Next Article in Journal
The Long Non-Coding RNA HOXA-AS2 Promotes Proliferation of Glioma Stem Cells and Modulates Their Inflammation Pathway Mainly through Post-Transcriptional Regulation
Next Article in Special Issue
Immune Checkpoint Blockade via PD-L1 Potentiates More CD28-Based than 4-1BB-Based Anti-Carbonic Anhydrase IX Chimeric Antigen Receptor T Cells
Previous Article in Journal
Lifespan Extension of Podospora anserina Mic60-Subcomplex Mutants Depends on Cardiolipin Remodeling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Plant Kunitz Inhibitors and Their Interaction with Proteases: Current and Potential Pharmacological Targets

by
Camila Ramalho Bonturi
1,
Ana Beatriz Silva Teixeira
1,
Vitória Morais Rocha
1,
Penélope Ferreira Valente
1,
Juliana Rodrigues Oliveira
1,
Clovis Macêdo Bezerra Filho
1,
Isabel Fátima Correia Batista
2,* and
Maria Luiza Vilela Oliva
1,*
1
Departamento de Bioquímica, Universidade Federal de São Paulo, Escola Paulista de Medicina, Rua Três de Maio, 100, São Paulo 04044-020, SP, Brazil
2
Center of Excellence in New Target Discovery (CENTD), Instituto Butantan, Av. Vital Brasil, 1500, São Paulo 05503-900, SP, Brazil
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(9), 4742; https://doi.org/10.3390/ijms23094742
Submission received: 1 April 2022 / Revised: 19 April 2022 / Accepted: 21 April 2022 / Published: 25 April 2022
(This article belongs to the Special Issue State-of-the-Art Molecular Oncology in Brazil)

Abstract

:
The action of proteases can be controlled by several mechanisms, including regulation through gene expression; post-translational modifications, such as glycosylation; zymogen activation; targeting specific compartments, such as lysosomes and mitochondria; and blocking proteolysis using endogenous inhibitors. Protease inhibitors are important molecules to be explored for the control of proteolytic processes in organisms because of their ability to act on several proteases. In this context, plants synthesize numerous proteins that contribute to protection against attacks by microorganisms (fungi and bacteria) and/or invertebrates (insects and nematodes) through the inhibition of proteases in these organisms. These proteins are widely distributed in the plant kingdom, and are present in higher concentrations in legume seeds (compared to other organs and other botanical families), motivating studies on their inhibitory effects in various organisms, including humans. In most cases, the biological roles of these proteins have been assigned based mostly on their in vitro action, as is the case with enzyme inhibitors. This review highlights the structural evolution, function, and wide variety of effects of plant Kunitz protease inhibitors, and their potential for pharmaceutical application based on their interactions with different proteases.

Graphical Abstract

1. Introduction

Protease inhibitors (PIs) are widely distributed in nature, including in animals, plants, and microorganisms. Particularly abundant in plants, PIs account for around 1–10% of the total protein in storage organs, such as seeds and tubers [1,2]. Classified into families based on their structural features, plant inhibitors generally share cysteine patterns, overall 3D structure, and mechanism of action. The most traditional classification system sorts the inhibitors into different families, including Kunitz, Bowman–Birk, Kazal, Potato I, Potato II, and Serpins [3]. The Rawlings classification is based on similarities in the amino acid sequence, grouping the inhibitors according to their structures [4,5,6]. Consolidated aspects include molecular mass, cysteine content, and disulfide bonds; for example, members of the serine protease inhibitor family (Serpin family) exhibit well-conserved structures composed of nine α-helices and three β-sheets [7]. The Bowman–Birk family are 8–10 kDa multifunctional double-headed inhibitors (two distinct reactive sites) linked by seven disulfide bonds [8], and the Kazal family (containing at least one Kazal inhibitory domain) is widespread among microorganisms, invertebrates, vertebrates, and plants, and is characterized by six highly conserved cysteine residues forming three intradomain disulfide bonds [9].
Another important family of protease inhibitors is the plant Kunitz-type, which has a molecular weight of 20–22 kDa, and may be composed of one or two polypeptide chains linked by one or two disulfide bonds [6]. Their structures may lack cysteine residues [10,11] or contain up to five Cys residues.
Along with the study of the structural and kinetic aspects of protease inhibitors, the analysis of their interaction with proteolytic enzymes has contributed to the understanding of the phenomena involved in the control of proteolysis (Graphical Abstract). Systematic studies with isolated plant inhibitors have shown that structurally characterized proteins can be used as useful models, especially to study the variability of enzyme specificity, helping us understand the mechanism of cellular action of proteases in pathophysiological phenomena and the development of microorganisms.

2. Plant Kunitz Inhibitors

Although widely distributed in nature, plant Kunitz inhibitors (I3) are especially abundant in the Fabaceae (Leguminosae) family, which comprises the Mimosoideae, Caesalpinioideae, and Faboideae/Papilionoideae subfamilies [10]. The Soybean Kunitz trypsin inhibitor (SKTI) was the first member of this family to be identified [12], and was reported by M. Kunitz; inhibitors subsequently discovered that share similar characteristics are therefore named Kunitz inhibitors. Generally, Kunitz inhibitors are defense proteins, which exhibit different activities as highlighted in this review, including antibacterial and antifungal activities, and act on inflammation, coagulation, thrombosis, and cancer.
Plant Kunitz inhibitors generally exhibit highly conserved primary structures. These inhibitors usually have a reactive site located in the region that binds to the enzyme, and the formation of the enzyme–inhibitor complex occurs in 1:1 stoichiometry. Their reactive sites are frequently composed of Arg and Lys; however, they may occasionally contain Glu, Ala, or Val [6]. These inhibitors can be single- or double-polypeptide-chain proteins, with variation in the number of cysteine residues and disulfide bridge patterns. The Bauhinia bauhinioides cruzipain inhibitor (BbCI) [13] and Bauhinia rufa trypsin inhibitor (BrTI) [14] are single-chain inhibitors lacking cysteine residues. The B. bauhinioides kallikrein inhibitor (BbKI) is also a single-chain inhibitor; however, it contains one cysteine residue [11,15]. The B. rufa elastase inhibitor (gBrEI) [16], Swartzia pickellii trypsin inhibitor (SWTI) [17], Entada scandens trypsin inhibitor (ESTI) [18], and Inga laurina trypsin inhibitor (IlTI) [19] are single-chain inhibitors with two cysteine residues. Although the Copaifera langsdorffii inhibitor (ClI) also contains two cysteines, it is a double-chain inhibitor, with the chains linked through noncovalent bridges [20]. The classic Kunitz-type inhibitors, soybean trypsin inhibitor (STI) [21], and Bauhinia variegata var. candida trypsin inhibitor (BvTI-C) [22] are single-polypeptide-chain proteins that possess four cysteines. The Enterolobium contortisiliquum trypsin inhibitor (EcTI) [23,24] and Acacia karroo trypsin inhibitor (AkTI) [25] also contain four cysteines; however, they are double-polypeptide-chain proteins linked by two S–S bridges. Canavalia lineata subtilisin inhibitor type II (CLSI-II) is a similar double-chain inhibitor linked by two S–S bridges, but contains five Cys residues, four of which are linked with bridges and one is a free cysteine [26]. These profiles are summarized in Figure 1.

3. Recombinant Kunitz Inhibitors

Natural compounds such as plant proteins have been used for a long time to improve human health. However, the yield of protease inhibitors obtained from natural plant sources is generally low. With the advent of genetic engineering, it is now possible to obtain large amounts of these inhibitors as recombinant proteins at a low cost, while maintaining or even improving their inhibitory activities.
The recombinant B. bauhinioides kallikrein inhibitor (rBbKI) and recombinant B. bauhinioides cruzipain inhibitor (rBbCI) were obtained by heterologous expression and production in Escherichia coli [27]. While BbCI (a cruzipain and elastase inhibitor) and BbKI (effective in inhibiting coagulation and fibrinolysis enzymes) share 82% sequence similarity, their reactive sites show slight differences. Based on this information, rBbKI was constructed by mutating the R64 residue in the BbKI reactive site to A64, as found in BbCI. This single mutation modified the specificity of BbKI, turning it into a weak plasma kallikrein inhibitor (Kiapp = 2.0–98 nM) and a strong plasmin inhibitor (Kiapp = 33–2.6 nM) [11]. Thus, the recombinant inhibitor can prevent fibrin breakup and maintain blood clot stability, which opens up new avenues for its application as a therapeutic protein (Table 1).
The same research group obtained rBbKIm based on the primary sequence of BbKI, but incorporated the RGD/RGE (Arg–Gly–Asp/Arg–Gly–Glu) motifs from BrTI. To this end, the mutations V21E, S24A, H25R, H27D, A28G, E127D, and Q130E were introduced into BbKI. The importance of these residues in the defense against predatory insects has been demonstrated [28]. This inhibitor also interfered with the viability of prostate cancer DU145 and PC3 cells, leading to cell death via the release of cytochrome c and caspase-3 activation [29] (Table 2).
Moreover, the Ligusticum chuanxiong subtilisin/alpha-amylase inhibitor (LASI), expressed in E. coli Rosetta and obtained from the rhizome of L. chuanxiong (frequently used in Chinese medicine to treat headaches, cardiovascular diseases, rheumatic joint pain, menstrual disorders, and painful swelling), was also obtained as a recombinant protein. LASI inhibits mammalian alpha-amylase and bacterial subtilisin (which may be involved in the protection of plants against microorganisms) and also affects pest lepidopterans, which are all relevant for agricultural applications. Although LASI shares sequence similarities with other Kunitz protease inhibitors, such as Cynara cardunculus (68%), it has a unique reactive site; consequently, LASI inhibits subtilisin but does not affect trypsin or chymotrypsin. Furthermore, LASI inhibits alpha-amylase from Plutella xylostella, Helicoverpa armigera, and Spodoptera litura, thereby reducing the survival of these pests [30].
The insecticidal properties of the Cassia obtusifolia trypsin inhibitor (CoTI1) have also been described [31]. CoTI1 was used to investigate stress tolerance and increase the resistance of crops against pests. It is a potent trypsin inhibitor (2.490 UI/mg activity) and also inhibits trypsin-like proteases from H. armigera, S. litura, and S. exigua (1.770 UI/mg, 1.510 UI/mg, and 1.210 UI/mg, respectively). The amino acids L84, R86, and T88 in CoTI1 were identified, using in silico analysis, as being involved in trypsin inhibition, which was further confirmed by the loss of inhibitory activity (54%, 90%, and 53%, respectively) on mutating these residues to alanine.
Inhibition of human neutrophil elastase (HNE) is known to be beneficial in modulating inflammatory lung diseases. The Caesalpinia echinata elastase inhibitor (CeEI), which inhibits HNE, has been used to study pulmonary inflammation [32]. Native CeEI is a potent inhibitor of HNE (Ki 1.9 nM), cathepsin G (CatG, Ki 3.6 nM), and protease 3 (Ki 3.7 μM). Two recombinant isoinhibitors of CeEI, rCeEI-4, and rCeEI-5 were cloned and expressed in E. coli. rCeEI-5 inhibited HNE 6.2 times more efficiently than CeEI, while rCeEI-4 exhibited 4.2-fold higher inhibition of chymotrypsin and 5.3-fold higher inhibition of HNE than CeCI. Synthetic, smaller functional peptides that inhibit HNE rCeEI-36 (Ki = 0.3 nM) and rCeEI-46 (Ki = 8.8 nM) were also obtained from the CeEI-4 sequence.
Apios americana, a legume tuber plant widely distributed across North America, is commonly used to make bread in Japan [33]. This plant produces two Kunitz-type protease inhibitors, Apios americana protease inhibitor 1 (AKPI-1) and Apios americana protease inhibitor 2 (AKPI-2), which were cloned and expressed as recombinant proteins. rAKPI-2, characterized as chymotrypsin (Ki 6 µM) and trypsin inhibitor (Ki 0.4 µM), loses its property of trypsin inhibition when heated at 70 °C; however, the chymotrypsin inhibitory activity is maintained at 80% of its initial value at the same temperature.
Although most recombinant inhibitors are produced in bacterial hosts, it is also possible to do so with yeast, as achieved by Bunyatang et al. [34]. The PKPI protease inhibitor, a bifunctional inhibitor that acts on protease and amylase, was cloned and transferred into Pichia pastoris to produce a full-length protease inhibitor based on the Hevea brasiliensis inhibitor (HbASI), which showed high activity against alpha-amylase (from Aspergillus oryzae, IC50 200 µg), subtilisin A (IC50 380 µg), and trypsin (IC50 325 µg), but no activity against chymotrypsin and human alpha-amylase. In addition, HbASI showed good stability over a wide temperature and pH range, and inhibited mycelial growth of Phytophthora palmivora (0.2 µM), including 50% inhibition of zoospore (Table 2).

