search
for
 About Bioline  All Journals  Testimonials  Membership  News


Electronic Journal of Biotechnology
Universidad Católica de Valparaíso
ISSN: 0717-3458
Vol. 5, Num. 1, 2002, pp. 93-109

Electronic Journal of Biotechnology, Vol. 5, No. 1, April, 2002

REVIEW ARTICLE

Plant protease inhibitors in control of phytophagous insects

Paulraj K. Lawrence*1 and Kripa Ram Koundal 2

1Department of Biological Sciences, Texas Tech University, Box No 43131 Lubbock, TX 79409-3131, USA Tel: 806 742 2740 Fax: 806 742 2963 E-mail: p.k.lawrence@ttu.edu
2National Research Center on Plant Biotechnology, Indian Agricultural Research Institute, PUSA, New Delhi, India-110012, India, Tel: 91 011 5711554 Fax: 91 011 5766420 E-mail: krk_pbio@iari.ernet.in
* Corresponding author

Received August 20, 2001
Accepted April 1, 2002

Code Number: ej02015

Abstract

Plant proteinase inhibitors (PIs) have been well established to play a potent defensive role against predators and pathogens. Although diverse endogenous functions for these proteins has been proposed, ranging from regulators of endogenous proteinases to act as storage proteins, evidence for many of these roles is partial, or confined to isolated examples. On the other hand, many PIs have been shown to act as defensive compounds against pests by direct assay or by expression in transgenic crop plants, and a body of evidence for their role in plant defense has been accumulated consistently. The role and mechanism of action for most of these inhibitors are being studied in detail and their respective genes isolated. These genes have been used for the construction of transgenic crop plants to be incorporated in integrated pest management programmes. This article describes the classes of protease inhibitors, their regulation and genes used to construct transgenic plants against phytophagous insects.

Keywords: crop pests, protease inhibitors, transgenic plants.

Article

World wide crop losses without the use of pesticides and other non-chemical control strategies is estimated to be about 70% of crop production, amounting to US $ 400 billion. The world wide pre-harvest losses due to insect pests, despite, the use of insecticides is 15% of total production representing over US $ 100 billion (Krattiger, 1997). The annual cost of insect control itself amounts to US $ 8 billion, thus warranting urgent economical control measures. Many of the crop varieties developed in the past 30 years were high yielders, but had poor storage characteristics (Kerin, 1994). Insect pests are capable of evolving to biotypes that can adapt to new situations, for instance, they overcome the effect of toxic materials or bypass natural or artificial plant resistance, which further confounds the problem (Roush and McKenzie, 1987). Under these circumstances, provision of food to the rapidly expanding population has always been a challenge facing mankind. This problem is more acute in the tropics and sub-tropics, where the climate provides a highly condusive environment for a wide range of insects and necessitates massive efforts to suppress the population densities of different pests in order to achieve an adequate supply of food. In developing countries, the problem of competition from insect pests is further complicated with a rapid annual increase in the human population (2.5-3.0 percentage) in comparison to a 1.0 percent increase in food production. In order to feed the ever expanding population, crop protection plays a vital and integral role in the modern day agricultural production to minimise yield losses. Currently, the crop protection practice in such agricultural systems relies exclusively on the use of agrochemicals, although a few specific cases do exist where inherent varietal resistance and biological control have been successfully employed. The exclusive use of chemical pesticides not only results in rapid build-up of resistance to such compounds, but their non-selectivity affects the balance between pests and natural predators, and is generally in favour of pests (Metcalf, 1986). Therefore, an integrated pest management (IPM) programme, comprising a combination of practices including the judicious use of pesticides, crop rotation, field sanitation and above all exploitation of inherently resistant plant varieties would provide the best option (Meiners and Elden, 1978). The last option includes the use of transgenic crops, expressing foreign insecticidal genes which could make a significant contribution to sustainable agriculture and thus could be an important component of IPM (Boulter, 1993). The production of transgenic crops has seen rapid advances during the last decade with the commercial introduction of Bt transgenics, but the major concern with these crops has been the development of resistance by pest and public acceptability. Hence, there has been a need to discover new effective plant genes which would offer resistance/protection against these pests. Protease inhibitors (PIs) are one of the prime candidates with highly proven inhibitory activity against insect pests and also known to improve the nutritional quality of food.

Plant protease inhibitors

The possible role of protease inhibitors (PIs) in plant protection was investigated as early as 1947 when, Mickel and Standish observed that the larvae of certain insects were unable to develop normally on soybean products. Subsequently the trypsin inhibitors present in soybean were shown to be toxic to the larvae of flour beetle, Tribolium confusum (Lipke et al. 1954). Following these early studies, there have been many examples of protease inhibitors active against certain insect species, both in in vitro assays against insect gut proteases (Pannetier et al. 1997; Koiwa et al, 1998) and in in vivo artificial diet bioassays (Urwin et al. 1997; Vain et al. 1998). The term “protease” includes both “endopeptidases” and “exopeptidases” whereas, the term “proteinase” is used to describe only “endopeptidases” (Ryan, 1990). Several non-homologous families of proteinase inhibitors are recognized among the animal, microorganisms and plant kingdom. Majority of proteinase inhibitors studied in plant kingdom originates from three main families namely leguminosae, solanaceae and gramineae (Richardson, 1991).

These protease inhibitor genes have practical advantages over genes encoding for complex pathways i.e. by transferring single defensive gene from one plant species to another and expressing them from their own wound inducible or constitutive promoters thereby imparting resistance against insect pests (Boulter, 1993). This was first demonstrated by Hilder et al. 1987 by transferring trypsin inhibitor gene from Vigna unguiculata to tobacco, which conferred resistance to wide range of insect pests including lepidopterans, such as Heliothis and Spodoptera, coleopterans such as Diabrotica, Anthonomnous and orthoptera such as Locusts. Further, there is no evidence that it had toxic or deleterious effects on mammals. Many of these protease inhibitors are rich in cysteine and lysine, contributing to better and enhanced nutritional quality (Ryan, 1989). Protease inhibitors also exhibit a very broad spectrum of activity including suppression of pathogenic nematodes like Globodera tabaccum, G. pallida, and Meloidogyne incognita by CpTi (Williamson and Hussey, 1996), inhibition of spore germination and mycelium growth of Alternaria alternata by buckwheat trypsin/chymotrypsin (Dunaevskii et al. 1997) and cysteine PIs from pearl millet inhibit growth of many pathogenic fungi including Trichoderma reesei (Joshi et al. 1998). These advantages make protease inhibitors an ideal choice to be used in developing transgenic crops resistant to insect pests. Further, transformation of plant genomes with PI-encoding cDNA clones appears attractive not only for the control of plant pests and pathogens, but also as a means to produce PIs, useful in alternative systems and the use of plants as factories for the production of heterologous proteins (Sardana et al. 1998). These inhibitor families that have been found are specific for each of the four mechanistic classes of proteolytic enzymes, and based on the active amino acid in their “reaction center” (Koiwa et al. 1997), are classified as serine, cysteine, aspartic and metallo-proteases.

Serine proteinase inhibitors

The role of serine PIs as defensive compounds against predators is particularly well established, since the major proteinases present in plants, used for processes such as protein mobilization in storage tissues, contain a cysteine residue as the catalytically active nucleophile in the enzyme active site. Serine proteinases are not used by plants in processes involving large scale protein digestion, and hence the presence of significant quantities of inhibitors with specificity towards these enzymes in plants cannot be used for the purposes of regulating endogenous proteinase activity (Reeck et al. 1997). In contrast, a major role for serine PIs in animals is to block the activity of endogenous proteinases in tissues where this activity would be harmful, as in case of pancreatic trypsin inhibitors found in mammals. The serine class of proteinases such as trypsin, chymotrypsin and elastase, which belong to a common protein superfamily, are responsible for the initial digestion of proteins in the gut of most higher animals (GarciaOlmedo et al. 1987). In vivo they are used to cleave long, essentially intact polypeptide chains into short peptides which are then acted upon by exopeptidases to generate amino acids, the end products of protein digestion. These three types of digestive serine proteinases are distinguished based on their specificity, trypsin specifically cleaving the C-terminal to residues carrying a basic side chain (Lys, Arg), chymotrypsin showing a preference for cleaving C-terminal to residues carrying a large hydrophobic side chain (Phe, Tyr, Leu), and elastase showing a preference for cleaving C-terminal to residues carrying a small neutral side chain (Ala, Gly) (Ryan, 1990). Inhibitors of these serine proteinases have been described in many plant species, and are universal throughout the plant kingdom, with trypsin inhibitors being the most common type. At least, part of this bias can be accounted for by the fact that (mammalian) trypsin is readily available and is the easiest of all the proteinases to assay using synthetic substrates, and hence is used in screening procedures. Because of these reasons the members of the serine class of proteinases have been the subject of intense research than any other class of proteinase inhibitors. Such studies have provided a basic understanding of the mechanism of action (Huber and Carrell, 1989) that applies to most serine proteinase inhibitor families and probably to the cysteine and aspartyl proteinase inhibitor families as well. All serine inhibitor families from plants are competitive inhibitors and all of them inhibit proteinases with a similar standard mechanism (Laskowski and Kato, 1980).

Serine proteinases have been identified in extracts from the digestive tracts of insects from many families, particularly those of lepidoptera (Houseman et al. 1989) and many of these enzymes are inhibited by proteinase inhibitors. The order lepidoptera, which includes a number of crop pests, the pH optima of the guts are in the alkaline range of 9-11 (Applebaum, 1985) where, serine proteinases and metallo-exopeptidases are most active. Additionally, serine proteinase inhibitors have anti-nutritional effects against several lepidopteran insect species (Shulke and Murdock, 1983; Applebaum, 1985). Purified Bowman-Birk trypsin inhibitor (Brovosky, 1986) at 5% of the diet inhibited growth of these larvae but SBTI (Kunitz, 1945), another inhibitor of bovine trypsin, was less effective when fed at the same levels.

Broadway and Duffey (1986a) compared the effects of purified SBTI and potato inhibitor II (an inhibitor of both trypsin and chymotrypsin) on the growth and digestive physiology of larvae of Heliothis zea and Spodoptera exigua and demonstrated that growth of larvae was inhibited at levels of 10% of the proteins in their diet. Trypsin inhibitors at 10% of the diet were toxic to larvae of the Callosobruchus maculatus (Gatehouse and Boulter, 1983) and Manduca sexta (Shulke and Murdock, 1983).

Recent X-ray crystallography structure of winged bean, Psophocarpus tetragonolobus Kunitz-type double headed alpha-chymotrypsin shows 12 anti-parallel beta strands joined in a form of beta trefoil with two reactive site regions (Asn 38-Leu 43 and Gln 63-Phe 68) at the external loops (Ravichandaran et al. 1999; Mukhopadhyay, 2000). Structural analysis of the Indian finger millet (Eleusine coracana) bifunctional inhibitor of alpha-amylase/trypsin with 122 amino acids has shown five disulphide bridges and a trypsin binding loop (Gourinath et al. 2000). These structural analysis would greatly help in “enzyme engineering” of the native PIs to a potent form, against the target pest species than the native PIs.

