Lytic Polysaccharide Monooxygenases (LPMOs) in ... - ACS Publications

optimally harness the potential of LPMOs in biomass processing, given the ..... stopped-flow spectrophotometry.80 They showed that Cu(I) reoxidation t...
2 downloads 0 Views 2MB Size
Subscriber access provided by ROBERT GORDON UNIVERSITY

Review

Lytic polysaccharide monooxygenases (LPMOs) in enzymatic processing of lignocellulosic biomass Piotr Chylenski, Bastien Bissaro, Morten Sørlie, Åsmund K. Røhr, Aniko Varnai, Svein J. Horn, and Vincent G.H. Eijsink ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.9b00246 • Publication Date (Web): 22 Apr 2019 Downloaded from http://pubs.acs.org on April 22, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 59 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 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

ACS Catalysis

1

Lytic Polysaccharide Monooxygenases (LPMOs) in Enzymatic

2

Processing of Lignocellulosic Biomass

3 4

Piotr Chylenski, Bastien Bissaro, Morten Sørlie, Åsmund K. Røhr, Anikó Várnai, Svein J. Horn

5

& Vincent G.H. Eijsink*

6 7

Norwegian University of Life Sciences (NMBU), Faculty of Chemistry, Biotechnology and Food

8

Science, P.O. Box 5003, N-1432 Ås, Norway

9 10

*For

correspondence; E-mail: [email protected]

11 12 13

Abstract: The discovery of lytic polysaccharide monooxygenases (LPMOs) has revolutionized

14

enzymatic processing of polysaccharides, in particular recalcitrant insoluble polysaccharides such

15

as cellulose. These monocopper enzymes display intriguing and unprecedented catalytic

16

chemistry, which make them highly valuable in industrial bioprocessing, but also generate

17

considerable challenges in terms of scientific understanding and optimal implementation. One

18

issue of particular interest is the fact that both molecular oxygen and hydrogen peroxide can drive

19

LPMO reactions. Here we review recent insights into the catalytic mechanism of LPMOs derived

20

from structural, spectroscopic and functional studies. We then turn to the question of how one can

21

optimally harness the potential of LPMOs in biomass processing, given the current knowledge of

22

their catalytic mechanism. Finally, we review recent, more applied studies that have addressed the

23

importance of LPMOs in enzymatic conversion of lignocellulosic biomass, and discuss how the

24

impact of these powerful enzymes could be improved.

25

1 ACS Paragon Plus Environment

ACS Catalysis 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 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

26

Keywords: LPMO, cellulose, copper, biomass, hydrogen peroxide, monooxygenase,

27

peroxygenase

28

2 ACS Paragon Plus Environment

Page 2 of 59

Page 3 of 59 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 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

ACS Catalysis

29

1. Introduction

30

The discovery of oxidative cleavage of polysaccharides in 2010 by enzymes today referred to as

31

Lytic Polysaccharide Monooxygenases (LPMOs; sometimes referred to as PMOs) shed light on

32

two intriguing earlier findings.1 In 1974, Eriksson et al. showed that the degradation of cellulose

33

by enzymes present in the supernatants of cultures of cellulose-degrading fungi was more efficient

34

if the reactions were conducted under aerobic conditions, leading them to suggest that oxidative

35

processes played a role.2 In 2005, Vaaje-Kolstad et al. showed that proteins known as “chitin-

36

binding proteins” or family 33 carbohydrate-binding modules (CBM33s) boost the efficiency of

37

canonical hydrolytic enzymes involved in chitin degradation, namely chitinases.3 Then, in 2010,

38

Vaaje-Kolstad showed that these chitin-binding proteins are enzymes that use reducing power and

39

O2 to oxidatively cleave glycosidic bonds in chitin.1 In the meantime, it had been shown that

40

proteins at the time classified as family 61 glycoside hydrolases (GH61) are structurally similar to

41

CBM33s4 and that they boost the activity of cellulases.5,6 Indeed, in 2011 oxidative cleavage of

42

cellulose by CBM33s7 and GH61s8-11 was demonstrated.

43

Following the initial discovery of chitin- and cellulose-active LPMOs, various LPMOs

44

were found to be active on other substrates, namely soluble cello-oligosaccharides,12 various

45

hemicelluloses13-15 and starch.16,17 The discovery of the catalytic function of CBM33s and GH61s

46

led to their re-classification in the Carbohydrate Active enZymes (CAZy) database as auxiliary

47

activities, belonging to family 10 (AA10) and 9 (AA9), respectively.18 The CAZy database, where

48

proteins are classified according to sequence similarity,19 currently contains five additional LPMO

49

families: AA11, AA13, AA14, AA15 and AA16. It is worth noting that LPMOs belonging to

50

different families are not necessarily very different in terms of function. For example, chitin-active

51

LPMOs appear in families AA10, AA11 and AA15 and while sequence similarities generally are

52

low (hence multiple families in CAZy), most of these chitin-active LPMOs can be identified by a

53

single hidden markov model (PF03067).20

54

LPMOs are monocopper enzymes (Figure 1) that bind copper in a characteristic “histidine

55

brace”,9, 21 similar to that of the particulate methane monooxygenase (pMMO).22,23 Initially, there

56

was some confusion about the nature of the catalytic metal,1,6 which is likely due to the ubiquitous

57

presence of minor amounts of copper combined with the high affinity of LPMOs for this metal

58

ion.9,24 The original reaction scheme for the LPMO reaction entails that each catalytic cycle

59

requires two externally delivered electrons and one molecule of O2 (Figure 2A) and several 3 ACS Paragon Plus Environment

ACS Catalysis 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 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

60

possible catalytic mechanisms have been proposed (see below).8,25,26 In Nature, electrons may be

61

delivered by a wide variety of small molecule reductants as well as by flavoproteins such as

62

cellobiose dehydrogenase (see below).1,11,27-29

63

Importantly, in 2016, Bissaro et al. showed that LPMOs can use H2O2 as co-substrate and

64

that controlled supply of H2O2 leads to reaction rates that are orders of magnitude higher than rates

65

commonly observed in O2-driven reactions.30 In the H2O2 reaction scheme (Figure 2B), a priming

66

reduction of the LPMO, from the Cu(II) to the Cu(I) form is followed by multiple catalytic cycles

67

using H2O2 as co-substrate.31 In this reaction scheme, the consumption of reductant is sub-

68

stoichiometric relative to the amount of generated products, whereas the O2-driven reaction

69

requires stoichiometric amounts of reductant. Bissaro et al. have suggested that O2-driven reactions

70

may not occur at all, and that the (low) rates observed for O2-driven reactions reflect the rate of

71

H2O2 formation in the reaction mixture,31,32 but this remains controversial.33 Nonetheless, the claim

72

that LPMOs can be made to run much faster than previously observed by supply of H2O2 has been

73

confirmed by multiple laboratories.32,34-39 Clearly, next to scientific implications, these recent

74

findings may have implications for the application of LPMOs in biomass processing, as discussed

75

below.

76

LPMOs are able to act on the surfaces of insoluble substrates, thus improving the

77

accessibility for canonical hydrolases (e.g. chitinases and cellulases) in the most recalcitrant parts

78

of the substrate that otherwise would have been degraded much more slowly or not be degraded at

79

all.40-43 While the boosting effect of LPMOs on the activity of hydrolytic enzyme cocktails varies

80

in magnitude, these enzymes have already found their way to industrial application. LPMOs are

81

part of modern commercial cellulose cocktails for biomass processing44,45 and it is well

82

documented that they make a considerable contribution to the efficiency of these cocktails.37,46-50

83

Since LPMOs are of major scientific and industrial interest it is worthwhile taking a closer

84

look at how they work and how they best can be applied.

85 86 87

2. LPMO Catalysis

88

The exact nature of the catalytic mechanism of LPMOs remains a subject of some controversy,

89

especially in the light of the recent discovery of the peroxygenase activity of LPMOs. An in-depth

90

discussion of structural features and putative mechanistic routes of LPMOs is beyond the scope of 4 ACS Paragon Plus Environment

Page 4 of 59

Page 5 of 59 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 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

ACS Catalysis

91

the present review, and these aspects have been extensively covered in previous

92

reviews.25,26,32,40,51-56 However, it is necessary to mention the key structural and mechanistic

93

characteristics pertaining to LPMO catalysis, particularly in the light of recent mechanistic

94

discoveries and their implications for biomass processing.

95 96

2.1 Uniqueness of the LPMO structure

97

The overall structure of LPMOs makes these enzymes uniquely suited for performing catalysis on

98

polysaccharides that are embedded in the crystalline lattice and otherwise unavailable for attack

99

by canonical glycoside hydrolases. Despite low sequence identity between LPMOs, both within

100

and in between families, all LPMOs share the key feature of having a relatively flat, solvent-

101

exposed substrate-binding surface that includes two conserved histidines that coordinate the

102

catalytically crucial single copper atom in a structural arrangement known as a “histidine brace”

103

(Figure 1C and D).9,21,54,57 A similar T-shaped histidine brace coordinating a single copper atom

104

also occurs in the copper transport protein CopC58 and in particulate methane monooxygenases

105

(pMMOs, Figure 1E).23 The surroundings of the histidines in the catalytic centers of LPMOs vary

106

both among and within LPMO families.53,54 In LPMOs from AA families 9, 11 and 13 the

107

relatively buried proximal axial copper coordination position (Figure 3) is occupied by tyrosine,

108

whereas in most AA10 LPMOs phenylalanine occupies this position, although tyrosine also

109

occurs.59 In fungal LPMOs, the N-terminal histidine is post-translationally methylated at Nε2

110

(Figure 1C), and current data suggests that this modification has little effect on catalytic properties

111

but may help protecting the enzyme from auto-catalytic oxidative damage.60 The core of the LPMO

112

structure has an immunoglobulin- or fibronectin type III-like structure consisting of a distorted β-

113

sandwich fold of typically 8–10 β-strands, connected by several helices and loops that generate

114

structural diversity among LPMOs and varying topologies of the substrate binding surface and the

115

catalytic center.25,53,54,61,62

116

5 ACS Paragon Plus Environment

ACS Catalysis 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 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

117 118

Figure 1. Structural features of a fungal and a bacterial LPMO and of pMMO. Panels A and

119

B show cartoon representations of TaAA9A (PDB ID code 2YET), whereas panel C-E shows the

120

copper binding active sites of TaAA9A, ScAA10C (PDB ID code 4OY7), and Methylocystis sp.

121

pMMO (PDB ID code 3RFR). In panels A and B, the side chains of residues that are part of the

122

(putative) substrate-binding surface of TaAA9A are shown with cyan carbons, whereas residues

123

that are also shown in panel C have green carbons. Copper atoms are shown as orange spheres.

124

Close interactions with the copper ion are shown as sticks, with distances in Å, whereas dashed

125

magenta lines indicate other relevant (potential) contacts.

126 127 128 129 130 131 132 6 ACS Paragon Plus Environment

Page 6 of 59

Page 7 of 59 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 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

ACS Catalysis

133 134

Figure 2. Reaction scheme for LPMO catalytic scenarios. Panel A shows the scheme of a

135

monooxygenase reaction originally proposed by Vaaje-Kolstad et al. in 2010.1 Panel B shows the

136

scheme of a peroxygenase reaction proposed by Bissaro et al.31 Note that the two scenarios differ

137

in terms of the consumption of reductant, which is stoichiometric, relative to the amount of

138

products formed, in scheme A and sub-stoichiometric in scheme B.

139 140 141

LPMOs have varying substrate specificities and vary in their oxidative regioselectivity.

142

LPMOs may exclusively oxidize either the C1 or the C4 carbon in the scissile glycosidic bond,

143

whereas others produce mixtures of C1- and C4-oxidized products (Figure 4).12,63 Similar

144

variations have been found for LPMOs acting on xyloglucan, whereas for the abundantly described

145

chitin-active LPMOs only C1-oxidized products have been detected so far. In spite of attempts to

146

unravel the determining factors of oxidative regioselectivity59,63,66 and substrate specificity of

147

LPMOs,66,67 these factors remain largely unknown. Current data do indicate that productive

148

binding of substrate depends on multiple interactions involving a larger part of the substrate-

149

binding surface interacting with multiple polysaccharide chains in the substrate.24,39,62 68

7 ACS Paragon Plus Environment

ACS Catalysis 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 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

150 151

Figure 3. Schematic overview geometries referred to in the text. (A) The typical LPMO

152

active site with three nitrogen atoms from the histidine brace coordinating the copper ion. One of

153

the axial coordinating positions, the one pointing towards the protein core, also referred to as

154

proximal, is usually blocked by a Tyr-OH group or a Phe side chain. Possible interaction sites for

155

other molecules at the equatorial (eq) or (distal) axial (ax) position are indicated. (B) Example of

156

superoxide that is equatorially bound in a side-on conformation to copper. (C) Example of

157

superoxide that is equatorially bound in an end-on conformation to copper.

8 ACS Paragon Plus Environment

Page 8 of 59

Page 9 of 59 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 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

ACS Catalysis

158

159 160 161

Figure 4. Oxidized sugar moieties produced during LPMO-catalyzed degradation of

162

cellulose. The primary oxidation products are a lactone or a 4-ketoaldose, which are in equilibrium

163

with their hydrated forms, an aldonic acid and a 4-gemdiol, respectively.