4. Defense Proteins

Plants can detect events locally and transmit defense signals to build up systemic acquired resistance (SAR). A Kunitz protease inhibitor in the tea plant Camellia sinensis (CsKPI1) was shown to be potentially involved in SAR based on its gene expression, which was strongly induced by insects through localized feeding, and in neighboring undamaged leaves. Jasmonic acid and CsKPI1 were detected simultaneously in damaged leaves; in addition, the jasmonic acid content also increased in neighboring leaves, suggesting its role as a vital hormonal signal that induces the expression of CsKPI1 in SAR [67].
Based on their ability to function as defense proteins that protect against predator attacks, protease inhibitors have been used as insecticidal agents that act by inhibiting digestive proteases in insects, thereby controlling the availability of amino acids for the growth and development of larvae (Figure 2, Table 2). The Kunitz-type Cassia leiandra trypsin inhibitor (ClTI) inhibits midgut digestive proteases (50% activity reduction at 4.65 µM) and larvae development (after 10 days at a final concentration of 15.4 µM), and delays adult emergence, in Aedes aegypti, causing 44% mortality. These findings suggest that ClTI has the potential to reduce the A. aegypti population without increasing the risk of developing insect resistance [43]. Similarly, alocasin, purified from the rhizome of Alocasia macrorrhiza, increased the mortality rate of A. aegypti larvae and adult females (IC50 = 0.17 mg/mL) by its action against the midgut enzymes of this insect. However, higher inhibitor concentrations lead to a reduction in the inhibitory effect, which may be attributed to the production of insect proteases insensitive to the inhibitor upon prolonged exposure [38]. In another study, proteolytic activity was restored on the 15th day after hatching of Anticarsia gemmatalis larvae exposed to SKTI, after a decrease in the total proteolytic activity of trypsin and similar enzymes until the 12th day [68]. On the other hand, the Adenanthera pavonina Kunitz-type inhibitor (ApKTI) acts through a noncompetitive mechanism, thereby offering an advantage over such adaptation mechanisms as its affinity for trypsin is not affected by amino acid mutations in the enzyme active site, unlike other competitive protease inhibitors. Thus, ApKTI showed insecticidal activity even after 15 days of chronic exposure when included in the artificial diet of Plodia interpunctella larvae [36], and achieved a 60% reduction in the survival of neonatal Anticarsia gemmatalis larvae [37]. According to Sasaki et al. [69], A. aegypti larvae also do not use regulatory mechanisms to overcome the effects of ApKTI, and the inhibitor demonstrated degeneration of the microvilli of epithelial cells of the posterior midgut region, hypertrophy of the gastric cecal cells, and an augmented ectoperitrophic space in larvae. Interestingly, although trypsin and chymotrypsin-like serine proteases are involved in the initial protein digestion of Lepidoptera, ILTI (from Inga laurina) inhibits trypsin in the midgut of S. frugiperda more effectively compared to other protease inhibitors, such as SKTI and Inga vera trypsin inhibitor (IVTI). Neither SKTI nor ILTI was active against chymotrypsin. Notably, ILTI was not degraded by enzymes in the midgut of insects, and was stable after elimination in feces, showing resistance to microbial metabolism [59]. IVTI exerted midgut inhibition in S. frugiperda, C. cephalonica, H. virescens, and H. zea of 83%, 80%, 70%, and 64%, respectively [60]. More recently, other inhibitors have been characterized, such as RsKI from Rhynchosia sublobata, which is active against lepidopteran gut proteases, including those of Achaea janata (IC50 = 1.95 µg) and Helicoverpa armigera (IC50 = 59 µg) [66], and the inhibitor CpPI 63, isolated from Cajanus platycarpus seeds, a wild relative of pigeonpea [70].
While protease inhibitors are important in defending against predators, they can also selectively assist in mutualistic relationships. For example, the ant Pseudomyrmex ferrugineus uses Acacia sp. hollow thorns as a nesting space, and while the Acacia sp. provides food bodies and extrafloral nectar for ant larvae, the ant offers protection against herbivores that compete for vegetation. Thus, protease inhibitors extracted from Acacia sp. food bodies are activated and convert the nutritive food reward into something difficult to digest for potential or optional exploiters, but not for mutualistic consumers, whose biochemistry takes advantage of this reward [71].
Cowpeas are an important food source for millions of people worldwide. Many studies have used Callosobruchus maculatus [72] and Zabrotes subfasciatus [73] as models to evaluate the effect of protease inhibitors because these beetles cause serious damage to the grain, making it unfit for human consumption. The Pithecellobium dumosum Kunitz inhibitor (PdKI), purified from the seeds of the P. dumosum tree, was evaluated against cysteine and serine proteases extracted from the larvae of coleopterans C. maculatus and Z. subfasciatus and the lepidopterans Alabama argillacea and Telchin licus. It completely inhibited bovine trypsin at 100 nM, with 50% inhibition against bovine chymotrypsin. PdKI showed specificity for serine proteases, especially against those from phytophagous insects, such as Coleoptera and Lepidoptera [74].
Many factors are involved in the adaptability of insects to protease inhibitors, such as the alternative use of other classes of proteases, overexpression of proteases to minimize the inhibitory effect, or the degradation of the inhibitors by nontarget proteases, which are insensitive to the inhibitors. Thus, despite the potential applicability of protease inhibitors as effective compounds for the protection of crops against herbivory, most recent studies have pointed to the importance of multitrophic approaches, taking into account not only target insect proteases, but also proteases from other organisms in the food chain, including the plant itself [75].
Termites are one of the main wood-destroying pests that can invade and attack the wooden building structures in urban environments. While synthetic pesticides are traditionally used to prevent termite attacks, these compounds tend to harm humans and the environment. As an alternative, the trypsin inhibitor isolated from Cassia grandis seeds (CgTI) can be employed, which exhibits termiticidal activity against Nasutitermes corniger. However, the activity has been found to vary between termite castes, with workers (IC50 = 0.685 mg/mL) being more resistant than soldiers (IC50 = 0.765 mg/mL) [42], which can be attributed to variation in enzymatic activities and digestive tracts between the castes. Similar to these results, EcTI was able to induce the death of worker termites, with an IC50 of 0.242 mg/mL, whereas it did not exert significant inhibition in soldier termites. Notably, the molecule from the intestinal extract that binds to EcTI is a chitinase from T. reesei, a symbiotic fungus present in the digestive tract of termites that is important in the cellulose degradation process. As this cellulose is exploited for the production of bioethanol, EcTI could be applied as a biotechnological tool in addition to being a natural active termiticide [50].
In the same line of investigation, two distinct recombinant protease inhibitors of B. bauhinioides origin, rBbKI (serine protease inhibitor) and rBbCI (cysteine protease inhibitor) were evaluated for their termiticidal activity. Despite sharing 82% structural similarity, the two inhibitors showed differing specificity for termite intestinal enzymes, and could be interesting tools for the characterization of N. corniger enzymes. rBbKI showed termiticidal activity in workers, with an LC50 of 0.9 mg/mL after four days, and did not affect the survival of soldiers, whereas rBbCI did not show any termiticidal activity on N. corniger [64]. In another study, Araucaria angustifolia pine nut extract was shown to exhibit termiticidal activity against N. corniger workers and soldiers at all tested concentrations [76].
Finally, the bifunctional lectin and Kunitz inhibitor AEL, isolated from Abelmoschus esculentus, was shown to exhibit activity against the fruit fly Ceratitis capitata and the nematodes Meloidogyne incognita and Meloidogyne javanica, which infest plant roots. In the pupal stage of M. incognita, 2 mg/mL AEL showed higher toxicity than organophosphates commonly used in agriculture, exemplifying the potential of protease inhibitors as promising insecticidal molecules [35] (Table 2).

5. Antibacterial and Antifungal Activities

Fungi and bacteria release proteases into the media during infections, which facilitates their defense and survival. Inhibitors of these proteases can therefore hinder pathogen metabolism to improve human health [77]. Plant inhibitors from the Kunitz family may be important in controlling fungal and bacterial growth, as demonstrated by the inhibition of proteases secreted by Bacillus sp. (Figure 2) and Aspergillus flavus induced by extracts from seeds of Cassia tora (L.) syn Senna tora (L.) Roxb [78]. In addition, the isolated protein fraction from Erythrina poeppigiana interferes with the germination of A. flavus spores [79]. ClTI from C. leiandra seeds inhibited the growth of Candida albicans, Candida tropicalis, Candida parapsilosis, and Candida krusei, with values of 5.13 μM, being most potent for C. albicans, with an IC50 of 2.10 μM [80].
Some inhibitors show more restricted specificity, such as the Inga edulis trypsin inhibitor (IeTI). Although it has no action against C. albicans, it is a potent inhibitor of C. tropicalis, as well as Candida buinensis (87% inhibition with 400 μg/mL IeTI). The antifungal activity was found to be mediated by the alteration of the plasma membrane, which diverts energy metabolism from physiological functions to repair it, thereby inducing morphological changes that can affect cell viability or even trigger apoptosis through complexation with vital enzymes in this pathway (Figure 2) [58]. The same mechanism was described in studies with Enterolobium timbouva trypsin inhibitor (EtTI), suggesting that this inhibitor affects the integrity of the plasma membrane of yeasts, inducing the generation of intracellular ROS due to mitochondrial activation to maintain membrane potential, and consequently affecting the organelle structure. In addition, ROS generation is the main cause of DNA damage in yeast, and the mode of action of EtTI appears to involve the activation of apoptosis via a different pathway than traditional antifungal drugs [56,57]. This mechanism is also similar to that used by Pseudostellaria heterophylla trypsin inhibitor (PhTI), a 20.5 kDa thermostable inhibitor that has been reported to show potent activity against the pathogenic fungi Candida gloeosporioides and Fusarium oxysporum. The inhibitor also exhibits high stability, suggesting its potential for use as an antifungal protein in the food industry and agriculture [63].