Cysteine proteinase inhibitors

Isolation of the midgut proteinases from the larvae of cowpea weevil, C. maculatus (Kitch and Murdock, 1986; Campos et al. 1989) and bruchid Zabrotes subfaceatus (Lemos et al. 1987) confirmed the presence of cysteine mechanistic class of proteinase inhibitors. Similar proteinases have been isolated from midguts of the flour beetle Tribolium castaneum, Mexican beetle Epilachna varivestis (Murdock et al. 1987) and the bean weevil Ascanthoscelides obtectus (Wieman and Nielsen, 1988). Cysteine proteinases isolated from insect larvae are inhibited by both synthetic and naturally occurring cysteine proteinase inhibitors (Wolfson and Murdock, 1987). In a study of the proteinases, from the midguts of several members of the order coleopteran, 10 of 11 beetle species representing 11 different families, had gut proteinases that were inhibited by p-chloromercuribenzene sulfonic acid (PCMBS), a potent sulphydryl reagent (Murdock et al. 1988) indicating that the proteinases were of the cysteine mechanistic class. The optimum activity of cysteine proteinases is usually in the pH range of 5-7, which is the pH range of the insect gut that use cysteine proteinases (Murdock et al. 1987). Another puzzling aspect of studies with C. maculatus is the apparent effects of certain members of Bowman-Birk trypsin inhibitor family on the growth and development of these larvae. Although cysteine proteinase is primarily responsible for protein digestion in C. maculatus, it is not clear, how the cowpea and soybean Bowman-Birk inhibitors are exert their anti-nutritional effects on this organism. Advances in enzymology has revealed the existence of a variety of cysteine proteinases resulting in their classification into several families namely papain, calpin and asparagines specific processing enzyme (Turk and Bode, 1991). Cystanins have also been characterized from potato (Waldron et al. 1993), ragweed (Rogers et al. 1993), cowpea (Fernandes et al. 1993) papaya (Song et al. 1995) and avacado (Kimura et al. 1995).

The rice cysteine proteinase inhibitors are the most studied of all the cysteine PIs which is proteinaceous in nature (Abe and Arai, 1985) and highly heat stable (Abe et al. 1987). Recent three dimensional structure analysis of oryzacystatin OC-I by Tanokura’s group (Nagata et al. 2000), using NMR has showed a well defined main body consisting of amino acids from Glu 13 - Asp 97 and an alpha helix with five stranded anti parallel beta-sheet, while the N terminus (Ser 2-Val 12) and C terminus (Ala 98-Ala 102) are less defined. Further, analysis has demonstrated OC-I to be similar to chicken cystatin which belongs to type-2 animal cystatin. Another rice cystatin named as OC-II, with a putative target binding motif gln-x-val-x-gly shares similar motif with OC-I but has a different inhibition constant (Ki) value (Arai et al. 1991; Kondo et al. 1991).

Aspartic and metallo-proteinase inhibitors

Knowledge on the role of aspartic proteinases in insect digestion is limited than that of cysteine proteinases. In species of six families of the order hemiptera, aspartic proteinases (cathepsin D-like proteinases) were found along with cysteine proteinases (Houseman and Downe, 1983). The low pH of midguts of many members of coleoptera and hemiptera provides more favourable environments for aspartic proteinases (pH optima ~ 3-5) than the high pH of most insect guts (pH optima ~ 8-11) (Houseman et al. 1987) where the aspartic and cysteine proteinases would not be active. No aspartic proteinases have been isolated from coleoptera but Wolfson and Murdock 1987 demonstrated that pepstatin, a powerful and specific inhibitor of aspartyl proteinases, strongly inhibited proteolysis of the midgut enzymes of Colorado potato beetle, Leptinotarsa decemlineata indicating that an aspartic proteinase was present in the midgut extracts. Potato tubers possess an aspartic proteinase inhibitor, cathepsin D (Mares et al. 1989) that shares considerable amino acid sequence identity with the trypsin inhibitor SBTI from soybeans. Plants have also evolved at least two families of metallo-proteinase inhibitors, the metallo-carboxypeptidase inhibitor family in potato (Rancour and Ryan, 1968), and tomato plants (Graham and Ryan, 1981) and a cathepsin D inhibitor family in potatoes (Keilova and Tomasek, 1976).

The cathepsin D inhibitor (27 kDa) is unusual as it inhibits trypsin and chymotrypsin as well as cathepsin D, but does not inhibit aspartyl proteases such as pepsin, rennin or cathepsin E. The inhibitors of the metallo-carboxypeptidase from tissue of tomato and potato are polypeptides (4 kDa) that strongly and competitively inhibit a broad spectrum of carboxypeptidases from both animals and microorganisms, but not the serine carboxypeptidases from yeast and plants (Havkioja and Neuvonen, 1985). The inhibitor is found in tissues of potato tubers where it accumulates during tuber development along with potato inhibitor I and II families of serine proteinase inhibitor. The inhibitor also accumulates in potato leaf tissues along with inhibitor I and II proteins in response to wounding (Graham and Ryan, 1981; Hollander-Czytko et al. 1985). Thus, the inhibitors accumulated in the wounded leaf tissues of potato have the capacity to inhibit all the five major digestive enzymes i.e. trypsin, chymotrypsin, elastase, carboxypeptidase A and carboxypeptidase B of higher animals and many insects (Hollander-Czytko et al. 1985). Aspartic PIs have been recently been isolated from sunflower (Park et al. 2000), barley (Kervinen et al. 1999) and cardoon (Cyanara cardunculus) flowers named as cardosin A (Frazao et al. 1999).

The detailed structural analysis of prophytepsin, a zymogen of barley aspartic proteinase shows a pepsin like bilobe and a plant specific domain. The N terminal has 13 amino acids necessary for inactivation of the mature phytepsin (Kervinen et al. 1999), and the aspartic PI cardosin A from cardoon shows regions of glycolylations at Asn-67 and Asn 257. The Arg-Gly-Asp sequences recogonizes the cardosin receptor which is found in a loop between two-beta strands on the molecular surface (Frazao et al. 1999).

Mechanism of toxicity

The mechanism of action of these proteinase inhibitors has been a subject of intense investigation (Barrett, 1986; MacPhalen and James, 1987; Greenblatt et al. 1989). Knowledge on mechanisms of protease action and their regulation in vitro, and in vivo, in animals, plants, microorganisms and more recently in viruses have contributed to many practical applications for inhibitor proteins in medicine and agriculture.

Baker et al. 1984 showed that the secretion of proteases in insect guts depends upon the midgut protein content rather than the food volume. The secretion of proteases has been attributed to two mechanisms, involving either a direct effect of food components (proteins) on the midgut epithelial cells, or a hormonal effect triggered by food consumption (Applebaum, 1985). Models for the synthesis and release of proteolytic enzymes in the midguts of insects proposed by Birk and Applebaum, 1960, Brovosky, 1986 reveal that ingested food proteins trigger the synthesis and release of enzymes from the posterior midgut epithelial cells. The enzymes are released from membrane associated forms and sequestered in vesicles that are in turn associated with the cytoskeleton. The peptidases are secreted into the ectoperitrophic space between the epithelium, as a particulate complex (Eguchi et al. 1982), from where the proteases move transversely into the lumen of the gut, where the food proteins are degraded. PIs inhibit the protease activity of these enzymes and reduce the quantity of proteins that can be digested, and also cause hyper-production of the digestive enzymes which enhances the loss of sulfur amino acids (Shulke and Murdock, 1983) as a result of which, the insects become weak with stunted growth and ultimately die.

The digestive proteolytic enzymes in the different orders of commercially important insect pests belong to one of the major classes of proteinases predominantly. Coleopteran and hemipteran species tend to utilize cysteine proteinases (Murdock et al. 1987) while lepidopteran, hymenopteran, orthopteran and dipteran species mainly use serine proteinases (Ryan, 1990; Wolfson and Murdock, 1990). Examples from both of these classes of proteinases have been shown to be inhibited by their cognate proregions (Taylor et al. 1995). The effect of class specific inhibitors on the pest digestive enzymes is not always a simple inhibition of proteolytic activity, but recent studies have indicated the reverse may happen. It would appear that there are often two populations of digestive enzymes in target pests, those that are susceptible to inhibition and those that are resistant (Michaud et al. 1996; Bown et al. 1997), and some insects respond to ingestion of plant PIs such as soybean trypsin inhibitor (Broadway and Duffey, 1986b) and oryzacystatin (Michaud et al. 1996) by hyper-producing inhibitor-resistant enzymes.

The mechanism of binding of the plant protease inhibitors to the insect proteases appear to be similar with all the four classes of inhibitors. The inhibitor binds to the active site on the enzyme to form a complex with a very low dissociation constant (107 to 1014 M at neutral pH values), thus effectively blocking the active site. A binding loop on the inhibitor, usually "locked" into conformation by a disulphide bond, projects from the surface of the molecule and contains a peptide bond (reactive site) cleavable by the enzyme (Terra et al. 1996; Walker et al. 1998). This peptide bond may be cleaved in the enzyme inhibitor complex, but cleavage does not affect the interaction, so that a hydrolyzed inhibitor molecule is bound similar to an unhydrolyzed one. The inhibitor thus directly mimics a normal substrate for the enzyme, but does not allow the normal enzyme mechanism of peptide bond cleavage to proceed to completion ie., dissociation of the product (Walker et al. 1998). It would also appear that insect digestive trypsins do not fall into the classification of peptidase hydrolases, as defined by inhibition spectra. It has been shown, notably, that the trypsin like digestive proteases of several lepidopteran species are inhibited by (l-3-trans carboxiran-2-carbonyl)-l-leu-agmatin (E-64) (Lee and Anstee, 1995; Novillo et al. 1997) an inhibitor generally considered to be specific to cysteine proteinases (Dunn, 1989). Thus, true interactions will become clear only when we have protein crystals and X-ray diffraction data for the structure of insect enzyme/inhibitor complexes. Further, specificity of the inhibitor enzyme interaction is primarily determined by the specificity of proteolysis determined by the enzyme (Blancolabra et al. 1996).

Regulation of proteinase inhibitors

Plant proteinase inhibitor proteins that are known to accumulate in response to wounding have been well characterized. Earlier research on tomato inhibitors has shown that the protease inhibitor initiation factor (PIIF), triggered by wounding/injury switches on the cascade of events leading to the synthesis of these inhibitor proteins (Melville and Ryan, 1972; Bryant et al. 1976), and the newly synthesized PIs are primarily cytosolic (Hobday et al. 1973; Meige et al. 1976).