164 165 166 167 168

2.2 Catalytic mechanism

169

Several possible mechanisms for LPMO catalysis have been proposed and reviewed

170

(Figures 2 and 5; see below).25,26,31 The first experimental data shedding light on the key players

171

of the reaction was described by Vaaje-Kolstad et al.1 In this study, the chitin-active CBP21

172

(SmAA10A) was incubated with -chitin in the presence of an external reductant and molecular

173

oxygen (Figure 2A). Analysis of end products of the reaction using mass spectrometry and HPLC

174

revealed the formation of aldonic acids with a degree of polymerization (DP) of two and higher.

175

Experiments in presence of

176

increased mass of 2 Da compared to the standard setup of

177

suggested that LPMOs use activated oxygen to perform oxidation of a C1-carbon of an N-

178

acetylglucosamine moiety in the chitin chain and that a water molecule takes part in the formation

179

of the aldonic acid. Using similar methods, similar conclusions were reached by Beeson et al. for

180

a cellulose-active AA9 LPMO.69 Since then, most mechanistic studies have been centered around

181

the role of oxygen as co-substrate, and on identification of the key oxidative oxygen species.

18O

16O

2/H2

and

16O

18O

2/H2

both yielded aldonic acids with an 16O

16O.

2/H2

These results clearly

182

In 2012, Li et al. described the crystal structure of an LPMO from Neurospora crassa with

183

an ellipsoid-shaped weak electron density peak at the solvent-exposed distal axial copper 9 ACS Paragon Plus Environment

ACS Catalysis 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 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

Page 10 of 59

184

coordination position with an end-on (1) configuration to the copper ion.68 The electron density

185

was modeled as a dioxygen species with an unrestrained bond length of 1.16 Å. The authors

186

proposed this to be consistent with a superoxide species weakly coordinated to Cu(II), as bond

187

lengths for oxygen or superoxide are 1.2-1.3 Å, whereas the bond length in a peroxide would be

188

close to 1.49 Å. Since then, the possibility of axial oxygen activation has been explored further by

189

Kim et al. (see below),70 but recent work suggests that oxygen activation happens in the equatorial

190

plane (Figure 3).57,71-74

191

In a 2013 review, Hemsworth et al. suggested that Cu(III)-OH (or possibly Cu-(III)-

192

peroxide or “cupryl”) may be the reactive form of oxygen.75 One reason was the isolation of a

193

Cu(III)-OH

194

diisopropylphenyl)-2,6-pyridinedicarboxamideCuOH]) with a N3 T-shape coordination geometry

195

of the copper, akin to the histidine brace.76 The other reason was that the same active species had

196

been discussed in relation to C-H bond activation in methane by particulate methane

197

monooxygenase (pMMO).77 In relation to this, it is worth noting that it has been suggested that the

198

N-terminal amino group can be deprotonated.72 Such deprotonation could possibly facilitate

199

formation of a Cu(III) intermediate.

intermediate

of

a

copper

model

compound

([Bu4N][(N,N´-bis(2,6-

200

Kim et al. used the structure of an LPMO from Thermoascus aurantiacus (TaAA9A;

201

Protein Data Bank, PDB ID code 2YET, Figure 1) to build an active site cluster model and

202

performed density functional theory (DFT) calculations to compare two possible reaction paths for

203

hydrogen abstraction.70 One tested mechanism employed a η1-superoxo intermediate ([LPMO-

204

Cu(II)-OO) which abstracts a substrate hydrogen, while the other alternative entailed formation

205

of a copper-oxyl radical (LPMO-Cu(II)-O) that abstracts a hydrogen and subsequently

206

hydroxylates the substrate via an oxygen-rebound mechanism (Figure 5). The results predicted

207

that oxygen binds end-on (η1) to copper at the axial position, and that a copper-oxyl–mediated,

208

oxygen-rebound mechanism is energetically preferred. While axial binding of the oxygen is not

209

supported by subsequent studies, the formation of an oxyl intermediate is (Figure 5).

210 211 212 213

10 ACS Paragon Plus Environment

Page 11 of 59 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 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

ACS Catalysis

214 215 216

Figure 5. Schematic summaries of proposed mechanisms for hydrogen atom abstraction by

217

an LPMO using O2 (top) or H2O2 (bottom) as co-substrate. Note that alternative scenarios have

218

been proposed for both O2- (e.g. refs 25-26, 55) and H2O2-driven (ref 31) catalysis. Also note that,

219

while scenarios for H2O2-driven catalysis generally imply that the copper stays reduced in between

220

catalytic cycles, this is not the case for all proposed scenarios for O2-driven catalysis, some of

221

which entail that the LPMO ends up in the oxidized from at the end of one cycle. The copper-oxyl

222

intermediate (LPMO-Cu(II)-O), which is a favoured intermediate state in most modelling studies

223

of catalysis, also could be a copper-oxo (LPMO-Cu(III)=O), which bears resemblance to

224

Compound I formed by cytochrome P450s, discussed below. See text for more details. These

225

schemes are based on the work of Bertini et al.78 and Wang et al.79

226 227 11 ACS Paragon Plus Environment

ACS Catalysis 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 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

228 229

Kjærgård et al. investigated O2 reactivity with the same LPMO, TaAA9A, using EPR and

230

stopped-flow spectrophotometry.80 They showed that Cu(I) reoxidation takes place with a

231

minimum rate of > 0.15 s-1. Based upon reported redox potentials of ∼275 mV vs. the normal

232

hydrogen electrode (NHE)24,61 and the potential of the one-electron reduction of O2 to superoxide

233

(−165 mV vs. NHE), the rate of an outer-sphere electron transfer was calculated to be ∼4.5 × 10−4

234

s−1, using Marcus theory, which was ∼103 slower than the rate of Cu(I) reoxidation derived from

235

the EPR and stopped-flow data. The authors concluded that the single-electron transfer from

236

LPMO-Cu(I) to O2 was likely to proceed via an inner-sphere pathway involving rapid formation

237

of LPMO-Cu(II)-OO. Further, the coordination distances of the copper N-atom ligands were

238

derived from K-edge XANES (X-ray Absorption Near Edge Structure) and EXAFS (Extended X-

239

Ray Absorption Fine Structure) experiments, allowing for comparison with the geometry-

240

optimized active site models. Addition of an O2 molecule to the optimized structure with a reduced

241

copper resulted in a superoxide bound equatorially to the Cu(II) ion in an end-on fashion (Figure

242

3). According to the authors, it was not possible to stabilize a side-on bound Cu-O2 structure in the

243

AA9 site.

244

EPR spectroscopy studies by Borisova et al. showed that substrate binding leads to altered

245

g-values and additional superhyperfine coupling patterns, and these authors suggested that the

246

observed changes in the spectrum originated from a change in water coordination of the copper

247

upon substrate binding.81 This latter hypothesis was strengthened by spectroscopic and

248

computational results in a recent study of the chitin-active LPMO SmAA10A.39 Crystal structures

249

of LPMO-oligosaccharide complexes described in a seminal paper by Frandsen et al. showed a

250

chloride ion, considered to act as a superoxide analogue, bound in the equatorial position (Cu-Cl,

251

2.3 Å).57 Frandsen et al.57 observed changes in EPR spectra similar to what was observed by

252

Borisova et al.81 and suggested that this was due to the presence of a chloride nucleus leading to a

253

significant decrease in the observed gz value of the EPR spectra (from gz = 2.28 in the absence of

254

substrate, to 2.27 in the presence of substrate and low chloride concentrations, to 2.23 in the

255

presence of substrate and high chloride concentrations). It should be noted, however, that Borisova

256

et al. observed the same type of shift at low chloride concentrations.81 Importantly, the structures

257

of the complexes described by Frandsen et al. showed that the C6 hydroxymethyl group of the

258

glucose unit bound to subsite +1 was close to the copper ion (C-Cu, 3.8 Å) blocking access to the 12 ACS Paragon Plus Environment

Page 12 of 59

Page 13 of 59 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 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

ACS Catalysis

259

axial copper binding site.57 Frandsen et al. also pointed at the potential interdependence of binding

260

of substrate and chloride to the copper ion, which would imply that binding of the oxygen co-

261

substrate and a polysaccharide substrate may act in a concerted manner. In this way, production of

262

the reactive oxygen species may be controlled by the presence of a substrate.57,82

263

In a recent study, Bissaro et al. showed that LPMO reactions can be driven by H2O2.30,31

264

By controlling H2O2 supply, stable and fast reaction kinetics were achieved and the LPMOs

265

worked in the absence of O2, whereas the reductant was consumed in amounts that were sub-

266

stoichiometric relative to products formed (Figure 2). Experiments with labeled hydrogen peroxide

267

(H218O2) in the presence of a ten-fold molar surplus of O2, showed that the oxygen atom that is

268

introduced into the polysaccharide chain came from H2O2 and not O2 (Figure 6). Introduction of

269

18O

270

and chitin substrates (Figure 6). The authors suggested several possible reaction mechanisms,

271

including mechanisms involving hydroxyl radicals. One plausible mechanism entails the

272

conversion of H2O2 by LPMO-Cu(I) to a water molecule and a Cu(II)-O intermediate, which also

273

can be a Cu(III)=O. From here on, the reaction would proceed via a rebound mechanism, as

274

described above (Figure 5), implying hydrogen abstraction from the substrate, followed by

275

merging of the resulting Cu(II)-associated hydroxide with the substrate radical, leading to

276

hydroxylation of the substrate and regeneration of the Cu(I)-center ready for the next cycle. The

277

most recent studies on the catalytic mechanism of LPMOs have taken the possibility of H2O2-

278

driven LPMO reactions into account, albeit to varying extents.

from H218O2 into the final product was shown for both AA9s and AA10s and for both cellulose

279

13 ACS Paragon Plus Environment

ACS Catalysis 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 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

280 281

Figure 6. MALDI-ToF analysis of products generated in competition experiments with O2

282

and labeled H2O2. (A) MALDI-ToF MS spectrum showing products (DP6 cluster) released upon

283

incubation of Avicel (10 g.L-1) with ScAA10C (0.5 µM) in presence of a low amount of ascorbic

284

acid (AscA; 10 µM), O2 (200-250 M) and varying concentrations of H218O2 (25-100 M). The

285

reactions were started by addition of AscA. (B-D) MALDI-ToF MS spectra showing products

286

released from Avicel by ScAA10C (0.5 µM) after 4 min reaction (B), from PASC by PcAA9D (1

287

µM) after 15 min reaction (C) and from β-chitin by SmAA10A (0.5 µM) after 60 minutes of 14 ACS Paragon Plus Environment

Page 14 of 59

Page 15 of 59 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 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

ACS Catalysis

288

reaction (D). All reactions were carried out under normal aerobic conditions, which means that the

289

concentration of (non-labeled) 16O2 in solution was in the range of 200-250 µM. Reactions shown

290

in panel B were carried out in the presence of 100 µM H216O2 (purple line) or H218O2 (orange line)

291

and 1 mM AscA. Reactions shown in panel C were carried out in the presence of 200 µM H216O2

292

(purple line) or H218O2 (orange line) and 100 µM AscA. Reactions shown in panel D were carried

293

out in the presence of 100 µM H216O2 (purple line) or H218O2 (orange line) and 10 µM AscA. The

294

spectra show that when using H218O2, the characteristic signals for sodium adducts of the aldonic

295

acid form of an oxidized cellohexaose (m/z 1029.7 & 1051.7) or chitohexaose (m/z 1076.0 &

296

1298.0) shift by +2 Da. Abbreviations: DP, degree of polymerization; Nat, native; Lac, oxidized,

297

lactone form; Ald, oxidized, aldonic acid form. Nb. All MS spectra show sodium adducts of the

298

native (Nat), lactone (Lac) and the aldonic acid (Ald) form for the DP6 cluster. Adapted with

299

permission from reference 31. Copyright 2017, Springer Nature.

300 301 302 303

Bertini et al. used DFT calculations on a large active-site model for a fungal AA9 LPMO

304

and with a celloheptaose unit as a substrate mimic to calculate the energies of possible LPMO

305

mechanisms.78 A key finding was that binding of O2 to the T-shaped LPMO-Cu(I) active site

306

resulted in formation of a LPMO-Cu(II)-OO intermediate and a distorted tetrahedral geometry of

307

the Cu atom, a conformation that has not been observed in other computational studies. The

308

presence of the substrate did not change this geometry, but O2 binding was further favored with

309

4.0 kcal/mol. Starting at the LPMO-Cu(II)-OO superoxo intermediate, Bertini et al. then

310

calculated energies for several possible reaction mechanisms.78 The most favored mechanism

311

involved a proton coupled electron transfer to form LPMO-Cu(II)-OOH in the presence of the

312

substrate followed by a second coupled electron transfer and a loss of water to form a LPMO-

313

Cu(II)-O intermediate, that abstracts a hydrogen from the substrate (Figure 5). The final steps of

314

the proposed mechanism are similar to what has been described earlier by Kim et al., i.e. employing

315

an oxygen-rebound mechanism.70 The authors indicated that oxidation of the substrate by H2O2

316

via a Cu(II)-O intermediate would also be possible.