6. Inflammation

Microbial invasion triggers characteristic inflammatory reactions in the host, such as the continuous migration and degranulation of neutrophils, which release host proteases, such as HNE and CatG [81]. These enzymes are important in the inflammatory response; CatG induces monocyte and neutrophil chemotaxis and activates chemokines [82], while HNE activates the epidermal growth factor (EGF) receptor and induces the expression of IL-8 [83], for example, in emphysema. Although the role of proteases in emphysema is widely accepted, it remains unclear which cells and/or proteases have key functions in disease development and progression. Serine, cysteine, and/or metalloproteinase classes have been reported as the proteases most likely to be involved in the pathogenesis of chronic obstructive pulmonary disease (COPD) and may be the reason for the successful employment of protease inhibitors, as reported for the inhibitors applied in experimental in vivo models of elastase-induced lung injury [84,85] and an asthma model [86], with a reduction in the expression of TNF-α, MMP-9, MMP-12, TIMP-1, eNOS, and iNOS cells in the airways and alveolar walls with inhibitor treatment. However, a decrease in the volume ratio of 8-iso-PGF2, collagen, and elastic fibers in the airways and alveolar walls indicates that these proteins reduce pulmonary inflammation, remodeling, oxidative stress, and mechanical changes.
Crataeva tapia bark lectin (CrataBL) is another inhibitor used in this experimental model [45,46]. The inhibitor reduced the constriction of the airways and lung parenchyma. The effect was associated with a significant reduction in exhaled nitric oxide, and led to the control of the inflammatory response, with a decrease in bronchoalveolar lavage; the number of neutrophils, lymphocytes, and eosinophils; and the mean alveolar diameter. CrataBL was also able to decrease the number of TNF-α positive cells in the lung parenchyma and pulmonary airways, and macrophages in the lung parenchyma. All these investigations suggest that protease inhibitors may have therapeutic potential for the treatment of COPD and provide a better understanding of the involvement of proteases in this pathology.
Bauhinia inhibitors also exhibit protective effects on gastric mucosa [39]. The authors evaluated the in vivo effects of three Kunitz inhibitors on the prevention of gastric mucosal damage: Bauhinia mollis elastase inhibitor (BmEI), B. mollis trypsin inhibitor (BmTI), and B. bauhinioides cruzipain inhibitor (BbCI). These effects were compared with those of ranitidine and saline solution (vehicle). The fraction exhibiting inhibitory activity against HNE (Kiapp 2.8 nM) (Figure 2) and CatG (Kiapp 1.0 nM) exerted a significant antiulcer effect at 3 mg/kg. The same protective effect was observed after pretreatment with the BbCI inhibitor, which inhibits HNE (Kiapp 2.6 nM), CatG (Kiapp 160.0 nM), and the cysteine protease cathepsin L (Kiapp 0.2 nM). In rats, the administration of both BmEI and BbCI resulted in antiulcer effects, indicating the likely inhibition of elastase by both inhibitors, and confirming the relevance of elastase inhibition associated with the action of neutrophil proteases in the regulation of proinflammatory processes and stress-induced gastric mucosal injuries (Table 2).

7. Coagulation and Thrombosis

The activation of the blood coagulation cascade, essentially mediated by serine proteases with specificity for arginine residues, was one of the first biological events in which the importance of reactions involving proteolysis was shown. Arterial and venous thrombosis, consequences of dysfunction of this system, constitute the final triggering process of complications and mortality, not only from diseases of the cardio–circulatory system, but also from other systemic conditions, such as cancer and diabetes, trauma, and even physiological events, such as pregnancy and the postpartum period [87].
Plasma kallikrein (PKa) is involved in numerous proteolytic reactions related to the heart and blood vessels, such as the intrinsic coagulation pathway, kallikrein–kinin system, fibrinolytic system, renin–angiotensin system, complement system pathways involving vascular responses of hemostasis, and pathological conditions, such as cardiovascular and vascular disease [88,89,90].
Furthermore, although PKa does not induce platelet aggregation, it can potentiate platelet aggregation with a low concentration of ADP (2 μM) by binding to platelet integrins through their KGD and KGE motifs and mediating PAR-1 cleavage [91]. The role of coagulation factors XIIa and XIa, and more recently PKa, in thrombus formation, dissociated from hemostasis, has been highlighted by many authors [88,89,90,92]. These investigations show how these proteases contribute to vascular diseases and thrombosis in the intravascular compartment, and are attractive targets for therapeutic inhibition in the treatment and/or prevention of thrombosis.
Numerous plant-derived inhibitors acting on different proteases (released by leukocytes) in the coagulation cascade have been described in more recent studies (Figure 2), such as EcTI, a multispecific inhibitor that acts on PKa, plasmin, human neutrophil elastase, and factor XIIa. It does not, however, inhibit thrombin, human plasmin, or factor Xa (FXa) [93,94].
The BbKI/rBbKI inhibitor from Bauhinia species, whose inhibition constants on human PKa and rat plasma kallikrein are 2.4 nM and 5.2 nM, respectively, is the only inhibitor capable of acting on tissue kallikrein (Kiapp 20 µM). In addition to human plasma kallikrein, BbKI inhibits plasmin. However, this inhibitor does not act on FXa, unlike the inhibitor BuXI from B. ungulata, which is the only inhibitor from Bauhinia that inhibits this enzyme (Kiapp 14 nM), distinguishing itself from the BvTI with which it shares a structural similarity of approximately 70%. In addition to FXa, BuXI inhibits FXIIa (Kiapp 74 nM), PKa (Kiapp 6.9 nM), and plasmin (Kiapp 76 nM). BvTI inhibited plasmin (Kiapp 2.9 nM) more efficiently than BuXI, in contrast to PKa (Kiapp 80 nM) and FXIIa (Kiapp 110 nM), and did not inhibit FXa. The differences in the reactive site are shown to be important to determine the specificity of action. For example, methionine residues at positions 59 and 67 in BuXI play an important role in the interaction with FXa, and do not interfere with trypsin complex formation [10,95].
Some of these inhibitors showed antithrombotic action in different experimental models of arterial and venous thrombosis (Figure 3) [47,49,65,96,97], with little or no interference in bleeding time [47,65], suggesting a potential therapeutic value in inhibiting the development of thromboses with a lower hemorrhagic effect (Table 1).