The current evidence suggests that the production of the inhibitors occurs via. the octadecanoid (OD) pathway, which catalyzes the break down of linolenic acid and the formation of jasmonic acid (JA) to induce protease inhibitor gene expression (Koiwa et al. 1997). There are four systemic signals responsible for the translocation of the wound response, which includes systemin, abscisic acid (ABA), hydraulic signals (variation potentials) and electrical signals (Malone and Alarcon, 1995). These signal molecules are translocated from the wound site through the xylem or phloem as a consequence of hydraulic dispersal. The plant systemin an 18-mer peptide has been intensely studied from wounded tomato leaves which strongly induced expression of protease inhibitor (PI) genes. Transgenic plants expressing prosystemin antisense cDNA exhibited a substantial reduction in systemic induction of PI synthesis, and reduced capacity to resist insect attack (McGurl et al. 1994). Systemin regulates the activation of over 20 defensive genes in tomato plants in response to herbivorous and pathogenic attacks. The polypeptide activates a lipid-based signal transduction pathway in which linolenic acid, is released from plant membranes and converted into an oxylipin signaling molecule, jasmonic acid (Ryan, 2000). A wound-inducible systemin cell surface receptor with an M(r) of 160,000 has also been identified and the receptor regulates an intracellular cascade including, depolarization of the plasma membrane and the opening of ion channels thereby increasing the intracellular Ca(2+), which activates a MAP kinase activity and a phospholipase A(2). These rapid changes, play a vital role leading to the intracellular release of linolenic acid from membranes and its subsequent conversion to JA, a potent activator of defense gene transcription (Ryan, 2000). The oligosaccharides, generated from the pathogen-derived pectin degrading enzymes i.e. polygalacturonase (Bergey et al. 1999) and the application of systemin as well as wounding have been shown to increase the jasmonate levels in tomato plants. Application of jasmonate or its methyl ester, methyl jasmonate, strongly induces local and systemic expression of PI genes in many plant species, suggesting that jasmonate has an ubiquitous role in the wound response (Wasternack and Parthier, 1997). Further, analysis of a potato PI-IIK promoter has revealed a G-box sequence (CACGTGG) as jasmonate-responsive element (Koiwa et al. 1997). The model developed for the wound-induced activation of the proteinase inhibitor II (Pin2) gene in potato (Solanum tuberosum) and tomato (Lycopersicon esculentum) establishes the involvement of the plant hormones, abscisic acid and jasmonic acid (JA) as the key components of wound signal transduction pathway (Titarenko et al. 1997). Recently, it has been shown that the defense signaling in suspensions of cultured cells of Lycopersicon peruvianum by peptide systemin, chitosan and by beta-glucan elicitor from Phytophtora megasperma, is inhibited by the polysulfonated naphtyl urea compound suramin, a known inhibitor of cytokine and growth factor receptor interactions in animal cells (Stratmann et al. 2000a). Levels of ABA have been shown to increase in response to wounding, electrical signal, heat treatment or systemin application in parallel with PI induction (Koiwa et al. 1997). Abscisic acid originally thought to be involved in the signaling pathway is now believed to weakly induce the mRNAs of wound response proteins and a concentration even as high as 100 mM induced only low levels of proteinase inhibitor as compared to systemin or jasmonic acid (Birkenmeiner and Ryan, 1998), suggesting the localized role of ABA.

However, it is evident that wound induction and pathogen defense pathways overlap considerably. Expression of wound and JA inducible genes can be positively and negatively regulated by ethylene or salicylic acid (SA), both of which are components of the pathogen-induced signaling pathway (Bent, 1996; Delaney et al. 1994). The expression of thionins in Arabidopsis (Epple et al. 1995) and lectin II in Griffonia simplicifolia  (Zhu-Salzman et al. 1998) was elicited by JA but suppressed by ethylene, showing their opposite effects on the wound signaling pathway.

Plants sometimes specifically forego one type of defense response for another. Salicylic acid (SA) and its methyl ester (Me-SA) are both defense compounds that potently induce systemic acquired resistance of plants against pathogenic microorganisms (Hunt et al. 1996). However, in response to spider mite infestation, lima bean plants release Me-SA which functions as a volatile attractant of the predatory mite Amblyseius potentillae (Dicke et al. 1990). At the same time, SA itself negatively regulates the OD pathway through inhibition of SA biosynthesis and activity (Korth and Dixon, 1997), indicating that SA may suppress the plant defense response through attenuation of the OD pathway, but its methyl ester positively affects plant defense through another defense mechanism involving tritrophic plant herbivore interaction (Moura and Ryan, 2001). Different jasmonic acid-dependent and independent wound signal transduction pathways have been identified recently and partially characterized. Components of these signalling pathways are mostly similar to those implicated in other signalling cascades which include reversible protein phosphorylation steps, calcium/calmodulin-regulated events, and production of active oxygen species (León et al. 2001).

Stintzi et al. (2001) using biochemical genetic approach demonstrated that cyclopentenone precursor of JA, 12-oxo-phytodienoic acid (OPDA), as a physiological signal eliciting defensive response and resistance in the absence of JA. Studies on the effect of UV radiation on early signaling events in the response of young tomato plants to ultraviolet-C (< 280 nm) and UVB/UVA (280-390 nm) radiation induces a 48 kDa myelin basic protein kinase activity in leaves (Stratmann et al. 2000b). In the case of barley plants, ethylene increased the activity of both cell wall bound peroxidases types (ionically and covalently bound), comparable with infestation, which suggests that ethylene is involved in the oxidative responses of (Argandoña et al. 2001)

Studies on the induction of PI proteins have indicated a de novo synthesis of proteins such as a Boman Birk protease inhibitor (OsBBPI) from rice which was found to be rapidly induced in seedling leaf in response to cut, exogenous jasmonic acid (JA), and two potent protein phosphatase 2A (PP2A) inhibitors, in a light/dark, time and dose dependent manner but was completely inhibited by cycloheximide (Rakwal et al. 2001).

Structure of protease inhibitor genes

The gene size and coding regions of the inhibitors are generally small with no introns (Boulter, 1993) and many of these inhibitors are products of multigene families (Ryan, 1990). Bowman-Birk type double-headed protease inhibitors are assumed to have arisen by duplication of an ancestral single headed inhibitor gene and subsequently diverged into different classes i.e. trypsin/trypsin (T/T), trypsin/chymotrypsin (T/C) and trypsin/elastase (T/E) inhibitors (Odani et al. 1983). The mature proteins comprise a readily identifiable ‘core’ region, covering the invariant cysteine residues and active center serines, which are bound by highly variable amino and carboxy-terminal regions. There is a core region of 62 amino acids, both between and within the different classes of inhibitor, within cowpea and with other leguminosae, including azuki bean (Ishikawa et al. 1985), lima bean (Stevens et al. 1974), mung bean (Zhang et al. 1982) and soybean (Odani and Ikenaka, 1976). The average number of amino acid replacements in this region from all pair-wise comparisons show that the differences between the different classes of inhibitor within a species (around 16.5/62 residues) are much greater than the differences within a class between different species (around 11/62 residues). Considering that 18 of the residues in this region are obligatorily invariant for proteins to be classified as Bowman-Birk type inhibitors, these are very high rates of amino acid substitutions. This highlights the problems likely to be encountered in attempting to draw conclusions about the evolutionary history of the rapidly diverging, multigenic protein families from sequences which may be paralogous rather than orthologous. Corrected divergence between pair-wise combinations of sequences calculated according to the method of Perler et al. 1980 revealed that the average divergence between trypsin-specific and chymotrypsin-specific second domains (about 36%) is very similar to that between the first and second domains (about 40%). On an “evolutionary clock” model this would imply that the gene duplication leading to T/T and T/C families occurred very close to the duplication, leading to the appearance of the double-headed inhibitors and that the number of silent substitutions has reached saturation in all these genes (Hilder et al. 1989).

Analysis of the winged bean Kunitz chymotrypsin inhibitor (WCI) protein shows that it is encoded by a multigene family that includes four putative inhibitor coding genes and three pseudogenes. The structural analysis of the WCI genes indicates that an insertion at a 5' proximal site occurred after duplication of the ancestral WCI gene and that several gene conversion events subsequently contributed to the evolution of this gene family (Habu et al. 1997). The 5' region of the pseudogene, WCI-P1 contains frameshift mutations, an indication that the 5' region of the WCI-P1 gene may have spontaneously acquired new regulatory sequences during evolution. Since, gene conversion is a relatively frequent event and the homology between the WCI-P1 and the other inhibitor genes WCI-3a/b is disrupted at a 5' proximal site by remnants of an inserted sequence, the WCI-P1 gene appears to be a possible intermediate that could be converted into a new functional gene with a distinct pattern of expression by a single gene-conversion event (Habu et al. 1997). Molecular evolution of wip-1 genes from four Zea species show significant heterogeneity in the evolutionary rates of the two inhibitory loops, in which one inhibitory loop is highly conserved, whereas the second is diverged rapidly. Because these two inhibitory loops are predicted to have very similar biochemical functions, the significantly different evolutionary histories suggest that these loops have different ecological functions (Tiffin and Gaut, 2001).

The 3’ends of the sequences are comprised of alternating purine-rich and pyrimidine-rich traits of nucleotide. The first of these purine-rich trait occurs at the C-terminal region of the coding sequence. Mutations within this region (deletion/additions or base changes) give rise to substitution amongst the A/G rich codons, Asp, Glu, Lys, Asn and termination codons and also certain specific motifs within them appear to be relatively conserved, therefore likely to be of functional significance. This obviously applies to the cononical polyadenylation signals in the third and second purine-rich traits and the 3’region has regulatory elements which are dependent on a higher order of structure than the specific base sequence. X-ray crystallography of Bowman-Birk inhibitors suggests that this termini has no role in the interaction of the inhibitor with its target enzyme (Suzuki et al. 1987). Substitutions and deletion/additions appear to be very feasible in this region, provided that there is a cleavable serine or asparagine residue within 10-20 amino acids of the first cysteine (Ryan, 1990). The inhibitors are synthesized as precursors from which the leader sequence is cleaved and a long trait of leader-encoding sequence is present in soybean genomic clones (Hilder et al. 1989). There is no significant homology in this region to other seed-expressed protein leader sequences, other than a high representation of hydrophobic residues. Multiple potential initiator codons are a common feature of legume seed protein genes exemplifying the high degree of evolutionary novelty which appears to be tolerated within such seed specific secondary compound genes (Hilder et al. 1989). Analysis of oryzacystatin OC-I has revealed the presence of two introns; the first a 1.4 kbp region between Ala 38 and Asn 39 and a second region of 372 bp in the 3’ non coding region (Kishimoto et al. 1994) and a second oryzacystatin, OC-II present on chromosome 5 also has introns in the same positions (Kondo et al. 1991) thus suggesting deviation from the earlier PIs which lacked introns.