317

Wang et al.79 explicitly addressed H2O2-dependent catalysis by LPMOs, testing reaction

318

pathways suggested by Bissaro et al.31 They combined small model DFT calculations, classical 15 ACS Paragon Plus Environment

ACS Catalysis 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 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

319

molecular dynamics simulations, and quantum mechanical/molecular mechanical calculations to

320

detail the mechanism of a C4 oxidizing LPMO in the presence of H2O2 and cellotriose as model

321

substrate.79 Their key findings were that there is an efficient mechanism to break the O–O bond in

322

H2O2 via a one-electron transfer from the LPMO-Cu(I) to form an HO• radical and a Cu(II)-OH

323

species. This radical is stabilized by hydrogen bonding interactions with the enzyme. Moreover,

324

the calculations showed that the formed radical preferred to abstract a hydrogen atom from the

325

Cu(II)-OH species, to form a Cu(II)-O species, rather than directly abstracting a hydrogen from

326

the substrate. The authors thus concluded that it is the Cu(II)-O species that oxidizes the C4 carbon

327

in the scissile glycosidic bond. It is worth noting that the formation of hydroxyl radicals through

328

the reaction of H2O2 with reduced transition metals is well known from biomass conversion, where

329

the Fenton reaction, usually involving iron, is thought to play a role in the degradation of

330

lignocellulosic material by brown-rot fungi.83 While the Fenton reaction seems rather unspecific,

331

LPMOs may have found a way to harness, control and direct the power of hydroxyl radicals within

332

the confinement of an enzyme-substrate complex.30,31,79 The formation of such powerful oxidative

333

species brings the risk of auto-catalytic damage to the enzymes, as is discussed in section 2.7.

334

Important modelling studies were conducted by Hedegård and Ryde.73,74 In their 2017

335

study, they calculated bond-dissociation energies (BDE) for a number of possible LPMO

336

intermediates to determine if these are sufficiently high to activate the C1-H or C4-H bonds in

337

cellulose, with calculated BDEs of 423 and 434 kJ/mol, respectively.73 The calculated BDEs for

338

Cu(II)-oxyl or a Cu(III)-oxo suggested that the reaction with C1-H or C4-H bonds will result in

339

reaction energies of −25 and −34 kJ/mol, respectively. The reaction with a Cu(III)-hydroxide was

340

47 kJ/mol, but this species was nevertheless considered as a possible candidate for hydrogen atom

341

abstraction. This study also suggested that O–O bond breaking occurs prior to hydrogen

342

abstraction from the substrate.73

343

In their 2018 follow-up study, Hedegård and Ryde employed the same crystal structure of

344

an LPMO-cello-oligosaccharide complex as used in the study by Wang et al., discussed above.74,79

345

They used a QM/MM approach to analyze the catalytic mechanism and identify the species

346

abstracting a hydrogen atom from the polysaccharide substrate. Calculations were performed for

347

both O2 and H2O2 as co-substrates. The calculations showed that when O2 is the co-substrate, a

348

Cu(II)-OO complex is formed upon reaction of Cu(I) with O2 (as shown by Kjærgaard et al.80)

349

and that protonation of this complex by a nearby histidine residue with a concomitant electron 16 ACS Paragon Plus Environment

Page 16 of 59

Page 17 of 59 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 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

ACS Catalysis

350

transfer leads to cleavage of the O–O bond, dissociation of water and formation of an Cu(II)-O

351

intermediate, possibly protonated, that can react with substrate. Analyses with H2O2 as the co-

352

substrate showed that the reaction with LPMO-Cu(I) led to formation of an oxyl or hydroxyl

353

complex, both sufficiently reactive to abstract a hydrogen from the polysaccharide substrate.

354

Comparative calculations showed that the reaction is more favorable with H2O2 as the co-substrate

355

compared to O2. H2O2 was found to interact with the proton of a nearby histidine as well as a with

356

a nearby glutamine residue,74 in accordance with structural data.71

357

All in all, most studies indicate that Cu(II)-oxyl or a closely related species (Cu(III)=O or

358

a protonated oxyl) is the reactive oxygen species and that this species may be generated in both

359

O2- and H2O2-driven mechanisms. For a further discussion of the relevance of these two

360

mechanisms, see section 2.8.

361 362

2.3 Kinetics of LPMO action

363

Despite the increasing number of studies published on LPMOs, detailed kinetic analyses of LPMO

364

action remain scarce (see Table 1 for a summary of kinetic data). Vaaje-Kolstad et al. reported an

365

oxidation rate of 1 min-1 (0.017 s-1) for degradation of -chitin by (bacterial) CBP21 (SmAA10A)

366

in reactions with O2 as co-substrate.1 NcAA9C is a (fungal) AA9 enzyme showing activity on

367

xyloglucan, cellulose and cellodextrins. Again, with O2 as the co-substrate, measured degradation

368

rates for NcAA9C were 0.11 s-1 for xyloglucan and 0.06 s-1 and 0.03 s-1 for an oligomeric

369

xyloglucan and cellulose substrate, respectively.13 Similar values were obtained by Borisova et al.

370

for the same LPMO.81 The pathogen Vibrio cholerae expresses a four-domain AA10-type LPMO

371

(GbpA), which is a virulence factor and not likely involved in biomass processing. Loose et al.

372

showed that the initial reaction rate for GbpA on -chitin nanofibers was 2.7 min-1 (0.045 s-1).84

373

Frandsen et al. analyzed the kinetics of LsAA9A-catalyzed oxidation of a soluble oligomeric

374

substrate analogue of (Glc)4.57 The results were interpreted using a classical Michaelis – Menten

375

approach, to yield a kcat of 0.11 s-1 and a Km of 43 M with respect to the carbohydrate substrate.

376

Importantly, this study was the first to provide an efficiency constant (kcat / Km of 2.6 • 103 M-1 s-

377

1)

378

and a reductant, such as ascorbic acid, typically at a concentration of 1 mM. Generally, the

379

observed reaction rates were slow, varying from considerably below 1 s-1 to below 1 min-1, as

380

recently summarized by Bissaro et al.32

for an LPMO system.57 In all these studies, reactions were typically run in the presence of O2

17 ACS Paragon Plus Environment

ACS Catalysis 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 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

381

The first detailed kinetic characterization of H2O2–driven LPMO catalysis was done for

382

degradation of chitin by SmAA10A by Kuusk et al.35 Central to the study was the use of [14C]-

383

labeled chitin, which provided convenient and sensitive detection of released soluble products.

384

Also here, the results were interpreted using a classical Michaelis – Menten approach to yield a

385

kcat of 6.7 s-1, which is two orders of magnitude higher than previously reported (apparent) rate

386

constants for O2-driven reactions. These analyses yielded Km values of 0.58 mg/mL and 2.8 M

387

for chitin and H2O2, respectively. It is worth noting that the kcat/ Km for H2O2 is 2 • 106 M-1 s-1,

388

which is a value commonly seen for peroxygenases.85-86 Of note, the kinetic analyses by Kuusk et

389

al. suggested that LPMO-catalyzed oxidation of chitin in the presence of H2O2 follows a ternary

390

mechanism.35

391

In a subsequent study, Hangasky et al. assessed the kinetics of MtAA9E-catalyzed

392

oxidation of (Glc)6 evaluating both O2 and H2O2 as co-substrates.33 Studies with O2 at a

393

concentration of 208 M and with varying concentrations of (Glc)6, yielded an apparent kcat of

394

10.1 min-1 (0.17 s-1) and a Km of 32 M with respect to (Glc)6, resulting in a kcat/ Km of 5 • 103 s-1

395

M-1. These values are very similar to those obtained by Frandsen et al.57 in their study of LsAA9A-

396

catalyzed oxidation of (Glc)4 (see above). Another series of experiments, with a constant (Glc)6

397

concentration of 1 mM and varying O2 concentrations (0 to 800 M), yielded an apparent kcat of

398

17 min-1 (0.28 s-1) and a Km of 230 M with respect to O2, corresponding to a kcat/ Km of ca. 1 •

399

103 s-1M-1. When H2O2 was used as co-substrate, enzyme rates were much higher and rates

400

corresponding to 50% of added

402

H2O2 had been consumed, the authors calculated observed rate constants (kobs) of 285 – 916 min-1

403

(4.8 - 15 s-1) for concentrations of H2O2 ranging from 12.5 to 100 M. No attempt was made to

404

calculate a Km with respect to H2O2. A plot of kobs vs. H2O2 concentration in a Michaelis – Menten

405

plot using data in Table S9 of this study33 yields a Km of 53 M suggesting a kcat/ Km of 3 • 105 M-1

406

s-1.

407

Of note, currently available kinetic data (Table 1), derived from multiple laboratories, show

408

that H2O2-driven LPMO reactions are orders of magnitude faster and show much better catalytic

409

efficiencies (kcat/ Km), compared to O2-driven LPMO reactions. Bissaro et al. have suggested that

410

the very low rates of LPMO catalysis observed under “standard” conditions (O2 + reductant) reflect

411

the rate of the generation of the “true” LPMO substrate, H2O2, in the reaction mixtures,30-32 18 ACS Paragon Plus Environment

Page 18 of 59

Page 19 of 59 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 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

ACS Catalysis

412

whereas others claim that LPMOs can indeed use O2 directly, and that O2 is the natural co-

413

substrate.33 This current controversy is addressed in some more detail in section 2.8.

414 415 416 Table 1. Kinetics of LPMO reactionsa Enzyme IDb

Reductant

Substratec

O-sourced

kcat (s-1)

KM (M)e

kcat/KM (s-1.M1)f

Apparent oxidative rate (s-1)

Ref.

LsAA9A

AscA (5 mM)

FRET subst. (10-100 µM) Glc6 (0-400 mM) Glc6 (1 mM) Glc6 (1 mM) XG14 (0.2 mM) Glc5 (0.2 mM)

O2 (atm.)

0.11

43 / -

2.6103

-

(57)

0.17

32 / -

5103

-

(33)

0.28

- / 230

1103

-

(33)

3105 g

4.8 - 15

(33)

MtAA9E

NcAA9CCBM1

AscA (2 mM)

AscA (1 mM)

Tamarind XG (5 g·L-1)

O2 (208 µM) O2 (0-800 M) H2O2

O2 (atm.)

PASC (5 g·L-1)

SmAA10Ah

417 418 419 420 421 422 423 424 425 426 427 428 429 430 431

VcAA10BX-YCBM73i

nd

nd

nd

0.06

nd

nd

nd

0.03

nd

nd

nd

0.11

nd

nd

nd

0.11

(13)

Reduced Glutathione (1 mM)

-chitin (0.45 g·L-1)

O2 (atm.)

nd

nd

nd

0.017

(1)

AscA (100 µM)

CNW

H2O2

6.7

0.58 mg.mL-1 / 2.8 M

2106

-

(35)

AscA (1 mM)

-chitin nanofibers (5 g·L-1)

O2 (atm.)

nd

nd

nd

0.045

(84)

a

See reference 32 for a more extensive review of LPMO reaction rates and reaction conditions. Abbreviations: Ls, Lentinus similis; Nc, Neurospora crassa; Sm, Serratia marcescens; Mt, Myceliophthora thermophila (new name Thermothelomyces thermophila); Vc, Vibrio cholerae. c Abbreviatons: FRET subst. = fluorescence-labeled cellotetraose57; CNW, chitin nanowhiskers; PASC, phosphoric acid-swollen cellulose; XG, xyloglucan; XG14, xyloglucan oligomer. d O-source refers to the molecule (i.e. O or H O ) from which the oxygen used in the (per)oxygenase reaction presumably was 2 2 2 derived. Abbreviation: atm., atmospheric. e K values (in µM, unless stated otherwise) for the substrate and the O-source compound, respectively. M f Note that the K may refer to the substrate or the (oxygen containing) co-substrate. M g This value is an estimate made on the basis of data presented in Ref 33; see text for details. h Also known as CBP21 i GbpA is a four-domain protein where X and Y denote unknown domains related to the flagellin protein p5 and pili-binding chaperone FimC, respectively165. nd: not determined b

432 433

2.4 LPMOs among other oxidoreductases

19 ACS Paragon Plus Environment

ACS Catalysis 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 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

434

Another family of monooxygenases that can act as peroxygenases to activate C-H bonds are

435

cytochromes P450 (CYP). CYPs are found in all kingdoms of life and have a vast variety of

436

substrates. The active site of CYP contains a heme prosthetic group where the heme-iron is bound

437

to the protein through a cysteine thiolate ligand. In its resting state, the iron has a +3 charge. The

438

catalytic cycle begins with reductive activation of oxygen to form a ferryl–oxo with an porphyrin

439

π-cation radical known as Compound I, which is the hydroxylating species in an oxygen rebound

440

mechanism (Figure 7).87,88 Due to the vast number of CYPs and substrates, there are large

441

variations in observed catalytic rate constants for the various monooxygenase reactions, ranging

442

from 0.1 to 3 s-1.89,90

443

20 ACS Paragon Plus Environment

Page 20 of 59

Page 21 of 59 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 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

ACS Catalysis

444 445

Figure 7. A catalytic cycle of oxygen activation by cytochromes 450 and its H2O2-shunt. The

446

upper left state represents the resting state of the enzyme. Note that compound I bears resemblance

447

with the closely related Cu(III)-oxo and Cu(II)-oxyl states that may be formed during LPMO

448

catalysis. Adapted with permission from reference 87. Copyright 2014, Springer.

