digital.csic.esdigital.csic.es/bitstream/10261/137065/1/biotechnology... · web viewstructural...

52
Structural traits and catalytic versatility of the lipases from the Candida rugosa-like family: A review Jorge Barriuso, María Eugenia Vaquero, Alicia Prieto*, and Mª Jesús Martínez* Centro de Investigaciones Biológicas, Consejo Superior de Investigaciones Científicas, Ramiro de Maeztu 9, 28040 Madrid, Spain *Corresponding authors: María Jesús Martínez ([email protected] ); Alicia Prieto ([email protected]) Abstract Lipases and sterol esterases are enzymes with broad biotechnological applications, which catalyze the hydrolysis or synthesis of long-chain acylglycerols and sterol esters, respectively. In this paper, we review the current knowledge on the so-called Candida rugosa-like family of enzymes, whose members display in most cases affinity against the two substrates mentioned above. The family includes proteins with the α/β- hydrolase folding, sharing conserved motifs in their sequences, and common structural features. We will go through their production and purification, relate their described structures and catalytic activity, and discuss the influence of the hydrophobic character of these lipases on their aggregation 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31

Upload: vohanh

Post on 14-Apr-2018

216 views

Category:

Documents


0 download

TRANSCRIPT

Structural traits and catalytic versatility of the lipases from the Candida rugosa-like

family: A review

Jorge Barriuso, María Eugenia Vaquero, Alicia Prieto*, and Mª Jesús Martínez*

Centro de Investigaciones Biológicas, Consejo Superior de Investigaciones Científicas, Ramiro

de Maeztu 9, 28040 Madrid, Spain

*Corresponding authors: María Jesús Martínez ([email protected]); Alicia Prieto

([email protected])

Abstract

Lipases and sterol esterases are enzymes with broad biotechnological applications,

which catalyze the hydrolysis or synthesis of long-chain acylglycerols and sterol esters,

respectively. In this paper, we review the current knowledge on the so-called Candida rugosa-

like family of enzymes, whose members display in most cases affinity against the two substrates

mentioned above. The family includes proteins with the α/β-hydrolase folding, sharing

conserved motifs in their sequences, and common structural features. We will go through their

production and purification, relate their described structures and catalytic activity, and discuss

the influence of the hydrophobic character of these lipases on their aggregation state and

activity. On the basis of the few crystal structures available, the role of each of the functional

areas in catalysis will be analyzed. Considering the particular characteristics of this group, we

propose their classification as “Versatile Lipases” (EC 3.1.1.x).

Key words: Lipase, sterol-esterase, Candida rugosa, biocatalysts, hydrophobic enzymes.

1

2

3

4

5

6

7

8

9

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

1. Introduction

Lipases, also known as triacylglycerol lipases (EC. 3.1.1.3), act on ester bonds of

several compounds, with acylglycerols as their natural substrates. These enzymes catalyze the

hydrolysis of triglycerides to produce free fatty acids, diglycerides and/or monoglycerides under

aqueous conditions, but they can also carry out synthesis reactions, such as esterification and

transesterification, in the presence of organic solvents (Houde et al., 2004; Reis et al., 2009).

On the other hand, sterol esterases (EC 3.1.1.13) are defined as enzymes that hydrolyze

sterol esters releasing free sterols and fatty acids in aqueous media, being also able to perform

synthesis reactions in the presence of organic solvents (Barba Cedillo et al., 2013; Morinaga et

al., 2011). Most of the known sterol esterases have been reported to have both lipase and sterol

esterase activity (Calero-Rueda et al., 2002; Maeda et al., 2008; Vaquero et al., 2016).

Lipases and sterol esterases are carboxylic ester hydrolases (EC 3.1.1) that share the

α/β-hydrolase fold (Grochulski et al., 1993; Holmquist, 2000; Nardini and Dijkstra, 1999).

Hence, their catalytic machinery includes the residues of the catalytic triad (serine, histidine and

aspartic or glutamic acid) and the oxyanion hole, a pocket in the active site involved in

catalysis. They compose a very diverse group of ubiquitous enzymes in nature and are

represented from microbes to plants and animals. Nevertheless, bacterial and fungal lipases are

of special interest as they are easily produced and applicable for industrial processes due to their

versatility and stability to harsh conditions (Gupta et al., 2015; Jaeger and Eggert, 2002; Schmid

and Verger, 1998; Singh and Mukhopadhyay, 2012).

These enzymes can perform a variety of reactions, have wide substrate specificity and

good selectivity, and are regio- and stereoselective. The most important lipases from a

commercial point of view belong to yeasts, such as Candida rugosa (synonym Candida

cylindracea) and Candida antarctica, or filamentous fungi, such as Aspergillus niger, Humicola

lanuginosa, Mucor miehei, and Rhizopus species (Bornscheuer, 2002; Domínguez de María et

al., 2006; Gupta et al., 2015; Hasan et al., 2006; Singh and Mukhopadhyay, 2012). Among

them, some of the characterized lipase isoenzymes from C. rugosa show also sterol esterase

activity, as those from other microorganisms, still not produced at commercial level, which we

will see across this review.

Esterol esterases and lipases are included in ESTHER

(http://bioweb.ensam.inra.fr/esther), a broad database that collects very complete information on

the members of the α/β-hydrolase fold superfamily (Lenfant et al., 2013received). In addition, a

sequence-based comparatively simplified version is available in the Lipase Engineering

Database (LED) (http://www.led.uni-stuttgart.de/). LED lists sequences of all the available

microbial enzymes with lipase activity, putative or not, and provides links to 22 published lipase

structures. This database serves as a bioinformatics tool for the systematic analysis of sequence,

structure and function of diverse lipases, and for designing variants with optimized properties

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

66

67

68

(Fischer and Pleiss, 2003). Based on different characteristics, such as the presence of specific

conserved motifs in their amino acidic sequence, microbial enzymes with lipase activity are

grouped under several classes or subclasses in these databases. Yeast and fungal lipases fall into

five different subclasses: Yarrowia lipolytica-like lipase, C. rugosa-like lipase, filamentous

fungi lipases, C. antarctica lipase B-like and C. antarctica lipase A-like. Moreover, on the basis

of their sequence, structure, and function, these enzymes are classified in LED into the GX,

GGGX and Y classes.

In this review we will focus on a very versatile group of enzymes that generally show activity

towards acylglycerols and sterol esters, the so-called C. rugosa-like or abH03.01 lipase family.

Only a few members of this group have been characterized, although recent findings have

shown their potential for engineering future industrial applications. The characterized proteins

belong to ascomycetes and basidiomycetes, and little is known about their phylogenetic

affiliation and evolution (Barriuso et al., 2013). According to LED, the 336 protein sequences

currently assigned to this family belong to the class GGGX. However, most of them correspond

to hypothetical proteins and, according to this sequence-based classification, some enzymes can

be wrongly ascribed to this group. This was the case of the hypothetical protein EstA from A.

niger that, once expressed, showed to have structural characteristics and substrate preferences

different to those characteristic of the C. rugosa-like family (Bourne et al., 2004). The enzymes

described within this group are glycoproteins with a common overall structure. Their active

sites are hidden under a mobile region denominated lid or flap that, in a lipid-water interface or

in the presence of substrates or inhibitors, rearranges its position leaving an open gate to the

active center. Then, the position of the lid marks the difference between the open (active) or

closed (inactive) forms of these proteins. Other interesting properties of these catalysts as their

temperature and pH stabilities (T50= 40–60 °C, and pH 4-10), or their stereo- and

regioselectivity (Colton et al., 2011; Lee et al., 2007; Palocci et al., 2007; Vaquero et al.,

2015a), make them very attractive, and the current or proposed application of these catalysts

affect a wide range of industrial sectors, such as biofuels, oleochemical, food, detergents,

cosmetics, pharmaceutical, textile and paper industry (Hasan et al., 2006; Houde et al., 2004;

Reetz, 2002; Singh and Mukhopadhyay, 2012). Excellent reviews covering the biotechnological

applications of lipases, including the best known enzymes from the C. rugosa-like family, have

been recently published (Gupta et al., 2015; Stergiou et al., 2013). Here, we will revise their

structural and catalytic properties, their production and characterization in different hosts, and

the applicability of tools as genome mining for in silico search of novel catalysts or protein

engineering for tailoring enzyme activity.

2. Ecophysiological role

69

70

71

72

73

74

75

76

77

78

79

80

81

82

83

84

85

86

87

88

89

90

91

92

93

94

95

96

97

98

99

100

101

102

103

104

Fungi producing C. rugosa-like enzymes share a plant-associated habitat, and most of

them have been isolated from natural soils where plant material is frequent (Tomizuka et al.,

1966). In these cases, the ecological role of the extracellular lipases can be related to the attack

of esters from the epicuticular waxes and cuticle (Juniper and Jeffree 1983), improving the

accessibility to cell-wall polysaccharides that can serve as carbon sources for plant-associated

fungi. In pathogenic species these enzymes may act also on such components, located in the

surface of leaves, facilitating fungal colonization (de Vries et al., 1997; Doss, 1999).

The only studied lipase of this family from a basidiomycete is produced by Pleurotus

sapidus, a white-rot ligninolytic fungus (Zorn et al., 2005), while several have been

characterized form saprophytic ascomycetes, isolated from forest, agricultural, or composting

soils. Among them, we can mention C. rugosa, a non-sporogenic imperfect hemiascomycete, a

number of filamentous fungi, as the thermophilic Melanocarpus albomyces (Kontkanen et al.,

2006a), isolates from Trichoderma (Schuster and Schmoll, 2010received) or A. niger (Hu et al.,

2011), and the dimorphic ascomycetous yeast-like fungi Geotrichum candidum and Ophiostoma

piceae (Yan et al., 2007). The habitat of G. candidum is often associated with all kinds of soft

plant tissues, but it is ubiquitous in soils and pupal galleries of bark beetles (de Hoog and Smith,

2004). The environment of the wood-staining fungus O. piceae is much more restricted. This

species disseminates its spores using bark beetles as vectors and lives as saprobe in the

superficial layers of conifers’ sapwood. There, it metabolizes wood lipids, releasing hydrolysis

products as glycerol that seem to play an important role in the formation of the dark pigments

responsible for the “blue stain” (Eagen et al., 1997), which causes severe losses to the forestry

industry (Calero-Rueda et al., 2002; Haridas et al., 2013). Some extracellular fungal lipases are

also gaining attention for their potential role as virulence factors in relation to colonization,

adhesion, biofilm formation and pathogenesis (Gupta et al., 2015). Plant pathogens as Fusarium

solani are also producers of these enzymes (Vaquero et al., 2015a), and the causal agent of the

bunch rot disease of grapes, Botritys cinerea, releases lipases during the early phases of

parasitism induced by the components of the cuticle, being capable of entering its host directly

through an undamaged plant cuticle (Comménil et al., 1999).