Developing insect resistant transgenic plants expressing Pls

A large number of protease inhibitor genes with distinct modes of action have been isolated from a wide range of crop species. Development of transgenic crops have come a long way from the first transgenic developed by Hilder et al. 1987. Considering the high complexity of protease/inhibitor interactions in host pest systems and the diversity of proteolytic enzymes used by pests and pathogens to hydrolyze dietary proteins or to cleave peptide bonds in more specific processes (Graham et al. 1997), the choice of an appropriate proteinase inhibitor (PI) or set of PIs represents a primary determinant in the success or failure of any pest control strategy relying on protease inhibition. Firstly, the choice of suitable PIs should be based on a detailed understanding of the biological system assessed. Based on our current knowledge about the use of specific inhibitors in the study and control of various metabolic pathways, many PIs have been used to create transgenic crop plants as shown in Table 1 and many more inhibitors are also being isolated with divergent modes of action against different pest species (Table 2). Resistant biotypes of insects may evolve after prolonged exposure to selection pressure that is mediated by an insecticidal protein or plant resistance gene (Sparber, 1985). Unless the biotechnology strategy is designed and implemented to overcome these problems, it will become ineffective in due course like any pesticide based management strategy.

Second point to consider would be the targeted expession of PIs in response to pest attack. This will be controlled by using inducible promoters, such as those of PI-II85 and TobRB7, that are activated at the site of invasion by pests, pathogen and nematodes, respectively (Opperman et al.1994). An ideal promoter should be highly responsive to invasion of the host plant by a pest, or regulated by inducers just prior to pest attack. The promoter should be sufficiently active to mediate a substantial defense, specially localized to the site of pest invasion. Suitable promoters such as those regulated in response to pest invasion can be identified using promoter trapping techniques (Babiychuk et al. 1997).

Despite these promising developments, the general usefulness of recombinant PIs in plant protection still remains to be demonstrated. The inhibitory spectrum of PIs is usually limited to proteases in one of several mechanistic classes, leaving free proteases in the surrounding medium after inhibition (Barrett, 1994). Due to a progressive adaptation of plant pests to the continuous occurrence of PIs in the diet, the inhibitory spectrum of protein inhibitors against the extracellular proteases of several pests is even more limited, being often restricted to the family level (Michaud et al. 1995b; Visal et al. 1998). Non-target proteases, that may allow metabolic compensation of inhibited proteolytic functions, (Jongsma and Bolter, 1997) may also challenge the structural integrity of several PIs and thus potentially affect their effectiveness in vivo (Michaud, 1997).

Recently it has been shown that the presence of large amounts of inhibitors including soybean Kunitz inhibitor (Bown et al.1997) in the diets of economical pests has made insects to adapt and produce proteases which are insensitive to the action of host plant inhibitors and the ingested PIs activate these genes (Dennis et al. 1994). As a result, pest control using PIs in transgenic plants requires the isolation of inhibitors that are active towards these insensitive proteases (Jongsma et al.1996). One can search for active inhibitors among naturally-occurring peptides (Gruden et al. 1998) or can engineer inhibitors in such a way that they will acquire activity against the "PI insensitive" protease. Engineering of inhibitors can be performed in two distinct ways: 1) based on structures of the inhibitor-protease complex, predictions can be made on mutations that will enhance binding (Urwin et al. 1995), but lack of data on these complexes for insect proteases makes this procedure rather tough. However, it may be more appropriate to simply generate large arrays of mutants in the region of the inhibitor protein contacting the protease. The powerful method of phage display can subsequently be used to select the strongly binding mutants. Large number of PIs have been subjected to phage display as a result of which inhibition constant (Ki) for target protease which was initially poor (millimolar to micromolar) have been greatly improved. Some of the examples belonging to serine class are serpin (Pannekoek et al. 1993) and kazal from human (Rottgen and Collins, 1995), E.coli ecotin (Wang et al. 1995), proteinase inhibitor II (PI-II) from potato (Jongsma et al. 1995), chicken cystanin (Tanaka et al. 1995) and soybean phytocystanin (Koiwa et al. 1998) belonging to cysteine class. The improvement of plant PIs by phage display is still an infant stage to be commercially important.

Insect midgut contains an estimated 1020 different proteases (Bown et al. 1997) which are differentially regulated and all cannot be inhibited by plant’s PIs (Broadway, 1997). Therefore, to achieve an effective pest control strategy it is very important to achieve different inhibitors expression in a concerted manner.

Concluding Remarks

The availability of diverse genes from different plant species makes it a possibility to use one or more genes in combination, whose products are targeted at different biochemical and physiological processes within the insect. These packages will not only contain protease inhibitor genes but also lectins, alpha-amylase inhibitors, or other plant genes encoding insecticidal proteins. This technology may not replace the use of chemical pesticides in near future but effectively complement it. The use of recombinant PIs may also be an attractive way to protect plants from fungal, bacterial and viral pathogens. Currently, two principal strategies are proposed to engineer effective pest control in plants: ectopic expression of pesticidal proteins, and induction of the plant natural defensive response. At present, screening gene pools without taxonomic constraint can help identify novel insecticidal determinants, but in future this approach will be augmented by directed in vitro molecular evolution (Koiwa et al. 1998). Given the number of pesticidal proteins that are involved in host plant defense, it is presumed that effective pest control by this strategy will result from the co-expression of numerous determinants, each of which could be custom engineered by directed molecular evolution to maximize its effectiveness against specific pests.

However, in future non-scientific issues such as regulatory approval, propriety rights and public perception will be decisive in releasing crop plants produced by genetic engineering using recombinant DNA technology.