449 450 21 ACS Paragon Plus Environment

ACS Catalysis 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 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

451

CYPs can short-cut the generation of Compound I by using hydrogen peroxide, which

452

reacts with the CYP in its resting (i.e. non-reduced, or Fe(III)) state in a so-called H2O2-shunt

453

pathway. However, this shunt pathway is generally inefficient. As an example, naphthalene

454

oxidation by H2O2 using CYP3A4, the most abundant P450 enzyme in human liver, has a kcat of

455

1.6 min-1 (0.03 s-1) vs. 43 min-1 (0.7 s-1) for O2 and NADPH supported oxidation.91 A general

456

hypothesis used to explain this low efficiency is that CYPs lack a residue near the active site that

457

can participate in general acid-base catalysis important for the formation of Compound I in H2O2-

458

dependent oxidation by classical heme peroxidases and peroxygenases.87 Moreover, the high

459

hydrophobicity of the heme pocket of CYPs is unfavorable for the formation of Compound I via

460

the H2O2 shunt pathway (as opposed to the “normal” O2-pathway) and the subsequent

461

peroxygenation. A plethora of studies show that the peroxygenase activity of CYP can be enhanced

462

either through site-directed mutagenesis or directed evolution,91,92 but there are not many mutants

463

that show activities that are higher than those obtained in the monooxygenase reaction.87

464

Moreover, peroxide-driven reactions usually require high concentrations of peroxide that promote

465

enzyme inactivation.93-95 The need for such high concentrations is reflected in low catalytic

466

efficiency constants with respect to H2O2. The oxidation of 7-benzyloxyquinoline by CYP3A4

467

with varying concentrations of H2O2 yielded a kcat of 7.6 min-1 (0.13 s-1) and a Km of 61 mM

468

yielding a kcat/Km of 2 M-1s-1.91 In general, CYP have high Km values for H2O2 in the millimolar

469

range (20-250).96-98 Importantly, a few CYPs do seem to be genuine peroxygenases and these

470

enzymes have structural features that place a general acid-base residue near the catalytic center.99

471

These enzymes use H2O2 to catalyze the hydroxylation of long alkyl chain fatty acids with high

472

catalytic activity and substrate specificity.95,99-102 Here, the Km with respect to H2O2 has been

473

determined to be as low as 21 M.103

474

Due to the newly discovered peroxygenase activity of LPMOs, it is tempting to compare

475

this class of enzymes with CYPs. Both enzyme classes are able to activate C-H bonds with both

476

O2 and H2O2 as co-substrates. While LPMOs seem highly substrate specific, CYPs tend to be

477

promiscuous. Similar kcat values are typically observed for the two enzymatic systems when O2 is

478

the co-substrate. This changes rather drastically when H2O2 is the co-substrate, which leads to

479

increased catalytic rates and efficiency constants for LPMOs, and decreased catalytic rates and

480

efficiency constants for CYPs. The most striking characteristic is the clear difference in the Km

481

values, where LPMOs have up to a 1,000-fold higher affinity for H2O2 than CYP. Combined, this 22 ACS Paragon Plus Environment

Page 22 of 59

Page 23 of 59 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 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

ACS Catalysis

482

results in LPMOs having up to a million-fold higher catalytic efficiency constants with respect to

483

H2O2 than CYPs. In contrast to CYPs, the catalytic centers of LPMOs contain residues that may

484

contribute to H2O2-driven catalysis through hydrogen bonding and acid/base functionality (Figure

485

1).39,71,79,104

486

Another clear difference pertains to the redox state of the enzymes: while both CYP and

487

LPMOs activate O2 in their reduced state (i.e. Fe(II) and Cu(I), respectively), activation of H2O2

488

requires the LPMO to be in its reduced Cu(I) form whereas CYP is in its oxidized Fe(III) form

489

when working in shunt mode. A similar shunt pathway in LPMOs (i.e. LPMO-Cu(II) + H2O2)

490

would entail the unlikely reduction of H2O2 by Cu(II), or the formation of a Cu(II)-hydroperoxo

491

intermediate requiring deprotonation of H2O2. While such a mechanism perhaps cannot be

492

excluded, fact is that H2O2-driven catalysis by LPMOs is multiple orders of magnitude faster when

493

the LPMO is first reduced and this situation leads to stable reaction kinetics and high turnover

494

numbers.31,37

495 496 497

2.5 The role and nature of reductants

498

The reduction of LPMO-Cu(II) to LPMO-Cu(I) is generally accepted as a necessary step preceding

499

catalysis. It has been demonstrated that this reduction step can be carried out by plethora of

500

reductants, including small, organic molecules such as ascorbic acid,1 reduced glutathione,1

501

cysteine,17,27 a wide range of plant biomass or fungal phenolic compounds,27,28 lignin and its

502

fractions,6,105-108 and oxidoreductases.8,11,109 The most studied enzymatic electron donor is cellobiose dehydrogenase (CDH).8,11,28,110-

503 504

112

505

terminal heme b-binding AA8 cytochrome domain (CYT) connected by a flexible linker to a C-

506

terminal

507

AA3_1).110,113,114 Some CDHs have an additional C-terminal CBM1 domain binding to

508

cellulose.115 The DH domain is the catalytic part of the enzyme, carrying out a two-electron

509

oxidation of cellobiose and several other oligosaccharides, via the reduction of the FAD cofactor.

510

The reoxidation of the FAD cofactor may be carried out via sequential inter-domain electron

511

transfer (IET) to a CYT domain or via reduction of a two-electron acceptor, including O2.

512

Reduction of O2 leads to the formation of H2O2.116-118 The reduced CYT domain will transfer single

CDHs are bi-modular flavocytochromes, so far only identified in fungi, containing an Nflavin

adenine

dinucleotide

(FAD)-dependent

23 ACS Paragon Plus Environment

dehydrogenase

domain

(DH;

ACS Catalysis 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 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

513

electrons to appropriate acceptors, which includes the LPMOs.110-112 Even though the existence of

514

a putative CDH “docking” site on AA9 LPMOs has been proposed,68 experimental data and

515

computational modelling suggest direct electron transfer at the active site of the LPMO.110,111

516

Recently, a pyrroloquinoline quinone (PQQ)-dependent pyranose dehydrogenase (PDH)

517

from Coprinopsis cinerea (CcPDH) has been shown to activate cellulose-active AA9 LPMOs.109

518

Notably, PDH does not belong to the superfamily of glucose-methanol-choline (GMC)

519

oxidoreductase, which includes CDH. CcPDH has a three-domain structure with an N-terminal

520

heme b-binding AA8 CYT domain, a central PQQ-dependent AA12 DH domain and a C-terminal

521

CBM1.119 The CcPDH shows oxidative activity towards ᴅ-glucosone (2-keto-ᴅ-glucose), L-fucose

522

and some rare pyranoses, and, in contrast to the CDH, does not oxidize cello-oligosaccharides or

523

glucose, which makes this enzyme a promising tool for mechanistic studies of LPMOs and for

524

lignocellulose biorefining.

525

LPMO reactions may also be fueled by photocatalytic systems that provide electrons.120,121

526

In one approach, Bissaro et al. showed that light-driven oxidation of water catalyzed by vanadium-

527

doped titanium dioxide, is capable of providing reducing equivalents to LPMOs and drive LPMO

528

reactions, albeit at low rate.121 In a very important study, Cannella et al. reported that exposure of

529

plant derived pigments (e.g. chlorophyllin) to low intensity light in the presence of ascorbic acid

530

speeds up LPMO catalytic rates by up to 100-fold.120 This effect was attributed by the authors to

531

the delivery of high redox potential electrons to the LPMO from the photo-excited pigment

532

molecule. The nature of the observed enhancement of LPMO catalysis remains controversial. In

533

contrast with claims made in the ground-breaking work by Cannella et al. (see also Möllers et

534

al.122), the authors of this review have proposed that the efficiency of the chlorophyllin-light

535

system may be due to the formation of hydrogen peroxide.30-31 Importantly, the study by Cannella

536

et al. showed that LPMOs could attain much higher catalytic rates than previously thought.120

537

Kracher et al. described the correlation between the reduction potential of the reductant and

538

the catalytic performance of LPMOs.28 Using variation in pH to manipulate redox potentials of

539

reductants, Frommhagen et al. showed a similar correlation.123 Several studies on O2-driven LPMO

540

reactions have shown that higher concentrations of reductant lead to higher catalytic rates and

541

higher consumption of O2.37,112,124 However, high concentrations of reductant also lead to faster

542

inactivation of the LPMO, as discussed below.

24 ACS Paragon Plus Environment

Page 24 of 59

Page 25 of 59 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 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

ACS Catalysis

543

Considering the discrepancy between the rate of LPMO reduction (in the milliseconds

544

range; Kracher et al.28) and the reported apparent catalytic rates (in the minute range), it is unlikely

545

that reduction of LPMO-Cu(II) complex is a rate limiting step of an O2-driven LPMO reaction.

546

Thus, in the O2-driven reaction mechanism of LPMOs (Figure 2A), the delivery of the second

547

electron must be the rate-limiting step. In case of the peroxygenase mechanism (Figure 2B),

548

delivery of a second electron is not required. In the latter scenario, the generation of H2O2, either

549

via reactions between the reductant and O2 or via superoxide formation by the reduced LPMO

550

itself,80 may be the rate-limiting step. Of note, Hegnar et al. recently showed a correlation between

551

the (pH-dependent) generation of H2O2 by an LPMO-reductant combination in the absence of

552

substrate, and LPMO activity in the presence of substrate.125

553 554

2.6 The role of the substrate

555

Copper-binding studies indicate that LPMOs bind copper with nanomolar affinities and,

556

importantly, that Cu(I) binds with higher affinity compared to Cu(II).9,24 The catalytic copper site

557

of LPMOs is highly accessible (Figure 1) and it is difficult to envisage how an enzyme with

558

basically no substrate-binding pocket and no confinement around a highly reactive Cu(I) species

559

can control its reactivity and specificity. Although this issue remains partly unresolved, it is clear

560

that the substrate plays a major role,57,126 since binding to substrate, especially binding to an

561

extended substrate surface, secludes the copper from the solution and creates confinement at the

562

active center (Figure 8).39,111

563 564

Figure 8. Binding of LPMO to an extended substrate surface. Panel A shows an experimentally

565

guided model of LPMO SmAA10A bound to crystalline chitin. Panel B shows the active site in

566

the complex and the positioning of substrate hydrogen atoms relative to the copper (green and red 25 ACS Paragon Plus Environment

ACS Catalysis 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 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

567

carbons in the protein show the (almost identical) situations with and without substrate,

568

respectively). The Cu-H1 distance is ~1 Å shorter than the Cu-H4 distance, rationalizing the C1

569

oxidizing activity of this LPMO. Panel C shows the heat mapped LPMO-chitin interaction surface

570

indicating that most interactions occur along carbohydrate residues +3 to -5, all belonging to a

571

single polysaccharide chain. The residues labeled a-h belong to polysaccharide chains adjacent to

572

the central chain. Adapted with permission from reference 39. Copyright 2018, American

573

Chemical Society.

574 575

Insight into substrate-binding to LPMOs comes from NMR studies with both crystalline and

576

soluble substrates,24,111 X-ray crystallographic studies of LPMOs in complex with soluble

577

substrates (see section 2.2),57,126 and modelling.39,62,68 In addition, EPR studies provide insight into

578

how substrate binding changes the electronics of the copper site (see section 2.2).39,57,81 Of note,

579

such electronic effects of substrate binding have been observed for both cellulose and chitin. The

580

structures of enzyme-substrate complexes described in a seminal paper by Frandsen et al.57 and in

581

a follow up paper by Simmons et al.126 show how substrate-binding generates small

582

rearrangements of the copper site. Interestingly, a comparison of the binding of cello-oligomers

583

and xylo-oligomers to the same LPMO showed that the different substrates create quite different

584

environments near the copper.126 While the biological relevance of these differences is unclear, for

585

example because the xylo-oligosaccharides may not be the true substrates, they do underpin that

586

substrate-binding has an effect on the copper environment.

587

In an experimentally guided modeling study, Bissaro et al. generated models of complexes

588

between chitin-active SmAA10A and crystalline chitin (Figure 8).39 In these models the T-shaped

589

Cu-3N active site assumed a position relative to the substrate that was compatible with catalysis.

590

The models further showed that multiple conserved amino acid residues, spread over the putative

591

substrate-binding surface and interacting with multiple polysaccharide chains, are involved in

592

binding and correct positioning of the substrate. It was also shown that small molecules such as

593

water, oxygen and hydrogen peroxide could enter the confined reaction cavity formed in the

594

enzyme-substrate interface through a narrow tunnel, while larger molecules, such as ascorbic acid,

595

may not be able to reach the active site when the enzyme sits on a surface of crystalline substrate.

596

It is important to note that a reduced LPMO that is bound to substrate differs a lot from a

597

reduced LPMO in solution. In the latter situation, the enzyme basically carries an exposed reduced 26 ACS Paragon Plus Environment

Page 26 of 59

Page 27 of 59 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 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

ACS Catalysis

598

transition metal that can engage in all kinds of (off-pathway) redox reactions (see next section).