Many other hypothetical proteins from this family have been described in the genomes

of fungi usually living as vegetal saprobes or parasites, such as ascomycetes from the genera

Alternaria, Pyrenophora and Neurospora, or basidiomycetes belonging to Postia, Laccaria and

Puccinia, although these enzymes have not yet been produced and characterized (Barriuso et al.,

2013; Fischer and Pleiss, 2003).

3. Production, purification and biochemical characterization of the C. rugosa-like enzymes

Due to their wide applicability, enzymes from this family have been produced in a

variety of hosts, from their native producers to recombinant bacteria, yeasts or filamentous

105

106

107

108

109

110

111

112

113

114

115

116

117

118

119

120

121

122

123

124

125

126

127

128

129

130

131

132

133

134

135

136

137

138

139

140

141

fungi. As explained in the following sections, it has been shown that the expression of a given

enzyme in different systems may induce changes in its properties (Barba Cedillo et al., 2012;

Kontkanen et al., 2006b; Vaquero et al., 2015b) and then the choice of the appropriate host is

not a minor decision. Once the functional protein has been successfully released to the

extracellular medium it must be purified, although the existence of several lipase isoforms with

similar chemical characteristics in some of the natural producers complicates their isolation as

single molecular species and generally requires a combination of techniques. In these cases, the

expression of recombinant proteins can make protein purification much easier.

Preparation of pure protein samples is generally accomplished using ultrafiltration,

precipitation and/or solubilization protocols, followed by one or several chromatographic

strategies as ion-exchange, hydrophobic interaction or size-exclusion chromatography (SEC). It

is important to keep in mind that the hydrophobic character of these enzymes facilitates the

formation of protein aggregates in solution and that in many occasions the C. rugosa-like

enzymes purified by SEC may correspond to aggregated forms, as well as the molecular mass

calculated from this technique.

3.1. The C. rugosa isoenzymes

The best known members of this family are some of the C. rugosa isoenzymes

(hereinafter, CRL), which have been profusely investigated since their discovery and even

today. This yeast secretes a variety of closely related enzymes (Benjamin and Pandey, 1998;

Ferrer et al., 2001; Lotti et al., 1994), commercialized as lipases or sterol esterases depending on

their catalytic efficiency. Nevertheless, comparison of data on the biochemical properties of

these isoenzymes, reported from different research groups, is difficult due to the use of different

methodologies, substrates, and even solvents for substrates solubilization, which is not a trivial

parameter for its influence on the interface where the enzyme exerts its action (Domínguez de

María et al., 2006; Pernas et al., 2009).

Tomizuka et al. (1966) described for the first time the isolation and purification of an

extracellular lipase from C. rugosa as a hydrophobic glycoprotein of 62 kDa with 4.2% neutral

sugars. The subsequent research of several groups showed inconsistency about the properties of

these enzymes and it was soon corroborated that the C. rugosa lipase commercialized by Sigma

and Amano consisted of a mixture of proteins with different catalytic properties. Veeraragavan

and Gibbs (1989) separated two lipases (I and II) and Rúa et al. (1993) demonstrated the

presence of isoenzymes A and B (currently known as isoforms CRL3 and CRL1, respectively),

which were later separated by hydrophobic interaction chromatography as a result of the higher

hydrophobicity of CRL3 against CRL1 (Rúa, 1994). Seven or more genes codify for different

lipase isoforms (Brocca et al., 1995), and five of the proteins (CRL1-CRL5) have been

characterized (Lotti and Alberghina, 1996; Lotti et al., 1994) . The expression levels of these

142

143

144

145

146

147

148

149

150

151

152

153

154

155

156

157

158

159

160

161

162

163

164

165

166

167

168

169

170

171

172

173

174

175

176

177

178

genes was analyzed measuring the abundance of lip mRNAs in non-induced C. rugosa cultures,

finding that it was in the order lip1, lip3, lip2, lip5, and lip4 (Lee et al., 1999). The expression of

lip1 and lip3 is high and constitutive, whereas for the remaining isoenzymes it depends on the

culture medium and conditions. Therefore, the expression profile of individual genes could be

modulated adding inducers such as olive oil or Tween 20 (Ferrer et al., 2001; Lee et al., 1999;

Pernas et al., 2000). Commercial cocktails of CRLs containing different proportions of each

isoform may have been obtained from industrial strains as a result of mutations or using

different substrates (Ferrer et al., 2001; López et al., 2004) and can be suitable for diverse

biotechnological applications. For instance, CRL1 is the major isoenzyme in the crude

commercialized by Sigma as lipase, whereas CRL3 is the principal isoenzyme in the crude

commercialized by Roche as cholesterol esterase.

For heterologous production, the use of universal hosts to express the native genes of C.

rugosa isoforms as functional lipases is unfeasible because this organism uses the non-universal

codon CUG-Leu for serine (Kawaguchi et al., 1989). This fact forced to resort to alternative

approaches. One of them, proposed by Mileto et al. (1998), consisted on using Candida maltosa

as host for its similar codon usage, but the recombinant protein secreted was heavily hyper

glycosylated. Brocca et al. (1998) carried out the total synthesis of a codon-optimized

nucleotidic sequence of the lip1 gene, and its functional over expression in Saccharomyces

cerevisiae and Pichia pastoris. Low expression levels were achieved in the first host, but not in

P. pastoris, that secreted a recombinant CRL1 with similar physico-chemical properties and

substrate specificity toward triglycerides of different chain length to the native CRL1. From this

finding, many groups have reported the heterologous expression of C. rugosa isoforms in P.

pastoris using the same strategy (Chang et al., 2006a; Chang et al., 2006b; Chang et al., 2005;

Ferrer et al., 2009; Lee et al., 2002; Lee et al., 2007; Lee et al., 2011; Tang et al., 2001; Yen et

al., 2010; Zhao et al., 2008). On the other hand, Tang et al. (2000) described the functional

expression of the lip4 gene in Escherichia coli, which synthesizes non-glycosylated protein,

after converting 19 CUG codons into a universal serine codon. The enzymatic activity of this

protein was similar to that of the glycosylated lipase from P. pastoris, although the CRL4 from

the yeast had higher thermal stability.

Isoenzymes CRL1 to CRL5 were separated from the crudes of C. rugosa by means of

several chromatographic techniques and characterized. They are glycoproteins with high

sequence identity (77-88%), molecular masses around 57-61 kDa, and isoelectric points in the

range of 4.5-5.7. Despite these similarities, their stability to temperature and pH, and their

substrate specificity are different (López et al., 2004; Mancheño et al., 2003). SEC analysis

showed that, in all cases, a proportion of the proteins’ population forms aggregates in solution,

whose molecular mass was estimated to be over 200 kDa even in the presence of detergents

(Pernas et al., 2000; Pernas et al., 2009). Nevertheless, among the CRLs currently characterized

179

180

181

182

183

184

185

186

187

188

189

190

191

192

193

194

195

196

197

198

199

200

201

202

203

204

205

206

207

208

209

210

211

212

213

214

215

only CRL3 is dimeric in its active form, while CRL1 and CRL2 are monomers (Kaiser et al.,

1994; Otero et al., 2005; Pernas et al., 2009; Pernas et al., 2001).

According to the literature, CRL1 and CRL5 showed the highest affinity for

triglycerides among the CRL isoenzymes. CRL1 displays the maximum lipolytic activity

towards middle-chain (C8 and C10) substrates (Chang et al., 2006a; López et al., 2004), and its

activity against p-nitrophenyl and cholesteryl esters was very poor when compared with CRL2,

CRL3 and CRL4 (Chang et al., 2006b; López et al., 2004; Mancheño et al., 2003). CRL5 seems

to be even less active than CRL1 against these substrates (Lee et al., 2011). On the other hand,

CRL2 had the highest cholesterol esterase activity (Mancheño et al., 2003) and its efficiency

towards p-nitrophenyl esters and triglycerides is intermediate between those of CRL1 and CRL3

(Chang et al., 2006b; López et al., 2004; Pernas et al., 2000), preferring long-chain fatty acids

(C12-C18:1) in all substrates tested. In spite of having the highest sequence homology (86%),

CRL3 and CRL1 differed drastically in their hydrolytic activity. CRL3 was the most active

isoenzyme on p-nitrophenyl esters regardless of the length of their acyl chain, though the best

results were reported for medium-chain substrates (Chang et al., 2006b; López et al., 2004;

Pernas et al., 2000). Its cholesterol esterase activity showed intermediate values between those

of CRL1 and CRL2 for long-chain substrates (Mancheño et al., 2003). Similarly, the CRL4

isoform displayed more activity against p-nitrophenyl esters than the commercial CRL lipase

(mostly containing CRL1), but especially towards C16 and C18 esters (Tang et al., 2001).

3.2. The G. candidum lipases

Several papers describe the purification to homogeneity of lipases produced by G.

candidum (GCL). Tsujisaka et al. (1973) applied ammonium sulfate precipitation, anion

exchange chromatography and two SEC steps to obtain a supposedly pure lipase. Nevertheless,

subsequent works demonstrated that a mixture of enzymes existed in that preparation (Shimada

et al., 1990; Shimada et al., 1989). Sugihara et al. (1990) isolated lipases I and II (hereinafter

GCL1 and GCL2) but, after incorporating an additional hydrophobic interaction

chromatography step, the same group reported four molecular forms of lipase, with GCL1

representing 84% of the total lipase activity (Sugihara et al., 1991). In parallel, Sidebottom et al.

(1991) purified and characterized four different GCLs from two G. candidum strains (two of

each), which were described as closely related monomeric glycoproteins with similar molecular

masses (54-62 kDa, from SDS-PAGE), pI (4.5-4.7), and around 8% N-linked carbohydrates. It

is worthy to mention that the strategy followed to purify each one of the crudes had to be

different. In one of them the two enzymes were separated after hydrophobic interaction and ion

exchange chromatography (Mono Q) while using the same protocol for the other one the

activity resulted irreversibly bound to the second column (a strong anion exchanger) and the

proteins were finally separated using a weak anion exchanger.

216

217

218

219

220

221

222

223

224

225

226

227

228

229

230

231

232

233

234

235

236

237

238

239

240

241

242

243

244

245

246

247

248

249

250

251

252

These isoenzymes have 86% sequence identity, but GCL2 has two potential N-

glycosylation sites, while GCL1 has three (Bertolini et al., 1995; Schrag and Cygler, 1993).

Their identity with the isoforms of C. rugosa is around 40% (Bertolini et al., 1995; Holmquist et

al., 1997a).

