Cyclotides: disulfide-rich peptide toxins in

Yen-Hua Huang, Qingdan Du and David J. Craik*

Institute for Molecular Bioscience, The University of Queensland, Brisbane Queensland 4072,

Australia

*Corresponding Author:

Professor David J. Craik

Institute for Molecular Bioscience,

The University of Queensland,

Brisbane, Qld, 4072,

Tel: 61-7-3346 2019

Fax: 61-7-3346 2101 e-mail: [email protected]

1

1 Abstract

2 Cyclotides are a -derived family of peptides that comprise approximately 30 amino acid residues,

3 a cyclic backbone and a cystine knot. Due to their unique structure, cyclotides are exceptionally stable

4 to heat or proteolytic degradation and are tolerant to amino acid substitutions in their backbone loops

5 between conserved cysteine residues. Their toxicity to insect pests and their make-up of natural amino

6 acids has led to their applications in eco-friendly crop protection. Furthermore, their stability and cell

7 penetrating properties make cyclotides ideal scaffolds for bioactive epitope grafting. This article gives

8 a brief overview of cyclotide discovery, characterization, distribution, synthesis and mode of action

9 mechanisms. We focus on their toxicities to insect pests and their medical and agricultural

10 applications.

11 Keywords

12 Cyclotides; Stability; Insecticidal peptide; Cytotoxicity; Cell penetrating peptide; Applications

2

13 1 Introduction

14 Cyclotides are disulfide-rich peptides from plants that have been known now for nearly two decades

15 (Craik et al., 1999). They typically comprise 30 amino acids and contain the characteristic features of

16 a head-to-tail cyclized backbone and a knotted arrangement of three disulfide bonds. They occur in a

17 wide range of plant tissues, including , , stems and roots, and are thought to be present

18 as natural defense molecules, particularly against insects and nematodes. For this reason, they can be

19 regarded as plant toxins, i.e., molecules that are toxic to certain plant pests. However, most interest in

20 cyclotides has so far related to their exceptional stability and potential as a framework in drug design

21 rather than to their natural pesticidal functions. Here we provide an overview on the discovery,

22 structures and potential biotechnological applications of cyclotides. There have been a number of

23 recent reviews on the discovery and distribution (Gruber, 2010; Weidmann & Craik, 2016),

24 biosynthesis (Conlan et al., 2012; Craik et al., 2018; Qu et al., 2017), biological activities (Craik, 2012;

25 Daly et al., 2011; Göransson et al., 2012), mode of action (Henriques & Craik, 2012; Henriques &

26 Craik, 2017), and applications in drug design (Camarero, 2017; Craik & Du, 2017; de Veer et al.,

27 2017; Gould & Camarero, 2017; Gould et al., 2011; Poth et al., 2013; Wang & Craik, 2018) of

28 cyclotides as well as a recent analysis of publication trends in the cyclotide field (Kan & Craik, 2018).

29 The main aim of this article is to explore the context of cyclotides as toxins with interesting

30 biotechnological applications.

31 1.1 History of discovery

32 The first cyclotide discovered was kalata B1 which attracted attention due to its use in an indigenous

33 medicine in Africa (Gran, 1970, 1973a). In this usage, leaves from the Rubiaceae family plant

34 Oldenlandia affinis were boiled by Congolese women to make a tea that was ingested during labor,

35 resulting in accelerated child birth. This discovery was made by a Norwegian physician, Lorents Gran,

36 who observed the accelerated labor while serving on a Red Cross relief mission in the Congo in the

37 1960s. After taking some of the plant material back to Norway he discovered that the uterotonic

38 ingredient was a peptide, which he and colleagues partially characterized and found to comprise

39 approximately 30 amino acids (Gran, 1973a). At that stage, neither the disulfide knotted arrangement

40 nor the head-to-tail cyclized backbone was apparent, but Gran noted that the peptide was very stable 3

41 and was one of a number of peptides of similar molecular weight derived from an extract of the plant

42 leaves (Gran, 1973b).

43 The three-dimensional structure of kalata B1 was not elucidated until the mid-1990s, when mass

44 spectrometry and NMR studies were used to delineate the cyclic backbone and the knotted

45 arrangement of disulfide bonds, respectively (Saether et al., 1995). Around the same time, several

46 independent groups reported peptides of similar size with pharmaceutically relevant activities, which

47 had been discovered in the course of plant natural products screening programs. A report from a group

48 at Merck Research Laboratories described cyclopsychotride A, a macrocyclic peptide from a South

49 American plant from the Rubiaceae family, which had neurotensin antagonistic properties (Witherup

50 et al., 1994). In another report, a group at the National Cancer Institute, USA, noted the anti-HIV

51 activities of a series of macrocyclic peptides, which they named circulins because of their head-to-tail

52 cyclic backbone (Gustafson et al., 1994). The circulins were derived from the bark of a Tanzanian

53 tree, also from the Rubiaceae family. Earlier, the discovery of another macrocyclic peptide of similar

54 size from a violet plant ( family) was reported by an Austrian group, found while looking

55 for saponins with hemolytic activity (Schöpke et al., 1993).

56 Following these initial reports, two groups set up systematic discovery programs to see whether

57 similar peptides occur in related plants. Our group at The University of Queensland (Australia) found

58 a number of examples in plants of Rubiaceae and Violaceae families (Craik et al., 1999), while a

59 group in Sweden focused on plants from the Violaceae family (Claeson et al., 1998; Göransson et al.,

60 1999). These combined discoveries in the late 1990s led to the conclusion that there were indeed

61 many other members of this peptide family and the name “cyclotide” was coined to refer to plant-

62 derived head-to-tail cyclic peptides that contain a cystine knot motif (see Figure 1) (Craik et al., 1999).

63 1.2 Sequences and nomenclature of cyclotides

64 Currently more than 400 sequences of cyclotides have been reported and are documented in a database

65 called CyBase (www.cybase.org.au), which also contains data on other families of naturally occurring

66 cyclic peptides (Wang et al., 2008b). Figure 1 shows some example sequences from the three currently

67 defined subfamilies of cyclotides and a representative structure of prototypic examples from each

68 subfamily. 4

69

70 Figure 1: Selected sequences and three-dimensional structures from the three subfamilies of 71 cyclotides. A. An alignment of selected cyclotides sequences for which a solution structure has been 72 published. The sequences are aligned starting from the presumed N-terminal cleavage point in the 73 linear precursors. Cyclotides are classified into three subfamilies and listed chronologically according 74 to the publication date of their NMR derived structures. All cyclotides have six loops separated by six 75 conserved cysteines (numbered with Roman numerals I to VI) and are head-to-tail cyclized. The 76 cysteines are highlighted in yellow and the disulfide connectivities (I to IV, II to V and III to VI) are 77 shown at the top of the table with a black line. B. The solution structures (from left to right) of kalata 78 B1 (Rosengren et al., 2003), kalata B8 (Daly et al., 2006) and MCoTI-II (Felizmenio-Quimio et al., 79 2001) from the Möbius, bracelet and trypsin inhibitor subfamilies, respectively. The disulfide bonds 80 are indicated by yellow ball-and-stick structures, the beta sheets by blue arrows and the 310-helix turn 81 by red ribbon. 82 Initially, cyclotides were classified into two major subfamilies: the bracelet and Möbius

83 subfamilies, based on the absence or presence, respectively, of a cis-proline in loop 5 of the circular

84 backbone (the loops being defined as the backbone segments between successive Cys residues, as

85 shown in Figure 1). The origin of this nomenclature is quite straightforward: when all of the backbone

86 peptide linkages of a cyclic peptide exist in the usual trans arrangement, the backbone can be thought

87 of as bracelet-like; however, a three-dimensional structure containing a cis-proline results in a 180°

88 conceptual twist in the circular backbone, which can thus be topologically regarded as a Möbius strip

5

89 (Jennings et al., 2005). A third, smaller, subfamily of cyclotides was introduced upon discovery of

90 MCoTI-I and MCoTI-II from the tropical vine Momordica cochinchinensis (Hernandez et al., 2000).

91 Although these two trypsin inhibitor peptides are dissimilar from other cyclotides in sequence, they

92 share the common structural elements of three interlocked disulfide bridges and a cyclic backbone,

93 leading to their classification as cyclotides (Felizmenio-Quimio et al., 2001). They are alternatively

94 referred to as cyclic knottins (Chiche et al., 2004).

95 1.3 Cyclic cystine knot scaffold and structures of cyclotides

96 In addition to the structural elucidation of the prototypic cyclotide, kalata B1, numerous other

97 members of the family have been structurally characterized. Figure 1A shows a sequence alignment

98 of the 18 structurally characterized cyclotides. To date, NMR structures have been determined for:

99 eight cyclotides from the Möbius subfamily, including kalata B1 (Saether et al., 1995), kalata B2

100 (Jennings et al., 2005), cycloviolacin O14 (Ireland et al., 2006a), kalata B7 (Shenkarev et al., 2008),

101 varv F (Wang et al., 2009b), vhl-2 (Daly et al., 2010), Cter M (Poth et al., 2011a), and kalata B12

102 (Wang et al., 2011); ten cyclotides from the bracelet subfamily, including circulin A (Daly et al.,

103 1999a), cycloviolacin O1 (Rosengren et al., 2003), palicourein (Barry et al., 2004), vhr1 (Trabi &

104 Craik, 2004), circulin B (Koltay et al., 2005), tricyclon A (Mulvenna et al., 2005), vhl-1 (Chen et al.,

105 2005), kalata B8 (Daly et al., 2006), cycloviolacin O2 (Wang et al., 2009a), and kalata B5 (Plan et al.,

106 2010); and three cyclotides from the trypsin inhibitor subfamily MCoTI-II (Felizmenio-Quimio et al.,

107 2001), MCoTI-I (Kwon et al., 2018) and MCoTI-V (Mylne et al., 2012).

108 As can be seen in the sequence alignment in Figure 1A, loops 1 and 4 are highly conserved in

109 terms of both their number of residues and amino acid composition (Daly et al., 2009). The

110 conservation of these two loops is probably due to the fact that they form the central part of the

111 embedded cystine knot in cyclotides. Additionally, a highly conserved glutamic acid present in loop

112 1 has been shown to take part in a hydrogen-bond network that is important in stabilizing the structure

113 (Rosengren et al., 2003). In fact, this glutamic acid is almost completely conserved in both Möbius

114 and bracelet subfamilies, with the exception of kalata B12 (Plan et al., 2007), where the glutamic acid

115 is substituted for an aspartic acid in the corresponding position. Loop 1 in the trypsin inhibitor

6

116 subfamily lacks the Glu residue and contains a larger number of residues than the corresponding loop

117 in the Mobius and bracelet cyclotides.

118 Despite the sequence diversity in the other four loops (i.e. loops 2, 3, 5, and 6), cyclotides from

119 the three subfamilies all share the same structural features of the cyclic backbone and a cystine knot

120 core formed from three disulfide bonds (Figure 1B). The cystine knot motif is surrounded by a small

121 β-hairpin which, in most cyclotides, is combined with a third β-strand to form a distorted triple-

122 stranded β-sheet. This so-called cyclic cystine knot (CCK) is a special case of a common motif known

123 as the inhibitor cystine knot (ICK) scaffold (Craik et al., 2001), which is found in a wide range of

124 proteins from plants and animals. Overall, the combination of a head-to-tail cyclic backbone and a

125 cystine knot makes cyclotides extremely resistant to proteolytic breakdown, to high temperatures and

126 to chemical denaturants such as urea or guanidine (Colgrave & Craik, 2004). This stability might

127 explain why they are excellent insecticidal agents in plants, as this stability is presumably important

128 for protein-based natural products that accumulate in leaves without breakdown under harsh

129 environmental conditions. Stability is also the reason why cyclotides have attracted attention from a

130 drug design perspective.