8. Cancer

In cancer, the overexpression or activation of a protease provides a proteolytically active environment that surrounds the tumor. Importantly, tumor cells overexpress these enzymes to influence the local blood supply, cell extravasation through vessels, and migration through the cell matrix during metastasis.
The presence of protease inhibitors in legume seeds and cereals, in addition to epidemiological studies that identify legumes as possible protective agents in reducing the incidence of certain cancers in the vegetarian population, has prompted several studies to include protease inhibitors for blocking tumor propagation in vivo and in vitro (Figure 3, Table 2), although the mechanisms of inhibitor action are not fully understood [99]. Many proteins found in soybeans are being used for therapeutic purposes, one of which is a Kunitz-type protease inhibitor, with results indicating their ability to inhibit the proliferation of cancer cells. The 24.2 kDa Kunitz-type trypsin inhibitor KTI-A demonstrated antiproliferative activity (IC50 of approximately 6.6 µM) in the melanoma cell line, B16F1. Therefore, soybeans can be used to obtain bioactive molecules according to the requirements of pharmaceutical and other industries [62].
Sporamin, a Kunitz-type trypsin inhibitor found in sweet potatoes (Ipomoea potatoes), was shown to exert anticancer effects, as verified in colorectal cancer lineages and reaffirmed in the study xenografted in mice, where the inhibitor suppressed the growth of colorectal tumor nodules [61]. The study focused on changes in the liver, as it is the first metastatic target of colorectal cancer, and cancer cells can induce the liver to secrete endocrine factors that facilitate tumor growth and transformation. Sporamin (infused intragastrically, 0.5 g/kg body weight/day for three weeks) was found to restore normal liver structure and downregulate the expression and secretion of β-catenin and VEGF in the liver, which subsequently inhibits the transcription of downstream genes involved in cancer progression and angiogenesis [61].
EcTI is cytotoxic to colorectal cancer cells (HCT116 and HT29), breast cancer cells (SkBr-3 and MCF-7), human leukemia cell lines (K562 and THP-1), and mouse fibrosarcoma (L929), without affecting normal human primary fibroblasts or human mesenchymal stem cells (hMSCs) [51,52,53,54]. Its action was investigated in more detail in the cellular events of a human gastric cancer line, Hs746T, and a nontumorigenic fibroblast cell line isolated from the human amniotic fluid [52]. EcTI treatment altered the morphology of Hs746T cells and blocked cell adhesion, migration, and invasion of gastric cancer cells, especially on the type I collagen matrix, impairing the formation of invadopodia. EcTI also decreased plasmin-induced pro-MMP-9 activation and activated pro-MMP-2 mediated by MT1-MMP [93], which is a unique characteristic of the inhibitor that appears to play an important role in its anticancer properties.
EcTI activity has also been studied in a glioblastoma (GBM) model, an aggressive brain tumor with poor overall survival [53]. Treatment of GBM remains a great challenge in oncology, and advanced cell therapies using normal human neural cells and hMSCs appear promising in preclinical experimental conditions, and are being studied as tools for cancer treatment because of their tropism to tumors and immunomodulatory ability. In GBM progression, proteases and their inhibitors are involved in the degradation of the extracellular matrix (ECM) as well as the regulation of other processes by proteolytic trimming and activation or inactivation of several signaling proteins. The effects of EcTI on GBM (U87 cells) and MSCs were compared with those on direct co-culture (MSC/U87) [53]. They showed that metabolic activity was less affected by EcTI in p53-wild type U87 cells than in MSC monoculture, although the metabolic rate of mixed co-culture was significantly reduced at lower EcTI concentrations. Under co-culture conditions, EcTI potentiated MSC-induced cell cycle arrest, probably due to a considerable increase in the expression of p53 and p21, and lower D1 level, but did not affect apoptosis. In addition, EcTI also enhanced Ca2+ signaling mediated via bradykinin receptor 2 in association with nitric oxide release, which highly impaired tumor cell proliferation and invasion. The mechanism did not seem to involve changes in cell adhesion, but rather the downregulation of β1 integrin signaling associated with p-FAK in U87 cells, both promoting inhibition of invasion. Finally, some cytokines were downregulated, suggesting that EcTI-mediated inhibition of signaling may be mediated by cytokines. Thus, in co-cultured MSC/U87 cells, EcTI impaired metabolic activity and proliferation, and reduced invasion, possibly through changes in cytokine secretion. In this context, plant-derived proteins potentiate the anticancer effects induced by MSCs, as demonstrated in the GBM U87 cell line.
The effect of EcTI has also been evaluated in the triple-negative breast cancer cell line, MDA-MB-231. Decreased tumor cell viability and proliferation, as well as collagen I-mediated cell adhesion and the action of MMP-2 and MMP-9, enzymes that play an important role in the degradation of the extracellular matrix and promote tumor invasion and metastasis, have been reported. The inhibitor also reduced Akt and ERK levels, thereby enhancing the detrimental effects on cell viability and proliferation. NFκB, which is overexpressed in cancer and correlates with increased cell survival, was significantly reduced. Reduction in the levels of the proinflammatory cytokines IL-6 and IL-8 has also been reported, which, in addition to being expressed in cells that are resistant to several chemotherapeutic agents, are associated with NF-κB activation. In addition, there was an alteration in the integrin/FAK/SRC, BAX, and BCL-2 pathways, confirming the action of EcTI in carcinogenic processes [54].
Other protease inhibitors have also been studied for their roles in cancer progression. The effect of the recombinant form of the kallikrein inhibitor, rBbKI, on the viability of solid tumors was evaluated in the MKN-28 and Hs746T (gastric cancer), HCT116 (colorectal cancer), SkBr-3 and MCF-7 (breast cancer), and THP-1 and K562 (leukemia) cell lines. rBbKI inhibited cell viability in the Hs746T line more effectively than in nonmetastatic lines, such as MKN-28, HT-29, and HCT 116 [51]. In addition, the chimeric inhibitor rBbKIm [29], which has a peptide sequence containing the RGD motif of another Bauhinia inhibitor [14], has gained attention in tumor cell studies for its significant deleterious effect on prostate cancer. The potential mechanism involves the inhibition of tissue kallikreins, as these enzymes are important for the activation and development of this disease [55,100,101]. Ferreira et al. [29] demonstrated the differential action of rBbKIm on androgen-independent prostate cancer lines, PC3 and DU145, compared to normal amniotic fluid fibroblast cells and human endothelial cells (HUVECs). Treatment of the conditioned medium from the culture of DU145 and PC3 cells with rBbKIm for 48 h inhibited approximately 95% of the hydrolysis of HD-Pro-Phe-Arg-pNa by serine proteases of the kallikrein family. The importance of selective inhibitors of proteases with trypsin-like activity secreted by these cells may be reflected in cellular signaling events triggered by proteolysis, such as the activation of protease-activated receptors (PARs). PAR activation promotes cancer cell proliferation and invasion, and the release of angiogenic factors [101,102], and kallikreins are enzymes that cleave these receptors with a specificity similar to that of thrombin [103]. Thus, rBbKIm probably leads to the deficiency of these cellular events in prostate cancer via its direct inhibitory activity. In contrast, the RGD/RGE motif sequences contribute to recognizing integrins, regulating the invasive potential of metastatic tumor cells [104], and the increase in PAR1 level with the progression of prostate cancer [105]. Bilgin et al. [106] showed that the Kunitz-type inhibitor BbKI induced calcium release from the endoplasmic reticulum and consequently lowered the intracellular calcium concentration, which demonstrated a dose-dependent effect on the proliferation of HUVECs. In addition, BbKI was also found to cause hyperpolarization in the cell membrane, although the antiproliferative effect is independent of this property. Furthermore, rBbKI, a recombinant form of BbKI, affected L929 cell viability and cell adhesion to fibronectin, and it showed an interaction with the nucleus/cell surface of fibrosarcoma cells; however, rBbKI did not interfere with the cell cycle [55]. As highlighted previously, the trypsin inhibitor isolated from Bauhinia rufa (BrTI) has an atypical sequence containing the RGD and RGE motifs involved in cell adhesion [14]. The modified inhibitor rBbKIm, comprising RGD/RGE sequences [28], also interfered with the viability of prostate cancer cell lines DU145 and PC3, leading to cell death through the release of cytochrome c from, and caspase-3 activation in, the mitochondria [29].
BPLTI, isolated from seeds of B. purpurea L., is another protease inhibitor characterized from the genus Bauhinia. This inhibitor is a 20 kDa single polypeptide chain that shows relatively high thermal (0–60 °C) and pH (3–10) stability, and demonstrates antiproliferative and proapoptotic activities in human hepatocellular carcinoma HepG2 cells, as well as in MCF-7, CNE-1, and CNE-2 cells, but not in immortal human hepatic WRL 68 cells, L1210, HNE-2 tumor cells, or normal NP69 cells. Cell death tests revealed morphological changes in the nuclear structure of BPLTI-treated HepG2 cells, which reduced the number of cells and induced apoptosis proportional to the inhibitor dose. BPLTI also increased IL-1β transcript levels in a dose-dependent manner, and IL-2 and TNF-α levels at a specified time point. IL-1β and IL-2 levels increased from 4 to 16 h, while TNF-α levels began to decline after 4 h. However, BPLTI did not induce IFN-γ [40]. The selective action of plant protease inhibitors on tumor cells has also been demonstrated in a study by Shamsi et al. [41], in which a small trypsin and chymotrypsin Kunitz-type inhibitor (CCPI, MW~12 kDa) was characterized from pigeonpea (Cajanus cajan L.) seeds. CCPI acts on the adenocarcinoma human alveolar basal epithelial cell line A549, but does not affect the nontumor HEK cells. In another study, Karray et al. [44], investigated the highly stable PD inhibitor from Conyza dioscoridis, which showed cytotoxic effects on three human tumor cell lines (MDA-MB-231, HCT-116, and LoVo) without affecting HUVECs. Bonturi et al. [48] described CrataBL as an important bifunction protein (lectin and FXa inhibitor) in the glioblastoma tumor microenvironment, affecting U87 growth and invasion.
It is important to compare the effect of protease inhibitors with others in the Kunitz-type family to demonstrate that the effect is dependent on a specific protease. For example, the effect of rBbKIm was compared with the classical plant inhibitor of the Kunitz family, SbTI. In contrast to the inert effect of SbTI, rBbKIm was more selective, causing the death of prostate cancer cells [29]. This selectivity is promising for developing antitumor therapies, as cytotoxic drugs generally do not distinguish normal tissue from metastatic tissue, making it important to develop new compounds targeting proteins that are differentially expressed in cancer cells compared to normal adult tissue cells. The effect of protease inhibitors on the viability of nontumorigenic cells should therefore be analyzed to indicate selectivity for tumor cells.

9. Concluding Remarks

We have gathered the main characteristics and effects of several protease inhibitors in this review. These inhibitors exert a wide range of effects, including antibacterial and insecticidal activity, inhibition of enzymes involved in the inflammation process, antithrombotic action, and the inhibition of tumor cell proliferation and invasion in vitro. Nevertheless, further studies are needed to enable a deeper understanding of their inhibitory mechanisms and facilitate their use as new therapeutic drugs.

Author Contributions

Writing—original draft, A.B.S.T., C.R.B., C.M.B.F., I.F.C.B., J.R.O., P.F.V. and V.M.R.; Writing—review and editing, C.R.B., I.F.C.B. and M.L.V.O.; Supervision, C.R.B. and M.L.V.O.; Project administration, Funding acquisition, and Conceptualization M.L.V.O. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP) [2017/06630-7 and 2019/22243-9]; Coordenação de Aperfeiçoamento de Pessoal de Nível Superior—Brasil (CAPES)—Finance Code 001, and Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq) [301721/2016-5].