The expression of recombinant enzymes from this organism made easier their

purification. Two works from the same research group (Bertolini et al., 1995; Vernet et al.,

1993) described the production, purification and characterization of GCL1 and GCL2 in a

glycosylation extension-deficient S. cerevisiae strain. Inclusion of a His-tag in the gene

sequences of the two isoenzymes before their expression allowed their easy purification by

metal-affinity chromatography. Substrate specificity tests were carried out with the purified

enzymes, demonstrating the higher specific activity of GCL1 for long chain, unsaturated fatty

acid substrates with a double bond at position 9 (cis configuration, as oleate esters) whereas

GCL2 was more efficient against saturated, short-chain fatty acid substrates. These enzymes

have also been expressed in P. pastoris under the control of the methanol-inducible AOX

promoter and purified using the same strategy reported by Bertolini (1995), with higher yields

(60-fold) than in S. cerevisiae (Holmquist et al., 1997b). Recently, another lipase from

Galactomyces candidus, the teleomorph of G. candidum, has been produced in the same host.

This protein showed high sequence identity with GCL1 and GCL2 (>80%) and the same

substrate preferences, but higher values of optimum pH and temperature (Yan et al., 2007).

3.3. The O. piceae sterol esterase/lipase

After a screening on ascomycetes, basidiomycetes and conidial fungi, O. piceae was

selected for secreting lipase and sterol esterase activities (Calero-Rueda et al., 2002). The

responsible for these activities resulted to be a single enzyme of broad substrate specificity that

was isolated from the liquid cultures of this dimorphic ascomycete, and has been carefully

studied and characterized (Fig. 4A). This enzyme (OPE) was purified in a single step by

hydrophobic interaction chromatography as a 56.5 kDa-glycoprotein with 8% of N-

glycosylation, and pI of 3.3, and it forms multi-aggregates in aqueous solution (Calero-Rueda et

al., 2002), as confirmed by sedimentation velocity analyses (Barba Cedillo et al., 2012).

According to its sequence the protein contains three predicted glycosylation sites and shows

high identity (~45%) to C. rugosa lipases (Calero-Rueda et al., 2009).

The enzyme was successfully expressed in different heterologous hosts, resulting in

different yields and catalytic properties. The best yields were reached upon expression in P.

pastoris, higher than those obtained with the saprophytic fungus (~30-fold) and in a shorter

growth period (Barba Cedillo et al., 2012; Vaquero et al., 2015b). This recombinant protein had

a molecular mass of 75 kDa, 28% N-linked carbohydrates and presented improved stability at

basic pH values. Its kinetic parameters were also better when compared with the native enzyme

253

254

255

256

257

258

259

260

261

262

263

264

265

266

267

268

269

270

271

272

273

274

275

276

277

278

279

280

281

282

283

284

285

286

287

288

289

(Fig. 4B) due to its enhanced solubility, as confirmed by the presence of monomers and dimers

in solution. This behavior was attributed to the presence of 6-8 extra amino acids in the N-

terminus of the recombinant form that affect its aggregation and solubility (Barba Cedillo et al.,

2012). Recombinant forms of OPE were later obtained in other hosts (Vaquero et al., 2015b). A

sterol esterase with similar molecular mass and glycosylation degree was successfully expressed

in the Generally Recognized as Safe yeast S. cerevisiae. The enzyme produced in this host

presented an intermediate aggregation state between those of the proteins secreted by P.

pastoris and the native producer. However, only an inactive form of this enzyme was produced

in E. coli, probably because this organism cannot perform posttranslational modifications

(glycosylation, formation of disulfide bridges, etc), which can affect the correct protein folding.

The native sterol esterase from O. piceae and its recombinant form expressed in P.

pastoris display very broad substrate specificity (Barba Cedillo et al., 2012; Calero-Rueda et al.,

2002; Vaquero et al., 2015b), with higher activity towards medium-chain fatty acid p-

nitrophenyl esters (C12) and triglycerides (C8-C12), while among cholesterol esters the

preference shifts for long, unsaturated (cis-9) fatty acid acyl chains (C18:1, C18:2). Comparison

of their activity with those of other lipases or sterol esterases, under the same reaction

conditions, revealed the high efficiency and versatility of both OPE forms against most of the

substrates assayed.

3.4. The M. albomyces sterol esterase/lipase

The thermophilic fungus M. albomyces secretes a sterol esterase (STE1) with high

activity toward cholesterol esters. The enzyme was purified using hydrophobic interaction and

anion exchange chromatography after extraction with 0.1% Triton X-100, since around 30% of

the total activity was mycelium-bound (Kontkanen et al., 2006a). The monomeric protein had

around 64 kDa molecular mass as determined by SDS-PAGE, 5% glycosylation, and pI 4.5, and

its sequence presented around 47% homology with CRL1-CRL5. On the other hand, the enzyme

eluted as a single peak around 238 kDa after SEC, suggesting that it could be tetrameric in

aqueous solution. The sterol esterase STE1 was produced in two heterologous hosts, P. pastoris

GS115 (under the AOX promoter control) and Trichoderma reesei. In the first case, poor yields

of extracellular protein were recovered (Table I) since it was intracellular in a large proportion

(Kontkanen et al., 2006b). Production in T. reesei gave similar yields as M. albomyces, but the

activity was again bound to the mycelium or detected as aggregates, probably due to the high

content of hydrophobic amino acid residues in the protein (41.1%).

The steryl esterase from M. albomyces prefers unsaturated, long-chain fatty acid

substrates and its specific activity increases with the number of unsaturations

(C18:3>C18:2>C18:1>C18:0). In the case of p-nitrophenyl esters this enzyme presented the

290

291

292

293

294

295

296

297

298

299

300

301

302

303

304

305

306

307

308

309

310

311

312

313

314

315

316

317

318

319

320

321

322

323

324

325

highest substrate specificity for those with a short-chain (C4) fatty acid moiety (Kontkanen et

al., 2006a).

3.5. The B. cinerea lipase

An extracellular lipase produced by the fungus B. cinerea was purified by ammonium

sulfate fractionation and chromatography (Comménil et al., 1995). The protein had pI 6.5 and

its molecular mass was 60 kDa if determined by PAGE-SDS and 72 kDa by SEC, which

suggests that it is monomeric in solution. The amino acid composition of the lipase disclosed its

high homology with C. rugosa, G. candidum and O. piceae enzymes (46%, 38% and 56%,

respectively) and that, as them, it is very hydrophobic. Reis et al. (2005) expressed this enzyme

in P. pastoris and confirmed that the recombinant protein had properties virtually identical to

those reported for the B. cinerea culture supernatants. This lipase has activity against p-

nitrophenyl butyrate, but it shows a marked predilection for unsaturated, long-chain fatty acid

substrates (Comménil et al., 1995).

3.6. Other sterol estases/lipases from ascomycetes

The C. rugosa-like enzymes from ascomycetes most recently reported in the literature

are three hypothetical proteins from Nectria haematococca (teleomorph of F. solani), T. reesei

and A. niger, respectively, selected from in silico search in public databases of fungal genomes

(Barriuso et al., 2013). Their sequences resulted to be very similar to those of the sterol

esterases secreted by O. piceae and M. albomyces. The three proteins were cloned, expressed

and produced as functional proteins in P. pastoris with different yields, and purified by

hydrophobic interaction chromatography (Vaquero et al., 2015a). They are glycoproteins with

molecular masses between 63-67 kDa and 6-20% N-linked carbohydrates. N. haematococca and

T. reesei were active against all substrates tested, although with very different efficiency. Both

enzymes hydrolyzed cholesterol esters of long and short-chain fatty acids with low efficiency,

while their catalytic efficiency towards long and short-chain fatty acid triglycerides was one

order of magnitude higher, and were very active against p-nitrophenyl esters of short-, medium-

and long-chain fatty acid, although they were less efficient than OPE, with which they were

compared under the same experimental conditions. N. haematococca and T. reesei resembled

OPE in terms of their broad substrate specificity, but in most cases the catalytic efficiency of

OPE was similar or superior. On the other hand, the enzyme from A. niger hydrolyzed p-

nitrophenyl esters (preferably of short-chain fatty acids) and triglycerides, but it could not act on

cholesterol esters under the conditions assayed, revealing that in spite of their versatility not all

enzymes of the C. rugosa-like family are active towards sterol esters.

3.7. The P. sapidus lipase

326

327

328

329

330

331

332

333

334

335

336

337

338

339

340

341

342

343

344

345

346

347

348

349

350

351

352

353

354

355

356

357

358

359

360

361

362

Only one protein belonging to the C. rugosa-like family has been characterized so far

from basidiomycetes. This enzyme, secreted by the white rot fungus P. sapidus, was purified by

ion exchange chromatography and SEC, has isoelectric point of 4.5 and molecular mass of 54

kDa, although under native conditions the measured value was 430 kDa. From this information,

the authors suggested that the active form of the protein is a homo-octamer (Zorn et al., 2005),

although this point is not confirmed by other means and the result can simply owe to enzyme

aggregation. The sequence homology of the native lipase with C. rugosa isoenzymes is around

43%. The heterologous expression of the His-tagged protein sequence in E. coli rendered high

production of the recombinant lipase, mainly as inclusion bodies, although a small part was

soluble and active (Zelena et al., 2009). Hence, refolding was necessary to express active

enzyme from inclusion bodies (Krugener et al., 2009). Purification by metal-affinity

chromatography gave a protein of similar molecular mass to the native form. Apparently,

neither the particular codon usage of P. sapidus nor the lacking of glycosylation in this host

restricted the active expression of P. sapidus lipase in E. coli.

This lipase hydrolyzes triglycerides and xanthophyll esters, and its ability to hydrolyze

steryl esters has been suggested but never tested (Zorn et al., 2005).

4. Structural traits of C. rugosa-like lipases

The members of this family contain a similar number of amino acids (Table 1), have a

conserved glycosylation site in Asn351, and share several structural features (Figs. 1 and 2)

(Ghosh et al., 1995; Grochulski et al., 1993; Gutiérrez-Fernández et al., 2014; Mancheño et al.,

2003; Schrag and Cygler, 1993). Their protein scaffold, with the α/β hydrolase fold, is

composed of an 11-stranded mixed β-sheet, a small and nearly perpendicular N-terminal 3-

stranded β-sheet, and 16 helices (17 in the case of GCL2). In addition, they form two disulfide

bridges and several salt bridges for the stabilization of the N- and C-terminus. Similarly to

serine-proteases, the CRL-like enzymes display a catalytic mechanism based on a transfer

charge system through the so-called “catalytic triad” which is formed by Ser, His and Glu

residues in all family members. This establishes a difference with most esterases and lipases that

have Asp as the acidic residue of the triad. The catalytic machinery includes also the oxyanion

hole, a pocket in the active site involved in catalysis. The nucleophilic serine is located in a very

sharp turn, the nucleophilic elbow, composed of the conserved residues GESAG (Fig. 1). Other

characteristic of these enzymes is the presence of a substrate-binding site located in a long

internal tunnel formed by aromatic and aliphatic residues (Figs. 2 and 3) that confer a highly

hydrophobic environment all along it (Domínguez de María et al., 2006; Gutiérrez-Fernández et

al., 2014; Mancheño et al., 2003; Pleiss et al., 1998). The access to the active site is covered by

an amphipathic α-helix that serves as a lid, fixed by one of the disulfide bonds.