131 1.4 Distribution of cyclotides in the plant kingdom, in individual plants, and within

132 individual tissues

133 So far cyclotides have been discovered in a range of from five major plant families, i.e. the

134 Rubiaceae, Violaceae, Cucurbitaceae, Solanaceae, and Fabaceae families (Koehbach et al., 2013a).

135 These represent many economically important plant families and hence cyclotides are potentially of

136 broad interest in plant science. Their distribution within these plant families is highly variable; so far

137 every Violaceous plant examined has been found to contain cyclotides, suggesting that cyclotides are

138 ubiquitous in this family (Burman et al., 2010a). Violaceae comprises approximately 930 species that

139 are distributed both in temperate and tropical zones around the world and include common ornamental

140 plants such as pansies. By contrast, cyclotides occur in less than 5% of the Rubiaceae (Gruber et al.,

141 2008) and in only a few members of each of the other plant families, i.e. in two species of the

142 Solanaceae (Poth et al., 2012; Zenoni et al., 2011) and the Cucurbitaceae family (Du et al., 2019;

143 Hernandez et al., 2000), and only one from the Fabaceae (Poth et al., 2011b) so far. Cyclotide- 7

144 containing plant families, the corresponding orders and the reported number of cyclotides from each

145 plant family are summarized in Figure 2 (Craik, 2013).

146

147 Figure 2: Distribution of cyclotides within angiosperms (flowering plants). Cyclotide-bearing 148 plant families and the corresponding orders are highlighted, with the number of cyclotides reported 149 in the literature listed next to the plant images. All cyclotides reported to date were found in core 150 eudicot plants, and only acyclic variants were discovered in a monocot plant from the Poaceae family 151 (labeled with one asterisk) (Nguyen et al., 2013). Small circular peptides distinct from cyclotides in 152 sequence were derived from the Asteraceae family (labeled with two asterisks). The phylogenetic 153 information of angiosperms was obtained from the Angiosperm Phylogeny Website 154 (http://www.mobot.org/MOBOT/research/APweb/). Figure adapted from a previously published article 155 (Craik, 2013). 156 To date, cyclotides have only been discovered in core eudicot plants. It is expected that as more

157 investigations take place, the number of plant families that are reported to contain cyclotides will

158 increase, as will the number of cyclotide-bearing species. However, at this stage it is clear that many

159 plants do not contain cyclotides, but those that do, contain them in large amounts. There remain many

8

160 interesting questions as to why some plants have evolved the ability to produce cyclotides but most

161 have apparently not.

162 In the cyclotide-bearing plants that have been examined so far, cyclotides occur in many,

163 indeed probably all, tissues. For example, they occur in leaves, petioles, stems, pedicels, flowers, and

164 roots (Trabi & Craik, 2004). A single plant may contain dozens to hundreds of cyclotides. For example,

165 hederacea has been reported to contain at least 66 cyclotides (Trabi & Craik, 2004) and O.

166 affinis has so far been found to contain 22 cyclotides (Plan et al., 2007). It is expected the number of

167 individual cyclotides known to occur in plants will increase dramatically as more transcriptome and/or

168 genomic studies are published (Koehbach et al., 2013a). Until recently, most discoveries have been

169 made at the peptide rather than nucleic acid level, but this is likely to change in future. Despite this

170 shift to nucleic acids based discovery, new MS-based sequencing approaches for cyclotides show

171 promise for ongoing discoveries at the peptide-level (Parsley et al., 2018).

172 Within a single plant, the distribution and type of cyclotides varies from tissue to tissue. For

173 example, one study showed that the cyclotides present in roots in V. hederacea were typically more

174 hydrophobic than those found in the leaves and flowers (Trabi & Craik, 2004). In only a few cases is

175 the same cyclotide found in multiple plants. For example, varv A is found in four different plants,

176 including O. affinis, Viola odorata, Viola tricolor, and Viola arvensis (Gruber et al., 2008). Similarly,

177 within an individual plant, some cyclotides occur in several tissues and others are specific to a given

178 tissue. For example, vhr1 occurs only in roots in V. hederacea but kalata B1 occurs in leaves, stems

179 and roots in O. affinis (Trabi & Craik, 2004).

180 Imaging studies suggest that there is a non-uniform distribution of cyclotides within a given

181 tissue type. For example, matrix-assisted laser desorption/ionization-mass spectrometric imaging

182 (MALDI-MSI) has been applied to assess the spatial distribution and relative abundances of

183 cyclotides within the leaves of Petunia x hybrida. The study revealed four distinct masses on a P.

184 hybrida , one of them corresponding to Phyb A, which was found in higher abundance within the

185 mid-vein tissue compared with the laminar and peripheral leaf tissues (Poth et al., 2012). Thus, that

186 report demonstrated that cyclotides associate with the vascular section of leaf tissues, suggesting that

187 these peptides may play a role in plant defense through modulating insect herbivory.

9

188 2 Biological activities of cyclotides

189 In addition to the reported uterotonic activity of kalata B1 (Gran, 1973a, 1973b), a wide range of other

190 biological activities, including hemolytic activity (Daly et al., 1999a; Ireland et al., 2006a; Schöpke

191 et al., 1993; Tam et al., 1999), neurotensin antagonistic (Witherup et al., 1994), anti-HIV (Bokesch et

192 al., 2001; Chen et al., 2005; Daly et al., 2006; Daly et al., 2004; Gustafson et al., 1994; Gustafson et

193 al., 2000; Hallock et al., 2000; Ireland et al., 2008; Wang et al., 2008a), anti-microbial (Fensterseifer

194 et al., 2015; Gran et al., 2008; Pranting et al., 2010; Tam et al., 1999), protease inhibitory (Hernandez

195 et al., 2000; Quimbar et al., 2013), insecticidal (Jennings et al., 2001), antitumor (Herrmann et al.,

196 2008; Lindholm et al., 2002; Svangård et al., 2004; Tang et al., 2010), antifouling (Göransson et al.,

197 2004), nematocidal (Colgrave et al., 2008a), molluscicidal (Plan et al., 2008), cell-penetrating

198 (Contreras et al., 2011; Greenwood et al., 2007), immunosuppressive (Gründemann et al., 2012;

199 Gründemann et al., 2013), and prolyl oligopeptidase inhibitory activities (Hellinger et al., 2015) has

200 been reported for cyclotides, as summarized in Table 1.

10

201 Table 1. Summary of host defense-related and/or biological activities first reported for cyclotides

Activity Exemplary cyclotides First report

Uterotonic kalata B1, B2 and B7 (Gran, 1973a, 1973b)

Hemolytic violapeptide I, circulin A (Schöpke et al., 1993)

and B, cyclopsychotride

A, kalata B1, varv A

Anti-HIV circulin A and B, (Gustafson et al., 1994)

palicourin, kalata B1,

vhl-1

Neurotensin antagonist cyclopsychotride A (Witherup et al., 1994)

Anti-microbial circulin A and B, (Tam et al., 1999)

cyclopsychotride A,

kalata B1, kalata B7,

cycloviolacin O2

Trypsin inhibitor MCoTI-I and MCoTI-II (Hernandez et al., 2000)

Insecticidal kalata B1 and B2 (Jennings et al., 2001)

Antitumor cycloviolacin O2, varv A (Lindholm et al., 2002)

and F

Antifouling cycloviolacin O2 (Göransson et al., 2004)

Cell internalization MCoTI-II, kalata B1 and (Greenwood et al., 2007)

MCoTI-I

Nematocidal kalata B1 and B2, (Colgrave et al., 2008a)

cycloviolacin O2

Molluscicidal cycloviolacin O1, kalata (Plan et al., 2008)

B1, B2 and B5

Immunosuppressive kalata B1 (Gründemann et al., 2012)

Prolyl oligopeptidase kalata B1, psysol 2 (Hellinger et al., 2015)

inhibition

11

202 2.1 Toxic activities

203 One of the first activities reported for cyclotides was the ability to cause lysis of human erythrocytes.

204 This hemolytic activity was initially observed in violapeptide I, the first cyclotide discovered from

205 the violet family (Schöpke et al., 1993). Additional macrocyclic peptides were discovered during the

206 course of screening other plants for a range of other biological activities, including anti-HIV and anti-

207 neurotensin activities. In one of the first such studies, Gustafson et al. reported the antiviral activities

208 of circulins A and B, bracelet cyclotides isolated from Chassalia parvifolia (Rubiaceae), which

209 demonstrated antiviral cytoprotective effects on various human immunodeficiency virus (HIV) strains

210 at concentrations ranging from 40 to 260 nM (Gustafson et al., 1994). Similarly, screens for

211 neurotensin antagonistic activity led to the discovery of cyclopsychotride A from Psychotria longipes

212 (Rubiaceae) (Witherup et al., 1994), which was later also reported to be active against Gram-positive

213 and Gram-negative bacteria (Tam et al., 1999).

214 The antimicrobial activities of cyclotides were first described by Tam et al. (Tam et al., 1999),

215 whereby synthetically derived kalata B1 was reported to be active against Staphylococcus aureus in

216 the absence of salt in the test solution. The inhibitory activity was abolished under physiological

217 conditions, i.e. in the presence of 100 mM NaCl, and the peptide was inactive against E. coli. By

218 contrast, a later study of native kalata B1 by Gran et al. (Gran et al., 2008) suggested that it had

219 antibiotic effects on E. coli under low or high-salt conditions. In addition to these apparently

220 conflicting reports for in vitro studies, a recent publication reported that cyclotides kalata B2 and

221 cycloviolacin O2 display in vitro antibacterial activities and in vivo efficacy against S. aureus in a

222 subcutaneous wound infection animal model (Fensterseifer et al., 2015). The reported salt dependence

223 for the antimicrobial activity (Tam et al., 1999) suggests that the initial interaction between cyclotides

224 and the microbial surface may be electrostatic, similar to that described for defensins (Oren & Shai,

225 1998). The similarity in size of cyclotides to disulfide-rich defensins is consistent with a functional

226 role of cyclotides in host-defense.

227 There is ongoing interest in the literature in the antimicrobial potential of cyclotides. A recent

228 report on the antibacterial activity of peptide-containing extracts from the flowers of the elderberry

229 tree (Sambucus nigra) noted some partial sequences consistent with known cyclotides but the active

12

230 components have not yet been definitely confirmed as cyclotides (Álvarez et al., 2018a). That study

231 and another from the same group suggested potential applications of cyclotides as antimicrobial in

232 aquaculture applications (Álvarez et al., 2018b). Another report demonstrated the use of cyclotides as

233 surface coatings to reduce biofilm formation (Cao et al., 2018). Underpinning such applications, there

234 has also been interest in using computational approaches to predict the preferential cyclotide scaffolds

235 for antimicrobial applications (Balaraman & Ramalingam, 2018) and the potential for the

236 development of drug resistance (Malik et al., 2017; Noonan et al., 2017).

237 Cytotoxic activity with the potential for antitumor applications is another noteworthy property

238 reported in early cyclotide studies. In one study, cycloviolacin O2, isolated from V. odorata, and varv

239 A and varv F, isolated from V. arvensis, displayed cytotoxic activities against ten human tumor cell

240 lines, including immortalized myeloma, T-cell leukemia, small cell lung cancer, lymphoma, and renal

241 adenocarcinoma cell lines, as well as primary ovarian carcinoma cells from cancer patients (Lindholm

242 et al., 2002). The cytotoxicity of varv A and cycloviolacin O2 on healthy human lymphocytes was

243 also evaluated in the same study. Compared to the normal lymphocytes, both peptides displayed ~9-

244 fold higher selectivity towards the leukemia cells (Lindholm et al., 2002). Other cyclotides isolated

245 from the Violaceae family were extensively tested against a variety of human cancer cells of varying

246 origin and were reported to have potent activities in vitro (Herrmann et al., 2008; Pinto et al., 2018;

247 Svangård et al., 2004; Tang et al., 2010), suggesting that these plant-derived peptides might be

248 promising cytotoxic compounds with potential in cancer treatment. Since cycloviolacin O2 showed

249 promising results in cytotoxicity testing against lymphoma, leukemia, small cell lung cancer and

250 colon carcinoma cell lines in vitro at concentrations in the low micromolar range, a follow-up study

251 on the antitumor effects of the peptide was done in vivo using mouse xenograft models with hollow

252 fibers containing various human cancer cell lines implanted subcutaneously (Burman et al., 2010b).