599

When bound to its correct substrate, the copper site is secluded from solvent and precisely

600

positioned relative to the to-be-oxidized carbon.39,57,79 It is thus important for a reduced LPMO to

601

bind to its substrate to prevent engagement in off-pathway reactions. Accordingly, it has been

602

shown that reduction of LPMOs increases their affinity for the substrate.36,82,127

603 604

2.7 LPMO inactivation

605

Progress curves for LPMO reactions are often non-linear (Figures 9 and 10). It is now clear

606

that LPMOs, like other metallo-enzymes (see section 2.4) are subject to oxidative damage if the

607

conditions are not right. This may explain the lack of linearity in progress curves, although in some

608

studies the slow disappearance of LPMO activity may also have been due to depletion of another

609

reactant, such as the reductant. Work by Loose et al. studying engineered variants of chitin-active

610

SmAA10A in O2-driven reactions,128 and Bissaro et al. studying cellulose-active ScAA10C in

611

H2O2-driven reactions,31 have shown that LPMO inactivation is accompanied by oxidative damage

612

of residues close to the copper-site.

613

Considering the ongoing discussion on the roles of O2 versus H2O2 in LPMO catalysis (see

614

section 2.8), it is of major importance to note that autocatalytic inactivation of LPMOs happens no

615

matter how the reaction is fueled, as clearly shown by the data presented in Figures 9 and 10.

616 617 618

27 ACS Paragon Plus Environment

ACS Catalysis 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 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

619 620

Figure 9. Activity of a bacterial chitin-active LPMO, SmAA10A, in O2-driven reactions.

621

Panel A shows degradation of -chitin under aerobic conditions using a fungal cellobiose

622

dehydrogenase for the supply of reducing equivalents. Higher concentrations of CDH increase the

623

initial activity of the LPMO (higher product levels at 4 h up to a CDH concentration of 3 M), but

624

reduce enzyme stability (reduced increase in product levels between 4h and 24h). Panel B shows

625

degradation of -chitin under aerobic conditions in the presence of varying amounts of ascorbic

626

acid. Higher concentrations of ascorbic acid give higher initial rates but enzyme inactivation

627

becomes noticeable at the highest concentration and the total yield relative to the amount of added

628

reductant decreases. All reactions were carried out at pH 6.0, with 10 mg/ml -chitin and 1 M

629

SmAA10A. The total amount of solubilized oxidized products was determined after enzymatically

630

converting soluble oligomers to a mixture chitobionic acid and N-acetylglucosamine. Adapted

631

with permission from reference 112. Copyright 2016, John Wiley & Sons.

632 633

28 ACS Paragon Plus Environment

Page 28 of 59

Page 29 of 59 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 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

ACS Catalysis

634 635

Figure 10. LPMO activity during degradation of Avicel with the LPMO-containing

636

commercial cellulase cocktail Cellic CTec2. Panel A shows the accumulation of C4-oxidized

637

LPMO products over time in reactions containing 5 mM ascorbic acid and with varying oxygen

638

concentrations in the air used to sparge the reaction mixture and present in the head space of the

639

bottle. Higher oxygen concentrations give higher initial yields but also faster LPMO inactivation.

640

Note that the C4-oxidized product monitored here is unstable, which explains why levels gradually

641

go down as production ceases. Panel B shows the accumulation of LPMO products over time in

642

reactions containing 1 mM ascorbic acid that were carried out in fermentors under anaerobic

643

conditions, with feeding of H2O2 at a rate that is indicated in the Figure in M h-1. In the absence

644

of oxygen and H2O2 the LPMO is not active, whereas increasing amounts of added H2O2 lead to

645

faster catalytic rates but also faster enzyme inactivation. All reactions were carried out at pH 5.0,

646

with 100 mg mL-1 Avicel and 4 mg Cellic CTec2 proteins per gram of Avicel. Adapted with

647

permission from reference 37. Copyright 2018, BioMed Central.

648 649 650 651

As briefly alluded to in sections 2.3 & 2.6, reduced LPMOs that are not bound to substrate

652

contain a solvent-exposed Cu(I) ion. Although, LPMOs seem to have evolved to stabilize this

653

Cu(I) in the absence of substrate (since they bind Cu(I) more strongly than Cu(II); Aachmann et

654

al.24), it is plausible that this reduced transition metal engages in reactions with O2 or H2O2 leading 29 ACS Paragon Plus Environment

ACS Catalysis 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 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

655

to the formation of reactive oxygen species, including, possibly, highly damaging species such as

656

hydroxyl radicals. In the absence of substrate, these species will react with something else close to

657

the catalytic center, e.g. the copper coordinating histidines, as has indeed been observed.31 Kinetic

658

studies on H2O2-driven catalysis by chitin-active SmAA10A indicated that the rate for H2O2-driven

659

enzyme inactivation is about 1000-fold lower than the rate for H2O2-driven cleavage of chitin.35 It

660

is likely that the ratio between these two rates will vary between LPMOs and LPMO-substrate

661

combinations.

662

In line with the above considerations, Bissaro et al. showed that the presence of substrate

663

protects LPMOs from oxidative inactivation.31 Similar conclusions can be drawn from data in

664

Hangasky et al.33 Subsequently, studies of the effects of mutating residues on the substrate-binding

665

surface and of removing CBMs have shown a clear correlation between the affinity of the LPMO

666

for its substrate and LPMO stability.66,128,129 Likewise, it has been shown that higher substrate

667

concentrations promote LPMO stability.130 Of note, these studies on the impact of substrate on

668

LPMO stability concern several LPMOs acting on several substrates. There is no doubt that the

669

presence of substrate has a major effect on protecting an LPMO from autocatalytic oxidative

670

damage.

671

Potential enzymatic redox partners of LPMOs such as CDH8,11,110,112 and the

672

pyrroloquinoline quinone-dependent pyranose dehydrogenase (CcPDH),109 described above, often

673

contain a cellulose-binding CBM. Interestingly, Várnai et al. found that removal of the CBM from

674

CcPDH reduced the efficiency of CcPDH-driven LPMO action.109 Although speculative, this

675

suggests that it is beneficial that activation of the LPMO happens in the vicinity of the substrate,

676

i.e. by a substrate-bound rather than a freely moving CcPDH. This would be in line with the notion

677

that substrate binding is crucial to prevent inactivation of activated LPMOs.

678

These stability issues and the role of substrate are crucially important from a practical point

679

of view. During industrial bioprocessing of lignocellulosic biomass, both the amount and the

680

nature of the substrate will change along the reaction, while the LPMO is inactivated. In fact, it is

681

conceivable that in “typical” bioprocessing reactions with LPMO-containing commercial cellulose

682

cocktails, the LPMOs are no longer active during the final phase of the process, where the most

683

recalcitrant part of the substrate remains and the LPMOs may be most needed. These issues are

684

further addressed in section 3 of this review.

685 30 ACS Paragon Plus Environment

Page 30 of 59

Page 31 of 59 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 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

ACS Catalysis

686 687

2.8 The nature of the oxygen co-substrate

688

The discovery, in 2016,30 that LPMOs display peroxygenase activity, and perhaps may not

689

act as monooxygenases at all, has created quite some stir. While the peroxygenase activity and the

690

high catalytic rates that may be achieved by harnessing this activity now have been confirmed in

691

multiple laboratories,31,33,35,38 there remains controversy as to the question whether LPMOs do use

692

O2 directly and thus carry out a monooxygenase reaction. Some, including the authors of this

693

review, argue that H2O2 likely is the only, or, at least, the only kinetically relevant, co-substrate,

694

whereas others claim that O2 not only is relevant but is the natural co-substrate of LPMOs.33

695

Bissaro et al. have recently reviewed existing data on LPMO functionality in light of what is known

696

about other H2O2 producing or consuming redox enzymes that are (potentially) involved in fungal

697

degradation of lignocellulosic biomass.32

698

It is worth noting that a possible mix up of an oxygenase activity and a perxoygenase

699

activity is not unprecedented. In 2013, Wang et al. showed that HppE, a non-heme mono-iron

700

epoxidase involved in the production of fosfomycin, which had been studied for more than a

701

decade,131 is a peroxidase, reducing H2O2, rather than an oxidase, reducing O2.132

702

In their original work on SmAA10A, Vaaje-Kolstad et al. showed that one heavy oxygen

703

atom, provided in the form of 18O2, was incorporated in the oxidized products.1 While this may

704

seem to demonstrate a monooxygenase reaction, advocates of the peroxygenase activity of LPMOs

705

would argue that under the conditions of the assay, O2 may first have been converted to H2O2,

706

which then was used by the LPMO. They would argue that this may happen under all reaction

707

conditions usually used in LPMO research, not in the least because non-substrate bound LPMOs

708

generate H2O2 themselves.12,133 The competition experiments shown in Figure 6 and discussed in

709

section 2.2 may be taken to support this view.

710

In support of the peroxygenase reaction, Bissaro et al. showed that LPMO activity is

711

inhibited in reactions where horseradish peroxidase (HRP) competes for H2O2,31 and similar results

712

have later been described by Hangasky et al. for reactions with insoluble substrates.33 Importantly,

713

Hangasky et al. showed that the inhibitory effect of HRP on LPMO activity becomes less

714

pronounced at higher substrate concentrations and they demonstrated that when using soluble

715

substrate, cellohexaose, the inhibitory effect of HRP is virtually absent.33 These observations were

716

taken to show that, under some conditions, H2O2 does not play a role in LPMO catalysis and that, 31 ACS Paragon Plus Environment

ACS Catalysis 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 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

717

thus, O2 is used directly in a monooxygenase (or “coupled”) reaction. The authors then concluded

718

that LPMOs display both a peroxygenase (or “uncoupled”) reaction and a monooxygenase reaction

719

and that the occurrence of these two mechanisms is substrate-dependent. While we cannot exclude

720

that this conclusion is correct, we would argue that there is a plausible alternative explanation for

721

the observations made by Hangasky et al.33: it is conceivable that at high substrate concentrations,

722

and in particular when using an easily diffusible, soluble substrate, the H2O2–driven LPMO

723

reaction is so fast that HRP cannot compete for the H2O2. For the same reasons, the absence of an

724

effect of (H2O2–consuming) catalase on LPMO catalysis under certain conditions, cannot be taken

725

to indicate that LPMOs do not use H2O2. Kinetics dictate that the LPMO can easily compete with

726

catalase, as recently discussed by Kuusk et al.36

727

In any case, available data show that H2O2-driven LPMO reactions are orders of magnitude

728

faster than O2-driven LPMO reactions, as outlined in section 2.3 (Table 1). The O2 mechanism

729

may exist and be biologically relevant, for example because O2 concentrations will be higher than

730

H2O2 concentrations in many eco-systems. Still, as shown by the competition experiments of

731

Figure 6 and the low KM values for H2O2, low concentrations of H2O2, potentially generated by

732

enzymes secreted specifically for this purpose,32 will lead to H2O2 consumption by LPMOs and

733

speed up LPMO reactions.

734

While this review is not the place to finally settle the issue of the true co-substrate, it is

735

important to point at a few claims that are propagated in the literature and that are clearly incorrect.

736

Firstly, it has been suggested that LPMOs are more prone to inactivation in H2O2-driven reactions

737

compared to O2-driven reactions. This is not true, as illustrated by Figures 9 and 10, above. Of

738

course the degree of inactivation will be affected by a multitude of factors that may vary between

739

experiments, such as the H2O2 concentration in H2O2-driven reactions or the reductant type and

740

concentration in O2-driven reactions, or the substrate used. So, differences in inactivation rates

741

may be observed, but these are most likely due to varying reaction conditions, rather than to

742

fundamentally different catalytic processes.

743

Secondly, it has been claimed, or at least suggested, that LPMOs become less specific if

744

the reaction is fueled by H2O2 rather than O2. For example, when studying H2O2-driven

745

degradation of cellohexaose by a fungal LPMO, Hangasky et al. detected minor amounts of

746

products carrying atypical oxidations, indicating a lack of specificity.33 While these results are

747

indisputable, there is no evidence that the observed lack of specificity is due to the participation of 32 ACS Paragon Plus Environment

Page 32 of 59

Page 33 of 59 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 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

ACS Catalysis

748

H2O2 in the reaction. It should also be noted that cellohexaose is not necessarily the natural or

749

optimal substrate of the LPMO used and its binding may thus not provide the active site

750

confinement that is needed to control and direct the strong oxidative power of the emerging

751

reactive oxygen species (see section 2.3). The elegant structural work by Simmons et al. shows

752

that different oligomeric substrates have different binding modes with clear effects on the spatial

753

arrangement near the copper site.126 Since the substrate is crucial in controlling the orientation of

754

the reactive oxygen species,39,57,79 it is conceivable that aspecific background reactions occur for

755

certain LPMO-substrate combinations. Of note, in many reaction setups, there will be a gradual

756

increase in oxidative damage in the catalytic center of the LPMO,31,128 which could also affect

757

specificity. We have characterized multiple LPMOs (fungal, bacterial), with varying oxidative

758

regioselectivity (C1/C4/C1&C4), acting on multiple substrates (cellulose, cello-oligosaccharides,

759

xyloglucan, chitin) in both O2 and H2O2-driven reactions. In our hands, well controlled O2 and

760

H2O2-driven reactions give the same overall product profiles indicating that there is no basis to

761

claim that LPMOs become less specific if they are fueled with H2O2.