Acknowledgments

We would like to thank Raquel Cisneiros ([email protected]) for the creation of the images.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Richardson, M. Seed storage proteins: The Enzyme Inhibitors. In Methods in Plant Biochemistry Seed Proteins; Springer: Dordrecht, The Netherlands, 1991; Volume 5, pp. 259–305. [Google Scholar]
  2. Ussuf, K.K.; Laxmi, N.H.; Mitra, R. Proteinase inhibitors: Plant-derived genes of insecticidal protein for developing insect-resistant transgenic plants. Curr. Sci. 2001, 80, 847–853. Available online: http://www.jstor.org/stable/2410573570 (accessed on 22 March 2020).
  3. Srikanth, S.; Chen, Z. Plant Protease Inhibitors in Therapeutics-Focus on Cancer Therapy. Front. Pharmacol. 2016, 7, 470. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Vergnolle, N. Protease inhibition as new therapeutic strategy for GI diseases. Gut 2016, 65, 1215–1224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Rawlings, N.D.; Barrett, A.J.; Thomas, P.D.; Huang, X.; Bateman, A.; Finn, R.D. The MEROPS database of proteolytic enzymes, their substrates and inhibitors in 2017 and a comparison with peptidases in the PANTHER database. Nucleic Acids Res. 2018, 46, D624–D632. [Google Scholar] [CrossRef] [PubMed]
  6. Bendre, A.D.; Ramasamy, S.; Suresh, C.G. Analysis of Kunitz inhibitors from plants for comprehensive structural and functional insights. Int. J. Biol. Macromol. 2018, 113, 933–943. [Google Scholar] [CrossRef]
  7. Benbow, H.R.; Jermiin, L.S.; Dohan, F.M. Serpins: Genome-Wide Characterisation and Expression Analysis of the Serine Protease Inhibitor Family in Triticum aestivum. G3 Genes|Genomes|Genet. 2019, 9, 2709–2722. [Google Scholar] [CrossRef] [Green Version]
  8. Xie, Y.; Ravet, K.; Pearce, S. Extensive structural variation in the Bowman-Birk inhibitor family in common wheat (Triticum aestivum L.). BMC Genom. 2021, 22, 218. [Google Scholar] [CrossRef]
  9. Pariani, S.; Contreras, M.; Rossi, F.R.; Sander, V.; Corigliano, M.G.; Simón, F.; Busi, M.V.; Gomez-Casati, D.F.; Pieckenstain, F.L.; Duschak, V.G.; et al. Characterization of a novel Kazal-type serine proteinase inhibitor of Arabidopsis thaliana. Biochimie 2016, 123, 85–94. [Google Scholar] [CrossRef]
  10. Oliva, M.L.V.; Silva, M.C.; Sallai, R.C.; Brito, M.V.; Sampaio, M.U. A novel subclassification for Kunitz proteinase inhibitors from leguminous seeds. Biochimie 2010, 92, 1667–1673. [Google Scholar] [CrossRef]
  11. Zhou, D.; Hansen, D.; Shabalin, I.G.; Gustchina, A.; Vieira, D.F.; De Brito, M.V.; Araujo, A.P.U.; Oliva, M.L.; Wlodawer, A. Structure of BbKI, a disulfide-free plasma kallikrein inhibitor. Acta Crystallogr. Sect. F Struct. Biol. Commun. 2015, 71 Pt 8, 1055–1062. [Google Scholar] [CrossRef] [Green Version]
  12. Kunitz, M. Crystallization of a Trypsin Inhibitor from Soybean. Science 1945, 101, 668–669. [Google Scholar] [CrossRef] [PubMed]
  13. Hansen, D.; Macedo-Ribeiro, S.; Veríssimo, P.; Im, S.Y.; Sampaio, M.U.; Oliva, M.L.V. Crystal structure of a novel cysteinless plant Kunitz-type protease inhibitor. Biochem. Biophys. Res. Commun. 2007, 360, 735–740. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Nakahata, A.M.; Bueno, N.R.; Rocha, H.A.; Franco, C.R.; Chammas, R.; Nakaie, C.R.; Jasiulionis, M.; Nader, H.B.; Santana, L.A.; Sampaio, M.U.; et al. Structural and inhibitory properties of a plant proteinase inhibitor containing the RGD motif. Int. J. Biol. Macromol. 2006, 40, 22–29. [Google Scholar] [CrossRef] [PubMed]
  15. Li, M.; Srp, J.; Gustchina, A.; Dauter, Z.; Mares, M.; Wlodawer, A. Crystal structures of the complex of a kallikrein inhibitor from Bauhinia bauhinioides with trypsin and modeling of kallikrein complexes. Acta Crystallogr. Sect. D Struct. Biol. 2019, 75, 56–69. [Google Scholar] [CrossRef]
  16. Sumikawa, J.T.; Nakahata, A.M.; Fritz, H.; Mentele, R.; Sampaio, M.U.; Oliva, M.L. A Kunitz-Type Glycosylated Elastase Inhibitor with One Disulfide Bridge. Planta Med. 2006, 72, 393–397. [Google Scholar] [CrossRef]
  17. Cavalcanti, M.D.S.M.; Oliva, M.L.; Fritzc, H.; Jochumc, M.; Mentelec, R.; Sampaiob, M.; Coelho, L.C.; Batista, I.F.; Sampaio, C.A. Characterization of a Kunitz Trypsin Inhibitor with One Disulfide Bridge Purified from Swartzia pickellii. Biochem. Biophys. Res. Commun. 2002, 291, 635–639. [Google Scholar] [CrossRef]
  18. Lingaraju, M.H.; Gowda, L.R. A Kunitz trypsin inhibitor of Entada scandens seeds: Another member with single disulfide bridge. Biochim. Biophys. Acta (BBA) Proteins Proteom. 2008, 1784, 850–855. [Google Scholar] [CrossRef]
  19. Macedo, M.L.; Garcia, V.A.; Freire, M.D.; Richardson, M. Characterization of a Kunitz trypsin inhibitor with a single disulfide bridge from seeds of Inga laurina (SW.) Willd. Phytochemistry 2007, 68, 1104–1111. [Google Scholar] [CrossRef]
  20. Krauchenco, S.; Nagem, R.A.; da Silva, J.A.; Marangoni, S.; Polikarpov, I. Three-dimensional structure of an unusual Kunitz (STI) type trypsin inhibitor from Copaifera langsdorffii. Biochimie 2004, 86, 167–172. [Google Scholar] [CrossRef]
  21. Song, H.K.; Suh, S.W. Kunitz-type soybean trypsin inhibitor revisited: Refined structure of its complex with porcine trypsin reveals an insight into the interaction between a homologous inhibitor from Erythrina caffra and tissue-type plasminogen activator. J. Mol. Biol. 1998, 275, 347–363. [Google Scholar] [CrossRef] [Green Version]
  22. Di Ciero, L.; Oliva, M.L.; Torquato, R.; Köhler, P.; Weder, J.K.; Novello, C.J.; Sampaio, C.A.; Oliveira, B.; Marangoni, S. The complete amino acid sequence of a trypsin inhibitor from Bauhinia variegata var. candida seeds. J. Protein Chem. 1998, 17, 827–834. [Google Scholar] [CrossRef] [PubMed]
  23. Batista, I.F.; Nonato, M.C.; Bonfadini, M.R.; Beltramini, L.M.; Oliva, M.L.; Sampaio, M.U.; Sampaio, C.A.; Garratt, R.C. Preliminary crystallographic studies of EcTI, a serine proteinase inhibitor from Enterolobium contortisiliquum seeds. Acta Crystallogr. Sect. D Biol. Crystallogr. 2001, 57, 602–604. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Zhou, D.; Lobo, Y.; Batista, I.F.; Marques-Porto, R.; Gustchina, A.; Oliva, M.L.; Wlodawer, A. Crystal Structures of a Plant Trypsin Inhibitor from Enterolobium contortisiliquum (EcTI) and of Its Complex with Bovine Trypsin. PLoS ONE 2013, 8, e62252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Patthy, A.; Molnár, T.; Porrogi, P.; Naude, R.; Gráf, L. Isolation and characterization of a protease inhibitor from Acacia karroo with a common combining loop and overlapping binding sites for chymotrypsin and trypsin. Arch. Biochem. Biophys. 2015, 565, 9–16. [Google Scholar] [CrossRef]
  26. Terada, S.; Katayama, H.; Noda, K.; Fujimura, S.; Kimoto, E. Amino Acid Sequences of Kunitz Family Subtilisin Inhibitors from Seeds of Canavalia lineata. J. Biochem. 1994, 115, 397–404. [Google Scholar] [CrossRef] [PubMed]
  27. Araújo, A.P.; Hansen, D.; Vieira, D.F.; De Oliveira, C.; Santana, L.A.; Beltramini, L.M.; Sampaio, C.A.; Sampaio, M.U.; Oliva, M.L. Kunitz-type Bauhinia bauhinioides inhibitors devoid of disulfide bridges: Isolation of the cDNAs, heterologous expression and structural studies. Biol. Chem. 2005, 386, 561–568. [Google Scholar] [CrossRef]
  28. Sumikawa, J.T.; Brito, M.V.; Araújo, A.P.U.; Macedo, M.L.; Oliva, M.L.V.; Miranda, A. Action of Bauhinia-derivated compounds on Callosobruchus maculatus development. Adv. Exp. Med. Biol. 2009, 611, 563–564. [Google Scholar] [CrossRef]
  29. Ferreira, J.G.; Diniz, P.M.M.; de Paula, C.A.A.; Lobo, Y.A.; Gamero, E.J.P.; Paschoalin, T.; Nogueira-Pedro, A.; Maza, P.K.; Toledo, M.S.; Suzuki, E.; et al. The Impaired Viability of Prostate Cancer Cell Lines by the Recombinant Plant Kallikrein Inhibitor. J. Biol. Chem. 2013, 288, 13641–13654. [Google Scholar] [CrossRef] [Green Version]
  30. Yu, J.-H.; Li, Y.-Y.; Xiang, M.; Zhu, J.-Q.; Huang, X.-H.; Wang, W.-J.; Tan, R.; Zhou, J.-Y.; Liao, H. Molecular cloning and characterization of α-amylase/subtilisin inhibitor from rhizome of Ligusticum chuanxiong. Biotechnol. Lett. 2016, 39, 141–148. [Google Scholar] [CrossRef]
  31. Liu, Z.; Zhu, Q.; Li, J.; Zhang, G.; Jiamahate, A.; Zhou, J.; Liao, H. Isolation, structure modeling and function characterization of a trypsin inhibitor from Cassia obtusifolia. Biotechnol. Lett. 2015, 37, 863–869. [Google Scholar] [CrossRef]
  32. Cruz-Silva, I.; Gozzo, A.J.; Nunes, V.A.; Tanaka, A.S.; Araujo, M.D.S. Bioengineering of an elastase inhibitor from Caesalpinia echinata (Brazil wood) seeds. Phytochemistry 2021, 182, 112595. [Google Scholar] [CrossRef] [PubMed]
  33. Liu, J.; Yonekura, M.; Kouzuma, Y. Purification, cDNA cloning and characterization of Kunitz-type protease inhibitors from Apios americana tubers. Biosci. Biotechnol. Biochem. 2020, 84, 563–574. [Google Scholar] [CrossRef] [PubMed]
  34. Bunyatang, O.; Chirapongsatonkul, N.; Bangrak, P.; Henry, R.; Churngchow, N. Molecular cloning and characterization of a novel bi-functional α-amylase/subtilisin inhibitor from Hevea brasiliensis. Plant Physiol. Biochem. 2016, 101, 76–87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. De Lacerda, J.T.J.G.; Lacerda, R.R.E.; Assunção, N.A.; Tashima, A.K.; Juliano, M.A.; dos Santos, G.A. New insights into lectin from Abelmoschus esculentus seeds as a Kunitz-type inhibitor and its toxic effects on Ceratitis capitata and root-knot nematodes Meloidogyne spp. Process Biochem. 2017, 63, 96–104. [Google Scholar] [CrossRef]