253 The maximum tolerated dose was determined for single-dose injection (1.5 mg/kg) and repeated-dose

254 injection (0.5 mg/kg) prior to a hollow fiber study and xenograft study, respectively. Animals

255 implanted with a hollow fiber encapsulated with tumor cells received a single dose of cycloviolacin

256 O2 at 1 mg/kg one day after the implantation. Xenografted animals were treated with cycloviolacin

257 O2 by intravenous injection at 0.5 mg/kg daily up to 14 days. With no significant antitumor affects

13

258 observed, repeated dosing of cycloviolacin O2 at 1 mg/kg was reported to give a local-inflammatory

259 reaction at the injection site fter 2-3 doses and acute toxicity was observed after administration of the

260 cyclotide at 2 mg/kg. This negative result might be due to a low distribution of peptide at the site of

261 the implants or to intrinsically weak in vivo antitumor activity of the peptide. A recent study on the

262 selectivity of cyclotides against cells further demonstrated that cycloviolacin O2, kalata B1 and kalata

263 B2 are toxic towards both non-transformed (skin and PBMC) and transformed (cervical cancer,

264 melanoma and leukemia) cells and they exert their effects through targeting cell membranes

265 containing phosphatidylethanolamine (PE) phospholipids, causing subsequent membrane disruption

266 (Henriques et al., 2014). The finding of cyclotides interacting preferentially with PE-phospholipids is

267 in agreement with an earlier report by Burman et al (Burman et al., 2011).

268 Cyclotides possess potent intrinsic insecticidal activities, and thus have potential applications

269 in agriculture, as first reported by Jennings et al (Jennings et al., 2001). In that study, purified kalata

270 B1 was incorporated into an artificial diet and fed to larvae of Helicoverpa punctigera, a serious pest

271 of cotton crops worldwide. Half of the kalata B1 treated larvae died and none of the survivors

272 developed past the first instar stage of larval development. A later study demonstrated that cyclotides

273 damage the gut epithelium of H. armigera by inducing disruption of the microvilli, causing blebbing

274 of epithelial cells, swelling of the columnar cells and ultimately rupture of the cells (Barbeta et al.,

275 2008).

276 Since cyclotides possess insecticidal activity, it has been of interest to examine their activity

277 against other agricultural pests to understand their potentially broader role as natural pesticides. The

278 insecticidal activities of kalata B1 and a suite of alanine mutants against adult Drosophila

279 melanogaster were assessed and the disruption of the development of first and second instar larvae

280 through to adult D. melanogaster was demonstrated (Simonsen et al., 2008). In another series of

281 studies conducted by Colgrave et al., a range of cyclotides was found to inhibit the larval development

282 of two economically important sheep gastrointestinal nematodes, Haemonchus contortus and

283 Trichostrongylus colubriformis (Colgrave et al., 2008b; Wang et al., 2008a), as well as canine and

284 human hookworms (Colgrave et al., 2009). Kalata B1 (IC50 = 2.2 µM) showed equipotent activity to

285 the commercially used anthelmintic drugs, levamisole (IC50 = 8.9 µM) and naphthalophos (IC50 = 7.5

14

286 µM) against the larval life stage, confirming a potential role for cyclotides as natural anthelmintic

287 control agents. In a later study, kalata B1, B6 and cycloviolacin O14 were shown to have a pronounced

288 effect on the viability of larval and adult life stages of the dog hookworm Ancylostoma caninum and

289 inhibited larval development of the human hookworm Necator americanus (Colgrave et al., 2009).

290 Subsequent investigations of other natural cyclotides extracted from V. odorata were reported

291 and identified examples with up to 18-fold greater potency than kalata B1 in larval development

292 assays against H. contortus, with the most potent cyclotide being cycloviolacin O2 (Colgrave et al.,

293 2008b). Modification of cycloviolacin O2 by acetylation of the two lysine residues to mask their

294 charge resulted in a marked decrease in anthelmintic activity for this Viola-derived peptide to a level

295 comparable to kalata B1 (Colgrave et al., 2008b). A correlation was also observed between the

296 number of charged residues present in cyclotide sequences and their anthelmintic activity, suggesting

297 that the net charge of a cyclotide is probably an important determinant of anthelmintic activity.

298 Most recently, the antifungal activities of extracts from Viola odorata were reported (Slazak et al.,

299 2018) and suggest a role for cyclotides from this plant against microbial pests. Specifically,

300 cycloviolacin O2, O3, O13, and O19 (cyO2, O3, O13 and O19) isolated from Viola odorata were

301 evaluated for activity against five model plant fungal pathogens, namely Fusarium oxysporum,

302 Fusarium graminearum, Fusarium culmorum, Mycosphaerella fragariae, and Botrytis cinerea, and

303 two Viola-derived pathogens, namely Colletotrichum utrechtense and Alternaria alternate. All tested

304 cyclotides displayed antifungal activity. CyO3 exhibited the most potent activity with minimal

305 inhibitory concentrations (MICs) ranging from 0.8 to 12.5 μM; while cyO13 exhibited the lowest

306 activities with MICs ranging from 3 to 25 μM. All cyclotides displayed low micromolar activity

307 against A. alternate, a fungal pathogen also isolated from Viola odorata. Figure 3 schematically

308 illustrates the range of host defense-related activities of cyclotides.

309

15

310 Figure 3: Examples of pesticidal and toxic activities of native cyclotides. Cyclotides have been 311 reported to possess potent activities against various pests, including the nematode H. contortus 312 (nematocidal)(Colgrave et al., 2008a), budworm H. punctigera (insecticidal) (Jennings et al., 2001), 313 rice pest Pomacea canaliculata (molluscicidal) (Plan et al., 2008), and fouling barnacles Balanus 314 improvisus (antifouling) (Göransson et al., 2004). The reported toxic effects of cyclotides against the 315 human immunodeficiency virus (anti-HIV) (Gustafson et al., 1994) and a range of cancer cell lines 316 (anti-tumor) (Lindholm et al., 2002) highlight the potential therapeutic and agrochemical applications 317 of these peptides. Therapeutic activities that are not linked with the toxic effects of cyclotides, e.g. 318 immunosuppressive activity (Gründemann et al., 2012), are not included in this figure.

319 2.2 Pharmaceutical activities

320 In addition to their toxic or host-defense activities, a range of naturally occurring cyclotides possess

321 activities of pharmacological and pharmaceutical relevance beyond the uterotonic, anti-HIV and

322 neurotensin antagonist activities noted already. In a recent study that nicely links the original

16

323 indigenous medical applications of the cyclotide-bearing plant O. affinis with the latest

324 pharmacological studies, the oxytocic activity of kalata B7, a Möbius cyclotide closely related to

325 kalata B1 was reported and the possible molecular target underlying the mechanism of the uterotonic

326 activity was revealed. Pharmacological studies showed that kalata B7 is a partial agonist of both the

327 oxytocin receptor and vasopressin V1A receptor and a structural analysis suggested that loop 3 of

328 kalata B7, which displays moderate sequence homology to human oxytocin, could be responsible for

329 the observed uterostimulant effects on uterine smooth muscle cells (Koehbach et al., 2013b).

330 The discovery of the anti-proliferative activity of cyclotides on primary activated human

331 lymphocytes suggested the potential use of these peptides as immunosuppressant drugs. The

332 inhibitory effects of kalata B1 on the proliferation of human peripheral blood mononuclear cells

333 (PBMC) at non-cytotoxic concentrations was first reported by Gründemann et al. (Gründemann et al.,

334 2012). This research led to further investigations of the immunosuppressive properties of cyclotides,

335 whereby an analog of kalata B1 [T20K] was shown to attenuate the interleukin-2 (IL-2) secretion and

336 the expression of IL-2 surface receptor on activated T-lymphocytes (Gründemann et al., 2013).

337 Further mechanistic studies, described in the same report, using several kalata B1 analogs with single

338 point mutations determined the cyclotide motif accountable for the anti-proliferative activity. The

339 recent progression of [T20K]kalata B1 to Phase I clinical trials for multiple sclerosis further illustrates

340 the potential applications of cyclotides in immunopharmacology and immunosuppression

341 (Gründemann et al., 2019).

342 2.3 Cell penetrating properties

343 Delivery of peptide-based drugs to intracellular targets is one of the holy grails of drug development.

344 Cyclotides triggered interest as potential frameworks for intracellular drug delivery after Greenwood

345 et al. reported the internalization of the cyclotide MCoTI-II into mammalian cells by endocytosis

346 (Greenwood et al., 2007). In that study, cells treated with biotinylated cyclotides were fixed and

347 stained with avidin-FITC before the internalization was evaluated using confocal fluorescence

348 microscopy. Although internalization studies in such systems need careful analysis to avoid possible

349 artifacts, the study provided the first indication of the cell penetrating ability of a cyclotide. In a later

350 study, the cellular uptake of fluorescently labeled MCoTI-II and kalata B1 were explored using live- 17

351 cell confocal imaging techniques and their affinity for phospholipids was examined on model

352 membrane systems by surface plasmon resonance or on PIP strips™ membranes (Cascales et al.,

353 2011). It was confirmed that MCoTI-II and kalata B1 are both able to penetrate cells but that they

354 probably cross cell membranes through different pathways. That study categorized these two peptides

355 and a smaller cyclic sunflower peptide, SFTI-1, as a new class of cell-penetrating peptides, referred

356 to as cyclic cell-penetrating peptides (Cascales et al., 2011). A contemporaneous study, using real

357 time confocal microscopy, showed that MCoTI-I, a cyclotide with ~95% sequence similarity to

358 MCoTI-II, internalized into HeLa cells predominantly through a temperature-dependent active

359 endocytic pathway (Contreras et al., 2011). Overall, these independent reports confirm that cyclotides

360 have the potential to be used as stable scaffolds for delivering therapeutically significant peptide

361 epitopes into cells and this topic is likely to be an active area of future cyclotide research.

362 3 Synthesis, structure-activity relationships and mode of action of cyclotides

363 Research on cyclotides has led to a number of impacts in the field of biological chemistry, including

364 the development of approaches to the chemical and biological synthesis of cyclic peptides, which has

365 opened up technologies for deriving structure-activity studies of cyclotides and for their applications

366 as drug design frameworks.

367 3.1 Chemical and biological syntheses of cyclotides

368 Approaches for the chemical synthesis of cyclotides were first described in the late 1990’s (Daly et

369 al., 1999b; Tam & Lu, 1998). In Tam and Lu’s report (Tam & Lu, 1998), the backbone cyclization of

370 circulin B and cyclopsychotride A was achieved by adapting native chemical ligation (Dawson et al.,

371 1994), where an N-terminal cysteine and a C-terminal thioester of peptide precursors synthesized

372 using Boc-chemistry solid phase peptide synthesis (SPPS) reacted to form a thioester that

373 subsequently underwent a spontaneous acyl transfer reaction to produce a native amide bond between

374 the two termini. In that study orthogonal protection of pairs of Cys residues was used to direct the

375 oxidation in a stepwise manner to form the desired disulfide connectivities. Daly et al. (Daly et al.,

376 1999b) used a similar approach for the synthesis of kalata B1 but without orthogonal protection of

377 Cys residues and showed it was possible to readily isolate the correctly folded product from the

18

378 mixture of disulfide isomers. Overall, the native chemical ligation-based synthesis method using Boc-

379 chemistry has been applied to the cyclization of a wide range of macrocyclic peptides and has proven

380 to be valuable in the routine production of cyclotides (Clark & Craik, 2010). Various Fmoc-based

381 SPPS methods to generate the thioester precursors of cyclotides for head-to-tail cyclization via native

382 chemical ligation (Gunasekera et al., 2013; Taichi et al., 2013), bacterial expression of recombinant

383 linear precursors of cyclotide followed by in vitro cyclization (Cowper et al., 2013), and a more direct

384 strategy for backbone cyclization using Fmoc-based SPPS (Cheneval et al., 2014) have also been

385 reported in recent years.