762 763 764

3. Harnessing LPMOs in biomass processing

765 766

3.1 Degradation of lignocellulosic biomass

767

Lignocellulosic biomass is mainly composed of lignin, cellulose and a variety of hemicelluloses.

768

These three polymers are interlinked in a complex matrix where their relative abundance varies

769

depending on the type of biomass. Native lignocellulosic biomass is generally very compact and

770

little accessible to enzymatic attack. Thus, biomass processing is initiated with some sort of

771

pretreatment that rips the fibers apart and makes the biomass accessible for enzymes.134

772

Commercial enzyme cocktails for saccharification of lignocellulosic biomass, such as Cellic

773

CTec® products from Novozymes and Accellerase® products from DuPont, contain enzyme

774

activities that degrade the polysaccharides, leaving lignin as a solid residue. Cellulose is of

775

particular interest due to its abundance in the biomass and due to its simple composition, being a

776

linear polysaccharide of β-1,4 linked D-glucose units.

777

It is generally believed that efficient cellulose degradation requires the presence of three

778

classes of hydrolytic enzymes: endo-1,4-β-glucanases, randomly cleaving internal glucose bonds, 33 ACS Paragon Plus Environment

ACS Catalysis 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 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

779

cellobiohydrolases, which attack the reducing or non-reducing ends of the cellulose polymer

780

releasing cellobiose in a processive manner, and β-glucosidases, converting cellobiose to glucose

781

(Figure 11).5,135 Although the compositions of modern commercial cellulase cocktails are not

782

publicly available, these cocktails likely contain all three enzyme types. In addition, these modern

783

cocktails contain one or more LPMOs,44,45 which cleave internal bonds and increase the overall

784

efficiency of the enzyme blend.37, 44-50 Harnessing the potential of LPMOs in industrial biomass

785

processing is not straightforward and puts new demands on process design. One obvious issue

786

concerns the fact that (controlled) addition of air or hydrogen peroxide is necessary. By acting on

787

the surfaces of insoluble substrates, LPMOs improve the accessibility for canonical cellulases,

788

perhaps particularly in the most recalcitrant parts of the substrate,41,42 and thus improve the overall

789

efficiency of cellulase cocktails.

790

34 ACS Paragon Plus Environment

Page 34 of 59

Page 35 of 59 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 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

ACS Catalysis

791

792 793 794

Figure 11. Degradation of cellulose by LPMOs and cellulases. The upper half of the Figure

795

shows LPMO reactions, whereas the lower half illustrates the various cellulases. For clarity,

796

chemical equations for the O2- and H2O2-dependent reaction schemes (indicated by “A” and “B”

797

in the Figure, respectively) that are also shown in Figure 2 appear beneath the figure. Productive

798

LPMO reactions start with step 0, i.e. reduction of the copper, after which an O2- (A) or an H2O2-

799

(B) driven reaction may occur. In H2O2-driven catalysis the copper stays reduced (blue) in between

800

catalytic cycles, whereas the redox state of the copper during and at the end of O2-driven catalytic 35 ACS Paragon Plus Environment

ACS Catalysis 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 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

801

cycles is less clear; see section 2.2 and Figure 5 for more details. The figure also shows multiple

802

redox side reactions: generation (5) and consumption (5’) of H2O2 in reactions involving the

803

reductant, H2O2 generation by the LPMO (3), and inactivation (4) of LPMOs that are not bound to

804

substrate (2). Next to LPMOs, multiple cellulases may contribute to cellulose degradation:

805

endoglucanases (EG), processive cellobiohydrolases moving into both possible directions (6;

806

CBH), and a beta-glucosidase (BG) converting oligomeric products, mostly dimers, to monomeric

807

glucose.

808 809 810

3.2. Breaking down the hemicellulose-cellulose matrix

811

In lignocellulosic biomass, even after an efficient pretreatment, cellulose remains embedded in a

812

hemicellulose-lignin matrix and there are strong indications in the literature that xylan and

813

glucomannan strongly associate with cellulose fibers (Figure 12).136-139 Indeed, accessory

814

hydrolytic enzymes such as xylanases and mannanases are known to have beneficial effects on the

815

saccharification of cellulose by cellulases, likely by increasing cellulose accessibility.49,140-143

816

Importantly, even minor amounts of remaining hemicellulose in a pretreated substrate may inhibit

817

cellulose conversion.142,143

818

While addition of specific hemicellulases to improve cellulose accessibility may be useful

819

in certain situations, additional hemicellulose-removing capacity may come from hemicellulolytic

820

side activities of one or more of the cellulases in standard cellulolytic enzyme cocktails.142,144,145

821

In particular, endoglucanases belonging to the GH7 family, such as Cel7B from Trichoderma

822

reesei, show activities on both xylan and glucomannan that are important in the saccharification

823

of softwood.144

824

Similarly to certain cellulases, substrate promiscuity has been reported for cellulose-active

825

LPMOs belonging to the AA9 family, where known substrates include xyloglucan, glucomannan,

826

mixed-linkage glucan13,146 and xylan.14,126 The ability of cellulose-active LPMOs to cleave

827

hemicelluloses seems independent of their oxidative regioselectivity. For example, xyloglucan is

828

cleaved by C1-oxidizing TtAA9E,120 C4-oxidizing NcAA9C,13 and C1/C4-oxidizing GtAA9A-

829

2.146 Although less well explored up to now, additional industrially relevant functionalities of

830

LPMOs could relate to their ability to remove hemicelluloses that contribute to biomass

831

recalcitrance by binding tightly to cellulose fibrils (Figure 12). Indeed, in a seminal paper, 36 ACS Paragon Plus Environment

Page 36 of 59

Page 37 of 59 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 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

ACS Catalysis

832

Couturier et al. recently described a novel LPMO family (AA14) that specifically acts on those

833

parts of xylan chains that are attached to cellulose fibrils and that promotes cellulose

834

saccharification (Table 2).15 Earlier, Frommhagen et al. had detected LPMO activity on xylan but

835

also in this case only when the xylan was present together with cellulose.14

836 837

838 839 840

Figure 12. Artist impression of the interaction of xylan with cellulose fibrils. The picture,

841

generated by the Dupree group,138,139,147 shows cellulose fibrils (yellow) that are held together by

842

xylan chains that interact with multiple fibrils. Some lignin is also shown. The xylan chains carry

843

multiple decorations, as indicated in the Figure. The boxes, labeled with letters, indicate several

844

substructures in the xylan that differ in their level and type of interaction with cellulose. The

845

recently discovered xylan-active AA14 LPMOs are thought to act on parts of the xylan chains that 37 ACS Paragon Plus Environment

ACS Catalysis 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 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

846

interact with cellulose.15 Reproduced with permission from reference 139. Copyright 2016,

847

Biochemical Society.

848 849 850

3.3 The effects of LPMOs on cellulose conversion

851

The cost of cellulolytic enzyme cocktails has been greatly reduced in the past two decades,5, 44-45

852

but still remains one of the primary bottlenecks for successful commercialization of lignocellulose-

853

derived fuels and chemicals.45,148 Major enzyme producers have been looking into LPMOs and

854

their implementation in enzymatic degradation of biomass for about 15 years.5,44,45

855

Several studies preceding the discovery of the oxidative nature of LPMOs, reported enhanced

856

saccharification of cellulose by utilizing these proteins, which at the time were called GH61s. For

857

example, Merino and Cherry observed that supplementing a Trichoderma reesei secretome, which

858

harbours little GH61 proteins,149,150 with a secretome from Thielavia terrestris increased

859

degradation of pretreated corn stover (PCS), and reduced required enzyme loadings.5 This positive

860

effect was shown to be due the action of GH61 proteins present in the T. terrestris secretome.

861

Similar observations were reported by Harris et al., who showed that addition of a GH61 from

862

Thermoascus aurantiacus to a cellulase cocktail reduced the enzyme loading necessary to reach

863

90% conversion of PCS two-fold.6 Harris et al. also showed a beneficial effect of adding a

864

recombinantly produced GH61 from T. terrestris.6 Interestingly, both Merino and Cherry and

865

Harris et al. noted that none of the tested GH61s exhibited synergistic effects with canonical

866

cellulases during saccharification of “clean” cellulose substrates such as filter paper and Avicel,5,6

867

which in hindsight can be attributed to the lack of lignin and lignin-derived compounds that

868

provide the reducing power that is needed to drive the LPMO reaction.

869

In an early study, Cannella et al. compared the saccharification performance of a mixture of

870

Celluclast, an LPMO-poor149,150 cellulase cocktail, and Novozym 188, a -glucosidase

871

preparation, with the performance of Cellic CTec2, during enzymatic decomposition of

872

hydrothermally pretreated wheat straw.46 They observed increased cellulose saccharification with

873

the CTec2 preparation and detected oxidized glucose and cellobiose in the reaction mixtures.

874

Table 2 provides an overview of studies that have assessed the impact of individual LPMOs

875

on the efficiency of cellulase cocktails in cellulose saccharification. For example, Hu et al. reported

876

that spiking of LPMO-poor Celluclast, with TaAA9A resulted in an 18–63% increase in the 38 ACS Paragon Plus Environment

Page 38 of 59

Page 39 of 59 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 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

ACS Catalysis

877

conversion of cellulose from pretreated corn stover, poplar and loblolly pine, depending on the

878

substrate and type of substrate pretreatment (Table 2).48

879

In a landmark study, Müller et al. analyzed the effect of TaAA9A on the saccharifying

880

efficiency of the Celluclast/Novozym 188 mixture (Table 2), while simultaneously assessing the

881

impacts of reductants and oxygen, and with quantification of LPMO activity through detection of

882

LPMO products.50 Next to showing a beneficial effect of the LPMO, the results showed a clear

883

correlation between the overall saccharification efficiency of the enzyme cocktail and the

884

generation of oxidized sugars. Furthermore, this study showed that LPMO activity and the LPMO

885

effect on saccharification efficiency were absent in reactions run under anaerobic conditions.

886

When working with pure cellulose (e.g. Avicel) the LPMO effect depended on the presence of

887

externally added reductant, whereas such addition was not necessary when working with (lignin-

888

rich) steam-exploded birch.50

889

In a follow-up study, Chylenski et al. showed that optimal use of Cellic CTec3 in the

890

depolymerization of heavily delignified sulfite-pulped Norway spruce, required the addition of

891

oxygen and reductant. A beneficial effect of adding TaAA9A to a Celluclast and β-glucosidase

892

mixture was also found for saccharification of wheat straw.151 Du et al. showed similar boosting

893

effects for an AA9 LPMO from the filamentous fungus Penicillum oxalicum, using alkali-

894

pretreated wheat straw, alkali-pretreated corn stover and delignified corncob residues (Table 2).152

39 ACS Paragon Plus Environment

ACS Catalysis 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 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

896

Table 2. The effect of LPMOs on the efficiency of lignocellulolytic cocktails Added enzyme ID

Base enzyme cocktaila

LPMO (% w/w)a – total protein content (mg/g of biomass)

Reaction conditions (pH/ T, C/ Incubation time, h)

Biomass (DM, % w/v)

T. reesei secretome

20% - 5 mg/g

5.0/50/168

AP-CS (4.7%)

30 (69-89) 63 (43-70) 33 (70-93) 18 (57-67) 23 (65-80) 32 (53-70) 23 (39-48) 33 (64-85) 22 (63-77)

TtAA9E PoxAA9 CgAA9 GtAA9A

Relative increase in yield (%)b

Ref

(6)

Celluclast

5% - 20 mg/g

4.8/50/48

CelluclastN188 Celluclast (80%) + BG (10%) T. reesei secretome

15% - 5 mg/g 15% - 8 mg/g

5.0/50/20 5.0/50/48

OP-CS (2%) OP-P (2%) OP-LP (2%) SE-CS (2%) SE-P (2%) SE-LP (2%) SE-B (10%) SP-NS (5%)

10% - 3 mg/g

5.0/50/96

SE-WS (10%)

44 (39-56)

(151)

20% - 5 mg/g

5.0/50/168

AP-CS (4.7%)

20 (69-83)

(6)

20% - 5 mg/g

4.8/48/72

1 mg/g – nr

5.0/50/48

Alk-WS (2%) Alk-CS (2%) DCCR (2%) Alk-RS (1%) AP-RS (1%) Alk-WS (2%) AP-P (0.5%) AP-Pi (0.5%)

33 (67-89) 30 (63-82) 44 (54-78) 15 (0.83-0.96)* 33 (0.24-0.32)* 30 (4.9-6.4)* 38 (0.34-0.47) 82 (0.22-0.4)

TaAA9A

897 898 899 900 901 902 903 904 905 906 907 908 909 910

Page 40 of 59

Pox cellulase cocktail Celluclast (0.9 FPU/g) Celluclast

(48)

(50) (47)

(152) (153)

10% - 2 mg/g 5.0/50/48 (154) 87% - 7.65 mg/g PcoAA14A CL847 40% - 1.65 T. reesei 5.2/45/24c AP-Pi (0.5%) 25 (0.20-0.25) (15) mg/g secretome AP-P (0.5%) 32 (0.34-0.45) 87% - 7.66 PcoAA14B mg/g AP-Pi (0.5%) 68 (0.22-0.37) a When the total protein load was not reported (nr) the amount of added enzyme cocktail is expressed in filter paper units (FPU) per gram of dry matter (DM) in the column “Base enzyme cocktail”. In this case, the quantity of oxidoreductase is given in mg/g of DM instead of % (w/w) of total protein load in the column “Oxidoreductase-total protein content”. b Increase in biomass hydrolysis yield relative to the reference reaction where the LPMO is either absent or repressed (e.g. anaerobic conditions). The absolute conversion yields (expressed in % of maximum theoretical yield or in g/L of sugars when marked with *), obtained without and with LPMO addition, are given in between brackets in italics. c In this case the LPMO were first incubated with the biomass and AscA for 76 h prior to 24 h further incubation with the cellulase cocktail. Abbreviations: Alk, alkali-pretreated; AP, acid pretreated; BG, -glucosidase; OP, organosolv pretreated; SE, steam-exploded; SP, sulfite pulped; B, birchwood; CS, corn stover; DCCR, delignified corncob residue; LP, lodgepole pine; MCC, microcrystalline cellulose; NS, Norway spruce; P, poplar; Pi, Pine; RS, ricestraw; WS, wheatstraw Cg, Chaetomium globosum; Gt, Gloeophyllum trabeum; Pco, Pycnoporus coccineus; Pox, Penicillium oxalicum JU-A10; Ta, Thermoascus aurantiacus; Tt, Thielavia terrestris.