  36. De Oliveira, C.; Flores, T.M.D.O.; Cardoso, M.H.; Oshiro, K.G.N.; Russi, R.; de França, A.; Dos Santos, E.A.; Franco, O.L.; De Oliveira, A.S.; Migliolo, L. Dual Insecticidal Effects of Adenanthera pavonina Kunitz-Type Inhibitor on Plodia interpunctella is Mediated by Digestive Enzymes Inhibition and Chitin-Binding Properties. Molecules 2019, 24, 4344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Meriño-Cabrera, Y.; Mendes, T.A.D.O.; Castro, J.; Barbosa, S.L.; Macedo, M.; Oliveira, M.G.A. Noncompetitive tight-binding inhibition of Anticarsia gemmatalis trypsins by Adenanthera pavonina protease inhibitor affects larvae survival. Arch. Insect Biochem. Physiol. 2020, 104, e21687. [Google Scholar] [CrossRef] [PubMed]
  38. Vajravijayan, S.; Pletnev, S.; Pletnev, V.Z.; Nandhagopal, N.; Gunasekaran, K. Crystal structure of a novel Kunitz type inhibitor, alocasin with anti-Aedes aegypti activity targeting midgut proteases. Pest Manag. Sci. 2018, 74, 2761–2772. [Google Scholar] [CrossRef] [PubMed]
  39. Valois, M.V.; de Oliveira, C.; Lapa, A.J.; Souccar, C.; Oliva, M.L.V. Bauhinia Protease Inhibitors Attenuate Gastric Ulcer by Blocking Neutrophil Enzymes. Planta Med. 2020, 87, 169–176. [Google Scholar] [CrossRef] [PubMed]
  40. Fang, E.F.; Bah, C.S.F.; Wong, J.H.; Pan, W.L.; Chan, Y.S.; Ye, X.J.; Ng, T.B. A potential human hepatocellular carcinoma inhibitor from Bauhinia purpurea L. seeds: From purification to mechanism exploration. Arch. Toxicol. 2011, 86, 293–304. [Google Scholar] [CrossRef] [PubMed]
  41. Fatima, S.; Shamsi, T.N.; Parveen, R.; Ahamad, S. Structural and biophysical characterization of Cajanus cajan protease inhibitor. J. Nat. Sci. Biol. Med. 2017, 8, 186–192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Brandão-Costa, R.; Araújo, V.F.; Porto, A. CgTI, a novel thermostable Kunitz trypsin-inhibitor purified from Cassia grandis seeds: Purification, characterization and termiticidal activity. Int. J. Biol. Macromol. 2018, 118 Pt B, 2296–2306. [Google Scholar] [CrossRef] [PubMed]
  43. Dias, L.P.; Oliveira, J.T.; Rocha-Bezerra, L.C.; Sousa, D.O.; da Costa, H.P.S.; Araujo, N.M.; Carvalho, A.F.; Tabosa, P.M.; Monteiro-Moreira, A.C.; Lobo, M.D.; et al. A trypsin inhibitor purified from Cassia leiandra seeds has insecticidal activity against Aedes aegypti. Process Biochem. 2017, 57, 228–238. [Google Scholar] [CrossRef]
  44. Karray, A.; Alonazi, M.; Smaoui, S.; Michaud, P.; Soliman, D.; Ben Bacha, A. Purification and Biochemical Characterization of a New Protease Inhibitor from Conyza dioscoridis with Antimicrobial, Antifungal and Cytotoxic Effects. Molecules 2020, 25, 5452. [Google Scholar] [CrossRef] [PubMed]
  45. Oliva, L.V.; Almeida-Reis, R.; Theodoro-Junior, O.; Oliveira, B.M.; Leick, E.A.; Prado, C.M.; Brito, M.V.; Correia, M.T.D.S.; Paiva, P.M.; Martins, M.A.; et al. A plant proteinase inhibitor from Crataeva tapia (CrataBL) attenuates elastase-induced pulmonary inflammatory, remodeling, and mechanical alterations in mice. Process Biochem. 2015, 50, 1958–1965. [Google Scholar] [CrossRef] [Green Version]
  46. Bortolozzo, A.S.S.; Rodrigues, A.P.D.; Arantes-Costa, F.M.; Saraiva-Romanholo, B.M.; De Souza, F.C.R.; Brüggemann, T.R.; De Brito, M.V.; Ferreira, R.D.S.; Correia, M.T.D.S.; Paiva, P.M.G.; et al. The Plant Proteinase Inhibitor CrataBL Plays a Role in Controlling Asthma Response in Mice. BioMed Res. Int. 2018, 2018, 9274817. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Salu, B.R.; Ferreira, R.S.; Brito, M.V.; Ottaiano, T.F.; Cruz, J.W.M.; Silva, M.C.C.; Correia, M.T.S.; Paiva, P.M.; Maffei, F.H.A.; Oliva, M.L.V. CrataBL, a lectin and Factor Xa inhibitor, plays a role in blood coagulation and impairs thrombus formation. Biol. Chem. 2014, 395, 1027–1035. [Google Scholar] [CrossRef]
  48. Bonturi, C.R.; Silva, M.C.C.; Motaln, H.; Salu, B.R.; Ferreira, R.D.S.; Batista, F.P.; Correia, M.T.D.S.; Paiva, P.M.G.; Turnšek, T.L.; Oliva, M.L.V. A Bifunctional Molecule with Lectin and Protease Inhibitor Activities Isolated from Crataeva tapia Bark Significantly Affects Cocultures of Mesenchymal Stem Cells and Glioblastoma Cells. Molecules 2019, 24, 2109. [Google Scholar] [CrossRef] [Green Version]
  49. Salu, B.R.; Pando, S.C.; De Brito, M.V.; Medina, A.F.; Odei-Addo, F.; Frost, C.; Naude, R.; Sampaio, M.U.; Emsley, J.; Maffei, F.H.A.; et al. Improving the understanding of plasma kallikrein contribution to arterial thrombus formation using two plant protease inhibitors. Platelets 2018, 30, 305–313. [Google Scholar] [CrossRef]
  50. Ferreira, R.; Napoleão, T.H.; Silva-Lucca, R.A.; Silva, M.; Paiva, P.M.G.; Oliva, M.L.V. The effects of Enterolobium contortisiliquum serine protease inhibitor on the survival of the termite Nasutitermes corniger, and its use as affinity adsorbent to purify termite proteases. Pest Manag. Sci. 2019, 75, 632–638. [Google Scholar] [CrossRef] [PubMed]
  51. Nakahata, A.; Mayer, B.; Neth, P.; Hansen, D.; Sampaio, M.; Oliva, M. Blocking the Proliferation of Human Tumor Cell Lines by Peptidase Inhibitors from Bauhinia Seeds. Planta Med. 2013, 79, 227–235. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. De Paula, C.A.A.; Coulson-Thomas, V.J.; Ferreira, J.G.; Maza, P.K.; Suzuki, E.; Nakahata, A.M.; Nader, H.B.; Sampaio, M.U.; Oliva, M.L.V. Enterolobium contortisiliquum Trypsin Inhibitor (EcTI), a Plant Proteinase Inhibitor, Decreases in Vitro Cell Adhesion and Invasion by Inhibition of Src Protein-Focal Adhesion Kinase (FAK) Signaling Pathways. J. Biol. Chem. 2012, 287, 170–182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Bonturi, C.R.; Motaln, H.; Silva, M.C.C.; Salu, B.R.; De Brito, M.V.; Costa, L.D.A.L.; Torquato, H.F.V.; Nunes, N.N.D.S.; Paredes-Gamero, E.J.; Turnšek, T.L.; et al. Could a plant derived protein potentiate the anticancer effects of a stem cell in brain cancer? Oncotarget 2018, 9, 21296–21312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Lobo, Y.A.; Bonazza, C.; Batista, F.P.; Castro, R.A.; Bonturi, C.R.; Salu, B.R.; Sinigaglia, R.D.C.; Toma, L.; Vicente, C.M.; Pidde, G.; et al. EcTI impairs survival and proliferation pathways in triple-negative breast cancer by modulating cell-glycosaminoglycans and inflammatory cytokines. Cancer Lett. 2020, 491, 108–120. [Google Scholar] [CrossRef] [PubMed]
  55. Im, S.Y.; Bonturi, C.R.; Nakahata, A.M.; Nakaie, C.R.; Pott, A.; Pott, V.; Oliva, M.V. Differences in the Inhibitory Specificity Distinguish the Efficacy of Plant Protease Inhibitors on Mouse Fibrosarcoma. Plants 2021, 10, 602. [Google Scholar] [CrossRef] [PubMed]
  56. De Oliveira, C.F.; Oliveira, C.T.; Taveira, G.B.; Mello, E.D.O.; Gomes, V.M.; Macedo, M.L.R. Characterization of a Kunitz trypsin inhibitor from Enterolobium timbouva with activity against Candida species. Int. J. Biol. Macromol. 2018, 119, 645–653. [Google Scholar] [CrossRef] [PubMed]
  57. Mehmood, S.; Imran, M.; Ali, A.; Munawar, A.; Khaliq, B.; Anwar, F.; Saaed, Q.; Buck, F.; Hussain, S.; Saeed, A.; et al. Model prediction of a Kunitz-type trypsin inhibitor protein from seeds of Acacia nilotica L. with strong antimicrobial and insecticidal activity. Turk. J. Biol. 2020, 44, 188–200. [Google Scholar] [CrossRef] [PubMed]
  58. Dib, H.X.; De Oliveira, D.G.L.; De Oliveira, C.F.R.; Taveira, G.B.; Mello, E.D.O.; Verbisck, N.V.; Chang, M.; Corrêa, D., Jr.; Gomes, V.M.; Macedo, M.L.R. Biochemical characterization of a Kunitz inhibitor from Inga edulis seeds with antifungal activity against Candida spp. Arch. Microbiol. 2018, 201, 223–233. [Google Scholar] [CrossRef] [Green Version]
  59. Machado, S.W.; De Oliveira, C.; Zério, N.G.; Parra, J.; Macedo, M. Inga laurina trypsin inhibitor (ILTI) obstructs Spodoptera frugiperda trypsins expressed during adaptive mechanisms against plant protease inhibitors. Arch. Insect Biochem. Physiol. 2017, 95, e21393. [Google Scholar] [CrossRef] [PubMed]
  60. Bezerra, C.S.; de Oliveira, C.F.R.; Machado, O.L.T.; de Mello, G.S.V.; Pitta, M.G.R.; Rêgo, M.J.B.M.; Napoleão, T.H.; Paiva, P.M.G.; Ribeiro, S.D.F.F.; Gomes, V.M.; et al. Exploiting the biological roles of the trypsin inhibitor from Inga vera seeds: A multifunctional Kunitz inhibitor. Process Biochem. 2016, 51, 792–803. [Google Scholar] [CrossRef]
  61. Yang, C.; Zhang, J.-J.; Zhang, X.-P.; Xiao, R.; Li, P.-G. Sporamin suppresses growth of xenografted colorectal carcinoma in athymic BALB/c mice by inhibiting liver β-catenin and vascular endothelial growth factor expression. World J. Gastroenterol. 2019, 25, 3196–3206. [Google Scholar] [CrossRef]
  62. Roy, U.K.; Lavignac, N.; Rahman, A.M.; Nielsen, B.V. Purification of lectin and Kunitz trypsin inhibitor from soya seeds. J. Chromatogr. Sci. 2018, 56, 436–442. [Google Scholar] [CrossRef] [PubMed]
  63. Cai, X.; Xie, X.; Fu, N.; Wang, S. Physico-Chemical and Antifungal Properties of a Trypsin Inhibitor from the Roots of Pseudostellaria heterophylla. Molecules 2018, 23, 2388. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Ferreira, R.S.; Brito, M.V.; Napoleão, T.H.; Silva, M.; Paiva, P.; Oliva, M. Effects of two protease inhibitors from Bauhinia bauhinoides with different specificity towards gut enzymes of Nasutitermes corniger and its survival. Chemosphere 2019, 222, 364–370. [Google Scholar] [CrossRef] [PubMed]
  65. Brito, M.V.; de Oliveira, C.; Salu, B.R.; Andrade, S.A.; Malloy, P.M.; Sato, A.C.; Vicente, C.P.; Sampaio, M.U.; Maffei, F.H.; Oliva, M.L.V. The Kallikrein Inhibitor from Bauhinia bauhinioides (BbKI) shows antithrombotic properties in venous and arterial thrombosis models. Thromb. Res. 2014, 133, 945–951. [Google Scholar] [CrossRef] [PubMed]
  66. Mohanraj, S.S.; Gujjarlapudi, M.; Lokya, V.; Mallikarjuna, N.; Dutta-Gupta, A.; Padmasree, K. Purification and characterization of Bowman-Birk and Kunitz isoinhibitors from the seeds of Rhynchosia sublobata (Schumach.) Meikle, a wild relative of pigeonpea. Phytochemistry 2019, 159, 159–171. [Google Scholar] [CrossRef] [PubMed]