363

364

365

366

367

368

369

370

371

372

373

374

375

376

377

378

379

380

381

382

383

384

385

386

387

388

389

390

391

392

393

394

395

396

397

398

Only five enzymes from the C. rugosa-like lipase family have been structurally

characterized so far, and 14 three-dimensional structures from these lipases are currently

available. Crystals of C. rugosa CRL1 (PDBs: 1CRL, 1TRH, 1LPP, 1LPN, 1LPM, 1LPO) were

obtained in both its open and closed conformations (Grochulski et al., 1993; Grochulski et al.,

1994a; Grochulski et al., 1994b), while CRL2 (PDB: 1GZ7) was crystallized in its closed state

(Mancheño et al., 2003) and CRL3 (PDBs: 1CLE, 1LPS, 1LLF) in an open, dimeric

conformation (Ghosh et al., 1995; Pletnev et al., 2003) (Fig. 4D). As previously explained,

although these three isoforms have sequence identities of 77-88%, some amino acid changes at

strategic positions may be responsible for their different catalytic properties and substrate

specificity. The lipase GCL2 of G. candidum was crystallized in its closed conformation (PDB:

1TGH) (Schrag and Cygler, 1993) and very recently, the crystal structure of the O. piceae sterol

esterase/lipase (OPE) was elucidated in the open (dimer Fig. 4D) and closed (monomer, Fig.

4A) conformations (PDBs: 4BE4, 4BE9 and 4UPD) (Gutiérrez-Fernández et al., 2014).

4.1. The substrate-induced movement of the lid causes the transition to an open, active state

The flap region is one of the less conserved areas across the C. rugosa-like members,

and its structure is different in the crystallized enzymes, even within CRL isoforms. In CRLs,

the flap contains one α-helix with 10 residues and one 310 α-helix with 3 residues (Domínguez

de María et al., 2006; Mancheño et al., 2003), in OPE it has one α-helix with 6 residues and two

310 α-helices with 8 residues (Gutiérrez-Fernández et al., 2014), and in GCL2 it is formed by 3

α-helices with 12, 4, and 10 residues, respectively (Schrag and Cygler, 1993). The resolution of

four structures in their closed conformation allowed a more quantitative description of the

conformational changes that are likely associated with the transition to the active form of the

protein in water. In the closed conformation (Figs. 2 and 4), the amphiphilic α-helix is

positioned in such a way that the hydrophilic side of the lid faces the solvent, and the

hydrophobic side is directed towards the protein core (Grochulski et al., 1993; Gutiérrez-

Fernández et al., 2014; Schrag and Cygler, 1993). The transition to an open form in an aqueous

environment results from the displacement of the flap in the presence of a substrate, leaving a

large hydrophobic area exposed around the active site that contributes to recognition and

binding of substrates (Mancheño et al., 2003). This change increases the total hydrophobic

surface of the protein and opens the access to an internal tunnel. The movement of the flap of

CRL1 was analyzed by comparing its structure in the open and closed conformations,

concluding that the transition to the open state is associated exclusively to a rearrangement of

the secondary structure of the lid and to the cis-trans isomerization of the Pro92 peptidic bond

(Grochulski et al., 1994a), while the stem region is conformationally restricted by the presence

of a disulfide bridge (Cys60 – Cys97) and an ionic interaction (Glu96 and Arg37). These

structural components are also present in GCL2 and OPE, in which the disulfide bridge may

399

400

401

402

403

404

405

406

407

408

409

410

411

412

413

414

415

416

417

418

419

420

421

422

423

424

425

426

427

428

429

430

431

432

433

434

435

serve as a hinge for lid’s opening. The flap displacement in OPE is the largest (~30 Å) among

the related enzymes (Gutiérrez-Fernández et al., 2014). The highest degree of sequence

dissimilarity between CRLs is found in the flap and in the loop connecting strand β3 and helix

α3,4. This loop makes numerous contacts with the hydrophobic face of the flap and, in the case

of CRL3, it directly interacts with the sterol moiety of the substrate (Ghosh et al., 1995;

Grochulski et al., 1993; Mancheño et al., 2003).

4.2. S ubstrate-binding site and catalytic machinery: the internal tunnel, the catalytic triad, and

the oxyanion hole

It has already been mentioned that the catalytic machinery is highly conserved within

the members of the C. rugosa-like family. The substrate-binding site is located in a long and

extremely hydrophobic internal tunnel (Gutiérrez-Fernández et al., 2014; Holmquist, 2000;

Mancheño et al., 2003; Pleiss et al., 1998) and, once a compound enters this region, catalysis

takes place in two consecutive steps. In the acylation phase the hydroxyl group of the catalytic

Ser initiates the nucleophilic attack of the substrate´s carbonyl group, forming a first tetrahedral

intermediate negatively charged that is stabilized by specific residues of the oxyanion hole.

Then, the alcohol moiety of the substrate accepts a proton and is released, while the acidic part

forms a covalent acyl-enzyme intermediate that is finally broken by one water, alcohol, acid, or

ester molecule.

According to the crystal structures available, variations in substrate specificity observed

within this group of enzymes could be related to amino acid changes in these regions. For

example, few differences among the highly homologous CRL1 to CRL3 isoenzymes affecting

this area, namely one residue of the flap, three of the substrate binding pocket area, and two of

the tunnel’s mouth, are partly responsible for their variations in substrate specificity and

lipase/esterase preference. In the substrate-binding site area, the change of Ser450 in CRL1 for

Ala in CRL3 may prevent direct steric conflict to accommodate the strongly hydrophobic

cholesteryl moiety in the hydrophobic cleft, deviating enzyme specificity from triglycerides to

cholesteryl esters (Ghosh et al., 1995). OPE, with relevant sterol esterase activity, has also Ala

in this position, while in CRL2 and GCL2 it is occupied by Gly (Gutiérrez-Fernández et al.,

2014; Mancheño et al., 2003; Schrag and Cygler, 1993). In addition, comparison of the

structures of the three CRL isoenzymes disclosed that two regions of different hydrophobicity

could be differentiated along the internal tunnel (Mancheño et al., 2003). At the entrance, in

close proximity to the catalytic triad and the substrate-binding site, there is a phenylalanine-rich

region that was associated to substrate recognition and then to the lipase/esterase character of

the isoenzymes since their Phe content in this area (CRL1>CRL3≈CRL5>CRL2≈CRL4) is

inversely correlated to their activity on cholesterol esters. The aliphatic-rich region, at the

bottom of the tunnel, does not vary among CRL isoforms. OPE contains similar amount of Phe

436

437

438

439

440

441

442

443

444

445

446

447

448

449

450

451

452

453

454

455

456

457

458

459

460

461

462

463

464

465

466

467

468

469

470

471

472

as CRL2 in the first part of the tunnel, which agrees with its sterol esterase activity (Gutiérrez-

Fernández et al., 2014). Nevertheless, this enzyme is far superior to the commercial sterol

esterase of C. rugosa (Calero-Rueda et al., 2009). One of the reasons for this enhanced

efficiency has been related to the size and orientation of the tunnel (Fig. 4C) that is wide and

follows a straight trajectory in OPE (Gutiérrez-Fernández et al., 2014), while in CRL enzymes it

is narrower and bent, forming a 150º angle due to a change on Ser377 for Tyr (Mancheño et al.,

2003). A computational search for intramolecular tunnels among these enzymes and other

members of the abH03.01 family (Barriuso et al., 2013) showed that lipases from M. albomyces,

N. haematococca, T. reesei, A. niger, B. cinerea and P. sapidus have cavities comparable to

those from CRL isoforms or GCL2, although oriented towards different regions of the protein

surface (Fig. 3). The first half of the tunnels followed the same pattern in all members.

According to the structural models, the enzyme from A. niger has a bent, L-shaped internal

tunnel, similar to those of CRLs, which may be related to its lower specific activity compared

with the extracellular lipases/sterol esterases from O. piceae, N. haematococca and T. reesei.

These three enzymes have a quite straight internal tunnel, but the wider and straighter tunnel of

OPE may influence its better activity.

During catalysis, the tetrahedral intermediate could be stabilized by the main-chain

amide groups of residues located at the oxyanion hole. In CRL-like enzymes this region belongs

to the class GGGX- and in the CRLs, OPE, and GCL2 the positions Gly123, Gly124 and

Ala210 are conserved (Gutiérrez-Fernández et al., 2014; Monecke et al., 2016; Schrag and

Cygler, 1993). The possibility of these proteins having an exit tunnel has been suggested in

several works. Structural analysis of CRL3-substrate complexes showed that the fatty acid-

chain appears imbedded in a deep hydrophobic cleft of the protein that goes from the internal

catalytic site to the C-terminus, near the outer surface of the enzyme (Ghosh et al., 1995) and

the results from computational analysis of CRL1 reported by Foresti and Ferreira (2004) also

supported that this protein may have an exit tunnel. Similarly, the existence of a 30 Å-long

tunnel in OPE that connected the substrate-binding pocket with the outer surface of the protein

on the opposite side was demonstrated from its crystal structure (Gutiérrez-Fernández et al.,

2014).

4.3. Proteins’ surface: h ydrophobic patches, tertiary structure and N-linked carbohydrates

The proteins of the family abH03.01 possess several hydrophobic superficial areas that

generally locate in the surroundings of the lid and the catalytic pocket. These hydrophobic

patches have been related to interfacial activation, lid’s movement, substrate stabilization, and

protein aggregation. When opening, the flap of CRL isoforms interacts with a superficial

hydrophobic patch formed by residues 439 to 459 (Grochulski et al., 1994a). CRL2 contains a

second hydrophobic patch, located near the mouth of the substrate tunnel (Val296, Leu297,

473

474

475

476

477

478

479

480

481

482

483

484

485

486

487

488

489

490

491

492

493

494

495

496

497

498

499

500

501

502

503

504

505

506

507

508

509

Leu344 and Phe345), that prevents strong interactions of the substrate in the binding pocket,

modifying its catalytic behavior (Mancheño et al., 2003). OPE holds a hydrophobic patch

located close to the lid (Ile407, Phe408, Phe458, Pro459 and Phe460), with a unique Phe-Pro-

Phe pattern that is not present in other C. rugosa-like enzymes. Transition to the active form of

OPE involves two main structural rearrangements. The interaction of two of the superficial Phe

with two Leu residues of the lid, together with a conformational change in the α16-α17 loop,

induce stabilization of the substrate into a groove that leads to the active site (Gutiérrez-

Fernández et al., 2014). The composition of this loop differs in other family members. In the

structures of CRLs and GCLs a Trp and Phe residue, respectively, are filling this cavity and

then the channel does not exist, being a unique characteristic of OPE.