911 912

The discovery of the peroxygenase activity of LPMOs sheds new light on what may be

913

happening during enzymatic bioprocessing of cellulose. Depending on the substrate and the

914

process setup, access to oxygen, oxygen consumption, and the chemical or enzymatic formation

915

and consumption of hydrogen peroxide will vary. When using “real” biomass, containing lignin

916

and sugars and compounds formed during harsh pretreatments, a lot of redox chemistry may 40 ACS Paragon Plus Environment

Page 41 of 59 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 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

ACS Catalysis

917

happen (e.g. ref 155). In principle, the peroxygenase activity of LPMOs, and the much higher

918

efficiency of the peroxygenase reaction, open up new possibilities for harnessing and controlling

919

LPMO activity during degradation of lignocellulosic biomass. This potential became indeed

920

apparent from studies on the digestion of a “clean” substrate, Avicel, with LPMO-containing Cellic

921

CTec2 under anaerobic conditions (Figure 10B).31,37 Controlled pumping of low amounts of H2O2

922

into anaerobically operated bioreactors allowed tight control of LPMO activity throughout the

923

duration of the hydrolysis and revealed that LPMO activity is a direct function of the amount of

924

H2O2 added. Moreover, by controlling H2O2 supply, it was possible to develop protocols that gave

925

higher saccharification yields compared to those obtained previously in similar reactions run under

926

standard aerobic conditions.

927 928 929

3.4 Challenges in the application of LPMOs in biomass processing

930

In saccharification reactions run under aerobic conditions and with real biomass, there is

931

usually enough reducing power in the substrate to fuel the LPMO. Oxygen will either be used

932

directly, in what would be a monooxygenase reaction, or be converted to H2O2, through direct

933

reactions with reductants or catalyzed by LPMOs (Figure 11), and then used in what would be a

934

peroxygenase reaction, as discussed above. Either way, supply of oxygen is needed, which is

935

expensive in large industrial reactors. To achieve high oxygen transfer rates a high airflow must

936

be maintained and also usually a high stirrer speed, which is difficult in reactors with high dry

937

matter content.156 While oxygen supply to some extent can be controlled and monitored using an

938

oxygen electrode, controlling H2O2 concentrations is easier, since H2O2 is a liquid that is readily

939

soluble in water.

940

In industrial settings, the presence of reductants in the form of lignin and other compounds

941

naturally present in the biomass or formed during biomass pretreatment may be difficult to control.

942

Both too much and too little reducing power may be problematic. Too much reducing power may

943

lead to depletion of oxygen or H2O2 (Figure 11), whereas too little reducing power may lead to

944

impaired LPMO reduction.

945

For “clean” substrates, with no or very little lignin, the situation is manageable, and LPMO

946

activity can be controlled directly by varying the concentration of reductant, varying the O2

947

concentration or by controlled addition of H2O2, as shown in a recent study by Müller, Chylenski 41 ACS Paragon Plus Environment

ACS Catalysis 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 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

948

et al. and illustrated by Figure 10, which is derived from this study.37 Using Avicel as a substrate

949

and LPMO-containing Cellic CTec2 under anaerobic conditions, Müller, Chylenski et al. showed

950

that the degree of LPMO activity could be perfectly controlled by controlling the addition of H2O2

951

(Figure 10B) and by optimizing the conditions, they reached unprecedented high conversion

952

levels.37

953

The same study also showed that when using lignin-rich, and industrially more realistic

954

substrates, the situation became much less manageable. In these cases, including unpublished work

955

from our laboratory, running reactions under standard aerobic conditions usually is at least as

956

efficient, or even better, than running reactions with added H2O2. Likely side reactions involving

957

reductants (Figure 11, step 5) play a key role, as also suggested by recent work by Peciulyte et

958

al.155

959

One important finding in the study by Müller, Chylenski et al. is that, under most

960

conditions, the LPMOs in Cellic CTec2 became deactivated during the course of the

961

saccharification reaction.37 As discussed in detail in section 2.7, substrate depletion may be one

962

cause and it is indeed conceivable that depletion of LPMO binding sites occurs as the substrate is

963

depolymerized. As discussed in section 2.7 and illustrated in Figure 11 (step 4), there needs to be

964

balance in the system: if there are more reduced LPMOs in the system than there are productive

965

binding sites, and if there is sufficient supply of the oxygen co-substrate, non-substrate bound

966

LPMOs will suffer from auto-catalytic inactivation. Of note, in reactions with added H2O2 or with

967

conditions that lead to abundant in situ H2O2 formation, H2O2 will start accumulating, which may

968

damage the other enzymes in the cocktail, as has indeed been observed.37,151

969

A final important result from the work by Müller, Chylenski et al. is the observation that

970

even under optimal conditions for H2O2-driven degradation of Avicel (i.e. a “clean”, well-

971

controlled system), the LPMOs in Cellic CTec2 seemed to run at much lower rates than the rates

972

coming out of recent kinetic studies.33,35,37 When assuming an LPMO (TaAA9A) content of 15 %

973

in Cellic CTec2,50 the LPMO activity under optimized conditions was in the order of 1 – 5 min-1

974

(0.07 s-1), which is about 100 times less than the highest LPMO rates appearing in literature.33,35

975

This may be taken to indicate that only a small fraction of the LPMOs in Cellic Ctec2 were actually

976

engaged in substrate cleavage under these conditions, which could mean that products such as

977

Cellic Ctec2 contain more LPMOs than needed when H2O2 is directly added to the reaction. In this

978

respect, Müller, Chylenski et al. showed that further boosting of LPMO activity, by increasing the 42 ACS Paragon Plus Environment

Page 42 of 59

Page 43 of 59 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 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

ACS Catalysis

979

H2O2 feed, did indeed lead to higher initial LPMO activity, but also to less efficient cellulose

980

conversion and rapid inactivation of the LPMO.37 It is possible that under these conditions, the

981

cellulases became limiting. If the cellulases are not able to uncover novel LPMO binding sites by

982

peeling off oxidized oligosaccharides from the substrate surface, LPMOs get involved in side

983

reactions and become inactivated.

984

If one accepts the idea that LPMO reactions are primarily driven by H2O2, it is conceivable

985

that the high content of LPMOs in today’s enzyme cocktails is a consequence of these cocktails

986

being designed to work with O2 as oxygen source. Under such conditions part of the LPMO pool

987

is needed to generate H2O2 from O2 in situ. When H2O2 is supplied directly, all LPMOs are

988

available for polysaccharide oxidation.

989 990 991

3.5. LPMOs and fermentative valorization of biomass-derived sugars

992

The monomeric sugars obtained from saccharification of lignocellulosic biomass may be

993

converted to a variety of products through fermentation and/or chemical processes.157

994

Fermentative production of ethanol (“second generation biofuel”) is one of the best known and

995

most explored conversions.158 Another well-known conversion concerns the production of lactic

996

acid which may be used to produce bioplastics.159,160 There are two principle ways to combine

997

enzymatic hydrolysis with fermentation: the processes can be run sequentially in a separate

998

hydrolysis and fermentation (SHF) process, or simultaneously in a so called simultaneous

999

saccharification and fermentation (SSF) process.161 Generally, if enzymes and fermenting

1000

organisms have similar pH and temperature optima, SSF processes seem to be more efficient

1001

because these are one-tank processes and because liberated sugar is immediately fermented, thus

1002

alleviating product inhibition of the enzymes.162

1003

If the efficiency of cellulase cocktails depends on LPMOs, the need of these enzymes for

1004

oxygen may create problems in SSF setups. In aerobic fermentations there will be direct

1005

competition between the needs of the microorganism and the need of the LPMO in the enzyme

1006

cocktail. In anaerobic fermentations, which are commonly used, e.g. for the production of ethanol

1007

or lactic acid, oxygen scavenging mechanisms of the microbes and process parameters (i.e.

1008

anaerobic conditions) will keep the LPMO from being active. This was first noted by Cannella and

1009

Jørgensen, who observed that production of ethanol from pretreated wheat straw using LPMO43 ACS Paragon Plus Environment

ACS Catalysis 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 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

1010

containing Cellic CTec2 and yeast resulted in a 20% higher ethanol yield in an SHF setup

1011

compared to an SSF setup.162 Cannella and Jørgensen suggested that that this difference was due

1012

to competition for oxygen between LPMOs and yeast in the SFF setup, where the yeast was

1013

believed to scavenge most of the oxygen, thus outcompeting the LPMOs. Based on current

1014

knowledge concerning the role of H2O2, another explanation would be that the yeast either

1015

inhibited in situ production of H2O2 or consumed in situ produced H2O2 faster than the LPMO. In

1016

a similar study, Müller et al. analyzed the production of lactic acid by using steam exploded birch

1017

as substrate.159 In this work, a thermophilic strain of Bacillus coagulans was used together with

1018

Cellic CTec2 to compare SSF and SHF processes carried out at 50 °C. Generally, the SHF setup

1019

yielded around 30% higher production of lactic acid, and this higher yield was associated with

1020

production of oxidized sugars, indicative of LPMO activity. So, the SHF setup was more efficient

1021

and this could be coupled to the absence of LPMO activity in the SSF setup.

1022

One exciting aspect of the recent finding that H2O2 can efficiently fuel LPMO activity is

1023

that the “oxygen battle” in SSF may be avoided. It would seem possible to run fermentations with

1024

controlled addition of H2O2 in amounts that are sufficient to drive LPMO action while being so

1025

low that they do not harm the microbe. Our own preliminary data indicate that there indeed may

1026

be considerable potential in this type of experimental approach.

1027 1028 1029

4. Conclusions and perspectives

1030

The discovery of LPMOs has revolutionized research on enzymatic biomass conversion and the

1031

industrial implementation thereof. Still, many questions remain unanswered, and several of these

1032

questions are not easy to address. One key challenge lies in the multitude of reactions that may

1033

take place in “typical” reaction mixtures, in particular when using industrial (i.e. co-polymeric &

1034

complex) substrates. Even experiments with relatively “clean” substrates and relatively well-

1035

defined reaction mixtures raise questions. For example, in the original work on chitin-active

1036

SmAA10A,1 the boosting effect of the activated LPMO on chitinase efficiency was much larger

1037

than any boosting effect ever published for a cellulose-active LPMO acting in concert with a

1038

cellulase. So far, while the contributions of cellulose-active LPMOs in cellulose cocktails are

1039

considerable and undeniable, these contributions are lower than what one could expect on the basis

1040

of the original findings for chitin conversion. One attractive potential implication is that further 44 ACS Paragon Plus Environment

Page 44 of 59

Page 45 of 59 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 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

ACS Catalysis

1041

improvement of enzymatic cellulose conversion may still be possible, perhaps by harnessing

1042

LPMO power in a better manner, based on the most recent insights outlined above.

1043

LPMOs are abundant and widely distributed in Nature and show a stunning sequence

1044

diversity.67,163,164 It seems likely that some LPMOs are involved in bacterial virulence and act on

1045

substrates that yet have to be discovered.165-167 Lignocellulosic biomass is a complex material not

1046

only consisting of cellulose fibrils but also containing other polymers, a variety of hemicelluloses

1047

and lignins, that may be heavily intertwined (Figure 12). It is possible that the abundance of

1048

LPMOs in lignocellulose-degrading fungi reflects the fact that various LPMOs attack various

1049

substructures in the substrate, a notion for which some evidence has recently been provided.15

1050

Considering the sequence diversity of the extended substrate-binding surfaces of LPMOs and the

1051

ability of these enzymes to act on liquid-solid interfaces, it is also conceivable that some may act

1052

on non-polysaccharide substrates such as lignin or other insoluble, recalcitrant polymers.

1053

It seems likely that the full potential of LPMOs has not yet been harnessed. Firstly, the best

1054

LPMOs, combinations of LPMOs or LPMO-cellulase combinations may not yet have been

1055

discovered. Secondly, as outlined above, there is a clear lack of understanding of how to best

1056

harness LPMO action, in laboratory experiments and industrial bioreactors alike. When it comes

1057

to industrial biorefining of lignocellulosic biomass, perhaps even pretreatment strategies may need

1058

reconsideration, since the content and nature of remaining lignin will affect the efficiency of

1059

LPMOs in subsequent saccharification reactions. The operational stability of LPMOs is a

1060

potentially major success factor that may potentially be engineered or optimized by selecting the

1061

best natural candidates.