  67. Zhu, J.; He, Y.; Yan, X.; Liu, L.; Guo, R.; Xia, X.; Cheng, D.; Mi, X.; Samarina, L.; Liu, S.; et al. Duplication and transcriptional divergence of three Kunitz protease inhibitor genes that modulate insect and pathogen defenses in tea plant (Camellia sinensis). Hortic. Res. 2019, 6, 126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Mendonça, E.G.; Barros, R.A.; Cordeiro, G.; Silva, C.R.; Campos, W.G.; Oliveira, J.A.; Oliveira, M.G.A. Larval development and proteolytic activity of Anticarsia gemmatalis Hübner (Lepidoptera: Noctuidae) exposed to different soybean protease inhibitors. Arch. Insect Biochem. Physiol. 2020, 103, e21637. [Google Scholar] [CrossRef]
  69. Sasaki, D.Y.; Jacobowski, A.C.; de Souza, A.P.; Cardoso, M.H.; Franco, O.L.; Macedo, M.L.R. Effects of proteinase inhibitor from Adenanthera pavonina seeds on short- and long term larval development of Aedes aegypti. Biochimie 2015, 112, 172–186. [Google Scholar] [CrossRef]
  70. Swathi, M.; Mishra, P.K.; Lokya, V.; Swaroop, V.; Mallikarjuna, N.; Dutta-Gupta, A.; Padmasree, K. Purification and Partial Characterization of Trypsin-Specific Proteinase Inhibitors from Pigeonpea Wild Relative Cajanus platycarpus L. (Fabaceae) Active against Gut Proteases of Lepidopteran Pest Helicoverpa armigera. Front. Physiol. 2016, 7, 388. [Google Scholar] [CrossRef] [Green Version]
  71. Orona-Tamayo, D.; Wielsch, N.; Blanco-Labra, A.; Svatos, A.; Farías-Rodríguez, R.; Heil, M. Exclusive rewards in mutualisms: Ant proteases and plant protease inhibitors create a lock-key system to protect Acacia food bodies from exploitation. Mol. Ecol. 2013, 22, 4087–4100. [Google Scholar] [CrossRef]
  72. Sanon, A.; Zakaria, I.; Clémentine, L.D.B.; Niango, B.M.; Honora, N.R.C. Potential of Botanicals to Control Callosobruchus maculatus (Col.: Chrysomelidae, Bruchinae), a Major Pest of Stored Cowpeas in Burkina Faso: A Review. Int. J. Insect. Sci. 2018, 10, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Mallqui, K.S.V.; Oliveira, E.E.; Guedes, R.N.C. Competition between the bean weevils Acanthoscelides obtectus and Zabrotes subfasciatus in common beans. J. Stored Prod. Res. 2013, 55, 32–35. [Google Scholar] [CrossRef]
  74. Rufino, F.P.; Pedroso, V.M.; Araujo, J.N.; de França, A.F.; Rabelo, L.M.; Migliolo, L.; Kiyota, S.; Santos, E.A.; Franco, O.L.; Oliveira, A.S. Inhibitory effects of a Kunitz-type inhibitor from Pithecellobium dumosum (Benth) seeds against insect-pests’ digestive proteinases. Plant Physiol. Biochem. 2013, 63, 70–76. [Google Scholar] [CrossRef] [PubMed]
  75. Rumpold, B.; Schlüter, O. Insect-based protein sources and their potential for human consumption: Nutritional composition and processing. Anim. Front. 2015, 5, 20–24. [Google Scholar] [CrossRef]
  76. Sallai, R.C.; Salu, B.R.; Silva-Lucca, R.A.; Alves, F.L.; Napoleão, T.H.; Paiva, P.M.G.; Ferreira, R.D.S.; Sampaio, M.U.; Oliva, M.L.V. Biotechnological Potential of Araucaria angustifolia Pine Nuts Extract and the Cysteine Protease Inhibitor AaCI-2S. Plants 2020, 9, 1676. [Google Scholar] [CrossRef]
  77. Müller, V.; Bonacci, G.; Batthyány, C.; Amé, M.V.; Carrari, F.; Gieco, J.; Asis, R. Peanut Seed Cultivars with Contrasting Resistance to Aspergillus parasiticus Colonization Display Differential Temporal Response of Protease Inhibitors. Phytopathology 2017, 107, 474–482. [Google Scholar] [CrossRef]
  78. Tripathi, V.R.; Kumar, S.; Garg, S.K. A study on trypsin, Aspergillus flavus and Bacillus sp. protease inhibitory activity in Cassia tora (L.) syn Senna tora (L.) Roxb. seed extract. BMC Complement. Altern. Med. 2011, 11, 56. [Google Scholar] [CrossRef] [Green Version]
  79. Barros, K.M.A.; Sardi, J.D.C.O.; Maria-Neto, S.; Macedo, A.J.; Ramalho, S.R.; de Oliveira, D.G.L.; Macedo, M.L.R. A new Kunitz trypsin inhibitor from Erythrina poeppigiana exhibits antimicrobial and antibiofilm properties against bacteria. Biomed. Pharmacother. 2021, 144, 112198. [Google Scholar] [CrossRef]
  80. Araújo, N.M.; Dias, L.P.; Costa, H.P.; Sousa, D.O.; Vasconcelos, I.M.; de Morais, G.A.; Oliveira, J.T. ClTI, a Kunitz trypsin inhibitor purified from Cassia leiandra Benth. seeds, exerts a candidicidal effect on Candida albicans by inducing oxidative stress and necrosis. Biochim. Biophys. Acta (BBA)-Biomembr. 2019, 1861, 183032. [Google Scholar] [CrossRef]
  81. Georgiadou, A.; Lee, H.J.; Walther, M.; van Beek, A.; Fitriani, F.; Wouters, D.; Kuijpers, T.W.; Nwakanma, D.; D’Alessandro, U.; Riley, E.M.; et al. Modelling pathogen load dynamics to elucidate mechanistic determinants of host–Plasmodium falciparum interactions. Nat. Microbiol. 2019, 4, 1592–1602. [Google Scholar] [CrossRef]
  82. Morais, S.B.; Figueiredo, B.C.; Assis, N.R.G.; Alvarenga, D.M.; De Magalhães, M.T.Q.; Ferreira, R.S.; Vieira, A.T.; Menezes, G.B.; Oliveira, S.C. Schistosoma mansoni SmKI-1 serine protease inhibitor binds to elastase and impairs neutrophil function and inflammation. PLoS Pathog. 2018, 14, e1006870. [Google Scholar] [CrossRef] [Green Version]
  83. Ganesan, S.; Unger, B.L.; Comstock, A.T.; Angel, K.A.; Mancuso, P.; Martinez, F.J.; Sajjan, U.S. Aberrantly activated EGFR contributes to enhanced IL-8 expression in COPD airways epithelial cells via regulation of nuclear FoxO3A. Thorax 2012, 68, 131–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Martins-Olivera, B.T.; Almeida-Reis, R.; Theodoro-Júnior, O.A.; Oliva, L.V.; Nunes, N.N.D.S.; Olivo, C.R.; de Brito, M.V.; Prado, C.M.; Leick, E.A.; Martins, M.D.A.; et al. The Plant-Derived Bauhinia bauhinioides Kallikrein Proteinase Inhibitor (rBbKI) Attenuates Elastase-Induced Emphysema in Mice. Mediat. Inflamm. 2016, 2016, 5346574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Almeida-Reis, R.; Theodoro-Junior, O.A.; Oliveira, B.T.M.; Oliva, L.V.; Toledo-Arruda, A.; Bonturi, C.R.; Brito, M.V.; Lopes, F.D.T.Q.S.; Prado, C.M.; Florencio, A.C.; et al. Plant Proteinase Inhibitor BbCI Modulates Lung Inflammatory Responses and Mechanic and Remodeling Alterations Induced by Elastase in Mice. BioMed Res. Int. 2017, 2017, 8287125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Rodrigues, A.; Bortolozzo, A.; Arantes-Costa, F.M.; Saraiva-Romanholo, B.M.; de Souza, F.; Brüggemann, T.R.; Santana, F.; de Brito, M.V.; Bonturi, C.R.; Nunes, N.; et al. A plant proteinase inhibitor from Enterolobium contortisiliquum attenuates airway hyperresponsiveness, inflammation and remodeling in a mouse model of asthma. Histol. Histopathol. 2019, 34, 537–552. [Google Scholar] [CrossRef]
  87. Rakashanda, S.; Qazi, A.K.; Majeed, R.; Rafiq, S.; Dar, I.M.; Masood, A.; Hamid, A.; Amin, S. Antiproliferative activity of Lavatera cashmeriana- protease inhibitors towards human cancer cells. Asian Pac. J. Cancer Prev. 2013, 14, 3975–3978. [Google Scholar] [CrossRef] [Green Version]
  88. Schmaier, A.H. Antithrombotic potential of the contact activation pathway. Curr. Opin. Hematol. 2016, 23, 445–452. [Google Scholar] [CrossRef] [Green Version]
  89. Kolte, D.; Shariat-Madar, Z. Plasma Kallikrein Inhibitors in Cardiovascular Disease: An Innovative Therapeutic Approach. Cardiol. Rev. 2016, 24, 99–109. [Google Scholar] [CrossRef]
  90. Simão, F.; Feener, E.P. The Effects of the Contact Activation System on Hemorrhage. Front. Med. 2017, 4, 121. [Google Scholar] [CrossRef] [Green Version]
  91. Ottaiano, T.F.; Andrade, S.S.; de Oliveira, C.; Silva, M.C.; Buri, M.V.; Juliano, M.A.; Girão, M.J.; Sampaio, M.U.; Schmaier, A.H.; Wlodawer, A.; et al. Plasma kallikrein enhances platelet aggregation response by subthreshold doses of ADP. Biochimie 2017, 135, 72–81. [Google Scholar] [CrossRef] [Green Version]
  92. Geddings, J.E.; Mackman, N. Comment on “Tissue factor expressed by microparticles is associated with mortality but not with thrombosis in cancer patients”. Thromb. Haemost. 2014, 111, 180–181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Nakahata, A.M.; Mayer, B.; Ries, C.; de Paula, C.A.A.; Karow, M.; Neth, P.; Sampaio, M.U.; Jochum, M.; Oliva, M.L.V. The effects of a plant proteinase inhibitor from Enterolobium contortisiliquum on human tumor cell lines. Biol. Chem. 2011, 392, 327–336. [Google Scholar] [CrossRef] [Green Version]
  94. Pathak, M.; Manna, R.; Li, C.; Kaira, B.G.; Hamad, B.K.; Belviso, B.D.; Bonturi, C.R.; Dreveny, I.; Fischer, P.M.; Dekker, L.V.; et al. Crystal structures of the recombinant β-factor XIIa protease with bound Thr-Arg and Pro-Arg substrate mimetics. Acta Crystallogr. Sect. D Struct. Biol. 2019, 75 Pt 6, 578–591. [Google Scholar] [CrossRef]
  95. Oliva, M.L.V.; Ferreira, R.D.S.; Ferreira, J.G.; De Paula, C.A.A.; Salas, C.E.; Sampaio, M.U. Structural and functional properties of kunitz proteinase inhibitors from leguminosae: A mini review. Curr. Protein Pept. Sci. 2011, 12, 348–357. [Google Scholar] [CrossRef] [PubMed]
  96. Decrem, Y.; Rath, G.; Blasioli, V.; Cauchie, P.; Robert, S.; Beaufays, J.; Frère, J.-M.; Feron, O.; Dogné, J.-M.; Dessy, C.; et al. Ir-CPI, a coagulation contact phase inhibitor from the tick Ixodes ricinus, inhibits thrombus formation without impairing hemostasis. J. Exp. Med. 2009, 206, 2381–2395. [Google Scholar] [CrossRef] [Green Version]
  97. Oliveira, C.; Valois, M.V.; Ottaiano, T.F.; Miranda, A.; Hansen, D.; Sampaio, M.U.; Oliva, M.L.V.; Maffei, F.H.D.A. The recombinant plant Bauhinia bauhinioides elastase inhibitor reduces rat thrombus without alterations in hemostatic parameters. Sci. Rep. 2021, 11, 13475. [Google Scholar] [CrossRef] [PubMed]
  98. Kearney, K.J.; Butler, J.; Posada, O.M.; Wilson, C.; Heal, S.; Ali, M.; Hardy, L.; Ahnström, J.; Gailani, D.; Foster, R.; et al. Kallikrein directly interacts with and activates Factor IX, resulting in thrombin generation and fibrin formation independent of Factor XI. Proc. Natl. Acad. Sci. USA 2021, 118, 218–239. [Google Scholar] [CrossRef]