386 Several enzyme-mediated approaches for cyclization of cyclotides have been explored,

387 including intein-mediated biosynthetic methods (Camarero et al., 2007; Jagadish et al., 2015; Kimura

388 et al., 2006) and sortase A-catalyzed backbone cyclization (Jia et al., 2014). The intein-mediated

389 backbone cyclization of kalata B1 was achieved by recombinantly expressing linear cyclotide

390 precursors fused to a Met and an engineered intein unit at their N and C termini, respectively. The

391 fusion proteins were cleaved by endogenous Met aminopeptidase and underwent intein-mediated

392 protein splicing in E. coli which resulted in linear cyclotide precursors with a C-terminal α-thioester

393 and an N-terminal Cys required for native chemical ligation-based cyclization in vitro (Kimura et al.,

394 2006). The intein-mediated in vivo production of cyclotides MCoTI-II and MCoTI-I was reported by

395 the same group, in live bacterial cells (Camarero et al., 2007) and yeast cells (Jagadish et al., 2015),

396 respectively. These reports demonstrated the applicability of recombinant expression of natively

397 folded cyclotides in microorganisms and the possibility of producing large combinatorial cyclotide-

398 based libraries for screening. More recently, a chemo-enzymatic approach was developed to

399 synthesize cyclic disulfide-rich peptides, including kalata B1, cyclic α-conotoxin Vc1.1 and SFTI,

400 whereby the chemically synthesized linear peptide precursors containing a sortase A recognition motif

401 at the C-terminus were cyclized by sortase A in vitro without significant perturbation to the overall

402 peptide fold (Jia et al., 2014).

403 In parallel with the development of new methodologies for the efficient synthesis of cyclotides,

404 there is a growing effort towards understanding the biosynthetic mechanism of these macrocyclic

405 peptides in plants. Since an Asn is highly conserved at the C-terminus of the cyclotide domain,

19

406 asparaginyl endopeptidase (AEP), a cysteine protease with specificity for Asn, has been implicated

407 in the cyclization of cyclotides in vivo (Saska et al., 2007). In this key study, kalata B1 was expressed

408 transiently in Nicotiana benthamiana transformed with the precursor of kalata B1, and a reduction in

409 the yield of backbone-cyclized kalata B1 was observed with AEP-gene silencing constructs, providing

410 an initial correlation between AEP activity and cyclization yield of cyclotides in plants (Saska et al.,

411 2007). A recent publication reported the discovery of the AEP homolog butelase 1 in the cyclotide-

412 propeptides with a His-Val sequence at the C terminus in vitro (Nguyen et al., 2014). Butelase 1 has

413 been shown to cyclize peptides of varying lengths (from 14 to 34 residues) and sequences, including

414 kalata B1, SFTI, cyclic conotoxin MrlA, and antimicrobial peptide histatin-3 at a low enzyme-to-

415 peptide ratio (1:400) within 48 min. This finding suggests that butelase 1 could be developed as an

416 alternative approach to complement the current chemical and biological methodologies in producing

417 macrocyclic peptides.

418 Another impact deriving from cyclotide research has been the development of methodologies

419 for crystalizing disulfide-rich peptides. Of the many structures of cyclotides published, until recently,

420 only one involved X-ray crystallography because cyclotides along with other disulfide-rich peptides

421 are notoriously difficult to crystallize. However, Wang et al. (Wang et al., 2014) demonstrated that

422 the use of racemic crystallography dramatically improved crystallization rates and determined crystal

423 structures for a series of cyclic disulfide-rich peptides, ranging from SFTI-1 (14 amino acids with one

424 disulfide bond) to a cyclized conotoxin (22 amino acids with two disulfide bonds) to kalata B1 (29

425 amino acids with three disulfide bonds). Although this technology was demonstrated for cyclic

426 molecules, it should be equally applicable to the crystallization of a wide range of acyclic disulfide-

427 rich peptides.

428 3.2 Structure-activity relationships

429 The ability to chemically synthesize cyclotides has facilitated a wide range of mutagenesis studies

430 and structure-activity relationship studies that reveal the importance of the circular backbone for

431 maintenance of cyclotide structural integrity. In one early study by Daly et al. (Daly & Craik, 2000),

432 six acyclic permutants of kalata B1, with the backbone opened in each of the six inter-cysteine loops,

433 were synthesized and their overall folds were compared with that of the cyclic native peptide. A native 20

434 fold could not be achieved in acyclic mutants having a break of the backbone in either loops 1 or 4,

435 which are the loops forming the embedded ring in the cystine knot. This result suggests that the cystine

436 knot is essential in stabilizing the intermediates formed during the oxidative folding of cyclotides.

437 The overall folds of the four other acyclic analogs of kalata B1, with a break in one of loops 2, 3, 5,

438 or 6, were found to be very similar to that of their parent peptide, showing that a cyclic backbone is

439 not essential for a native-like fold. Although these four acyclic analogs of kalata B1 retained a native-

440 like conformation, their lack of hemolytic activity suggests that the circular backbone is functionally

441 important (Daly & Craik, 2000). Furthermore, the three-dimensional solution structures of a synthetic

442 acyclic permutant of kalata B1 with most of loop 6 removed and a naturally occurring linear cyclotide,

443 violacin A (with a discontinuous loop 6), showed that a backbone discontinuity renders structures

444 more flexible than in their cyclic counterparts (Barry et al., 2003; Daly & Craik, 2000; Ireland et al.,

445 2006b). These combined findings indicate that the circular backbone is crucial to both the structure

446 rigidity and activity of cyclotides (Barry et al., 2003; Daly & Craik, 2000).

447 The ability to chemically synthesize cyclotides has also facilitated a wide range of

448 mutagenesis studies that have helped to delineate their mode of action, as described in the following

449 section.

450 3.3 The mode of action of cyclotides

451 The mode of action of cyclotides may vary depending on the particular biological activity but in

452 general is strongly dependent on their unique structural features. The cystine knot structural motif

453 effectively results in the exclusion of non-Cys side-chains from the core region of cyclotides,

454 promoting the surface exposure of hydrophobic residues, some of which are clustered together to form

455 a hydrophobic patch. Several characteristic biophysical properties derive from this surface-exposed

456 hydrophobic patch, including late elution on RP-HPLC and weak self-association. These properties

457 have potential implications for the mode of action of cyclotides as cytotoxic agents since they provide

458 clues as to how these molecules might interact with and form pores in membranes. In this regard, the

459 oligomerization and self-association properties of cyclotides have been investigated using analytical

460 ultracentrifugation techniques. For instance, kalata B2 was observed to self-associate into tetramers

461 and octamers (Nourse et al., 2004), but not dimers. In one model of the geometry of the tetramer 21

462 proposed in that study, the oppositely charged residues Arg-24 and Asp-25 in kalata B2 create an

463 exposed bipolar patch at one end of the molecule, which was postulated to facilitate intermolecular

464 ionic self-interaction and potentially play a role in the formation of membrane-spanning pores. A later

465 experimental NMR study suggested an alternative model for the self-association in solution based on

466 interaction between the hydrophobic patches of kalata B2 (Rosengren et al., 2013). The significance

467 of the solution-state oligomers to the function of cyclotides remains unknown, and similarly it is not

468 known if cyclotides form aggregates in their membrane bound states, but a wide range of studies do

469 suggest that membrane interactions are intimately associated with cyclotide functions.

470 Synthetic and mutagenesis-based studies have contributed significantly to defining the

471 membrane-interacting hypothesis for the mode of action of kalata B1. For instance, a comparison of

472 enantiomer forms of a peptide ligand provides a definitive means to indicate whether a chiral protein

473 receptor is involved in its biological function or whether it acts via a (largely achiral) membrane

474 disruption mechanism. Colgrave et al. (Colgrave et al., 2008a) showed that the nematocidal activity

475 of the mirror-image stereoisomer of kalata B1 was similar to the wild-type peptide, suggesting that

476 the mechanism of action is probably via membrane interaction rather than by a chiral (i.e. protein-

477 based) receptor interaction. The self-association behaviour of cyclotides and the comparable

478 bioactivity of the all D-enantiomer of kalata B1 to the native L-form are key pieces of evidence which

479 suggest that the mechanism of action may be via membrane interaction. The membrane-based

480 mechanism of action of cyclotides was supported by an early surface plasmon resonance study which

481 demonstrated that kalata B1 and B6 bind selectively to phosphatidylethanolamine (PE)-containing

482 model membranes (Kamimori et al., 2005). Further support for the membrane-binding properties of

483 native cyclotides derived from the observation that cycloviolacin O2 induced leakage of both calcein-

484 loaded HeLa cells and a lipid model in the form of palmitoyloleoylphosphatidylcholine (POPC)

485 liposomes (Svangård et al., 2007). NMR studies by Shenkarev et al. showed that the binding of kalata

486 B1 and kalata B7 to dodecylphosphocholine (DPC) micelles is modulated by both electrostatic and

487 hydrophobic interactions (Shenkarev et al., 2008; Shenkarev et al., 2006). Varv F was shown to bind

488 to DPC micelles and its overall conformation remained unchanged upon binding (Wang et al., 2009b).

22

489 The nematocidal activity of a suite of alanine mutants of kalata B1 was examined and the

490 residues critical for activity against helminths correlated with those significant for insecticidal activity

491 against D. melanogaster (Simonsen et al., 2008). Residues critical for the biological activities of

492 kalata B1 were clustered on one side of the molecule, named ‘the bioactive patch’. Since membrane

493 interaction involving oligomerization was speculated to be responsible for the insecticidal activities

494 of kalata B1, the whole suite of alanine mutants of kalata B1 was screened against a range of model

495 membranes encapsulated with self-quenching dye for their lytic activities (Huang et al., 2009). The

496 leakage study confirmed that the bioactive patch of kalata B1 plays a critical role in its lytic, as well

497 as its insecticidal and hemolytic activities. In addition, kalata B1 was observed to have a preference

498 for phospholipids containing PE headgroups compared to model membranes containing only

499 zwitterionic or anionic phospholipids (Huang et al., 2009). Results from patch-clamp experiments

500 suggested that kalata B1 induced leakage via pore formation on reconstituted asolectin (soybean

501 lecithin), when compared with a membrane-inactive mutant of kalata B1 (V25A) (Huang et al., 2009).

502 A later study of lysine mutants of kalata B1 revealed that a single lysine substitution on a face opposite

503 to the bioactive patch improved its nematocidal activity (Huang et al., 2010). Furthermore, Colgrave

504 et al. observed increasing uptake of the radiolabel [3H]inulin of ligated adult nematodes upon kalata

505 B1 treatment, providing evidence to support the conclusion that the anthelmintic effect of the

506 cyclotide was due to increased permeability of the external membrane of the nematodes (Colgrave et

507 al., 2010). Together, these various mutagenesis studies and electrophysiological recordings provide

508 mechanistic insights into how kalata B1 exerts its effects on different organisms.