References

  • ABE, M.; ABE, K.; KURODA, M. and ARAI, S. Corn kernel cysteine proteinase inhibitor as a novel cystatin superfamily member of plant origin. Molecular cloning and statement studies. European Journal of Biochemistry, November 1992, vol. 209, no. 3, p. 933-937.
  • ABE, K.; EMORI, Y.; KONDO, H.; SUZUKI, K. and ARAI, S. Molecular cloning of a cysteine proteinase inhibitor of rice (oryzacystatin). Homology with animal cystatins and transient expression in the ripening process of rice seeds. Journal of Biological Chemistry, December 1987, vol. 262, no. 35, p. 16793-16797.
  • ABE, K. and ARAI, S. Purification of a cysteine proteinase inhibitor from rice, Oryza sativa L. japonica. Agriculture Biology and Chemistry, 1985, vol. 49, p. 3349-3350.
  • ALPTETER, F.; DIAZ, I.; McAUSLANE, H.; GADDOUR, K.; CARBONERO, P. and VASIL, I.K. Increased insect resistance in transgenic wheat stably expressing trypsin inhibitor CMe. Molecular Breeding, 1999, vol. 5, p. 53-63.
  • APPLEBAUM, S.W. Biochemistry of digestion. In: KERKOT, G.A. and GILBERT, L.I., eds.  Comprehensive insect physiology; Biochemistry and Pharmacology. New York, Pergamon Press, 1985, vol. 4, p. 279-311.
  • ARAI, S.; WATANABE, H.; KONDO, H.; EMORI, Y. and ABE, K. Papain inhibitory activity of oryzacystatin, a rice seed cysteine proteinase inhibitor, depends on the central Gln-Val-Val-Ala-Gly region conserved among cystatin superfamily members. Journal of Biochemistry, 1991, vol. 109, p. 294-298.
  • ARGANDOÑA, V.H.; CHAMAN, M.; CARDEMIL, L.; MUÑOZ, O., ZUÑIGA, G.E. and CORCUERA, L.J. Ethylene production and peroxidase activity in aphid-infested barley. Journal of Chemical Ecology, 2001, vol. 27, p. 53-68.
  • BABIYCHUK, Elena; FUANGTHONG, Mayuree; VAN MONTAGU, Mare; INZE, Dirk and KUSHNIR, Sergei. Efficient gene tagging in Arabidopsis thaliana using a gene trap approach. Proceedings of the National Academy of Sciences United States of America, November 1997, vol. 94, p.12722-12727.
  • BAKER, J.E.; WOO, S.M. and MULLEN, M.A. Distribution of proteinases and carbohydrates in the midgut of the larvae of the sweet potato weevil Cyclas formicarius and response of proteinase to inhibitors from sweet potato. Entomologia  Experimentalis et Applicata, 1984, vol. 36, p. 97-105.
  • BARRETT, A.J. Classification of  peptidases. In: BARRETT, A.J, ed. Methods in Enzymology, New York,  Academic Press, 1994, vol. 244, p. 1-15.
  • BARRETT, A.J. The classes of proteolytic enzymes. In: DALLING, M.J ed. Plant proteolytic enzymes, Florida, CRC Press Inc., 1986, vol. 1, p. 1-16.
  • BENT, A.F. Plant disease resistance genes: Function meets structure. Plant Cell, 1996, vol. 8, p. 1757-1771.
  • BERGEY, Daniel R.; OROZCO-CARDENAS, Martha; DE MOURA, Daniel S. and RYAN, Clarence A. A wound- and systemin-inducible polygalacturonase in tomato leaves. Proceedings of the National Academy of Sciences United States of America, February 1999, vol. 96, no. 4, p.1756-1760.
  • BIRK, Y. and APPLEBAUM, Y. Effect of soybean trypsin inhibitors on the development and midgut proteolytic activity of Tribolium castaneum larvae. Enzymologia Acta Biocatalytica, November 1960, vol. 22, no. 5, p. 318-326.
  • BIRKENMEIER, Guy F. and RYAN, Clarence A. Wound signaling in tomato plants: evidence that ABA is not a primary signal for defense gene activation. Plant Physiology, June 1998, vol. 117, no. 2, p. 687-693.
  • BLANCOLABRA, A.; MARTÍNEZ-GALLARDO, N.A.; SANDOVAL-CARDOSO, L. and DÉLANO-FRIER, J. Purification and characterization of a digestive cathepsinD proteinase isolated from Tribolium castaneum larvae (Herbst). Insect Biochemistry and Molecular Biology, 1996, vol. 26, p. 95-100.
  • BROADWAY, R.M. Dietary regulation of serine proteinases that are resistant to serine proteinase inhibitors. Insect Biochemistry and Molecular Biology, 1997, vol. 43, p. 855-874.
  • BROADWAY, R.M. Are insects resistant to plant proteinase inhibitors? Journal of Insect Physiology, 1995, vol.  41, p. 107-116.
  • BROADWAY, R.M. and DUFFEY, S.S. The effect of dietary protein on the growth and digestive physiology of larval Heliothis zea and Spodoptera exigua. Journal of Insect Physiology, 1986a, vol.  32, p. 673-680.
  • BROADWAY, R.M. and DUFFEY, S.S. Plant proteinase inhibitors: mechanism of action and effect on the growth and digestive physiology of larval Heliothis zea and Spodoptera exigua. Journal of Insect Physiology, 1986b, vol. 32, p. 827-833.
  • BROVOSKY, D. Proteolytic enzymes and blood digestion in the mosquito Culex nigripalpus.  Archives of Insect Biochemistry and Physiology, 1986, vol. 3, p. 147-160.
  • BRYANT, J.; GREEN, T.R., GURUSADDAIAH, T. and RYAN, C.A. Proteinase inhibitor II from potatoes: isolation and characterization of its promoter components. Biochemistry, 1976, vol. 15, p. 3418-3424.
  • BOULTER, D. Insect pest control by copying nature using genetically engineered crops. Biochemistry, 1993, vol. 34, p. 1453-1466.
  • BOWN, D.P.; WILKINSON, H.S. and GATEHOUSE, J.A. Differentially regulated inhibitor sensitive and insensitive protease genes from the phytophagous insect pest, Helicoverpa armigera, are members of complex multigene families. Insect Biochemistry and Molecular Biology, July 1997, vol. 27, no.7, p. 625-638.
  • BURGESS, E.P.J.; STEVEN, P.S.; KEEN, G.K.; LAING, W.A. and CHRISTELLER, J.T. Effects of protease inhibitors and dietary protein level on the black field cricket Teleogryllus commodus. Entomologia Experimentalis et Applicata, 1991, vol. 61, p. 123-130.
  • CAMPOS, F.A.P.; XAVIER-FILHO, J.; SILVA, C.P. and ARY, M.B. Resolution and partial characterization of proteinases and a-amylases from midgets of larvae of the bruchid beetle Callosobruchus maculatus (F.). Comparative Biochemistry and Physiology-B, 1989, vol. 92, p. 51-57.
  • CARBONERO, P.; ROYO, J.; DÍAZ, I., GARCÍA-MAROTO, F.; GONZALEZ-HIDALGO, E.; GUTIERREZ, C. and CASANERA, P. Cereal inhibitors of insect hydrolases (a-amylases and trypsin): genetic control, transgenic expression and insect pests. In: Workshop on engineering plants against pests and pathogens (1st – 13rd January, 1993, Madrid, Spain). Instituto Juan March de Estudios Investigaciones, BRUENING, G.J.; GARCÍA-OLMEDO, F. and PONZ, F.J. eds., 1993.
  • CHARITY, Julia A.; ANDERSON, Merilyn A.; BITTISNICH, Dennis J.; WHITECROSS, Malcom and HIGGINS, T.J.V. Transgenic tobacco and peas expressing a proteinase inhibitor from Nicotiana alata have increased insect resistance. Molecular Breeding, 1999, vol. 5, no. 4, p. 357-365.
  • CHEN, M.S.; JOHNSON, B.; WEN, L.; MUTHUKRISHNAN, S.; KRAMER, K.J.; MORGAN, T. and REECK, G.R. Rice cystatin: bacterial expression, purification, cysteine proteinase inhibitory activity, and insect growth suppressing activity of a truncated form of the protein. Protein Expresssion and Purification, 1992, vol. 3, p. 41-49.
  • DELANEY, T.P.; UKNES, S.; VERNOOIJ, B.; FRIEDRICH, L.; WEYMANN, K.; NEGROTTO, D.; GAFFNEY, T.; GUT-RELLA, M.; KESSMANN, H.; WARD, E. and RYALS, J. A central role of salicylic acid in plant disease resistance. Science, 1994, vol.  266, p. 1247-1250.
  • DENNIS, M.S.; HERZKA, A. and LAZARUS, R.A. Potent and selective Kunitz domain inhibitors of plasma kallikrein designed by phage display. Journal of Biological Chemistry, October 1994, vol. 270, no. 50, p. 25411-25417.
  • DICKE, M.; BEEK, T.A.; POSTHUMUS, M.A.; BEN DOM, N.; VAN BOKHOVEN, H. and DE GROOT, A.E. Isolation and identification of volatile kairomone that affects acarine predator prey interactions. Journal of Chemical Ecology, 1990, vol. 16, p. 381-396.
  • DUAN, X.; LI, X.; XUE, Q.; ABO-EI-SAAD, M.; XU, D. and WU, R. Transgenic rice plants harboring an introduced potato proteinase inhibitor II gene are insect resistant. Nature Biotechnology, April 1996, vol. 14, no. 4, p. 494-498.
  • DUNAEVSKII, Y.E.; GLADYSHEVA, I.P.; PAVLUKOVA, E.B.; BELIAKOVA, G.A.; GLADYSCHEV, D.P.; PAPISOVA A.I.; LARIONOVA N.I. and BELOZERSKY M.A. The anionic protease inhibitor BBWI - 1 from buckwheat seeds. Kinetic properties and possible biological role. Physiologia Plantarum, 1997, vol. 100, p. 483-488.
  • DUNN, B.M. Proteolytic enzymes: A practical approach. Oxford,  IRL Press, 1989. 57-81 p. ISBN 0-19-963058-5.
  • EDMONDS, H.S.; GATEHOUSE, L.N.; HILDER, V.A. and GATEHOUSE, J.A. The inhibitory effects of the cysteine protease inhibitor, oryzacystatin, on digestive proteases and on larval survival and development of the southern corn rootworm (Diabrotica undecimpunctata howardi). Entomologia Experimentalis et Applicata, 1996, vol. 78, p. 83-94.
  • EGUCHI, M.; IWAMOTO, A. and YAMAGUHI, K. Interaction of proteases from the midgut lumen, epithelial and peritrophic membrane of the silkworm Bombyx mori L. Comparative Biochemistry and Physiology-A, 1982, vol. 72, p. 359-363.
  • EPPLE, P.; APRL, K. and BOHLMANN, H. An Arabidopsis thaliana thionin gene is inducible via a signal transduction pathway different from that for pathogenesis related protein. Plant Physiology, 1995, vol. 109, p. 813-820.
  • FERNANDES, K.V.S.; SABELLI, P.A.; BARRATT, D.H.P.; RICHARDSON, M.; XAVIER-FILHO, J. and SHEWRY, P.R. The resistance of cowpea seeds to bruchid beetles is not related to level of cysteine proteinase inhibitors. Plant Molecular Biology, October 1993, vol. 23, no. 1, p. 215-219.
  • FRAZAO, Carlos; BENTO, Isabel; COSTA, Julia; SOARES, Claudio M.; VERISSIMO, Paula; FARO, Carlos; PIRES, Euclides; COOPER, Jon and CARRONDO, Maria A. Crystal structure of cardosin A, a glycosylated and Arg-Gly-Asp- containing aspartic proteinase from the flowers of Cyanara cardunculus L. Journal of Biological Chemistry, September 1999, vol. 274, no. 39, p. 27694-27701.
  • GARCIAOLMEDO SALCEDO, F.; SANCHEZ MONGE, G.; GOMEZ, R.L.; ROYO, J. and  CARBONERO, P. Plant proteinaceous inhibitors of proteinases and a-amylases. Oxford Survey Plant Molecular and Cell Biology, 1987, vol. 4, p. 275-334.