  99. Eatemadi, A.; Aiyelabegan, H.T.; Negahdari, B.; Mazlomi, M.A.; Daraee, H.; Daraee, N.; Eatemadi, R.; Sadroddiny, E. Role of protease and protease inhibitors in cancer pathogenesis and treatment. Biomed. Pharmacother. 2017, 86, 221–231. [Google Scholar] [CrossRef]
  100. Wilkinson, R.; Woods, K.; D’Rozario, R.; Prue, R.; Vari, F.; Hardy, M.; Dong, Y.; Clements, J.; Hart, D.N.J.; Radford, K. Human kallikrein 4 signal peptide induces cytotoxic T cell responses in healthy donors and prostate cancer patients. Cancer Immunol. Immunother. 2011, 61, 169–179. [Google Scholar] [CrossRef] [Green Version]
  101. Prassas, I.; Eissa, A.; Poda, G.; Diamandis, E.P. Unleashing the therapeutic potential of human kallikrein-related serine proteases. Nat. Rev. Drug Discov. 2015, 14, 183–202. [Google Scholar] [CrossRef]
  102. Ossovskaya, V.S.; Bunnett, N.W. Protease-Activated Receptors: Contribution to Physiology and Disease. Physiol. Rev. 2004, 84, 579–621. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Obiezu, C.V.; Diamandis, E.P. Human tissue kallikrein gene family: Applications in cancer. Cancer Lett. 2005, 224, 1–22. [Google Scholar] [CrossRef] [PubMed]
  104. Trabocchi, A.; Menchi, G.; Cini, N.; Bianchini, F.; Raspanti, S.; Bottoncetti, A.; Pupi, A.; Calorini, L.; Guarna, A. Click-Chemistry-Derived Triazole Ligands of Arginine−Glycine−Aspartate (RGD) Integrins with a Broad Capacity To Inhibit Adhesion of Melanoma Cells and Both in Vitro and in Vivo Angiogenesis. J. Med. Chem. 2010, 53, 7119–7128. [Google Scholar] [CrossRef] [PubMed]
  105. Tantivejkul, K.; Loberg, R.D.; Mawocha, S.C.; Day, L.L.; John, L.S.; Pienta, B.A.; Rubin, M.; Pienta, K. PAR1-mediated NFκB activation promotes survival of prostate cancer cells through a Bcl-xL-dependent mechanism. J. Cell. Biochem. 2005, 96, 641–652. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Bilgin, M.; Burgazli, K.M.; Rafiq, A.; Mericliler, M.; Neuhof, C.; Oliva, M.L.; Parahuleva, M.; Soydan, N.; Doerr, O.; Abdallah, Y.; et al. Effect of bauhinia bauhinioides kallikrein inhibitor on endothelial proliferation and intracellular calcium concentration. Eur. Rev. Med. Pharmacol. Sci. 2014, 18, 46–51. [Google Scholar] [PubMed]
Figure 1. Structural aspects of plant Kunitz inhibitors in terms of the number of cysteine residues (0, 1, 2, 4, or 5 Cys), S–S bridges (none, intrachain and/or interchain bridges), and polypeptide chain representation (single or double chains).
Figure 1. Structural aspects of plant Kunitz inhibitors in terms of the number of cysteine residues (0, 1, 2, 4, or 5 Cys), S–S bridges (none, intrachain and/or interchain bridges), and polypeptide chain representation (single or double chains).
Ijms 23 04742 g001
Figure 2. Uses of protease inhibitors as insecticides, and antibacterial and antifungal agents. The inhibitors are also active in mammal pathologies associated with inflammation, thrombosis, and tumorigenesis.
Figure 2. Uses of protease inhibitors as insecticides, and antibacterial and antifungal agents. The inhibitors are also active in mammal pathologies associated with inflammation, thrombosis, and tumorigenesis.
Ijms 23 04742 g002
Figure 3. Activity of plant Kunitz inhibitors on some intrinsic pathway clotting enzymes, and on fibrinolytic enzymes. Coagulation cascade scheme based on Kearney et al. [98].
Figure 3. Activity of plant Kunitz inhibitors on some intrinsic pathway clotting enzymes, and on fibrinolytic enzymes. Coagulation cascade scheme based on Kearney et al. [98].
Ijms 23 04742 g003
Table 1. Inhibitory Properties of some Plant Kunitz Inhibitors.
Table 1. Inhibitory Properties of some Plant Kunitz Inhibitors.
Structural CharacteristicsSerine Proteases Inhibition
(Ki nM)
Cysteine Proteases Inhibition
(Ki nM)
InhibitorCys Residues/S-S bonds/ ChainsDigestion EnzymesCoagulation & Fibrinolysis EnzymesInflammatory EnzymesInflammatory Enzymes
CysS-SChainsTrypsin
(Bovine)
Chymotrypsin
(Bovine)
FXIIaPKa
(Human)
PKa
(Rat)
FXaPlasminHNECat GCat LCat B
EcTI4220.91.182.055.0n.d.n.i.9.455n.d.n.d.n.d.
BbKI1010.62.7n.d.4.75.2n.i.33.0n.i.n.i.n.i.n.i.
rBbKI10120.0n.i.n.d.2.0n.d.n.d.33.0n.i.n.i.n.i.n.i.
rBbKI-R64A10125.0n.i.n.d.98.0n.d.n.d.2.6n.i.n.d.n.i.n.d.
BrTI0012.9n.i.n.d.141.3n.i.n.i.n.i.n.i.n.i.n.i.
BbCI001n.i.n.i.n.i.n.i.n.i.n.i.n.i.5.3160.00.2n.i.
rBbCI001n.i.n.i.n.d.n.i.n.d.n.d.n.i.1.7210.09.0n.i.
BuXI42121.028.074.06.9n.d.14.076.0n.i.n.i.n.d.n.d.
BvTI-C4216.9n.d.n.d.4.5n.d.n.i.n.d.n.d.n.d.n.d.n.d.
BvTI4211.61.211080.02.2n.i.2.9n.i.n.d.n.d.n.d.
n.i.—no detectable inhibition; n.d.—not determined.
Table 2. Functions of Some Kunitz Protease Inhibitors.
Table 2. Functions of Some Kunitz Protease Inhibitors.
Plant Kunitz InhibitorIdentificationAction
Abelmoschus esculentus lectin and Kunitz inhibitor AEL Toxicity against the fruit fly Ceratitis capitata and the nematodes Meloidogyne incognita and M. javanica [35]
Adenanthera pavonina Kunitz-type inhibitor ApKTI Insecticidal activity against Plodia interpunctella larvae [36] and reduces the survival of neonatal Anticarsia gemmatalis larvae [37]
Alocasia macrorrhiza Kunitz-type inhibitor Alocasin Increases the mortality rate of Aedes aegypti larvae and adult females [38]
Bauhinia bauhinioides cruzipain inhibitor and B. mollis elastase inhibitor BbCI and BmEI Anti-ulcer effects: regulation of pro-inflammatory processes and stress-induced gastric mucosal injuries [39]
Bauhinia purpurea L. trypsin inhibitor BPLTI Anti-proliferative and pro-apoptotic activities in HepG2, MCF-7, CNE-1, and CNE-2 cells [40]
Cajanus cajan L. Kunitz-type inhibitor CCPI Acts on the adenocarcinoma human alveolar basal epithelial cell line A549 [41]
Cassia grandis trypsin inhibitor CGTI Termiticidal activity against Nasutitermes corniger [42]
Cassia leiandra trypsin inhibitor ClTI Inhibits larvae development and delays adult emergence in A. aegypti causing mortality [43]
Conyza dioscoridis protease inhibitor PDInhibitor Cytotoxic effects on MDA-MB-231, HCT-116, and LoVo, without affecting HUVECs [44]
Crataeva tapia bark lectin CrataBL Controls the inflammatory response with a decrease in bronchoalveolar lavage, number of neutrophils, lymphocytes, eosinophils and mean alveolar diameter [45,46];
Impairs thrombus formation [47] and U87 cell microenvironment [48]
Delonix regia trypsin inhibitor and Acacia schweinfurthii trypsin inhibitor DrTI and AsTI Prolong the time for total carotid artery occlusion in mice and prevent arterial thrombus formation, without affecting bleeding time [49]
Enterolobium contortisiliquum trypsin inhibitor EcTI Induces the death of Nasutitermes corniger worker termites [50];
Cytotoxic to colorectal cancer cells (HCT116 and HT29), breast cancer cells (SkBr-3, MDA-MB-231 and MCF-7), human leukemia cell lines (K562 and THP-1), human gastric cancer (Hs746T), glioblastoma (U87) and mouse fibrosarcoma (L929), without affecting normal human primary fibroblasts and human mesenchymal stem cells (hMSCs) [51,52,53,54,55]
Enterolobium timbouva trypsin inhibitor EtTI Affects the integrity of the plasma membrane of yeasts, inducing the generation of intracellular ROS [56,57]
Recombinant Hevea brasiliensis amylase/subtilisin inhibitor rHbASI High activity against Aspergillus oryzae and inhibited mycelial growth of Phytophthora palmivora [34]
Inga edulis trypsin inhibitor IeTI Inhibits the pathogenic fungi Candida tropicalis and Candida buinensis [58]
Inga laurina trypsin inhibitor ILTI Inhibits trypsin midgut of lepidopteran pest: S. frugiperda [59]
Inga vera trypsin inhibitor IVTI Midgut inhibition of lepidopteran pests: S. frugiperda, C. cephalonica, H. virescens, and H. zea [60]
Ipomoea potatoes Kunitz-type trypsin inhibitor Sporamin Suppress the growth of colorectal tumor nodules in mice [61]
Kunitz-type trypsin inhibitor KTI-A Anti-proliferative activity in the melanoma cell line, B16F1 [62]
Pseudostellaria heterophylla trypsin inhibitor PhTI Potent activity against the pathogenic fungi Candida gloeosporioides and Fusarium oxysporum [63]
Recombinant B. bauhinioides kallikrein inhibitor rBbKI Termiticidal activity in workers termites (Nasutitermes corniger) [64];
Antithrombotic properties in venous and arterial thrombosis models [65];
Inhibits cell viability of MKN-28, HT-29, HCT 116, Hs746T [51] and L929 [55]
Recombinant B. bauhinioides kallikrein inhibitor modified rBbKIm Defense against Callosobruchus maculatus development [28]; Reduces cell viability in prostate cancer (DU145/PC3) leads to cell death [29]
Rhynchosia sublobata Kunitz isoinhibitor RsKI Actives against lepidopteran gut proteases: Achaea Janata and Helicoverpa armigera [66]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bonturi, C.R.; Silva Teixeira, A.B.; Rocha, V.M.; Valente, P.F.; Oliveira, J.R.; Filho, C.M.B.; Fátima Correia Batista, I.; Oliva, M.L.V. Plant Kunitz Inhibitors and Their Interaction with Proteases: Current and Potential Pharmacological Targets. Int. J. Mol. Sci. 2022, 23, 4742. https://doi.org/10.3390/ijms23094742

AMA Style

Bonturi CR, Silva Teixeira AB, Rocha VM, Valente PF, Oliveira JR, Filho CMB, Fátima Correia Batista I, Oliva MLV. Plant Kunitz Inhibitors and Their Interaction with Proteases: Current and Potential Pharmacological Targets. International Journal of Molecular Sciences. 2022; 23(9):4742. https://doi.org/10.3390/ijms23094742

Chicago/Turabian Style

Bonturi, Camila Ramalho, Ana Beatriz Silva Teixeira, Vitória Morais Rocha, Penélope Ferreira Valente, Juliana Rodrigues Oliveira, Clovis Macêdo Bezerra Filho, Isabel Fátima Correia Batista, and Maria Luiza Vilela Oliva. 2022. "Plant Kunitz Inhibitors and Their Interaction with Proteases: Current and Potential Pharmacological Targets" International Journal of Molecular Sciences 23, no. 9: 4742. https://doi.org/10.3390/ijms23094742

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

Article Metrics

Back to TopTop