In the same way as the content of the aromatic Phe in the tunnel seems to be related to

the higher sterol esterase or lipase activity, it has been suggested that the amount of

hydrophobic residues in the lid area of CRL isoenzymes (CRL2 > CRL3 > CRL1) influences

their affinity for very hydrophobic substrates as sterol esters (Domínguez de María et al., 2006;

Mancheño et al., 2003). This hypothesis also agrees with the data reported for OPE, which has

higher content in hydrophobic amino acids and improved sterol esterase activity than the CRL

isoforms (Gutiérrez-Fernández et al., 2014), and with the data from structural models for the

hypothetical proteins expressed by Vaquero et al. (2015b), since that from A. niger had the least

amount of hydrophobic amino acids in the lid region and was not active on sterol esters under

the assayed conditions.

Among the crystallized enzymes of this family, CRL1, CRL2 and GCL2 are monomeric

in their open, active state. Evidences for dimerization have been reported exclusively for the

open forms of CRL3 and OPE (Gutiérrez-Fernández et al., 2014; Otero et al., 2005; Pernas et

al., 2001). In the presence of a substrate or inhibitor, CRL3 monomers interact to form a tight

homodimer, with the flaps of the two monomers open and the active-site grooves from the two

subunits facing each other. This arrangement of the two protein molecules creates a 7.3 Å-

radius cavity composed of hydrophobic and aromatic residues (Pletnev et al., 2003), with four

openings to the surface (Fig. 4D). The organization of the OPE dimer is completely different

since the aliphatic residues from the lid of one monomer interact with the hydrophobic patch of

the opposite chain in such a way that contacts between monomers are scarce. This disposition

conforms a pacman-like structure (Fig. 4D), creating a wide hydrophobic opening (23 Å × 38

Å) exposed to the solvent where large and hydrophobic substrates can interact, allowing also the

quick release of the products upon catalysis (Gutiérrez-Fernández et al., 2014). This type of

dimeric organization can be related to the high activity of OPE towards many substrates.

The phenomenon of interfacial activation, consistent in the increase of the lipolytic

activity of certain enzymes upon binding to the lipid-water interface, has usually been related to

the displacement of the lid (Otero et al., 2005; Pernas et al., 2001). Nevertheless, some enzymes

510

511

512

513

514

515

516

517

518

519

520

521

522

523

524

525

526

527

528

529

530

531

532

533

534

535

536

537

538

539

540

541

542

543

544

545

546

of this family as CRL3, CRL4 and OPE don’t show interfacial activation (Calero-Rueda et al.,

2002) in spite of having lids that must be opened to allow the access to the active center.

Curiously, these three enzymes possess relevant esterase activity and two of them are dimeric

(there is not information on the tertiary structure of CRL4), with their flap participating at more

or less extent in dimer formation.

All members of this family are glycoproteins, and their amino acidic sequences contain

one conserved N-glycosylation site (Asn351 in CRLs) with the only exception of the T. reesei

lipase that bears a glycidic chain in a different position (Vaquero et al., 2015a). The presence of

carbohydrates usually enhances proteins’ stability (Hung et al., 2011; Tang et al., 2000) and,

particularly, the sugar chain attached at the conserved glycosylation site has shown to provide

additional stabilization to the open conformation of the flap in CRL1 (Brocca et al., 2000;

Grochulski et al., 1994a). However, sugars may not be indispensable in these proteins once

secreted, as deduced for full activity retention of the deglycosylated OPE (Vaquero et al.,

2015a). On the other hand, a non-glycosylated recombinant form of CRL4 expressed in E. coli

showed the same catalytic properties as the glycosylated protein expressed in P. pastoris (Tang

et al., 2001) while expression of non-glycosylated OPE in the prokaryotic host gave inactive

intracellular or extracellular protein (Vaquero et al., 2015b).

5. Protein engineering of C. rugosa-like lipases to decipher the role of structural areas in

enzyme´s function

As explained above, sterol esterases and lipases from the C. rugosa-like family usually

display activity on triglycerides, p-nitrophenyl esters and cholesterol esters (López et al., 2004;

Mancheño et al., 2003; Vaquero et al., 2015a), although their catalytic efficiency towards these

compounds is variable. Many studies have tried to relate the structural elements or key residues

of the CRL isoenzymes with their different catalytic properties. These structure-function studies

can be attempted only with the previous knowledge of the crystal structure of the protein, and

are mostly based in two approaches. One is the replacement of a whole region of a given

isoform for the same region of other isoenzyme with different properties. The other analyzes the

effects of changing specific residues, located in strategic positions around the active center, by

sequence-based directed mutagenesis. Then, the changes introduced in the new variants may

serve not only to gain information on the role of particular areas or amino acids of the proteins,

but also to tune the enzyme’s activity.

As the flap area is not very conserved among CRL isoenzymes, its role in modulating

their activity and selectivity was analyzed. Brocca et al. (2003) reported that swapping the lid of

CRL3 to CRL1 (their sequences differ in 6 out to 27 amino acids) was sufficient to confer

higher cholesterol esterase activity to the chimeric CRL1lid3, although without any specific

shift in its chain-length specificity. Particular residues of the lid of CRL3, namely Phe69,

547

548

549

550

551

552

553

554

555

556

557

558

559

560

561

562

563

564

565

566

567

568

569

570

571

572

573

574

575

576

577

578

579

580

581

582

583

Gly74, Thr76 and Gln88, seem to have a role in facilitating the access to the substrate-binding

pocket, allowing the recognition of very hydrophobic substrates as cholesterol esters. The flap

region of CRL2 differs from those of CRL1 and CRL3 in 11 and 9 positions (out of 27),

respectively, and it is more hydrophobic (Mancheño et al., 2003). Lid swapping in CRL4 (Akoh

et al., 2004) resulted in increased hydrolytic activities on tributyrin of the chimeric CRL4/lid2

and CRL4/lid3 by 14% and 32%, respectively, and affected also the substrate specificity of the

enzymes to various unsaturated fatty acids (C18:1>C18:2>>C18.0).

Similarly, the implication of the C-terminal region of the CRLs on enzymatic activity

was investigated by Hung et al. (2011) by generating a chimeric lipase made up with the C-

terminus of CRL1 on a CRL4 scaffold. While the native CRL4 possess high esterase activity,

low lipolytic activity, and lacks interfacial activation, the chimera showed to have similar

activity than commercial CRL on triolein, enhanced lipolytic activity over CRL4, and interfacial

activation.

Several studies tackle the effects derived from modification of amino acids located in

the tunnel region. Schmitt et al. (2002) studied the molecular basis of chain-length specificity of

CRL1, showing that placing a bulky amino acid in different areas of this channel restricts the

specificity of CRL1 towards substrates of different acyl-chain length. For example, mutants

P246F and L413F had a strong preference for short chain lengths whereas substrates longer than

C10 were hardly hydrolyzed. In the same way, increasing the bulkiness of the amino acid at

position 410 led to mutants that showed a reduced activity on substrates of chain lengths longer

than C14, since the protein activity sharply decreased as the acyl chain became long enough to

reach the mutated site. Similarly, in OPE the change of a Ile residue for a Trp at the end of the

tunnel (Fig. 4C) increased its specificity towards substrates of medium acyl-chain length and

abolished its activity against long acyl-chain length substrates (Gutiérrez-Fernández et al.,

2014). Similar results have recently been reported in LIP1 mutants of CRL expressed in P.

pastoris (Zhang et al., 2016).

In CRL4, residues 296 and 344 are located at a hydrophobic pocket, near the entrance to

the hydrophobic tunnel, and then they are good targets for single-residue mutations in structure-

function studies. Some experiments suggest that the small Ala residue at the 296 position of

CRL4 might be responsible for its lower hydrolytic activity towards triglycerides. The

replacement of A296 by Ile, a more hydrophobic residue that resembles those of other CRL

isoforms at the same position, increased the hydrophobic interaction with triglycerides, showing

highly improved activity against these substrates. On the other hand, the results of mutating

V344 indicated that a hydrophobic residue is required at this position for binding of the

medium- or long-chain triglycerides (Lee et al., 2007). Furthermore, it has been reported that

mutations in the substrate-binding site of CRL2 shifted the specificity of this enzyme from

short- to medium-chain triglycerides (Yen et al., 2010). The substitution of L132 for a smaller

584

585

586

587

588

589

590

591

592

593

594

595

596

597

598

599

600

601

602

603

604

605

606

607

608

609

610

611

612

613

614

615

616

617

618

619

620

side-chain amino acid could enhance the accommodation of medium- to long-chain triglycerides

in the active site, while the replacement of G450 by the hydrophobic Ala, which is the residue

present in other CRL isoforms at this position, might increase the hydrophobic interaction with

triglycerides resulting in improved catalytic activity.

Enantioselectivity is a very interesting characteristic of some enzymes of this group,

especially in the pharmacological context where they are used for the resolution of racemic

mixtures. Protein engineering helped to show that in CRLs specific phenylalanine residues of

the tunnel are responsible for this property. For example, the work from Berglund et al. (1995)

concluded that Phe345 is essential in CRL1 and CRL3 for providing high enantioselectivity

toward naproxen and 2-arylpropionic acid, while the adjacent Phe344 is not. Piamtongkam et al.

(2011) studied the enantioselectivity of CRL1, CRL3 and CRL4 in the resolution of

enantiomers of 2-bromo-arilacetic acid esters, an important class of chemical intermediates in

the pharmaceutical industry. All exhibited a high enantioselectivity, with CRL4 preferring the

R-enantiomer, while CRL1 and CRL3 showed S–enantioselectivity. The construction of CRL1

and CRL4 variants in position 296 revealed that the bulkier is the amino acid introduced, the

larger the selectivity towards the S-enantiomer.

Few structure-function studies concerning other enzymes from the C. rugosa-like

family have been published. Crucial elements were found to be located at the entrance and the

bottom of the active site cavity of the G. candidum lipase GCL1. Mutations affecting

exclusively target residues of the active site access increased the specific activity towards

trioctanoin, whereas amino acid replacements at the end of the tunnel gave active proteins

against both triolein and trioctanoin (Holmquist, 1998). Computational analysis of the GCLs

revealed that a region that shows high conformational flexibility, comprising residues 349 – 406

and the flap, is involved in substrate recognition and catalysis. These segments have a crucial

role for the high oleate preference of GCL1.