  • GATEHOUSE, A.M.R.; DAVIDSON, G.M.; NEWELL, C.A.; MERRYWEATHER, A.; HAMILTON, W.D.O.; BURGESS, E.P.J.; GILBERT, R.J.C. and GATEHOUSE, J.A. Trangenic potato plants with enhanced resistance to the tomato moth, Lacanobia oleracea: growth room trials. Molecular Breeding, 1997, vol.  3, p. 49-63.
  • GATEHOUSE, A.M.R. and BOULTER, D. Assessment of the antimetabolic effects of trypsin inhibitors from cowpea (Vigna  unguiculata) and other legumes on development of the bruchid beetle Callosobruchus maculatus. Journal of the Science of Food and Agriculture, 1983, vol. 34, p. 345-350.
  • GIRARD, C.; LE MÉTAYER, M.; BONADÉ-BOTTINO, M.; PHAM-DELÈGUE M.H. and JOUANIN, L. High level of resistance to proteinase inhibitors may be conferred by proteolytic cleavage in beetle larvae. Insect Biochemistry and Molecular Biology. 1998, vol. 28, p. 229-237.
  • GOURINATH, S.; ALAM, N.; SRINIVASAN, A.;  BETZEL, C. and SINGH, T.P. Structure of the bifunctional inhibitor of trypsin and alpha-amylase from ragi seeds at 2.2 A resolution. Acta Crystallography D Biology Crystallography, March 2000, vol. 56, no. 3, p. 287-293.
  • GRAHAM J.; GORDON, S.C. and MCNICOL, R.J. The effect of the CpTI gene in strawberry against attack by vine weevil (Otiorhynchus sulcatus F. Coleoptera: Curculionidae). Annals of Applied Biology, 1997, vol. 131, p. 133-139.
  • GRAHAM, J.S. and RYAN, C.A. Accumulation of  metallocarboxy-peptidase inhibitor in leaves of wounded potato plants. Biochemical and Biophysics Research Communication, 1997, vol. 101, p. 1164-1170.
  • GREENBLATT, H.M.; RYAN, C.A. and JAMES, M.N.G. Strucuture of the complex Streptomyces griseus proteinase B and polypeptide chymotrypsin inhibitor-I at 2.1 A resolution.  Journal of Molecular Biology, 1989, vol. 205, p. 201-228.
  • GRUDEN, K.; STRUKELJ, B.; POPOVIC, T.; LENARCIC, B.; BEVEC, T.; BRZIN, J.; KREGAR, I.; HERZOG-VELIKONJA, J.; STIEKEMA, W.J.; BOSCH, D. and JONGSMA M.A. The cysteine protease activity of Colorado potato beetle (Leptinotarsa decemlineata Say) guts, which is insensitive to potato protease inhibitors, is inhibited by thyroglobulin type1 domain inhibitors. Insect Biochemistry and Molecular Biology, 1998, vol. 28, p. 549-560.
  • HABU, Y.; PEYACHOKNAGUL, S.; SAKATA, Y.; FUKASAWA, K. and OHNO, T. Evolution of a multigene family that encodes the Kunitz chymotrypsin inhibitor in winged bean: a possible intermediate in the generation of a new gene with a distinct pattern of expression. Molecular General Genetics, 1997, vol. 254, no. 1, p. 73-80.
  • HAVKIOJA, E. and NEUVONEN, L. Induced long-term resistance to birch foliage against defoliators : defense or incidental. Ecology, 1985, vol. 66, p. 1303-1308.
  • HEATH, R.G.; MCDONALD, J.T.; CHRISTELLER, M.; LEE, K.; BATEMAN, J.; WEST, R.; VAN HEESWIJCK and ANDERSON, M.A. Proteinase inhibitors from Nicotiana alata enhance plant resistance to insect pests. Journal of Insect Physiology, 1997, vol. 9, p. 833-842.
  • HILDER, V.A.; BARKER, R.F.; SAMOUR, R.A.; GATEHOUSE, A.M.R.; GATEHOUSE, J.A. and BOULTER, D. Protein and cDNA sequences of Bowman-Birk protease inhibitors from the cowpea (Vigna unguiculata Walp.). Plant Molecular Biology, December 1989, vol. 13, no. 6, p. 701-710.
  • HILDER, V.A.; GATEHOUSE, A.M.R.; SHEERMAN, S.E.; BARKER, R.F. and BOULTER, D. A novel mechanism of insect resistance engineered into tobacco. Nature, 1987, vol. 300, p. 160-163.
  • HOBDAY, S.M.; THURMAeN, D.A. and BARBER, D.J. Proteolytic and trypsin inhibitory activities in extracts of germinating Pisum sativum seeds. Phytochemistry, 1973, vol. 12, p. 1041-1046.
  • HOLLANDER-CZYTKO, H.; ANDERSEN, J.L. and RYAN, C.A. Vacuolar localization of wound-induced carboxy peptidase inhibitor in potato leaves. Plant Physiology, 1985, vol. 78, p. 76-79.
  • HOUSEMAN, J.G.; DOWNE, A.E.R. and PHILOGENE, B.J.R. Partial characterization of proteinase activityin the larval midgut of the European corn borer Ostrinia nubilalis Hubner (Lepidoptera: Pyralidae). Canadian Journal of Zoology, 1989, vol. 67, p. 864-868.
  • HOUSEMAN, J.G.; CAMPBELL, F.C. and MORRISON, P.E. A preliminary characterization of digestive proteases in the posterior midgut of the stable fly Stomoxys calcitrans (L.) (Diptera: Muscidae). Insect Biochemistry, 1987, vol. 17, p. 213-218.
  • HOUSEMAN, J.G. and DOWNE, A.E.R. Cathepsin D-like activity in the posterior midgut of Hemipteran insects. Comparative Biochemistry and Physiology-B, 1983, vol. 75, p. 509-512.
  • HUBER, R. and CARRELL, R.W. Implications of the three-dimensional structure of alpha I -antitrypsin for structure and function of serpins. Biochemistry, 1989, vol. 28, p. 8951-8966.
  • HUNT, M.D.; NEUENSCHWANDER, U.H.; DELANEY, T.P.; WEYMANN, K.B.; FRIEDRICH, L.B.; LAWTON, K.A.; STEINER, H.Y. and RYALS, J.A. Recent advances in systemic acquired resistance research-a review. Gene, November 1996, vol. 179, no. 1, p. 89-95
  • ISHIKAWA, A.; OHTA, S.; MATSUOKA, K.; HATTORI, T. and NAKAMURA, K. A family of potato genes that encode Kunitz-type proteinase inhibitors: structural comparisons and differential expression.  Plant and Cell Physiology, 1994, vol. 35, p. 303-312.
  • ISHIKAWA, C.; WATANABE, K.; SAKATA, N.; NAKAGSAKI, C.; NAKAMURA, S. and TAKAHASHI, K. Azuki bean (Vigna angularis) protease inhibitors: isolation and amino acid sequences. Journal of Biochemistry, 1985, vol. 97, p. 55-70.
  • JOHNSON, R.; NARRAEZ, J.; AN, G. and RYAN, C.A. Expression of proteinase inhibitors I and II in transgenic tobacco plants: effects on natural defense against Manduca sexta larvae. Proceedings of the National Academy of Sciences United States of America, 1989, vol. 86, p. 9871-9875.
  • JOHNSTON, K.A.; GATEHOUSE, J.A. ANSTEE, J.H. Effects of soybean protease inhibitors on the growth and development of larval Helicoverpa armigera. Journal of Insect Physiology, 1993, vol. 39, p. 657-664.
  • JONGSMA, M.A. and BOLTER, C. The adaptation of insects to plant protease inhibitors. Journal of Insect Physiology, 1997, vol. 43, p. 885-895.
  • JONGSMA, M.A.; STIEKEMA, W.J. and BOSCH, D. Combatting inhibitor insensitive proteases of insect pests. Trends in Biotechnology, 1996, vol. 14, p. 331-333.
  • JONGSMA, M.A.; BAKKER, P.L.; STIEKEMA, W.J. and BOSCH, D. Phage display of a double-headed proteinase inhibitor: analysis of the binding domains of potato proteinase inhibitor II. Molecular Breeding, 1995, vol. 1, p. 181-191.
  • JOSHI, B.; SAINANI, M.; BASTAWADE, K.; GUPTA, V.S. and RANJEKAR, P.K. Cysteine protease inhibitor from pearl millet: a new class of antifungal protein. Biochemical and Biophysical Research Communications, 1998, vol. 246, p. 382-387.
  • KEILOVA, H. and TOMASEK, V. Isolation and properties of cathespin D inhibitor from potatoes. Collection of Czechoslovak Chemical Communication, 1976, vol. 41, p. 489-497.
  • KERIN, J. Opening address. In: Proceedings of the International Working Conference on Stored-product Protection. (6º, 17th-23rd April, 1994, Canberra, Australia). Volume 1, p. xix-xx.
  • KERVINEN, J.; TOBIN, G.J.; COSTA, J.; WAUGH, D.S.; WLODAWER, A. and ZDANOV, A. Crystal structure of plant aspartic proteinase prophytepsin: inactivation and vacuolar targeting. Journal of European Molecular Biology Organization, July 1999, vol. 18, no. 14, p. 3947-3955.
  • KIMURA, M.; IKEDA, T.; FUKUMOTO, D.; YAMASAKI, N. and YONEKURA, M. Primary structure of a cysteine proteinase inhibitor from the fruit of avocado (Persea americana Mill). Bioscience Biotechnology and Biochemistry, December 1995, vol. 59, no. 1, p. 2328-2329.
  • KISHIMOTO, N.; HIGO, H.; ABE, K.; ARAI, S.; SAITO, A. and HIGO, K. Identification of the duplicated segments in rice chromosomes 1 and 5 by linkage analysis of cDNA markers of known functions. Theoretical and Applied Genetics, 1994, vol. 88, p. 722-726.
  • KITCH, L.W. and MURDOCK, L.L. Partial characterization of a major gut thiol proteinase from larvae of Callosobruchus maculatus (F.). Archives of Insect Biochemistry and Physiology, 1986, vol.  3, p. 561-575.
  • KRATTIGER, Anatole F. Insect resistance in crops: a case study of Bacillus thuringiensis (Bt) and its transfer to developing countries. ISAAA Briefs, 1997, no. 2, p. 42.
  • KONDO, H.; ABE, K.; EMORI, Y. and ARAI, S. Gene organization of oryzacystatinII, a new cystatin superfamily member of plant origin, is closely related to that of oryzacystatinI but different from those of animal cystatins. FEBS Letters, 1991, vol. 278, p. 87-90.
  • KOIWA, K.; SHADE, R.E.; ZHU-SALZMAN, K.; SUBRAMANIAN, L.; MURDOCK, L.L.; NIELSEN, S.S.; BRESSAN, R.A. and HASEGAWA, P.M. Phage display selection can differentiate insecticidal activity of soybean cystatins. Plant Journal, 1998, vol. 14, p. 371-379.
  • KOIWA, H.; BRESSAN, R.A. and HASEGAWA, P.M. Regulation of protease inhibitors and plant defense. Trends in Plant Science, 1997, vol. 2, p. 379-384.
  • KORTH, K.L. and DIXON, R.A. Evidence for chewing insect specific molecular events distinct from a general wound response in leaves. Plant Physiology, 1997, vol. 115, p. 1299-1305.
  • KUNITZ, M. Crystallization of a trypsin inhibitor from soybean. Science, 1945, vol. 101, p. 668-669.
  • LASKOWSKI, M. JR. and KATO, I. Protein inhibitors of proteinases. Annual Review of Biochemistry, 1980, vol. 49, p. 685-693.
  • LAWRENCE, Paulraj K. and KOUNDAL, Kripa Ram. Agrobacterium tumefaciens mediated transformation of pigeonpea (Cajanus cajan (L.) Millsp) and molecular analysis of regenerated plants. Current Science, June 2001, vol. 80, no. 11, p. 1428-1432.
  • LAWRENCE, Paulraj, K.; NIRMALA, Jayaveeramuthu and KOUNDAL Kripa Ram.  Nucleotide sequence of a genomic clone encoding a cowpea (Vigna unguiculata L.) trypsin inhibitor. EJB Electronic Journal of Biotechnology [online]. 15 April 2001,  vol. 4, no. 1 [cited 16 April 2001]. Available from Internet: http://www.ejbiotechnology.info/content/vol4/issue1/full/4/index.html. ISSN  0717-3458.