509 Many other bioactivities of cyclotides correlate with lipid-binding properties, as supported by

510 detailed biophysical studies using surface plasmon resonance (Henriques et al., 2012; Henriques et

511 al., 2014; Henriques et al., 2011) and isothermal titration calorimetry (ITC) (Wang et al., 2012) on

512 model membranes. An extensive lipid binding study of kalata B1 and a range of its active and inactive

513 mutants using surface plasmon resonance suggested that kalata B1 preferred more rigid membranes

514 containing PE phospholipids and exerted its anti-HIV activities by disrupting the membrane envelope

515 of viral particles (Henriques et al., 2011). Furthermore, a titration of kalata B1 with PE, monitored

516 using 1H NMR chemical shifts, suggested that it interacted specifically with the PE headgroups

23

517 through residues that formed part of the bioactive patch (Henriques et al., 2011). Therefore, the

518 membrane-targeting properties of cyclotides against PE headgroups was proposed as the initial step

519 of their lytic action, followed by membrane insertion with the hydrophobic patch, which leads to local

520 membrane disturbances and eventually membrane disruption (Henriques et al., 2012). More recently,

521 the PE-targeting ability of kalata B1 was also suggested to be responsible for the observed cell

522 internalization of kalata B1 at concentrations lower than the cytotoxicity threshold (Henriques et al.,

523 2015).

524 Figure 4 summarizes our current understanding of the proposed mechanism of cell

525 internalization of kalata B1, which involves the binding of the bioactive patch to PE phospholipids in

526 cell membranes via electrostatic interactions, before the hydrophobic face of the cyclotide is inserted

527 into the core of the bilayer. The accumulation of cyclotide on the lipid bilayer causes local membrane

528 disturbances, which eventually leads to cell penetration.

529

530 Figure 4: A schematic representation of the cell internalization of kalata B1. The initial step of 531 cell internalization of kalata B1 is PE-targeting. The bioactive patch (highlighted in red) of kalata B1 532 binds to the headgroup of PE phospholipids in cell membranes via electrostatic interaction (1), 533 followed by the insertion of the hydrophobic face (highlighted in green) into the core of the bilayer 534 (2). The accumulation of cyclotide molecules on the lipid bilayer leads to local membrane 535 disturbances, which eventually leads to cell penetration (3) through: i) endocytosis or ii) membrane 536 translocation by an energy-independent process. Figure adapted from a previously published article 537 (Henriques et al., 2015). 538 24

539 Overall, membrane binding is fundamental for many reported functions of cyclotides,

540 including hemolytic, insecticidal, nematocidal, and anti-HIV activities, as well as cell internalization.

541 However, considering that some cyclotides have been reported to bind to several members of the G

542 protein-coupled receptor family for their oxytocic properties (Koehbach et al., 2013b), there is a

543 possibility that other cyclotides might also exert other activities via modulating cellular receptors

544 separately from or in addition to binding to membranes.

545

546 4 Applications

547 Cyclotides have a range of potential applications in agriculture and medicine based on their

548 exceptional stability and their tolerance to sequence modifications that allow “designer cyclotides” to

549 be made. In this section, we very briefly outline some of those applications.

550 4.1 Medical applications of cyclotides

551 One approach to medical application of cyclotides is to harness some of the toxic properties of natural

552 cyclotides for therapeutic applications, for example, as cytotoxic (anti-cancer agents) or as

553 nematocidal agents with applications for human parasites, such as hookworms. A second area of

554 medical applications is through “designer” cyclotides made by grafting a bioactive epitope into a

555 cyclotide sequence to introduce a new activity of therapeutic relevance not present in the original

556 cyclotide framework. The aim of all of these studies is to take advantage of the stability of the

557 cyclotide framework to stabilize the peptide epitope. These grafting applications have been widely

558 reviewed elsewhere, so we will not discuss them further here except to refer readers to recent reviews

559 on the topic (Camarero, 2017; Craik & Du, 2017; de Veer et al., 2017; Gould & Camarero, 2017;

560 Gould et al., 2011; Poth et al., 2013; Wang & Craik, 2018). There are more than 25 examples of

561 grafted cyclotides for a range of diseases, including cardiovascular disease, cancer, wound healing,

562 pain, inflammation and multiple sclerosis. So far, none of these examples has reached human clinical

563 trials but all are well exemplified by animal studies or at least in vitro testing.

25

564 4.2 Agricultural applications

565 Stimulated by their natural functions as endogenous pesticidal agents in certain plants,

566 cyclotides have attracted attention for potential applications in the protection of crop plants that

567 naturally do not contain them. Such applications include their incorporation into transgenic plants, a

568 topic that is outside the scope of the current article, as well as applications involving external

569 administration onto growing crops or harvested material. The most advanced application involving

570 external application is exemplified with the recent approval of SeroX, an extract from butterfly pea

571 (Clitoria ternatea), as a treatment for insect pests on cotton and macadamia nut crops in Australia.

572 This plant, from the Fabaceae family, contains more than 47 different cyclotides (Gilding et al., 2016),

573 of which the peptide Cter M, at least, has been shown to have insecticidal properties as an isolated

574 peptide (Poth et al., 2011a). The SeroX product is marketed as a spray for cotton at doses as low as

575 2L/hectare by its developer, Innovate Ag, based in Australia.

576 While we will not cover the alternative mode of delivery of cyclotides via the incorporation

577 of transgenes encoding cyclotides into crop plants here, it is useful to note that there have been a

578 number of recent advances in understanding the roles of asparaginyl endopeptidase enzymes in

579 facilitating cyclotide biosynthesis (Harris et al., 2015; Jackson et al., 2018). Additionally, the enzyme

580 kalatase A, which is responsible for the N-terminal processing of cyclotide precursors was recently

581 reported (Rehm et al., 2019). These studies will no doubt be useful in facilitating the production of

582 transgenic plants with high yields of pesticidal cyclotides, thereby engendering these plants with

583 similar levels of insect protection to natural cyclotide-producer plants.

584 5 Outlook and future studies

585 Overall, cyclotides have attracted a great deal of interest, not only for their natural insecticidal

586 activities and their potential as drug scaffolds but also because of their topologically unique structures.

587 These structures engender cyclotides with exceptional stability and thus, in principle, they have

588 advantages over less stable peptides in that they offer potential for a variety of formulation approaches

589 and are stable to long term storage, an important consideration both for pharmaceutical and

590 agricultural applications.

26

591 The cyclotide field is still relatively young and only a small number of groups are currently

592 studying these fascinating cyclic and knotted peptides. Their natural function as insecticidal or

593 nematocidal agents justifiably allows them to be regarded as toxins. Their mechanism of toxicity

594 appears to be mainly related to membrane-binding. Their membrane binding, however, is far from

595 non-specific, and cyclotides exhibit a remarkable preference for PE lipids compared to other lipid

596 types. It is not yet known whether it is this lipid specificity that controls the specificity of different

597 cyclotides for different organisms, but it seems to be a reasonable hypothesis. Also unknown at the

598 moment is why one plant produces so many cyclotides. Is it a strategy for the plant to try and avoid

599 the development of resistance by pests to the chemical defense? Or is it a strategy for simultaneously

600 targeting a wide variety of different pests? These questions will undoubtedly be answered in due

601 course, assisted by advances in technologies for the synthesis of cyclotides. For example, as we have

602 noted, it is now possible to make variants of cyclotides where individual residues or individual loops

603 can be replaced to explore structure-activity studies. Biological methods of producing cyclotides are

604 also improving and promise to accelerate the development of structure-activity relationships.

605 Amongst toxins cyclotides do not have quite the same caché as the deadly toxins from animal

606 venoms, but we hope that this article has convinced readers that they are toxins with a vast range of

607 potential applications in the pharmaceutical and agricultural fields.

608

609 Acknowledgements

610 Work in our laboratory on cyclotides is funded by grants from the Australian Research Council

611 (DP150100443) and the National Health and Medical Research Council (APP1084965 and

612 APP1060225). DJC is an Australian Research Council Laureate Fellow (FL150100146).

27

613 References

614 Álvarez, C. A., Barriga, A., Albericio, F., Romero, M. S., & Guzmán, F. 2018a. Identification of

615 peptides in flowers of sambucus nigra with antimicrobial activity against aquaculture

616 pathogens. Molecules, 23, e1033

617 Álvarez, C. A., Santana, P., Luna, O., Cárdenas, C., Albericio, F., Romero, M. S., & Guzmán, F.

618 2018b. Chemical synthesis and functional analysis of varvA cyclotide. Molecules, 23, e952

619 Balaraman, S., & Ramalingam, R. 2018. The structural and functional reliability of Circulins of

620 Chassalia parvifolia for peptide therapeutic scaffolding. J. Cell. Biochem., 119, 3999-4008

621 Barbeta, B. L., Marshall, A. T., Gillon, A. D., Craik, D. J., & Anderson, M. A. 2008. Plant cyclotides

622 disrupt epithelial cells in the midgut of lepidopteran larvae. Proc. Natl. Acad. Sci. U. S. A.,

623 105, 1221-1225

624 Barry, D. G., Daly, N. L., Bokesch, H. R., Gustafson, K. R., & Craik, D. J. 2004. Solution structure

625 of the cyclotide palicourein: Implications for the development of a pharmaceutical framework.

626 Structure, 12, 85-94

627 Barry, D. G., Daly, N. L., Clark, R. J., Sando, L., & Craik, D. J. 2003. Linearization of a naturally

628 occurring circular protein maintains structure but eliminates hemolytic activity. Biochemistry,

629 42, 6688-6695

630 Bokesch, H. R., Pannell, L. K., Cochran, P. K., Sowder, R. C., 2nd, McKee, T. C., & Boyd, M. R.

631 2001. A novel anti-HIV macrocyclic peptide from Palicourea condensata. J. Nat. Prod., 64,

632 249-250

633 Burman, R., Gruber, C. W., Rizzardi, K., Herrmann, A., Craik, D. J., Gupta, M. P., & Göransson, U.

634 2010a. Cyclotide proteins and precursors from the genus Gloeospermum: Filling a blank spot

635 in the cyclotide map of Violaceae. Phytochemistry, 71, 13-20

636 Burman, R., Stromstedt, A. A., Malmsten, M., & Göransson, U. 2011. Cyclotide-membrane

637 interactions: Defining factors of membrane binding, depletion and disruption. Biochim.

638 Biophys. Acta, 1808, 2665-2673

28

639 Burman, R., Svedlund, E., Felth, J., Hassan, S., Herrmann, A., Clark, R. J., Craik, D. J., Bohlin, L.,

640 Claeson, P., Göransson, U., & Gullbo, J. 2010b. Evaluation of toxicity and antitumor activity

641 of cycloviolacin O2 in mice. Biopolymers, 94, 626-634

642 Camarero, J. A. 2017. Cyclotides, a versatile ultrastable micro-protein scaffold for biotechnological

643 applications. Bioorg. Med. Chem. Lett., 27, 5089-5099

644 Camarero, J. A., Kimura, R. H., Woo, Y. H., Shekhtman, A., & Cantor, J. 2007. Biosynthesis of a

645 fully functional cyclotide inside living bacterial cells. Chembiochem, 8, 1363-1366

646 Cao, P., Yang, Y., Uche, F. I., Hart, S. R., Li, W.-W., & Yuan, C. 2018. Coupling plant-derived

647 cyclotides to metal surfaces: An antibacterial and antibiofilm study. Int. J. Mol. Sci., 19, 793

648 Cascales, L., Henriques, S. T., Kerr, M. C., Huang, Y. H., Sweet, M. J., Daly, N. L., & Craik, D. J.

649 2011. Identification and characterization of a new family of cell-penetrating peptides: Cyclic

650 cell-penetrating peptides. J. Biol. Chem., 286, 36932-36943

651 Chen, B., Colgrave, M. L., Daly, N. L., Rosengren, K. J., Gustafson, K. R., & Craik, D. J. 2005.

652 Isolation and characterization of novel cyclotides from Viola hederaceae: Solution structure

653 and anti-HIV activity of vhl-1, a leaf-specific expressed cyclotide. J. Biol. Chem., 280, 22395-