6. Conclusions and future prospects

Among extracellular lipases, those belonging to family C. rugosa-like share a number

of well defined structural traits, as the presence of a Glu residue instead of Asp in the catalytic

triad, the conserved GESAG sequence where the catalytic Ser is accommodated, or the GGGX

sequence in the oxyanion hole. They display wide substrate specificity and, in this sense, the

literature review evidences the difficulty for defining a given enzyme of this group as lipase,

aryl esterase, or sterol esterase, as many of them are able to act, more or less efficiently, on all

these substrates. The use of standardized protocols to determine their activity levels and

substrate specificity would enable a better comparison of the catalysts, since these parameters

are hardly comparable among the papers examined. The outcomes from structure-function

621

622

623

624

625

626

627

628

629

630

631

632

633

634

635

636

637

638

639

640

641

642

643

644

645

646

647

648

649

650

651

652

653

654

655

656

studies suggest that the morphology, dimensions, and amino acid composition of the internal

tunnel and the lid are crucial to the catalytic versatility of this lipase family.

Some of the C. rugosa-like enzymes have proved to be efficient catalysts from a

biotechnological perspective (Benjamin and Pandey, 1998) and for this reason there is a

renewed interest in discovering novel enzymes from this group or improving the activity of the

currently known. These enzymes are extremely hydrophobic, and although this feature causes in

many occasions unwanted effects, such as aggregation, it can also simplify their

chromatographic purification from other proteins in crude extracts. Nevertheless, when pools of

hydrophobic proteins are secreted (e.g. CRL isoforms) their purification to homogeneity is

difficult. Cloning and heterologous expression of a specific protein helps to solve this problem

since specific tags can be introduced in the recombinant species.

Among the systems tested so far for heterologous enzyme production within this family,

P. pastoris has proved to be the most promising host in terms of easy manipulation, yields/costs

and product quality. Since the recombinant enzyme does not necessarily have the same

properties as the native protein, its characteristics should be analyzed. With this technology,

hypothetical lipases from this group have also been cloned and produced to assess their true

structure and properties. Moreover, mutants for amino acids relevant to function, or chimeric

enzymes with changes in entire structural domains, can be designed on the scaffold of known

enzymes, and expressed in order to study the effect of such modifications or to get a catalyst

with enhanced properties for a particular application, expanding also the basic knowledge on C.

rugosa-like enzymes. Additionally, recent reports have shown the usefulness of in silico data

mining and its associated bioinformatics tools to unveil potential members of this family. The

increasingly high availability of fungal genomes makes feasible extending the search for the

conserved motifs and structural characteristics common to these versatile enzymes to

unexplored fungal groups. Further expression and analysis of the hypothetical proteins likely

related to this group may provide catalysts with new or improved capacities. Finally, and

considering the unique structural characteristics and substrate versatility of the enzymes

grouped under this common denomination, we propose their classification as “Versatile

Lipases” (EC 3.1.1.x).

Acknowledgements

This work was supported by the Spanish projects BIO2012-36372, BIO2015-68387-R, RTC-

2014-1777-3 from MINECO and S2013/MAE-2972 from Comunidad de Madrid.

657

658

659

660

661

662

663

664

665

666

667

668

669

670

671

672

673

674

675

676

677

678

679

680

681

682

683

684

685

686

687

688

689

690

691

References

Akoh, C.C., Lee, G.C., Shaw, J.F., 2004. Protein engineering and applications of Candida rugosa lipase isoforms. Lipids 39, 513-526.

Barba Cedillo, V., Plou, F.J., Martínez, M.J., 2012. Recombinant sterol esterase from Ophiostoma piceae: an improved biocatalyst expressed in Pichia pastoris. Microb Cell Fact 11, 73

Barba Cedillo, V., Prieto, A., Martínez, Á.T., Martínez, M.J. 2013. Acylation method for producing food and/or pharmaceutical compounds of interest using fungal sterol esterases. WO 2013/001126 A1.

Barriuso, J., Prieto, A., Martinez, M.J., 2013. Fungal genomes mining to discover novel sterol esterases and lipases as catalysts. BMC Genomics 14, 712

Benjamin, S., Pandey, A., 1998. Candida rugosa lipases: Molecular biology and versatility in biotechnology. Yeast 14, 1069-1087.

Berglund, P., Holmquist, M., Hult, K., Högberg, H.E., 1995. Alcohols as enantioselective inhibitors in a lipase catalysed esterification of a chiral acyl donor. Biotechnol Lett 17, 55-60.

Bertolini, M.C., Schrag, J.D., Cygler, M., Ziomek, E., Thomas, D.Y., Vernet, T., 1995. Expression and characterization of Geotrichum candidum lipase I gene - comparison of specificity profile with lipase II. Eur J Biochem 228, 863-869.

Bornscheuer, U.T., 2002. Microbial carboxyl esterases: classification, properties and application in biocatalysis. FEMS Microbiol Rev 26, 73-81.

Bourne, Y., Hasper, A.A., Chahinian, H., Juin, M., de Graaff, L.H., Marchot, P., 2004. Aspergillus niger protein EstA defines a new class of fungal esterases within the α/ hydrolase fold superfamily of proteins. Structure 12, 677-687.

Brocca, S., Persson, M., Wehtje, E., Adlercreutz, P., Alberghina, L., Lotti, M., 2000. Mutants provide evidence of the importance of glycosydic chains in the activation of lipase 1 from Candida rugosa. Protein Sci 9, 985-990.

Brocca, S., Schmidt-Dannert, C., Lotti, M., Alberghina, L., Schmid, R.D., 1998. Design, total synthesis, and fuctional overexpression of the Candida rugosa lip1 gene coding for a major industrial lipase. Protein Sci. 7, 1415-1422.

Brocca, S., Secundo, F., Ossola, M., Alberghina, L., Carrea, G. et al., 2003. Sequence of the lid affects activity and specificity of Candida rugosa lipase isoenzymes. Protein Sci. 12, 2312-2319.

Brocca, S., Grandori, R., Breviario, D., Lotti, M., 1995. Localization of lipase genes on Candida rugosa chromosomes. Current Genetics 28, 454-457.

Calero-Rueda, O., Barba, V., Rodriguez, E., Plou, F., Martínez, A.T., Martínez, M.J., 2009. Study of a sterol esterase secreted by Ophiostoma piceae: Sequence, model and biochemical properties. BBA-Proteins Proteomics 1794, 1099-1106.

Calero-Rueda, O., Plou, F.J., Ballesteros, A., Martínez, A.T., Martínez, M.J., 2002. Production, isolation and characterization of a sterol esterase from Ophiostoma piceae. BBA Proteins Proteomics 1599, 28-35.

692

693694

695696697

698699700

701702

703704

705706707

708709710

711712

713714715

716717718

719720721

722723724

725726

727728729

730731732

Chang, S.W., Lee, G.C., Shaw, J.F., 2006a. Codon optimization of Candida rugosa lip1 Gene for improving expression in Pichia pastoris and biochemical characterization of the purified recombinant LIP1 lipase. J Agr Food Chem 54, 815-822.

Chang, S.W., Lee, G.C., Shaw, J.F., 2006b. Efficient production of active recombinant Candida rugosa LIP3 lipase in Pichia pastoris and biochemical characterization of purified enzyme. J Agr Food Chem 54, 5831-5838.

Chang, S.W., Shieh, C.J., Lee, G.C., Shaw, J.F., 2005. Multiple mutagenesis of the Candida rugosa lip1 gene and optimum production of recombinant LIP1 expressed in Pichia pastoris. App Microbiol Biot 67, 215-224.

Colton, I.J., Yin, D., Grochulski, P., Kazlauskas, R., 2011. Molecular basis of chiral acid recognition by Candida rugosa lipase: X-ray xtructure of transition state analog and modeling of the hydrolysis of methyl 2-methoxy-2-phenylacetate. Adv Synth Catal 353, 2529-2544.

Comménil, P., Belingheri, L., Bauw, G., Dehorter, B., 1999. Molecular characterization of a lipase induced in Botrytis cinerea by components of grape berry cuticle. Physiol Mol Plant P 55, 37-43.

Comménil, P., Belingheri, L., Sancholle, M., Dehorter, B., 1995. Purification and properties of an extracellular lipase from the fungus Botrytis cinerea. Lipids 30, 351-356.

de Hoog, G.S., Smith, M.T., 2004. Ribosomal gene phylogeny and species delimitation in Geotrichum and its teleomorphs. Stud Mycol 50, 489-516.

de Vries, R.P., Michelsen, B., Poulsen, C.H., Kroon, P.A., van den Heuvel, R.H., Faulds, C.B. et al., 1997. The faeA genes from Aspergillus niger and Aspergillus tubingensis encode ferulic acid esterases involved in degradation of complex cell wall polysaccharides. Appl Environ Microbiol 63, 4638-4644.

Domínguez de María, P., Sanchez-Montero, J.M., Sinisterra, J.V., Alcántara, A.R., 2006. Understanding Candida rugosa lipases: An overview. Biotechnol Adv 24, 178-194.

Doss, R.P., 1999. Composition and enzymatic activity of the extracellular matrix secreted by germlings of Botrytis cinerea. Appl Environ Microbiol 65, 404-408.

Eagen, R., Brisson, A., Breuil, C., 1997. The sap-staining fungus Ophiostoma piceae synthesizes different types of melanin in different growth media. Can J Microbiol 43, 592-595.

Ferrer, P., Alarcón, M., Ramón, R., Benaiges, M.D., Valero, F., 2009. Recombinant Candida rugosa LIP2 expression in Pichia pastoris under the control of the AOX1 promoter. Biochem Eng J 46, 271-277.

Ferrer, P., Montesinos, J.L., Valero, F., Sola, C., 2001. Production of native and recombinant lipases by Candida rugosa - A review. Appl Biochem Biotechnol 95, 221-255.

Fischer, M., Pleiss, J., 2003. The Lipase Engineering Database: a navigation and analysis tool for protein families. Nucleic Acids Res 31, 319-321.

Foresti, M.L., Ferreira, M.L., 2004. Computational approach to solvent-free synthesis of ethyl oleate using Candida rugosa and Candida antarctica B lipases. I. Interfacial activation and substrate (ethanol, oleic acid) adsorption. Biomacromolecules 5, 2366-2375.

733734735

736737738

739740741

742743744745

746747748

749750

751752

753754755756

757758

759760

761762763

764765766

767768

769770

771772773

Ghosh, D., Wawrzak, Z., Pletnev, V.Z., Li, N.Y., Kaiser, R., Pangborn, W. et al., 1995. Structure of uncomplexed and linoleate-bound Candida cylindracea cholesterol esterase. Structure 3, 279-288.

Grochulski, P., Li, Y., Schrag, J.D., Bouthillier, F., Smith, P., Harrison, D. et al., 1993. Insights into interfacial activation from an open structure of Candida rugosa lipase. J Biol Chem 268, 12843-12847.

Grochulski, P., Li, Y., Schrag, J.D., Cygler, M., 1994a. Two conformational states of Candida rugosa lipase. Protein Sci 3, 82-91.