  • LEE, S.I.; LEE, S.H.; KOO, J.C.; CHUN, H.J.; LIM, C.O.; MUN, J.H.; SONG, Y.H. and CHO, M.J. Soybean Kunitz trypsin inhibitor (SKTI) confers resistance to the brown planthopper (Nilaparvata lugens Stal) in transgenic rice. Molecular Breeding, 1999, vol. 5, p. 1-9.
  • LEE, M.J. and ANSTEE, J.H. Endoproteases from the midgut of larval Spodoptera littoralis include a chymotrypsinlike enzyme with an extended binding site. Insect Biochemistry and Molecular Biology, 1995, vol. 25, p. 49-61.
  • LEMOS, F.J.A.; XAVIER-FILNO, J. and CAMPOS, F.A.P. Proteinases of the midgut of Savrotes subfasciatos larvae. Agricultural Biology Tecnoi, 1987, vol. 80, p. 46.
  • LEÓN, José.; ROJO, Enrique and SANCHEZ-SERRANO, José J. Wound signalling in plants. Journal of Experimental Botany, January 2001, vol. 52, no. 354, p. 1-9.
  • LEPLE, J.C.; BONADE-BOTTINO, M.; AUGUSTIN, S.; PILATE, G.; LE TÂN, V.D.; DELPLANQUE, A.; CORNU, D. and JOUANIN, L. Toxicity to Chrysomela tremulae (Coleoptera: Chrysomelidae) of transgenic poplars expressing a cysteine proteinase inhibitor. Molecular Breeding, 1995, vol. 1, p. 319-328.
  • LI, Y.E.; ZHU, Z.; CHEN, Z.X.; WU, X.; WANG, W. and LI, S.J. Obtaining transgenic cotton plants with cowpea trypsin inhibitor. Acta Gossypii Sinica, 1998, vol. 10, p. 237-243.
  • LIPKE, H.; FRAENKEL, G.S. and LIENER, I.E. Effects of soybean inhibitors on growth of Tribolium confusum. Journal of the Science of Food and Agriculture, 1954, vol. 2, p. 410-415.
  • MACINTOSH, S.C.; KISHORE, G.M.; PERLAK, F.J.; MARRONE, P.G.; STONE, T.B.; SIMS, S.R. and FUCHS, R.L. Potentiation  of  Bacillus thuringiensis  insecticidal activity by serine protease inhibitors. Journal of Agriculture and Food Chemistry, 1990, vol. 38, p. 50-58.
  • MACPHALEN, C.N. and JAMES, M.N.G. Crystal and molecular structure of the serine proteinase inhibitor CI-2 from barley seeds. Biochemistry, 1987, vol. 26, p. 261-269.
  • MALONE, M. and ALARCON, J.J. Only xylem-borne factors can account for systemic wound signalling in the tomato plant. Planta, 1995, vol. 196, p. 740-746.
  • MARES, M.; MELOUN, B.; PAVLIK, M.; KOSTKA, V. and BAUDYS, M. Primary structure of Cathepsin-D inhibitor from potatoes and its structural relationship to trypsin inhibitor family. FEBS Letters, 1989, vol. 251, p. 94-98.
  • MCGURL, B.; OROZCO-CARDENAS, M.; PEARCE, G. and RYAN, Clarence A. Overexpression of the prosystemin gene in transgenic tomato plants generates a systemic signal that constitutively induces proteinase inhibitor synthesis. Proceedings of the National Academy of  Sciences United States of America, 1994, vol. 91, p. 9799-9802.
  • MCMANUS, M.T. and BURGESS, E.P.J. Effects of the soybean (Kunitz) trypsin inhibitor on growth and digestive proteases of larvae of Spodoptera litura. Transgenic Research, 1995, vol. 8, p. 383-395.
  • MCMANUS, M.T., WHITE, D.W.R. AND MCGREGOR, P.G. Accumulation of chymotrypsin inhibitor in transgenic tobacco can affect the growth of insect pests. Transgenic Research, 1994, vol. 3, p. 50-58.
  • MEIGE, M.; MASCHERPA, J.; ROYER-SPIERER, A.; GRANG, A. and MEIGE, J. Analyse des crops proteiques isoles de Lablab Pupureus (L.) Sweet: localisation  intracellulaire des globulines proteases et inhibiteurs de la trypsire. Planta, 1976, vol. 131, p. 181-86.
  • MEINERS, J.P. and ELDEN, T.C. Resistance to insects and diseases in Phaseolus.  In:  Advances in legume science. International Legume Conference, (1978, Kew Surrey, UK). SUMMERFIELD, R.S and BUNTING, A.H.  eds., 1978. p. 359-364.
  • MELVILLE, J.C. and RYAN, C.A. Chymotrypsin inhibitor I from potato: large scale preparation and the characterization of its subunit components. Journal of Biological Chemistry, 1973, vol. 247, p. 3415-3453.
  • METCALF, R.L. The ecology of insecticides and the chemical control of insects.  In: KOGAN, M. ed.  Ecological theory and integrated pest management. New York, John Wiley and Sons, 1986, p. 251-297.
  • MICHAUD, D. Avoiding protease mediated resistance in herbivorous pests. Trends in Biotechnology, 1997, vol. 15, p. 4-6.
  • MICHAUD, D.; NGUYEN-QUOC, B.; VRAIN, T.C.; FONG, D. and YELLE, S. Response of digestive cysteine proteinases from the Colorado potato beetle (Leptinotarsa decemlineata) and the black vine weevil (Otiorynchus sulcatus) to a recombinant form of human stefin A. Archieves of Insect Biochemistry and Physiology, 1996, vol. 31, p. 451-64.
  • MICHAUD, D.; CANTIN, L. and VRAIN T.C. Carboxy terminal truncation of oryzacystatin II by oryzacystatin insensitive insect digestive proteinases. Archives of Biochemistry and Biophysics, October 1995a, vol. 322, no. 2, p. 469-474.  
  • MICHAUD, D.; BERNIERVADNAIS, N.; OVERNEY, S. and  YELLE, S. Constitutive expression of digestive cysteine proteinase forms during development of the Colorado potato beetle, Leptinotarsa decemlineata Say (Coleoptera: Chrysomelidae). Insect Biochemistry and Molecular Biology, 1995b, vol. 25, p. 1041-1048.
  • MICKEL, C.E. and STANDISH, J. Susceptibility of processed soy flour and soy grits in storage to attack by Tribolium castaneum. University of Minnesota Agricultural Experimental Station Technical Bulletin, 1947, vol. 178, p. 1-20.
  • MOURA, Daniel S. and RYAN, Clarence A. Wound-inducible proteinase inhibitors in pepper. Differential regulation upon wounding, systemin, and methyl jasmonate. Plant Physiology, May 2001, vol. 126, p. 289-298.
  • MUKHOPADHYAY, D. The molecular evolutionary history of an winged bean alpha-chymotrypsin inhibitor and modeling of its mutations through structural analysis. Journal of Molecular Evolution, 2000, vol. 50, p. 214-223.
  • MURDOCK, L.L.; SHADE, R.E. and POMEROY, M.A. Effects of E. 64 a cysteine proteinase inhibitor on cowpea weevil growth development and fecundity. Environmental  Entomology, 1988, vol. 17, p. 467-469.
  • MURDOCK, L.L.; BROOKHART, G.; DUNN, P.E.; FOARD, D.E. and KELLEY, S. Cysteine digestive proteinases in Coleoptera. Comparative Biochemistry and Physiology-B, 1987, vol. 87, p. 783-787.
  • NAGATA, K.; KUDO, N.; ABE, K.; ARAI, S. and TANOKURA, M. Three dimensional solution structure of oryzacystatin-I, a cysteine proteinase inhibitor of the rice, Oryza sativa L. japonica. Biochemistry, 2000, vol. 39, p. 14753-14760.
  • NEWELL, C.A.; LOWE, J.M.; MERRYWEATHER, A.; ROOKE, L.M. and HAMILTON, W.D.O. Transformation of sweet potato (Ipomea batatas (L.). Lam) with Agrobacterium tumefaciens and regeneration of plants expressing cowpea trypsin inhibitor and Snowdrop lectin. Plant Science, 1995, vol. 107, p. 215-227.
  • NOVILLO, C.; CASTAÑERA, P. and ORTEGO, F. Inhibition of digestive trypsin like proteases from larvae of several lepidopteran species by the diagnostic cysteine protease inhibitor E64. Insect Biochemistry and Molecular Biology, 1997, vol. 27, p. 247-254.
  • ODANI, S.; KOIDE, T. and ONO, T. The complete amino acid sequence of barley trypsin inhibitor.  Journal of Biological Chemistry, July 1983, vol. 258, no. 13, p. 7998-8003.
  • ODANI, S. and IKENAKA, T. The amino acid sequences of two soybean double-headed proteinase inhibitors and evolutionary considerations on the legume proteinase inhibition. Journal of Biochemistry, September 1976, vol. 80, no. 3, p. 641-643.
  • OPPERMAN, C.H.; TAYLOR, C.G. and CONKLING, M.A. Rootknot nematode directed expression of a plant root specific gene. Science, 1994, vol. 263, p. 221-223.
  • OPPERT, B.; MORGAN, T.D.; CULBERTSON, C. and KRAMER, K.J. Dietary mixtures of cysteine and serine proteinase inhibitors exhibit synergistic toxicity toward the red flour beetle, Tribolium castaneum. Comparative Biochemistry and Physiology, 1993, vol. 105C, p. 379-385.
  • ORR, G.L.; STRICKLAND, J.A. and WALSH, T.A. Inhibition of Diabrotica larval growth by a multicystatin from potato tubers. Journal of Insect Physiology, 1994, vol. 40, p. 893-900.
  • PANNEKOEK, H.; VAN MEIJER, M.; SCHLEEF, R.R.; LOSKUTOFF, D.J. and BARBAS, III. Functional display of human plasminogen activator inhibitor 1 (PAI-1) on phages: Novel perspectives for structure-function analysis by error-prone DNA synthesis. Gene, June 1993, vol. 128, no. 1, p. 135-140.
  • PANNETIER, C.; GIBAND, M.; COUZI, P.; LETAN, V.; MAZIER, M.; TOURNEUR, J. and HAU, B. Introduction of new traits into cotton through genetic engineering: insect resistance as example. Euphytica, 1997, vol. 96, p. 163-166.
  • PARK, H.; YAMANAKA, N.; MIKKONEN, A.; KUSAKABE, I. and KOBAYASHI, H. Purification and characterization of aspartic proteinase from sunflower seeds. Bioscience Biotechnology and Biochemistry, 2000, vol. 64, p. 931-939.
  • PERLER, F.; EFSTRATIADIS, A.; LOMEDICO, P.; GILBERT, N.; KOLODNER, R. and OODGSON, J. The evolution of genes: the chicken preproinsulin gene. Cell, June 1980, vol. 20, no. 2, p. 555-556.
  • RAKWAK, R.; AGARWAL, K.G. and JHA, N.S. Characterization of a rice (Oryza sativa L.) Bowman-Birk proteinase inhibitor: tightly light regulated induction in response to cut, jasmonic acid, ethylene and protein phosphatase 2A inhibitors. Gene, January 2001, vol. 263, no. 1-2, p. 189-198.
  • RANCOUR, J.M. and RYAN, C.A. Isolation of a carboxypeptidase B inhibitor from potatoes. Archives of Biochemistry and Biophysics, 1968, vol. 125, p. 380-382.
  • RAVICHANDARAN, S.; SEN, U.; CHAKRABARTI, C. and DATTAGUPTA, J.K. Cryocrystallography of a Kunitz type serine protease inhibitor: 90 K structure of winged bean chymotrypsin WCI) at 2.13 A resolution. Acta Crystallography D Biology Crystallography, 1999, vol. 55, p. 1814-1821.
  • RAWLINGS, N.D. and BARRETT, A.J. Evolutionary families of metallopeptidases. In: BARRETT, A.J., ed. Methods in Enzymology. New York, Academic Press, 1995, vol. 248, p. 183-228.
  • REECK, G.R.; KRAMER, K.J.; BAKER, J.E.; KANOST, M.R.; FABRICK J.A. and BEHNKE, C.A. Proteinase inhibitors and resistance of transgenic plants to insects. In: CAROZZI, N. and KOZIEL, M., eds. Advances in insect control: the role of transgenic plants. London, Taylor and Francis, 1997, p. 157-183.
  • RICHARDSON, M.J. Seed storage proteins: The enzyme inhibitors. In: RICHARDSON, M.J, ed. Methods in Plant Biochemistry, New York, Academic Press, 1991, p. 259-305.