654 22405

655 Cheneval, O., Schroeder, C. I., Durek, T., Walsh, P., Huang, Y. H., Liras, S., Price, D. A., & Craik,

656 D. J. 2014. Fmoc-based synthesis of disulfide-rich cyclic peptides. J. Org. Chem., 79, 5538-

657 5544

658 Chiche, L., Heitz, A., Gelly, J. C., Gracy, J., Chau, P. T., Ha, P. T., Hernandez, J. F., & Le-Nguyen,

659 D. 2004. Squash inhibitors: From structural motifs to macrocyclic knottins. Curr. Protein Pept.

660 Sci., 5, 341-349

661 Claeson, P., Göransson, U., Johansson, S., Luijendijk, T., & Bohlin, L. 1998. Fractionation protocol

662 for the isolation of polypeptides from plant biomass. J. Nat. Prod., 61, 77-81

663 Clark, R. J., & Craik, D. J. 2010. Native chemical ligation applied to the synthesis and bioengineering

664 of circular peptides and proteins. Biopolymers, 94, 414-422

665 Colgrave, M. L., & Craik, D. J. 2004. Thermal, chemical, and enzymatic stability of the cyclotide

666 kalata B1: The importance of the cyclic cystine knot. Biochemistry, 43, 5965-5975

29

667 Colgrave, M. L., Huang, Y. H., Craik, D. J., & Kotze, A. C. 2010. Cyclotide interactions with the

668 nematode external surface. Antimicrob. Agents Chemother., 54, 2160-2166

669 Colgrave, M. L., Kotze, A. C., Huang, Y. H., O'Grady, J., Simonsen, S. M., & Craik, D. J. 2008a.

670 Cyclotides: Natural, circular plant peptides that possess significant activity against

671 gastrointestinal nematode parasites of sheep. Biochemistry, 47, 5581-5589

672 Colgrave, M. L., Kotze, A. C., Ireland, D. C., Wang, C. K., & Craik, D. J. 2008b. The anthelmintic

673 activity of the cyclotides: Natural variants with enhanced activity. Chembiochem, 11, 1939-

674 1945

675 Colgrave, M. L., Kotze, A. C., Kopp, S., McCarthy, J. S., Coleman, G. T., & Craik, D. J. 2009.

676 Anthelmintic activity of cyclotides: In vitro studies with canine and human hookworms. Acta

677 Trop., 109, 163-166

678 Conlan, B. F., Colgrave, M. L., Gillon, A. D., Guarino, R., Craik, D. J., & Anderson, M. A. 2012.

679 Insights into processing and cyclization events associated with biosynthesis of the cyclic

680 peptide kalata B1. J. Biol. Chem., 287, 28037-28046

681 Contreras, J., Elnagar, A. Y., Hamm-Alvarez, S. F., & Camarero, J. A. 2011. Cellular uptake of

682 cyclotide MCoTI-I follows multiple endocytic pathways. J. Control. Release, 155, 134-143

683 Cowper, B., Craik, D. J., & Macmillan, D. 2013. Making ends meet: Chemically mediated

684 circularization of recombinant proteins. Chembiochem, 14, 809-812

685 Craik, D. J. 2012. Host-defense activities of cyclotides. Toxins (Basel), 4, 139-156

686 Craik, D. J. 2013. Joseph Rudinger Memorial Lecture: Discovery and applications of cyclotides. J.

687 Pept. Sci., 19, 393-407

688 Craik, D. J., Daly, N. L., Bond, T., & Waine, C. 1999. Plant cyclotides: A unique family of cyclic and

689 knotted proteins that defines the cyclic cystine knot structural motif. J. Mol. Biol., 294, 1327-

690 1336

691 Craik, D. J., Daly, N. L., & Waine, C. 2001. The cystine knot motif in toxins and implications for

692 drug design. Toxicon, 39, 43-60

693 Craik, D. J., & Du, J. 2017. Cyclotides as drug design scaffolds. Curr. Opin. Chem. Biol., 38, 8-16

30

694 Craik, D. J., Lee, M.-H., Rehm, F. B., Tombling, B., Doffek, B., & Peacock, H. 2018. Ribosomally-

695 synthesised cyclic peptides from plants as drug leads and pharmaceutical scaffolds. Bioorg.

696 Med. Chem., 26, 2727-2737

697 Daly, N. L., Chen, B., Nguyencong, P., & Craik, D. J. 2010. Structure and activity of the leaf-specific

698 cyclotide vhl-2. Aust. J. Chem., 63, 771-778

699 Daly, N. L., Clark, R. J., Plan, M. R., & Craik, D. J. 2006. Kalata B8, a novel antiviral circular protein,

700 exhibits conformational flexibility in the cystine knot motif. Biochem. J., 393, 619-626

701 Daly, N. L., & Craik, D. J. 2000. Acyclic permutants of naturally occurring cyclic proteins.

702 Characterization of cystine knot and beta-sheet formation in the macrocyclic polypeptide

703 kalata B1. J. Biol. Chem., 275, 19068-19075

704 Daly, N. L., Gustafson, K. R., & Craik, D. J. 2004. The role of the cyclic peptide backbone in the

705 anti-HIV activity of the cyclotide kalata B1. FEBS Lett., 574, 69-72

706 Daly, N. L., Koltay, A., Gustafson, K. R., Boyd, M. R., Casas-Finet, J. R., & Craik, D. J. 1999a.

707 Solution structure by NMR of circulin A: a macrocyclic knotted peptide having anti-HIV

708 activity. J. Mol. Biol., 285, 333-345

709 Daly, N. L., Love, S., Alewood, P. F., & Craik, D. J. 1999b. Chemical synthesis and folding pathways

710 of large cyclic polypeptides: Studies of the cystine knot polypeptide kalata B1. Biochemistry,

711 38, 10606-10614

712 Daly, N. L., Rosengren, K. J., & Craik, D. J. 2009. Discovery, structure and biological activities of

713 cyclotides. Adv. Drug Delivery Rev., 61, 918-930

714 Daly, N. L., Rosengren, K. J., Henriques, S. T., & Craik, D. J. 2011. NMR and protein structure in

715 drug design: Application to cyclotides and conotoxins. Eur. Biophys. J., 40, 359-370

716 Dawson, P. E., Muir, T. W., Clark-Lewis, I., & Kent, S. B. 1994. Synthesis of proteins by native

717 chemical ligation. Science, 266, 776-779

718 de Veer, S. J., Weidmann, J., & Craik, D. J. 2017. Cyclotides as tools in chemical biology. Acc. Chem.

719 Res., 50, 1557-1565

720 Du, J. Q., Chan, L. Y., Poth, A. G., & Craik, D. J. 2019. Discovery and characterization of cyclic and

721 acyclic trypsin inhibitors from Momordica dioica. J. Nat. Prod., 82, 293-300

31

722 Felizmenio-Quimio, M. E., Daly, N. L., & Craik, D. J. 2001. Circular proteins in plants: Solution

723 structure of a novel macrocyclic trypsin inhibitor from Momordica cochinchinensis. J. Biol.

724 Chem., 276, 22875-22882

725 Fensterseifer, I. C. M., Silva, O. N., Malik, U., Ravipati, A. S., Novaes, N. R. F., Miranda, P. R. R.,

726 Rodrigues, E. A., Moreno, S. E., Craik, D. J., & Franco, O. L. 2015. Effects of cyclotides

727 against cutaneous infections caused by Staphylococcus aureus. Peptides, 63, 38-42

728 Gilding, E. K., Jackson, M. A., Poth, A. G., Henriques, S. T., Prentis, P. J., Mahatmanto, T., & Craik,

729 D. J. 2016. Gene coevolution and regulation lock cyclic plant defence peptides to their targets.

730 New. Phytol., 210, 717-730

731 Göransson, U., Burman, R., Gunasekera, S., Stromstedt, A. A., & Rosengren, K. J. 2012. Circular

732 proteins from plants and fungi. J. Biol. Chem., 287, 27001-27006

733 Göransson, U., Luijendijk, T., Johansson, S., Bohlin, L., & Claeson, P. 1999. Seven novel

734 macrocyclic polypeptides from Viola arvensis. J. Nat. Prod., 62, 283-286

735 Göransson, U., Sjogren, M., Svangård, E., Claeson, P., & Bohlin, L. 2004. Reversible antifouling

736 effect of the cyclotide cycloviolacin O2 against barnacles. J. Nat. Prod., 67, 1287-1290

737 Gould, A., & Camarero, J. A. 2017. Cyclotides: Overview and biotechnological applications.

738 Chembiochem, 18, 1350-1363

739 Gould, A., Ji, Y., Aboye, T. L., & Camarero, J. A. 2011. Cyclotides, a novel ultrastable polypeptide

740 scaffold for drug discovery. Curr. Pharm. Des., 17, 4294-4307

741 Gran, L. 1970. An oxytocic principle found in Oldenlandia affinis DC. Medd. Nor. Farm. Selsk., 12,

742 173-180

743 Gran, L. 1973a. On the effect of a polypeptide isolated from "Kalata-Kalata" (Oldenlandia affinis DC)

744 on the oestrogen dominated uterus. Acta Pharmacol. Toxicol. (Copenh.), 33, 400-408

745 Gran, L. 1973b. Oxytocic principles of Oldenlandia affinis. Lloydia, 36, 174-178

746 Gran, L., Sletten, K., & Skjeldal, L. 2008. Cyclic peptides from Oldenlandia affinis DC. Molecular

747 and biological properties. Chem. Biodivers., 5, 2014-2022

32

748 Greenwood, K. P., Daly, N. L., Brown, D. L., Stow, J. L., & Craik, D. J. 2007. The cyclic cystine

749 knot miniprotein MCoTI-II is internalized into cells by macropinocytosis. Int. J. Biochem.

750 Cell Biol., 39, 2252-2264

751 Gruber, C. W. 2010. Global cyclotide adventure: A journey dedicated to the discovery of circular

752 peptides from flowering plants. Biopolymers, 94, 565-572

753 Gruber, C. W., Elliott, A. G., Ireland, D. C., Delprete, P. G., Dessein, S., Göransson, U., Trabi, M.,

754 Wang, C. K., Kinghorn, A. B., Robbrecht, E., & Craik, D. J. 2008. Distribution and evolution

755 of circular miniproteins in flowering plants. Plant Cell, 20, 2471-2483

756 Gründemann, C., Koehbach, J., Huber, R., & Gruber, C. W. 2012. Do plant cyclotides have potential

757 as immunosuppressant peptides? J. Nat. Prod., 75, 167-174

758 Gründemann, C., Stenberg, K. G., & Gruber, C. W. 2019. T20K: An immunomodulatory cyclotide

759 on its way to the clinic. Int. J. Pept. Res. Ther., 25, 9-13

760 Gründemann, C., Thell, K., Lengen, K., Garcia-Kaufer, M., Huang, Y. H., Huber, R., Craik, D. J.,

761 Schabbauer, G., & Gruber, C. W. 2013. Cyclotides suppress human T-lymphocyte

762 proliferation by an interleukin 2-dependent mechanism. PLoS One, 8, e68016

763 Gunasekera, S., Aboye, T. L., Madian, W. A., El-Seedi, H. R., & Göransson, U. 2013. Making ends

764 meet: Microwave-accelerated synthesis of cyclic and disulfide rich proteins via in situ

765 thioesterification and native chemical ligation. Int. J. Pept. Res. Ther., 19, 43-54