Grochulski, P., Bouthillier, F., Kazlauskas, R., Serreqi, A., Schrag, J., Ziomek, E. et al., 1994b. Analogs of reaction intermediates identify a unique substrate binding site in Candida rugosa lipase. Biochemistry-US 33, 3494-3500.

Gupta, R., Kumari, A., Syal, P., Singh, Y., 2015. Molecular and functional diversity of yeast and fungal lipases: Their role in biotechnology and cellular physiology. Prog Lipid Res 57, 40-54.

Gutiérrez-Fernández, J., Vaquero, M.E., Prieto, A., Barriuso, J., Martinez, M.J., Hermoso, J.A., 2014. Crystal structures of Ophiostoma piceae sterol esterase: Structural insights into activation mechanism and product release. J Struct Biol 187, 215-222.

Haridas, S., Wang, Y., Lim, L., Alamouti, S., Jackman, S., Massoumi Alamouti, S. et al., 2013. The genome and transcriptome of the pine saprophyte Ophiostoma piceae, and a comparison with the bark beetle-associated pine pathogen Grosmannia clavigera. BMC Genomics 14, 373-373.

Hasan, F., Shah, A.A., Hameed, A., 2006. Industrial applications of microbial lipases. Enzyme Microb Tech 39, 235-251.

Holmquist, M., 2000. Alpha/beta-hydrolase fold enzymes: structures, functions and mechanisms. Curr Protein Pept Sc 1, 209-235.

Holmquist, M., Tessier, D.C., Cygler, M., 1997b. High-level production of recombinant Geotrichum candidum lipases in yeast Pichia pastoris. Prot Expres Purif 11, 35-40.

Holmquist, M., Tessier, D.C., Cygler, M., 1997a. Identification of residues essential for differential fatty acyl specificity of Geotrichum candidum lipases I and II. Biochemistry-US 36, 15019-15025.

Holmquist, M., 1998. Insights into the molecular basis for fatty acyl specificities of lipases from Geotrichum candidum and Candida rugosa. Chem Phys Lipids 93, 57-65.

Houde, A., Kademi, A., Leblanc, D., 2004. Lipases and their industrial applications. Appl Biochem Biotechnol 118, 155-170.

Hu, H.L., van den Brink, J., Gruben, B.S., Wösten, H.A.B., Gu, J.D., de Vries, R.P., 2011. Improved enzyme production by co-cultivation of Aspergillus niger and Aspergillus oryzae and with other fungi. Int Biodeter Biodegr 65, 248-252.

Hung, K.S., Chen, S.Y., Liu, H.F., Tsai, B.R., Chen, H.W., Huang, C.Y. et al., 2011. C-Terminal region of Candida rugosa lipases affects enzyme activity and interfacial activation. J Agr Food Chem 59, 5396-5401.

Jaeger, K.E., Eggert, T., 2002. Lipases for biotechnology. Curr Opin Biotech 13, 390-397.

774775776

777778779

780781

782783784

785786787

788789790

791792793794

795796

797798

799800

801802803

804805

806807

808809810

811812813

814

Juniper, B.E., Jeffree, C.E.1983. Plant Surfaces, Eduard Arnold, Baltimore, MD.

Kaiser, R., Erman, M., Duax, W.L., Ghosh, D., Jörnvall, H., 1994. Monomeric and dimeric forms of cholesterol esterase from Candida cylindracea: primary structure, identity in peptide patterns, and additional microheterogeneity. FEBS Lett 337, 123-127.

Kawaguchi, Y., Honda, H., Taniguchi-Morimura, J., Iwasaki, S., 1989. The codon CUG is read as serine in an asporogenic yeast Candida cylindracea. Nature 341, 164-166.

Kontkanen, H., Reinikainen, T., Saloheimo, M., 2006b. Cloning and expression of a Melanocarpus albomyces steryl esterase gene in Pichia pastoris and Trichoderma reesei. Biotechnol Bioeng 94, 407-415.

Kontkanen, H., Tenkanen, M., Reinikainen, T., 2006a. Purification and characterisation of a novel steryl esterase from Melanocarpus albomyces. Enzyme Microb Tech 39, 265-273.

Krugener, S., Zelenaa, K., Zorn, H., Nimtz, M., Bergera, R.G., 2009. Heterologous expression of an extra-cellular lipase from the basidiomycete Pleurotus sapidus. J Mol Catal B 57, 16-21.

Lee, G.C., Lee, L.C., Sava, V., Shaw, J.F., 2002. Multiple mutagenesis of non-universal serine codons of the Candida rugosa lip2 gene and biochemical characterization of purified recombinant LIP2 lipase overexpressed in Pichia pastoris. Biochem J 366, 603-611.

Lee, G.-C., Tang, S.-J., Sun, K.-H., Shaw, J.-F., 1999. Analysis of the gene family encoding lipases in Candida rugosa by competitive reverse transcription-PCR. Appl Environ Microbiol 65, 3888-3895.

Lee, L.C., Chen, Y.T., Yen, C.C., Chiang, T.C.Y., Tang, S.J., Lee, G.C. et al., 2007. Altering the substrate specificity of Candida rugosa LIP4 by engineering the substrate-binding sites. J Agr Food Chem 55, 5103-5108.

Lee, L.C., Yen, C.C., Malmis, C.C., Chen, L.F., Chen, J.C., Lee, G.C. et al., 2011. Characterization of codon-optimized recombinant Candida rugosa lipase 5 (LIP5). J Agr Food Chem 59, 10693-10698.

Lenfant, N., Hotelier, T., Velluet, E., Bourne, Y., Marchot, P., Chatonnet, A., 2013received. ESTHER, the database of the a/-hydrolase fold superfamily of proteins: tools to explore diversity of functions. Nucleic Acids Res 41, D423-D429.

López, N., Pernas, M.A., Pastrana, L.M., Sánchez, A., Rúa, M.L., 2004. Reactivity of pure Candida rugosa lipase isoenzymes (Lip1, Lip2, and Lip3) in aqueous and organic media. Influence of the isoenzymatic profile on the lipase performance in organic media. Biotechnol Prog 20, 65-73.

Lotti, M., Alberghina, L., 1996. Candida rugosa lipase isozymes. In: Engineering of/with lipases. Malcata,F.X. NATO Advanced Science Institutes Series. Series E, Applied Sciences 317, 115-124.

Lotti, M., Tramontano, A., Longhi, S., Fusetti, F., Brocca, S., Pizzi, E. et al., 1994. Variability within the Candida rugosa lipases family. Protein Eng 7, 531-535.

Maeda, A., Mizuno, T., Bunya, M., Sugihara, S., Nakayama, D., Tsunasawa, S. et al., 2008. Characterization of novel cholesterol esterase from Trichoderma sp. AS59 with high ability to synthesize steryl esters. J Biosci Bioeng 105, 341-349.

815

816817818

819820

821822823

824825

826827828

829830831

832833834

835836837

838839840

841842843

844845846847

848849850

851852

853854855

Mancheño, J.M., Pernas, M.A., Martínez, M.J., Ochoa, B., Rua, M.L., Hermoso, J.A., 2003. Structural insights into the lipase/esterase behavior in the Candida rugosa lipases family: crystal structure of the lipase 2 isoenzyme at 1.97 Å resolution. J Mol Biol 332, 1059-1069.

Mileto, D., Brocca, S., Lotti, M., Takagi, M., Alquati, C., Alberghina, L., 1998. Characterization of the Candida rugosa lipase system and overexpression of the lip1 isoenzyme in a non-conventional yeast. Chem Phys Lipids 93, 47-55.

Monecke, P., Friedemann, R., Naumann, S., Csuk, R., 2016. Molecular modelling studies on the catalytic mechanism of Candida rugosa lipase. J Mol Model 4, 395-404.

Morinaga, N., Maeda, A., Mizuno, T., Bunya, M., Sugihara, S., Sugihara, A., 2011. Synthesis of fatty acid sterol esters using cholesterol esterase from Trichoderma sp. AS59. Enzyme Microb Tech 48, 498-504.

Nardini, M., Dijkstra, B.W., 1999. a/ hydrolase fold enzymes: the family keeps growing. Curr Opin Struc Biol 9, 732-737.

Otero, C., Fernandez-Perez, M., Hermoso, J.A., Ripoll, M.M., 2005. Activation in the family of Candida rugosa isolipases by polyethylene glycol. J Mol Catal B 32, 225-229.

Palocci, C., Falconi, M., Alcaro, S., Tafi, A., Puglisi, R., Ortuso, F. et al., 2007. An approach to address Candida rugosa lipase regioselectivity in the acylation reactions of trytilated glucosides. J Biotechnol 128, 908-918.

Pernas, M.A., Lopez, C., Pastrana, L., Rua, M.L., 2000. Purification and characterization of Lip2 and Lip3 isoenzymes from a Candida rugosa pilot-plant scale fed-batch fermentation. J Biotechnol 84, 163-174.

Pernas, M.A., Pastrana, L., Fucinos, P., Rua, M.L., 2009. Regulation of the interfacial activation within the Candida rugosa lipase family. J Phys Org Chem 22, 508-514.

Pernas, M.A., López, C., Rúa, M.L., Hermoso, J., 2001. Influence of the conformational flexibility on the kinetics and dimerisation process of two Candida rugosa lipase isoenzymes. FEBS Lett 501, 87-91.

Piamtongkam, R., Duquesne, S., Bordes, F., Barbe, S., André, I., Marty, A. et al., 2011. Enantioselectivity of Candida rugosa lipases (Lip1,Lip3, and LIp4) towards 2-bromo phenylacetic acid octyl esters controlled by a single amino acid. Biotechnol. Bioeng. 108, 1749-1756.

Pleiss, J., Fischer, M., Schimd, R.D., 1998. Anatomy of lipase binding sites: the scissile fatty acid binding site. Chem Phys Lip 93, 67-80.

Pletnev, V., Addlagatta, A., Wawrzak, Z., Duax, W., 2003. Three-dimensional structure of homodimeric cholesterol esterase-ligand complex at 1.4 angstrom resolution. Acta Crystallogr. Sect D -Biol Crystallogr 59, 50-56.

Reetz, M.T., 2002. Lipases as practical biocatalysts. Curr Opin Chem Biol 6, 145-150.

Reis, H., Pfiffi, S., Hahn, M., 2005. Molecular and functional characterization of a secreted lipase from Botrytis cinerea. Mol Plant Pathol 6, 257-267.

Reis, P., Holmberg, K., Watzke, H., Leser, M.E., Miller, R., 2009. Lipases at interfaces: a review. Adv Colloid Interface Sci 147-148, 237-250.

856857858859

860861862

863864

865866867

868869

870871

872873874

875876877

878879

880881882

883884885886

887888

889890891

892

893894

895896

Rua, L., Diaz-Mauriño, T., Fernández, V.M., Otero, C., Ballesteros, A., 1993. Purification and characterization of two distinct lipases from Candida cylindracea. Biochim Biophys Acta 1156, 181-189.