  • ROGERS, B.L.; POLLOCK, J.; KLAPPER, D.G. and GRIFFITH, I.J. Sequence of the proteinase inhibitor cystatin homolog from the pollen of Ambrosia artemisiifolia (short ragweed). Gene, 1993, vol. 133, p. 219-221.
  • ROTTGEN, P. and COLLINS, J. A human pancreatic secretory trypsin inhibitor presenting a hypervariable highly constrained epitope via monovalent phagemid display. Gene, October 1995, vol. 194, no. 2, p. 243-250.
  • ROUSH, R.T. and MCKENZIE, J.A. Ecological and genetics of insecticide and araricide resistance. Annual Review of Entomology, 1987, vol. 32, p. 361-381.
  • RYAN, Clarence A. The systemin signaling pathway: differential activation of plant defensive genes. Biochim Biophys Acta, March 2000, vol. 1477, no. 1-2, p. 112-121.
  • RYAN, Clarence A. Protease inhibitors in plants: genes for improving defenses against insects and pathogens. Annual Review of Phytopathology, 1990, vol. 28, p. 425-449.
  • RYAN, Clarence A. Insect-induced chemical signals regulating natural plant protection responses. In: DENNO, R.F. and MCCLURE M.S., eds. Variable plants and herbivores in natural and managed systems. New York, Academic Press, 1989, p. 43-60.
  • SANE, V.A.; NATH, P.; AMINUDDIN and SANE, P.V. Development of insect-resistant transgenic plants using plant genes: expression of cowpea trypsin inhibitor in transgenic tobacco plants. Current Science, 1997, vol. 72, p. 741-747.
  • SARDANA, R.K.; GANZ, P.R.; DUDANI, A.K.; TACKABERRY, E.S.; CHENG, X. and ALTOSAAR, I. Synthesis of recombinant human cytokine GMCSF in the seeds of transgenic tobacco plants. In: CUNNINGHAM, C. and PORTER A.J.R., eds. Recombinant proteins from plants. Production and isolation of clinically useful compounds. Totowa NJ, Humana Press, 1998, p. 77-87.
  • SHULKE, R.H. and MURDOCK, L.L. Lipoxygenase trypsin inhibitor and lectin from soybeans: effects on larval growth of Manduca sexta (Lepidoptera: Sphingidae). Environmental Entomology, 1983, vol. 12, p. 787-791.
  • SONG, I.; TAYLOR, M. and BAKER, K. Inhibition of cysteine proteinases by Carica papaya cystatin produced in Escherichia coli. Gene, September 1995, vol. 162, no. 2, p. 221-224.
  • SPARBER, S.B. Insect resistance to biological insecticide Bacillus thuringiensis. Science, 1985, vol. 229, p. 193-195.
  • STEVENS, F.C.; WOERY, S. and KRAHU, J. Structure-function relationship in lima bean proteinase inhibitor. In: FRITZ, H., TSCHESCHE, H., GREENE, L.J and TRUSCHEIT, E., eds. Proteinase inhibitors, 1974, Berlin, Springer-Verlag, p. 344-354.
  • STRATMANN, Johannes W.; SCHEER, Justin and RYAN, Clarence A. Suramin inhibits initiation of defense signaling by systemin, chitosan, and a beta-glucan elicitor in suspension-cultured Lycopersicon peruvianum cells. Proceedings of the National Academy of Sciences United States of America, August 2000a, vol. 97, no. 16, p. 8862-8867.
  • STRATMANN, Johannes W.; STELMACH, B.A.; WEILER, E.W. and RYAN, Clarence A. UVB/UVA radiation activates a 48 kDa myelin basic protein kinase and potentiates wound signaling in tomato leaves. Photochemistry and Photobiology, February 2000b, vol. 71, no. 2, p. 116-23.
  • STINTZI, Annick; WEBER, Hans.; REYMOND, Philippe; BROWSE, John and FARMER, Edward E. Plant defense in the absence of jasmonic acid: the role of cyclopentenones. Proceedings of the National Academy of Sciences United States of America, October 2001, vol. 98, no. 22, p. 12837-12842.
  • SUZUKI, A.; TSONOGA, Y.; TANAKA, I.; YAMANE, T.;  ASHIDA, T.; NORIOKA, S.; HARA, S. and IKENAKA, T. The structure of Bowman-Birk type protease inhibitor A-II from peanut (Arachis hypogaea) at 3.3A resolution. Journal of Biochemistry, 1987, vol. 101, p. 267-274.
  • TANAKA, A.S.; SAMPAIO, C.A; FRITZ H. and AUERSWALD, E.A. Functional display and expression of chicken cystatin using a phagemid system. Biochemical and Biophysical Research Communications, 1995, vol. 214, p. 389-395.
  • TAYLOR, M.A.J.; BAKER, K.C.; BRIGGS, G.S.; CONNERTON, I.F.; CUMMINGS, N.J.; PRATT, K.A.; REVELL, D.F.; FREEDMAN, R.B. and GOODENOUGH, P.W. Recombinant proregions from papain and papaya proteinase IV are selective high affinity inhibitors of the mature papaya enzymes. Protein Engineering, January 1995, vol. 8, no. 1, p. 59-62.
  • TERRA , W.R.; FERREIRA, C. and JORDAO, B.P. Digestive enzymes. In: LEHANE M. J., ed. Billin London, Chapman and Hall, 1996, p.153-194.
  • THOMAS, J.C.; ADAMS, D.G.; KEPPENNE, V.D.; WASMANN, C.C.; BROWN, J.K.; KANOSH, M.R. and BOHNERT, H.J. Proteinase inhibitors of Manduca sexta  expressed in transgenic cotton. Plant Cell Reports, 1995, vol. 14, p. 758-762.
  • THOMAS, J.C.; WASMANN, C.C.; ECHT, C.; DUNN, R.L.; BOHNERT, H.J. and MCCOY, T.J. Introduction and expression of an insect proteinase inhibitor in alfalfa (Medicago sativa L.). Plant Cell Reports, 1994, vol. 14, p. 31-36.
  • TIFFIN, Peter and GAUT, Brandon S. Molecular evolution of the wound- induced serine protease inhibitor wip1 In Zea and related genera. Molecular Biology and Evolution, 2001, vol. 18, p. 2092-20101.
  • TITARENKO, E.; ROJO, E.; LEON, J.; SANCHEZ-SERRANO, J.J. Jasmonic acid-dependent and -independent signaling pathways control wound-induced gene activation in Arabidopsis thaliana. Plant Physiology, 1997, vol. 115, p. 817-826.
  • TURK, V. and BODE, W. The cystatins: protein inhibitors of cysteine proteinases. FEBS Letters, 1991, vol. 285, p. 213-219.
  • URWIN, P.E.; LILLEY, C.J.; MCPHERSON, M.J. and ATKINSON, H.J. Resistance to both cyst and rootnot nematodes conferred by transgenic Arabidopsis expressing a modified plant cystatin. Plant Journal, August 1997, vol. 12, no. 2, p. 455-461.
  • URWIN, P.E.; ATKINSON, H.J.; WALLER, D.A. and MCPHERSON, M.J. Engineered oryzacystatinI expressed in hairy roots confers resistance to Globodera pallida. Plant Journal, July 1995, vol. 8, no. 1, p. 121-131.
  • VAIN, P.; WORLAND, B.; CLARKE, M.C.; RICHARD, G.; BEAVIS, M.; LIU, H.; KOHLI, A.; LEECH, M.; SNAPE, J. and CHRISTOU, P. Expression of an engineered cysteine proteinase inhibitor (Oryzacystatin-I delta D86) for nematode resistance in transgenic rice plants. Theoretical and Applied Genetics, 1998, vol. 96, p. 266-271.
  • VISAL, S.; TAYLOR, M.A.J. and MICHAUD, D. The proregion of papaya IV inhibits Colorado potato beetle digestive cysteine proteinases. FEBS Letters, September 1998, vol. 434, no. 3, p. 401-405.
  • WALDRON, C.; WEGRICH, L.M.; OWENS METRO, P.A. and WALSH, T.A. Characterization of a genomic sequence coding for potato multicystatin, of eight-domain cysteine proteinase inhibitor. Plant Molecular Biology, 1993, vol. 23, p. 801-812.
  • WALKER, A.J.; FORD, L.; MAJERUS, M.E.N.; GEOGHEGAN, I.E.; BIRCH, A.N.E.; GATEHOUSE, J.A. and GATEHOUSE, A.M.R. Characterisation of the midgut digestive proteinase activity of the two-spot ladybird (Adalia bipunctata L.) and its sensitivity to proteinase inhibitors. Insect Biochemistry and Molecular Biology, 1998, vol. 28, p. 173-180.
  • WANG, C.I.; YANG, Q. and CRAIK, C.S. Isolation of a high affinity inhibitor of urokinase-type plasminogen activator by phage display of ecotin. Journal of Biological Chemistry, May 1995, vol. 270, no. 20, p. 12250-12256.
  • WASTERNACK, C. and PARTHIER, B. Jasmonate-signalled  plant gene expression. Trends in Plant Science, 1997, vol. 2, p. 302-307.
  • WIEMAN, K.F. and NIELSEN, S.S. Isolation and partial characterization of a major gut proteinase from larval Acanthoscelides obtectus Say (Coleoptera: bruchidae). Comparative Biochemistry and Physiology-B, 1988, vol. 89, p. 419-426.
  • WILLIAMSON, V.M. and HUSSEY, R.S. Nematode pathogenesis and resistance in plants. Plant Cell, October 1996, vol. 8, p. 1735-1745.
  • WOLFSON, J.L. and MURDOCK, L.L. Diversity in digestive proteinase activity among insects. Journal of Chemical Ecology, 1990, vol. 16, p. 1089-1102.
  • WOLFSON, J.L. and MURDOCK, L.L. Suppression of larval Colorado beetle growth and development by digestive proteinase inhibitors. Entomologia Experimentalis et Applicata, 1987, vol. 44, p. 235-240.
  • XU, D.P.; XUE, Q.Z.; MCELROY, D.; MAWAL, Y.; HILDER, V.A. and WU, R. Constitutive expression of a cowpea trypsin inhibitor gene, CpTi, in transgenic rice plants confers resistance to two major rice insect pests. Molecular Breeding, 1996, vol. 2, p. 167-173.
  • YEH, K.W.; LIN, M.L.; TUAN, S.J.; CHEN, Y.M.; LIN, C.Y. and KAO, S.S.  Sweet potato (Ipomea batatas) trypsin inhibitors expressed in transgenic tobacco plants confer resistance against Spodoptera litura. Plant Cell Reporter, 1997, vol. 16, p. 696-699.
  • ZHANG, Y.S.; LOS, S.; TAN, F.L.; CHI, C.W.; XU, L.X. and ZHANG A.L. Complete amino acid sequence of mung bean trypsin inhibitor. Science Sinica, March 1982, vol. 25, no. 3, p. 268-277.
  • ZHAO, Y.; BOTELLA, M.A.; SUBRAMANIAN, L.; NIU, X.;  NIELSEN, S.S.; BRESSAN, R.A. and HASEGAWA, P.M. Two wound inducible soybean cysteine proteinase inhibitors have greater insect digestive proteinase inhibitory activities than a constitutive homolog. Plant Physiology, August 1996, vol. 111, no. 4, p. 1299-1306.
  • ZHU-SALZMAN, K.; SALZMAN, R.A.; KOIWA, H.; MURDOCK, L.L.; BRESSAN, R.A. and HASEGAWA, P.M. Ethylene negatively regulates local expression of plant defense lectin genes. Physiologia Plantarum, 1998, vol. 104, p. 365-372.

Note: EJB Electronic Journal of Biotechnology is not responsible if on-line references cited on manuscripts are not available any more after the date of publication.

Supported by UNESCO / MIRCEN network 

© 2002 by Universidad Católica de Valparaíso -- Chile


The following images related to this document are available:

Photo images

[ej02015t1.jpg] [ej02015t2.jpg]
Home Faq Resources Email Bioline
© Bioline International, 1989 - 2024, Site last up-dated on 01-Sep-2022.
Site created and maintained by the Reference Center on Environmental Information, CRIA, Brazil
System hosted by the Google Cloud Platform, GCP, Brazil