766 Gustafson, K. R., Sowder II, R. C., Henderson, L. E., Parsons, I. C., Kashman, Y., Cardellina II, J.

767 H., McMahon, J. B., Buckheit, J. R. W., Pannell, L. K., & Boyd, M. R. 1994. Circulins A and

768 B: Novel HIV-inhibitory macrocyclic peptides from the tropical tree Chassalia parvifolia. J.

769 Am. Chem. Soc., 116, 9337-9338

770 Gustafson, K. R., Walton, L. K., Sowder, R. C., Jr., Johnson, D. G., Pannell, L. K., Cardellina, J. H.,

771 Jr., & Boyd, M. R. 2000. New circulin macrocyclic polypeptides from Chassalia parvifolia. J.

772 Nat. Prod., 63, 176-178

773 Hallock, Y. F., Sowder, R. C., 2nd, Pannell, L. K., Hughes, C. B., Johnson, D. G., Gulakowski, R.,

774 Cardellina, J. H., 2nd, & Boyd, M. R. 2000. Cycloviolins A-D, anti-HIV macrocyclic peptides

775 from Leonia cymosa. J. Org. Chem., 65, 124-128

33

776 Harris, K. S., Durek, T., Kaas, Q., Poth, A. G., Gilding, E. K., Conlan, B. F., Saska, I., Daly, N. L.,

777 van der Weerden, N. L., Craik, D. J., & Anderson, M. A. 2015. Efficient backbone cyclization

778 of linear peptides by a recombinant asparaginyl endopeptidase. Nat. Commun., 6, 10199

779 Hellinger, R., Koehbach, J., Puigpinós, A., Clark, R. J., Tarragó, T., Giralt, E., & Gruber, C. W. 2015.

780 Inhibition of human prolyl oligopeptidase activity by the cyclotide psysol 2 isolated from

781 Psychotria solitudinum. J. Nat. Prod., 78, 1073-1082

782 Henriques, S. T., & Craik, D. J. 2012. Importance of the cell membrane on the mechanism of action

783 of cyclotides. ACS Chem. Biol., 7, 626-636

784 Henriques, S. T., & Craik, D. J. 2017. Cyclotide structure and function: The role of membrane binding

785 and permeation. Biochemistry, 56, 669-682

786 Henriques, S. T., Huang, Y. H., Castanho, M. A., Bagatolli, L. A., Sonza, S., Tachedjian, G., Daly,

787 N. L., & Craik, D. J. 2012. Phosphatidylethanolamine binding is a conserved feature of

788 cyclotide-membrane interactions. J. Biol. Chem., 287, 33629-33643

789 Henriques, S. T., Huang, Y. H., Chaousis, S., Sani, M. A., Poth, A. G., Separovic, F., & Craik, D. J.

790 2015. The Prototypic Cyclotide Kalata B1 Has a Unique Mechanism of Entering Cells. Chem.

791 Biol., 22, 1087-1097

792 Henriques, S. T., Huang, Y. H., Chaousis, S., Wang, C. K., & Craik, D. J. 2014. Anticancer and toxic

793 properties of cyclotides are dependent on phosphatidylethanolamine phospholipid targeting.

794 Chembiochem, 15, 1956-1965

795 Henriques, S. T., Huang, Y. H., Rosengren, K. J., Franquelim, H. G., Carvalho, F. A., Johnson, A.,

796 Sonza, S., Tachedjian, G., Castanho, M. A., Daly, N. L., & Craik, D. J. 2011. Decoding the

797 membrane activity of the cyclotide kalata B1: The importance of phosphatidylethanolamine

798 phospholipids and lipid organization on hemolytic and anti-HIV activities. J. Biol. Chem., 286,

799 24231-24241

800 Hernandez, J. F., Gagnon, J., Chiche, L., Nguyen, T. M., Andrieu, J. P., Heitz, A., Trinh Hong, T.,

801 Pham, T. T., & Le Nguyen, D. 2000. Squash trypsin inhibitors from Momordica

802 cochinchinensis exhibit an atypical macrocyclic structure. Biochemistry, 39, 5722-5730

34

803 Herrmann, A., Burman, R., Mylne, J. S., Karlsson, G., Gullbo, J., Craik, D. J., Clark, R. J., &

804 Göransson, U. 2008. The alpine violet, Viola biflora, is a rich source of cyclotides with potent

805 cytotoxicity. Phytochemistry, 69, 939-952

806 Huang, Y. H., Colgrave, M. L., Clark, R. J., Kotze, A. C., & Craik, D. J. 2010. Lysine-scanning

807 mutagenesis reveals an amendable face of the cyclotide kalata B1 for the optimization of

808 nematocidal activity. J. Biol. Chem., 285, 10797-10805

809 Huang, Y. H., Colgrave, M. L., Daly, N. L., Keleshian, A., Martinac, B., & Craik, D. J. 2009. The

810 biological activity of the prototypic cyclotide kalata b1 is modulated by the formation of

811 multimeric pores. J. Biol. Chem., 284, 20699-20707

812 Ireland, D. C., Colgrave, M. L., & Craik, D. J. 2006a. A novel suite of cyclotides from Viola odorata:

813 Sequence variation and the implications for structure, function and stability. Biochem. J., 400,

814 1-12

815 Ireland, D. C., Colgrave, M. L., Nguyencong, P., Daly, N. L., & Craik, D. J. 2006b. Discovery and

816 characterization of a linear cyclotide from Viola odorata: Implications for the processing of

817 circular proteins. J. Mol. Biol., 357, 1522-1535

818 Ireland, D. C., Wang, C. K., Wilson, J. A., Gustafson, K. R., & Craik, D. J. 2008. Cyclotides as natural

819 anti-HIV agents. Biopolymers, 90, 51-60

820 Jackson, M. A., Gilding, E. K., Shafee, T., Harris, K. S., Kaas, Q., Poon, S., Yap, K., Jia, H., Guarino,

821 R., Chan, L. Y., Durek, T., Anderson, M. A., & Craik, D. J. 2018. Molecular basis for the

822 production of cyclic peptides by plant asparaginyl endopeptidases. Nat. Commun., 9, 2411

823 Jagadish, K., Gould, A., Borra, R., Majumder, S., Mushtaq, Z., Shekhtman, A., & Camarero, J. A.

824 2015. Recombinant expression and phenotypic screening of a bioactive cyclotide against

825 alpha-Synuclein-Induced cytotoxicity in Baker's Yeast. Angew. Chem. Int. Ed. Engl., 54,

826 8390-8394

827 Jennings, C., West, J., Waine, C., Craik, D., & Anderson, M. 2001. Biosynthesis and insecticidal

828 properties of plant cyclotides: The cyclic knotted proteins from Oldenlandia affinis. Proc. Natl.

829 Acad. Sci. U. S. A., 98, 10614-10619

35

830 Jennings, C. V., Rosengren, K. J., Daly, N. L., Plan, M., Stevens, J., Scanlon, M. J., Waine, C.,

831 Norman, D. G., Anderson, M. A., & Craik, D. J. 2005. Isolation, solution structure, and

832 insecticidal activity of kalata B2, a circular protein with a twist: Do Mobius strips exist in

833 nature? Biochemistry, 44, 851-860

834 Jia, X., Kwon, S., Wang, C. I., Huang, Y. H., Chan, L. Y., Tan, C. C., Rosengren, K. J., Mulvenna, J.

835 P., Schroeder, C. I., & Craik, D. J. 2014. Semienzymatic cyclization of disulfide-rich peptides

836 using Sortase A. J. Biol. Chem., 289, 6627-6638

837 Kamimori, H., Hall, K., Craik, D. J., & Aguilar, M. I. 2005. Studies on the membrane interactions of

838 the cyclotides kalata B1 and kalata B6 on model membrane systems by surface plasmon

839 resonance. Anal. Biochem., 337, 149-153

840 Kan, M.-W., & Craik, D. J. (2018). Trends in Cyclotide Research In Cyclic Peptides: From

841 Bioorganic Synthesis to Applications (pp. 302-339, chapter 14): The Royal Society of

842 Chemistry.

843 Kimura, R. H., Tran, A. T., & Camarero, J. A. 2006. Biosynthesis of the cyclotide Kalata B1 by using

844 protein splicing. Angew. Chem. Int. Ed. Engl., 45, 973-976

845 Koehbach, J., Attah, A. F., Berger, A., Hellinger, R., Kutchan, T. M., Carpenter, E. J., Rolf, M.,

846 Sonibare, M. A., Moody, J. O., Wong, G. K., Dessein, S., Greger, H., & Gruber, C. W. 2013a.

847 Cyclotide discovery in Gentianales revisited-identification and characterization of cyclic

848 cystine-knot peptides and their phylogenetic distribution in Rubiaceae plants. Biopolymers,

849 100, 438-452

850 Koehbach, J., O'Brien, M., Muttenthaler, M., Miazzo, M., Akcan, M., Elliott, A. G., Daly, N. L.,

851 Harvey, P. J., Arrowsmith, S., Gunasekera, S., Smith, T. J., Wray, S., Göransson, U., Dawson,

852 P. E., Craik, D. J., Freissmuth, M., & Gruber, C. W. 2013b. Oxytocic plant cyclotides as

853 templates for peptide G protein-coupled receptor ligand design. Proc. Natl. Acad. Sci. U. S.

854 A., 110, 21183-21188

855 Koltay, A., Daly, N. L., Gustafson, K. R., & Craik, D. J. 2005. Structure of circulin B and implications

856 for antimicrobial activity of the cyclotides. Int. J. Pept. Res. Ther., 11, 99-106

36

857 Kwon, S., Duarte, J. N., Li, Z., Ling, J. J., Cheneval, O., Durek, T., Schroeder, C. I., Craik, D. J., &

858 Ploegh, H. L. 2018. Targeted delivery of cyclotides via conjugation to a nanobody. ACS Chem.

859 Biol., 13, 2973-2980

860 Lindholm, P., Göransson, U., Johansson, S., Claeson, P., Gullbo, J., Larsson, R., Bohlin, L., &

861 Backlund, A. 2002. Cyclotides: A novel type of cytotoxic agents. Mol. Cancer Ther., 1, 365-

862 369

863 Malik, S. Z., Linkevicius, M., Göransson, U., & Andersson, D. I. 2017. Resistance to the cyclotide

864 cycloviolacin O2 in Salmonella enterica caused by different mutations that often confer cross-

865 resistance or collateral sensitivity to other antimicrobial peptides. Antimicrob. Agents

866 Chemother., 61, e00684-00617

867 Mulvenna, J. P., Sando, L., & Craik, D. J. 2005. Processing of a 22 kDa precursor protein to produce

868 the circular protein tricyclon A. Structure, 13, 691-701

869 Mylne, J. S., Chan, L. Y., Chanson, A. H., Daly, N. L., Schaefer, H., Bailey, T. L., Nguyencong, P.,

870 Cascales, L., & Craik, D. J. 2012. Cyclic peptides arising by evolutionary parallelism via

871 asparaginyl-endopeptidase–mediated biosynthesis. Plant Cell, 24, 2765-2778

872 Nguyen, G. K., Lian, Y., Pang, E. W., Nguyen, P. Q., Tran, T. D., & Tam, J. P. 2013. Discovery of

873 linear cyclotides in monocot plant Panicum laxum of Poaceae family provides new insights

874 into evolution and distribution of cyclotides in plants. J. Biol. Chem., 288, 3370-3380

875 Nguyen, G. K., Wang, S., Qiu, Y., Hemu, X., Lian, Y., & Tam, J. P. 2014. Butelase 1 is an Asx-

876 specific ligase enabling peptide macrocyclization and synthesis. Nat. Chem. Biol., 10, 732-

877 738

878 Noonan, J., Williams, W., & Shan, X. 2017. Investigation of antimicrobial peptide genes associated

879 with fungus and insect resistance in maize. Int. J. Mol. Sci., 18, E1938

880 Nourse, A., Trabi, M., Daly, N. L., & Craik, D. J. 2004. A comparison of the self-association behavior

881 of the plant cyclotides kalata B1 and kalata B2 via analytical ultracentrifugation. J. Biol.