Rúa, M.L., 1994. Rapid purification of two lipase isoenzymes from Candida rugosa. Biotechnol Tech 8, 21-26.

Schmid, R.D., Verger, R., 1998. Lipases: Interfacial enzymes with atractive applications. Angew Chem Int Ed 37, 1609-1633.

Schmitt, J., Brocca, S., Schmid, R.D., Pleiss, J., 2002. Blocking the tunnel: engineering of Candida rugosa lipase mutants with short chain length specificity. Protein Eng 15, 595-601.

Schrag, J.D., Cygler, M., 1993. 1.8 Å Refined structure of the lipase from Geotrichum candidum. J Mol Biol 230, 575-591.

Schuster, A., Schmoll, M., 2010received. Biology and biotechnology of Trichoderma. Appl Microbiol Biotechnol 87, 787-799.

Shimada, Y., Sugihara, A., Iizumi, T., Tominaga, Y., 1990. cDNA cloning and characterization of Geotrichum candidum lipase II. J Biochemistry 107, 703-707.

Shimada, Y., Sugihara, A., Tominaga, Y., Iizumi, T., Tsunasawa, S., 1989. cDNA molecular cloning of Geotrichum candidum lipase. J Biochemistry 106, 383-388.

Sidebottom, M., Charton, E., Dunn, P.P.J., Mycock, G., Davies, C., Sutton, J.L. et al., 1991. Geotrichum candidum produces several lipases with markedly different sustrate specifities. Eur. J. Biochem. 202, 485-491.

Singh, A.K., Mukhopadhyay, M., 2012. Overview of fungal lipase: a review. Appl Biochem Biotechnol 166, 486-520.

Stergiou, P.Y., Foukis, A., Filippou, M., Koukouritaki, M., Parapouli, M., Theodorou, L.G. et al., 2013. Advances in lipase-catalyzed esterification reactions. Biotechnol Adv 31, 1846-1859.

Sugihara, A., Shimada, Y., Tominaga, Y., 1990. Separation and characterization of two molecular forms of Geotrichum candidum lipase. J. Biochem. (Tokyo) 107, 426-430.

Sugihara, A., Shimada, Y., Tominaga, Y., 1991. A novel Geotrichum candidum lipase with some preference for the 2-position on a triglyceride molecule. Appl Microbiol Biotechnol 35, 738-740.

Tang, S.J., Shaw, J.F., Sun, K.H., Sun, G.H., Chang, T.Y., Lin, C.K. et al., 2001. Recombinant expression and characterization of the Candida rugosa LIP4 lipase in Pichia pastoris: comparison of glycosilation, activity, and stability. Arch Biochem Biophys 387, 93-98.

Tang, S.J., Sun, K.H., Sun, G.H., hang, T.Y., Lee, G.C., 2000. Recombinant Expression of the Candida rugosa lip4 Lipase in Escherichia coli. Protein Expr Purif 20, 308-313.

Tomizuka, N., Ota, Y., Yamada, K., 1966. Studies on lipase from Candida cylindracea: amino acid composition, carbohydrate component and some physical properties. Agric. Biol. Chem. 30, 1090-1096.

897898899

900901

902903

904905906

907908

909910

911912

913914

915916917

918919

920921922

923924

925926927

928929930

931932

933934935

Tsujisaka, Y., Iwai, M., Tominaga, Y., 1973. Purification, crystallization and some properties of lipase from Geotrichum candidum Link. Agric. Biol. Chem. 37, 1457-1464.

Vaquero, M.E., Barriuso, J., Martinez, M.J., Prieto, A., 2016. Properties, structure, and applications of microbial sterol esterases. App Microbiol Biot 100, 2047-2061.

Vaquero, M.E., Barriuso, J., Medrano, F., Prieto, A., Martinez, M.J., 2015b. Heterologous expression of a fungal sterol esterase/lipase in different hosts: Effect on solubility, glycosylation and production. J Biosci Bioeng 120, 637-643.

Vaquero, M.E., Prieto, A., Barriuso, J., Martinez, M.J., 2015a. Expression and properties of three novel fungal lipases/sterol esterases predicted in silico: comparison with other enzymes of the Candida rugosa-like family. App Microbiol Biot 99, 10057-10067.

Veeraragavan, K., Gibbs, B.F., 1989. Detection and partial purification of two lipases from Candida rugosa. Biotechnol Lett 11, 345-348.

Vernet, T., Ziomek, E., Recktenwald, A., Schrag, J.D., Demontigny, C., Tessier, D.C. et al., 1993. Cloning and expression of Geotrichum candidum lipase II gene in yeast: probing of the enzyme active site by site-directed mutagenesis. J Biol Chem 268, 26212-26219.

Yan, J.Y., Yang, H.K., Xu, L., Yan, Y.J., 2007. Gene cloning, overexpression and characterization of a novel organic solvent tolerant and thermostable lipase from Galactomyces geotrichum Y05. J Mol Catal B 49, 28-35.

Yen, C.C., Malmis, C.C., Lee, G.C., Lee, L.C., Shaw, J.F., 2010. Site-specific saturation mutagenesis on residues 132 and 450 of Candida rugosa LIP2 enhances catalytic efficiency and alters substrate specificity in various chain lengths of triglycerides and esters. J Agr Food Chem 58, 10899-10905.

Zelena, K., Krugener, S., Lunkenbein, S., Zorn, H., Berger, R.G., 2009. Functional expression of the lipase gene Lip2 of Pleurotus sapidus in Escherichia coli. Biotechnol Lett 31, 395-401.

Zhang, X., Zhang, Y., Yang, G., Xie, Y., Xu, L., An, J. et al., 2016. Modulation of the thermostability and substrate specificity of Candida rugosa lipase1 by altering the acyl-binding residue Gly414 at the a-helix-connecting bend. Enzyme Microb Tech 82, 34-41.

Zhao, W., Wang, J.W., Deng, R.Q., Wang, X.Z., 2008. Scale-up fermentation of recombinant Candida rugosa lipase expressed in Pichia pastoris using the GAP promoter. J Ind Microbiol Biot 35, 189-195.

Zorn, H., Bouws, H., Takenberg, M., Nimtz, M., Getzlaff, R., Breithaupt, D.E. et al., 2005. An extracellular carboxylesterase from the basidiomycete Pleurotus sapidus hydrolyses xanthophyll esters. Biol Chem 386, 435-440.

936937

938939

940941942

943944945

946947

948949950

951952953

954955956957

958959960

961962963964

965966967

968969970971

Figure captions

Figure 1. Sequence alignment of the characterized enzymes from the C. rugosa-like family.

Consensus residues at a given position are highlighted in blue. C. rugosa lipases 1, 2, 3 and 4

(CRL1-4), O. piceae sterol esterase (OPE), M. albomyces sterol esterase (STE1), T. reesei lipase

(Trire2), A. niger lipase (Aspni5), N. haematococca lipase (Necha2), P. sapidus lipase (Lip2),

B. cinerea lipase (BCL1), G. candidum lipases 1 and 2 (GCL1 and 2). In a red boxes are

highlighted the lid region (between the two cysteines, positions 97-141), the conserved residues

from the catalytic elbow (GESAG, positions 251-255), the other two residues from the catalytic

triad (Gln at position 389 and His at position 503) and the conserved N-glycosylation site (Asn

at position 399).

Figure 2. Common structural features of the members of the C. rugosa-like family. 3D structure

of the lipase 3 from C. rugosa (CRL3) is presented as model. A) Cartoon representations of

CRL3. B) Surface representation of CRL3. The core of the protein is formed by with

antiparallel β-sheets that form the α/β-hydrolase fold, the α-helices forming the lid is

highlighted in magenta, and the intramolecular tunnel is represented by a mesh. The residue

depicted in the centre of the protein corresponds to the catalytic Ser.

Figure 3. Models of the intramolecular tunnels modeled using Caver 2.0 in the C. rugosa-like

lipases. Cartoon and surface representations are presented for all the proteins, the lid of each

enzyme is highlighted in magenta, and the intramolecular tunnels are represented by a mesh. A)

C. rugosa lipase 3 (CRL3), B) O. piceae sterol esterase (OPE), C) M. albomyces sterol esterase

(STE1), D) T. reesei lipase (Trire2), E) A. niger lipase (Aspni5), F) N. haematococca lipase

(Necha2), G) P. sapidus lipase, H) B. cinerea lipase (BCL1), I) G. candidum lipase 2 (GCL2).

Figure 4. Case study: O. piceae sterol esterase (OPE). A) Detail on the structure of OPE with a

large internal tunnel that exits opposite to the lid region near a glycosylation site. The lid

undergoes a great displacement (~30 Å) from its closed to its open conformation (Gutiérrez-

Fernández et al., 2014). B) Heterologous expression of OPE in E. coli renders a non-active

unfolded protein as shown by circular dichroism (far UV spectrum) (Vaquero et al., 2015b),

while expression in P. pastoris increases its catalytic properties due to a modification of the N-

terminus of the protein. C) Introduction of a bulky residue (Ile544Trp) at the end of the internal

tunnel of OPE modified its substrate specificity. The activity of OPE against p-nitrophenyl

esters of different acyl-chain length was modified, abolishing its activity against these long-

chain fatty acid esters. D) Monomers association in CRL3 and OPE to form a dimer in their

active form.

972973974

975

976

977

978

979

980

981

982

983

984

985

986

987

988

989

990

991

992

993

994

995

996

997

998

999

1000

1001

1002

1003

1004

1005

1006

1007

1008

Organism Habitat % id* Similarity % aac Reference

C. rugosa Soil 79-88a 88-89b 534 Lotti et al., 1994

G. candidum Cosmopolitan 41 57 544 Bertolini et al., 1995

O. piceae Wood 44 62 537 Calero-Rueda et al., 2009

M. albomyces Soil 43 58 558 Kontkanen et al., 2006a

B. cinerea Vine pathogen 46 61 558 Comménil et al., 1999

P. sapidus Wood pathogen 42 61 528 Zorn et al., 2005

T. reesei Soil 41 54 551 Vaquero et al., 2015a

N. haematococca Ubiquitous 40 54 570 Vaquero et al., 2015a

A. niger Soil 41 57 544 Vaquero et al., 2015a* Sequence identity with C. rugosa lipases a Sequence identity among the isoenzymes CRL1-4 b Structural similarity among the isoenzymes CRL1-4 c Number of amino acids in the mature protein

Table I. Known lipases/sterol esterases belonging to the C. rugosa-like family1009

1010

1011

1012

Fig 11013

1014

1015

Fig 21016

1017

1018

1019

1020

Fig 31021

1022

1023

1024

Fig 4102510261027

1028