regulation by abscisic acid of pathogenesis-related β-1,3glucanase genes in tobacco cell cultures. Plant Physiol. 1998; 117: 585-592.
[130].Akiyama T, Pillai MA. Molecular cloning, characterization and in vitro expression of a novel endo-1,3-β-glucanase upregulated by ABA and drought stress in rice (Oryza sativa L.). Plant Sci. 2001; 161: 1089-1098.
[131].Wu J, Khan AA, Shih CT, Shih DS. Cloning and sequence determination of a gene encoding an osmotin-like protein from
strawberry (Fragaria ananassa Ducth.). DNA Seq., 2001; 12: 447-453.
[132].Wu CT, Bradford KJ. Class I chitinase and β-1,3-glucanase are differentially regulated by wounding, methyl jasmonate,
ethylene, and gibberellin in tomato seeds and leaves. Plant Physiol. 2003; 133: 263-273.
[133].Fecht-Christoffers MM, Braun HP, Lemaitre-Guillier C, Van Dorsselaer A, Horst WJ. Effect of manganese toxicity on the
proteome of the leaf apoplast in cowpea. Plant Physiol. 2003; 133: 1935-1946.
[134].Sandermann H, Ernst D, Heller W, Langebartels C. Ozone: an abiotic elicitor of plant defense reactions. Trends Plant Sci.
1998; 3: 47-50.
[135].Thalmair M, Bauw G, Thiel S, Dohring T, Langebartels C, Sandermann H. Ozone and ultraviolet B effects on the defenserelated proteins β-1,3-glucanase and chitinase in tobacco. J. Plant Physiol. 1996; 148: 222-228.
©FORMATEX 2011
1053
Science against microbial pathogens: communicating current research and technological advances
______________________________________________________________________________
A. Méndez-Vilas (Ed.)
[136].Southerton SG, Deverall BJ. Histochemical and chemical evidence for lignin accumulation during the expression of resistance
to leaf rust fungi in wheat. Physiol. and Molecul. Plant Pathology. 1990; 36: 483-494.
[137].Lewis NG. A 20th century roller coaster ride: a short account of lignification. Current Opinion in Plant Biology. 1999; 2: 153162.
[138].Vance C P, Kirk TK, Sherwood R T. Lignification as a mechanism of disease resistance. Annual Review of Phytopathology.
1980; 18: 259-288.
[139] Nicholson RL, Hammerschmidt R. Phenolic compounds and their role in disease resistance. Annul. Rev. Phytopathol. 1992; 30:
369-389.
[140].Zeyen RJ, Carver T L W, Lyngkjaer M F. Epidermal cell papillae. In: Belanger, R. R. and Bushnell, W. R., eds. The powdery
mildew: a comprehensive treatise. APS Press, MN, USA, p. 107-125; 2002.
[141].Boerjan W, Ralph J, Baucher M. Lignin biosynthesis. Annu. Rev. Plant Biol. 2003, 54: 519-546.
[142].Lange BM, Lapierre C, Sandermann HJ. Elicitor induced spruce stress lignin. Plant Physiol. 1995; 108:1277-87.
[143].Ride JP. The role of cell wall alterations in resistance to fungi. Ann. Appl. Biol. 1978; 89: 302-306.
[144].Ride JP, Pearce RB. Lignification and papilla formation at site of attempted penetration of wheat leaves by nonpathogenic
fungi. Physiol. Plant Pathol. 1979; 5: 79-92.
[145].Keen NT, Littlefield L J. The possible association of phytoalexins with resistance gene expression in flax to Melamspora lini.
Physiol. Plant Pathol. 1979; 14: 265-280.
[146].Hammerschmidt R. Rapid deposition of lignin in potato tuber tissue as a response to fungi non-pathogenic on potato.
Physiological Plant Pathology. 1984; 24: 33-42.
[147].Bhuiyan NH, Selvaraj G, Wei Y, King J. Role of lignification in plant defense. Plant Signaling and Behavior, 2009; 4: 158159.
1054
©FORMATEX 2011
Download