882 Chem., 279, 562-570

883 Oren, Z., & Shai, Y. 1998. Mode of action of linear amphipathic alpha-helical antimicrobial peptides.

884 Biopolymers, 47, 451-463

37

885 Parsley, N. C., Kirkpatrick, C. L., Crittenden, C. M., Rad, J. G., Hoskin, D. W., Brodbelt, J. S., &

886 Hicks, L. M. 2018. PepSAVI-MS reveals anticancer and antifungal cycloviolacins in Viola

887 odorata. Phytochemistry, 152, 61-70

888 Pinto, M. E. F., Najas, J. Z. G., Magalhães, L. G., Bobey, A. F., Mendonça, J. N., Lopes, N. P., Leme,

889 F. M., Teixeira, S. P., Trovó, M., Andricopulo, A. D., Koehbach, J., Gruber, C. W., Cilli, E.

890 M., & Bolzani, V. S. 2018. Inhibition of breast cancer cell migration by cyclotides isolated

891 from Pombalia Calceolaria. J. Nat. Prod., 81, 1203-1208

892 Plan, M. R., Göransson, U., Clark, R. J., Daly, N. L., Colgrave, M. L., & Craik, D. J. 2007. The

893 cyclotide fingerprint in oldenlandia affinis: Elucidation of chemically modified, linear and

894 novel macrocyclic peptides. Chembiochem, 8, 1001-1011

895 Plan, M. R., Rosengren, K. J., Sando, L., Daly, N. L., & Craik, D. J. 2010. Structural and biochemical

896 characteristics of the cyclotide kalata B5 from Oldenlandia affinis. Biopolymers, 94, 647-658

897 Plan, M. R., Saska, I., Cagauan, A. G., & Craik, D. J. 2008. Backbone cyclised peptides from plants

898 show molluscicidal activity against the rice pest Pomacea canaliculata (golden apple snail). J.

899 Agric. Food Chem., 56, 5237-5241

900 Poth, A. G., Chan, L. Y., & Craik, D. J. 2013. Cyclotides as grafting frameworks for protein

901 engineering and drug design applications. Biopolymers, 100, 480-491

902 Poth, A. G., Colgrave, M. L., Lyons, R. E., Daly, N. L., & Craik, D. J. 2011a. Discovery of an unusual

903 biosynthetic origin for circular proteins in legumes. Proc. Natl. Acad. Sci. U. S. A., 108,

904 10127-10132

905 Poth, A. G., Colgrave, M. L., Philip, R., Kerenga, B., Daly, N. L., Anderson, M. A., & Craik, D. J.

906 2011b. Discovery of cyclotides in the fabaceae plant family provides new insights into the

907 cyclization, evolution, and distribution of circular proteins. ACS Chem. Biol., 6, 345-355

908 Poth, A. G., Mylne, J. S., Grassl, J., Lyons, R. E., Millar, A. H., Colgrave, M. L., & Craik, D. J. 2012.

909 Cyclotides associate with leaf vasculature and are the products of a novel precursor in petunia

910 (Solanaceae). J. Biol. Chem., 287, 27033-27046

38

911 Pranting, M., Loov, C., Burman, R., Göransson, U., & Andersson, D. I. 2010. The cyclotide

912 cycloviolacin O2 from Viola odorata has potent bactericidal activity against Gram-negative

913 bacteria. J. Antimicrob. Chemother., 65, 1964-1971

914 Qu, H., Smithies, B. J., Durek, T., & Craik, D. J. 2017. Synthesis and protein engineering applications

915 of cyclotides. Aust. J. Chem., 70, 152-161

916 Quimbar, P., Malik, U., Sommerhoff, C. P., Kaas, Q., Chan, L. Y., Huang, Y. H., Grundhuber, M.,

917 Dunse, K., Craik, D. J., Anderson, M. A., & Daly, N. L. 2013. High-affinity cyclic peptide

918 matriptase inhibitors. J. Biol. Chem., 288, 13885-13896

919 Rehm, F. B. H., Jackson, M. A., De Geyter, E., Yap, K., Gilding, E. K., Durek, T., & Craik, D. J.

920 2019. Papain-like cysteine proteases prepare plant cyclic peptide precursors for cyclization.

921 Proc. Natl. Acad. Sci. U. S. A., 116, 7831-7836

922 Rosengren, K. J., Daly, N. L., Harvey, P. J., & Craik, D. J. 2013. The self-association of the cyclotide

923 kalata B2 in solution is guided by hydrophobic interactions. Biopolymers, 100, 453-460

924 Rosengren, K. J., Daly, N. L., Plan, M. R., Waine, C., & Craik, D. J. 2003. Twists, knots, and rings

925 in proteins. Structural definition of the cyclotide framework. J. Biol. Chem., 278, 8606-8616

926 Saether, O., Craik, D. J., Campbell, I. D., Sletten, K., Juul, J., & Norman, D. G. 1995. Elucidation of

927 the primary and three-dimensional structure of the uterotonic polypeptide kalata B1.

928 Biochemistry, 34, 4147-4158

929 Saska, I., Gillon, A. D., Hatsugai, N., Dietzgen, R. G., Hara-Nishimura, I., Anderson, M. A., & Craik,

930 D. J. 2007. An asparaginyl endopeptidase mediates in vivo protein backbone cyclization. J.

931 Biol. Chem., 282, 29721-29728

932 Schöpke, T., Hasan Agha, M. I., Kraft, R., Otto, A., & Hiller, K. 1993. Hämolytisch aktive

933 komponenten aus Viola tricolor L. und Viola arvensis Murray. Sci. Pharm., 61, 145-153

934 Shenkarev, Z. O., Nadezhdin, K. D., Lyukmanova, E. N., Sobol, V. A., Skjeldal, L., & Arseniev, A.

935 S. 2008. Divalent cation coordination and mode of membrane interaction in cyclotides: NMR

936 spatial structure of ternary complex Kalata B7/Mn2+/DPC micelle. J. Inorg. Biochem., 102,

937 1246-1256

39

938 Shenkarev, Z. O., Nadezhdin, K. D., Sobol, V. A., Sobol, A. G., Skjeldal, L., & Arseniev, A. S. 2006.

939 Conformation and mode of membrane interaction in cyclotides. Spatial structure of kalata B1

940 bound to a dodecylphosphocholine micelle. FEBS J., 273, 2658-2672

941 Simonsen, S. M., Sando, L., Rosengren, K. J., Wang, C. K., Colgrave, M. L., Daly, N. L., & Craik,

942 D. J. 2008. Alanine scanning mutagenesis of the prototypic cyclotide reveals a cluster of

943 residues essential for bioactivity. J. Biol. Chem., 283, 9805-9813

944 Slazak, B., Kapusta, M., Strömstedt, A. A., Słomka, A., Krychowiak, M., Shariatgorji, M., Andrén,

945 P. E., Bohdanowicz, J., Kuta, E., & Göransson, U. 2018. How does the sweet violet (Viola

946 odorata L.) fight pathogens and dests - cyclotides as a comprehensive plant host defense

947 system. Front. Recent Dev. Plant Sci., 9, 1296

948 Svangård, E., Burman, R., Gunasekera, S., Lovborg, H., Gullbo, J., & Göransson, U. 2007.

949 Mechanism of action of cytotoxic cyclotides: Cycloviolacin O2 disrupts lipid membranes. J.

950 Nat. Prod., 70, 643-647

951 Svangård, E., Göransson, U., Hocaoglu, Z., Gullbo, J., Larsson, R., Claeson, P., & Bohlin, L. 2004.

952 Cytotoxic cyclotides from Viola tricolor. J. Nat. Prod., 67, 144-147

953 Taichi, M., Hemu, X., Qiu, Y., & Tam, J. P. 2013. A thioethylalkylamido (TEA) thioester surrogate

954 in the synthesis of a cyclic peptide via a tandem acyl shift. Org. Lett., 15, 2620-2623

955 Tam, J. P., & Lu, Y. A. 1998. A biomimetic strategy in the synthesis and fragmentation of cyclic

956 protein. Protein Sci., 7, 1583-1592

957 Tam, J. P., Lu, Y. A., Yang, J. L., & Chiu, K. W. 1999. An unusual structural motif of antimicrobial

958 peptides containing end-to-end macrocycle and cystine-knot disulfides. Proc. Natl. Acad. Sci.

959 U. S. A., 96, 8913-8918

960 Tang, J., Wang, C. K., Pan, X., Yan, H., Zeng, G., Xu, W., He, W., Daly, N. L., Craik, D. J., & Tan,

961 N. 2010. Isolation and characterization of cytotoxic cyclotides from Viola tricolor. Peptides,

962 31, 1434-1440

963 Trabi, M., & Craik, D. J. 2004. Tissue-specific expression of head-to-tail cyclized miniproteins in

964 Violaceae and structure determination of the root cyclotide Viola hederacea root cyclotide1.

965 Plant Cell, 16, 2204-2216

40

966 Wang, C. K., Clark, R. J., Harvey, P. J., Johan Rosengren, K., Cemazar, M., & Craik, D. J. 2011. The

967 role of conserved Glu residue on cyclotide stability and activity: a structural and functional

968 study of kalata B12, a naturally occurring Glu to Asp mutant. Biochemistry, 50, 4077-4086

969 Wang, C. K., Colgrave, M. L., Gustafson, K. R., Ireland, D. C., Göransson, U., & Craik, D. J. 2008a.

970 Anti-HIV cyclotides from the Chinese medicinal herb Viola yedoensis. J. Nat. Prod., 71, 47-

971 52

972 Wang, C. K., Colgrave, M. L., Ireland, D. C., Kaas, Q., & Craik, D. J. 2009a. Despite a conserved

973 cystine knot motif, different cyclotides have different membrane binding modes. Biophys. J.,

974 97, 1471-1481

975 Wang, C. K., & Craik, D. J. 2018. Designing macrocyclic disulfide-rich peptides for biotechnological

976 applications. Nat. Chem. Biol., 14, 417-427

977 Wang, C. K., Hu, S. H., Martin, J. L., Sjogren, T., Hajdu, J., Bohlin, L., Claeson, P., Göransson, U.,

978 Rosengren, K. J., Tang, J., Tan, N. H., & Craik, D. J. 2009b. Combined X-ray and NMR

979 analysis of the stability of the cyclotide cystine knot fold that underpins its insecticidal activity

980 and potential use as a drug scaffold. J. Biol. Chem., 284, 10672-10683

981 Wang, C. K., Kaas, Q., Chiche, L., & Craik, D. J. 2008b. CyBase: A database of cyclic protein

982 sequences and structures, with applications in protein discovery and engineering. Nucleic

983 Acids Res., 36, D206-210

984 Wang, C. K., King, G. J., Northfield, S. E., Ojeda, P. G., & Craik, D. J. 2014. Racemic and quasi-

985 racemic X-ray structures of cyclic disulfide-rich peptide drug scaffolds. Angew. Chem. Int.

986 Ed. Engl., 53, 11236-11241

987 Wang, C. K., Wacklin, H. P., & Craik, D. J. 2012. Cyclotides insert into lipid bilayers to form

988 membrane pores and destabilize the membrane through hydrophobic and

989 phosphoethanolamine-specific interactions. J. Biol. Chem., 287, 43884-43898

990 Weidmann, J., & Craik, D. J. 2016. Discovery, structure, function, and applications of cyclotides:

991 circular proteins from plants. J. Exp. Bot., 67, 4801-4812

41

992 Witherup, K. M., Bogusky, M. J., Anderson, P. S., Ramjit, H., Ransom, R. W., Wood, T., & Sardana,

993 M. 1994. Cyclopsychotride A, a biologically active, 31-residue cyclic peptide isolated from

994 Psychotria longipes. J. Nat. Prod., 57, 1619-1625

995 Zenoni, S., D’Agostino, N., Tornielli, G. B., Quattrocchio, F., Chiusano, M. L., Koes, R., Zethof, J.,

996 Guzzo, F., Delledonne, M., Frusciante, L., Gerats, T., & Pezzotti, M. 2011. Revealing

997 impaired pathways in the an11 mutant by high-throughput characterization of Petunia axillaris

998 and Petunia inflata transcriptomes. The Plant Journal, 68, 11-27

999

42