1062

Almost seventy years after Reese et al. suggested the existence of a “substrate-disrupting”

1063

factor (C1) in what is known as the C1-Cx hypothesis for enzymatic cellulose degradation,168 and

1064

almost 50 years after Eriksson et al. demonstrated that O2 plays a role in enzymatic cellulose

1065

conversion,2 we now have access to a multitude of substrate-disrupting LPMOs. We expect these

1066

enzymes to remain in the center of attention, due to their fascinating chemistry, their abundance in

1067

Nature, and their indisputable industrial importance.

1068 1069 1070 1071

Acknowledgements 45 ACS Paragon Plus Environment

ACS Catalysis 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 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

1072

We thank current and former team members for their contributions to our LPMO research, which

1073

has been supported by the Research Council of Norway, most recently through grants 240967,

1074

243663, 243950, 257622, 256766, 268002, 262853 and 270038. Additional support was received

1075

from the Marie-Curie FP7 COFUND People Programme, through the award of an AgreenSkills

1076

fellowship (under grant agreement n° 267196), and from the French Institut National de la

1077

Recherche Agronomique (INRA).

1078 1079 1080 1081

References

1082 1083 1084 1085 1086 1087 1088 1089 1090 1091 1092 1093 1094 1095 1096 1097 1098 1099 1100 1101 1102 1103 1104 1105 1106 1107 1108 1109 1110 1111

1. Vaaje-Kolstad, G.; Westereng, B.; Horn, S. J.; Liu, Z. L.; Zhai, H.; Sorlie, M.; Eijsink, V. G. H. An Oxidative Enzyme Boosting the Enzymatic Conversion of Recalcitrant Polysaccharides. Science 2010, 330, 219-222. 2. Eriksson, K. E.; Pettersson, B.; Westermark, U. Oxidation - Important Enzyme Reaction in Fungal Degradation of Cellulose. FEBS Lett. 1974, 49, 282-285. 3. Vaaje-Kolstad, G.; Horn, S. J.; van Aalten, D. M. F.; Synstad, B.; Eijsink, V. G. H. The Non-catalytic Chitin-binding Protein CBP21 from Serratia marcescens Is Essential for Chitin Degradation. J. Biol. Chem. 2005, 280, 28492-28497. 4. Karkehabadi, S.; Hansson, H.; Kim, S.; Piens, K.; Mitchinson, C.; Sandgren, M. The First Structure of a Glycoside Hydrolase Family 61 Member, Cel61B from Hypocrea jecorina, at 1.6 Å Resolution. J. Mol. Biol. 2008, 383, 144-154. 5. Merino, S. T.; Cherry, J. Progress and Challenges in Enzyme Development for Biomass Utilization. Adv. Biochem. Eng. Biotechnol. 2007, 108, 95-120. 6. Harris, P. V.; Welner, D.; McFarland, K. C.; Re, E.; Navarro Poulsen, J. C.; Brown, K.; Salbo, R.; Ding, H.; Vlasenko, E.; Merino, S.; Xu, F.; Cherry, J.; Larsen, S.; Lo Leggio, L. Stimulation of Lignocellulosic Biomass Hydrolysis by Proteins of Glycoside Hydrolase Family 61: Structure and Function of a Large, Enigmatic Family. Biochemistry 2010, 49, 3305-3316. 7. Forsberg, Z.; Vaaje-Kolstad, G.; Westereng, B.; Bunaes, A. C.; Stenstrom, Y.; MacKenzie, A.; Sorlie, M.; Horn, S. J.; Eijsink, V. G. H. Cleavage of Cellulose by a CBM33 Protein. Protein Sci. 2011, 20, 1479-1483. 8. Phillips, C. M.; Beeson, W. T.; Cate, J. H.; Marletta, M. A. Cellobiose Dehydrogenase and a Copper-dependent Polysaccharide Monooxygenase Potentiate Cellulose Degradation by Neurospora crassa. ACS Chem. Biol. 2011, 6, 1399-1406. 9. Quinlan, R. J.; Sweeney, M. D.; Lo Leggio, L.; Otten, H.; Poulsen, J. C.; Johansen, K. S.; Krogh, K. B.; Jorgensen, C. I.; Tovborg, M.; Anthonsen, A.; Tryfona, T.; Walter, C. P.; Dupree, P.; Xu, F.; Davies, G. J.; Walton, P. H. Insights into the Oxidative Degradation of Cellulose by a Copper Metalloenzyme that Exploits Biomass Components. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 15079-15084. 10. Westereng, B.; Ishida, T.; Vaaje-Kolstad, G.; Wu, M.; Eijsink, V. G. H.; Igarashi, K.; Samejima, M.; Stahlberg, J.; Horn, S. J.; Sandgren, M. The Putative Endoglucanase PcGH61D

46 ACS Paragon Plus Environment

Page 46 of 59

Page 47 of 59 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 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

1112 1113 1114 1115 1116 1117 1118 1119 1120 1121 1122 1123 1124 1125 1126 1127 1128 1129 1130 1131 1132 1133 1134 1135 1136 1137 1138 1139 1140 1141 1142 1143 1144 1145 1146 1147 1148 1149 1150 1151 1152 1153 1154 1155 1156 1157

ACS Catalysis

from Phanerochaete chrysosporium Is a Metal-dependent Oxidative Enzyme that Cleaves Cellulose. PLoS One 2011, 6, e27807. 11. Langston, J. A.; Shaghasi, T.; Abbate, E.; Xu, F.; Vlasenko, E.; Sweeney, M. D. Oxidoreductive Cellulose Depolymerization by the Enzymes Cellobiose Dehydrogenase and Glycoside Hydrolase 61. Appl. Environ. Microbiol. 2011, 77, 7007-7015. 12. Isaksen, T.; Westereng, B.; Aachmann, F. L.; Agger, J. W.; Kracher, D.; Kittl, R.; Ludwig, R.; Haltrich, D.; Eijsink, V. G. H.; Horn, S. J. A C4-oxidizing Lytic Polysaccharide Monooxygenase Cleaving Both Cellulose and Cello-oligosaccharides. J. Biol. Chem. 2014, 289, 2632-2642. 13. Agger, J. W.; Isaksen, T.; Varnai, A.; Vidal-Melgosa, S.; Willats, W. G. T.; Ludwig, R.; Horn, S. J.; Eijsink, V. G. H.; Westereng, B. Discovery of LPMO Activity on Hemicelluloses Shows the Importance of Oxidative Processes in Plant Cell Wall Degradation. Proc. Natl. Acad. Sci. U. S. A. 2014, 111, 6287-6292. 14. Frommhagen, M.; Sforza, S.; Westphal, A. H.; Visser, J.; Hinz, S. W. A.; Koetsier, M. J.; van Berkel, W. J. H.; Gruppen, H.; Kabel, M. A. Discovery of the Combined Oxidative Cleavage of Plant Xylan and Cellulose by a New Fungal Polysaccharide Monooxygenase. Biotechnol. Biofuels 2015, 8, 101. 15. Couturier, M.; Ladeveze, S.; Sulzenbacher, G.; Ciano, L.; Fanuel, M.; Moreau, C.; Villares, A.; Cathala, B.; Chaspoul, F.; Frandsen, K. E.; Labourel, A.; Herpoel-Gimbert, I.; Grisel, S.; Haon, M.; Lenfant, N.; Rogniaux, H.; Ropartz, D.; Davies, G. J.; Rosso, M. N.; Walton, P. H.; Henrissat, B.; Berrin, J. G. Lytic Xylan Oxidases from Wood-decay Fungi Unlock Biomass Degradation. Nat. Chem. Biol. 2018, 14, 306-310. 16. Vu, V. V.; Beeson, W. T.; Span, E. A.; Farquhar, E. R.; Marletta, M. A. A Family of Starch-active Polysaccharide Monooxygenases. Proc. Natl. Acad. Sci. U. S. A. 2014, 111, 1382213827. 17. Lo Leggio, L.; Simmons, T. J.; Poulsen, J. C.; Frandsen, K. E.; Hemsworth, G. R.; Stringer, M. A.; von Freiesleben, P.; Tovborg, M.; Johansen, K. S.; De Maria, L.; Harris, P. V.; Soong, C. L.; Dupree, P.; Tryfona, T.; Lenfant, N.; Henrissat, B.; Davies, G. J.; Walton, P. H. Structure and Boosting Activity of a Starch-degrading Lytic Polysaccharide Monooxygenase. Nat. Commun. 2015, 6, 5961. 18. Levasseur, A.; Drula, E.; Lombard, V.; Coutinho, P. M.; Henrissat, B. Expansion of the Enzymatic Repertoire of the CAZy Database to Integrate Auxiliary Redox Enzymes. Biotechnol. Biofuels 2013, 6, 41. 19. Henrissat, B. A Classification of Glycosyl Hydrolases Based on Amino Acid Sequence Similarities. Biochem. J. 1991, 280, 309-316. 20. El-Gebali, S.; Mistry, J.; Bateman, A.; Eddy, S. R.; Luciani, A.; Potter, S. C.; Qureshi, M.; Richardson, L. J.; Salazar, G. A.; Smart, A.; Sonnhammer, E. L L.; Hirsh, L.; Paladin, L.; Piovesan, D.; Tosatto, S. C E.; Finn, R. D. The Pfam Protein Families Database in 2019. Nucleic Acids Res. 2019, 47, D427-D432. 21. Ciano, L.; Davies, G. J.; Tolman, W. B.; Walton, P. H. Bracing Copper for the Catalytic Oxidation of C-H bonds. Nat. Catal. 2018, 1, 571-577. 22. Lieberman, R. L.; Rosenzweig, A. C. Crystal Structure of a Membrane-bound Metalloenzyme that Catalyses the Biological Oxidation of Methane. Nature 2005, 434, 177-182. 23. Cao, L. L.; Caldararu, O.; Rosenzweig, A. C.; Ryde, U. Quantum Refinement Does Not Support Dinuclear Copper Sites in Crystal Structures of Particulate Methane Monooxygenase. Angew. Chem. Int. Ed. 2018, 57, 162-166. 47 ACS Paragon Plus Environment

ACS Catalysis 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 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

1158 1159 1160 1161 1162 1163 1164 1165 1166 1167 1168 1169 1170 1171 1172 1173 1174 1175 1176 1177 1178 1179 1180 1181 1182 1183 1184 1185 1186 1187 1188 1189 1190 1191 1192 1193 1194 1195 1196 1197 1198 1199 1200 1201 1202 1203

24. Aachmann, F. L.; Sorlie, M.; Skjak-Braek, G.; Eijsink, V. G. H.; Vaaje-Kolstad, G. NMR Structure of a Lytic Polysaccharide Monooxygenase Provides Insight into Copper Binding, Protein Dynamics, and Substrate Interactions. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 1877918784. 25. Beeson, W. T.; Vu, V. V.; Span, E. A.; Phillips, C. M.; Marletta, M. A. Cellulose Degradation by Polysaccharide Monooxygenases. Annu. Rev. Biochem. 2015, 84, 923-946. 26. Walton, P. H.; Davies, G. J. On the Catalytic Mechanisms of Lytic Polysaccharide Monooxygenases. Curr. Opin. Chem. Biol. 2016, 31, 195-207. 27. Frommhagen, M.; Koetsier, M. J.; Westphal, A. H.; Visser, J.; Hinz, S. W. A.; Vincken, J. P.; van Berkel, W. J. H.; Kabel, M. A.; Gruppen, H. Lytic Polysaccharide Monooxygenases from Myceliophthora thermophila C1 Differ in Substrate Preference and Reducing Agent Specificity. Biotechnol. Biofuels 2016, 9, 186. 28. Kracher, D.; Scheiblbrandner, S.; Felice, A. K. G.; Breslmayr, E.; Preims, M.; Ludwicka, K.; Haltrich, D.; Eijsink, V. G. H.; Ludwig, R. Extracellular Electron Transfer Systems Fuel Cellulose Oxidative Degradation. Science 2016, 352, 1098-1101. 29. Garajova, S.; Mathieu, Y.; Beccia, M. R.; Bennati-Granier, C.; Biaso, F.; Fanuel, M.; Ropartz, D.; Guigliarelli, B.; Record, E.; Rogniaux, H.; Henrissat, B.; Berrin, J. G. Singledomain Flavoenzymes Trigger Lytic Polysaccharide Monooxygenases for Oxidative Degradation of Cellulose. Sci. Rep. 2016, 6, 28276. 30. Bissaro, B.; Røhr, Å. K.; Skaugen, M.; Forsberg, Z.; Horn, S. J.; Vaaje-Kolstad, G.; Eijsink, V. Fenton-type Chemistry by a Copper Enzyme: Molecular Mechanism of Polysaccharide Oxidative Cleavage. bioRxiv 2016, 097022. 31. Bissaro, B.; Røhr, A. K.; Müller, G.; Chylenski, P.; Skaugen, M.; Forsberg, Z.; Horn, S. J.; Vaaje-Kolstad, G.; Eijsink, V. G. H. Oxidative Cleavage of Polysaccharides by Monocopper Enzymes Depends on H2O2. Nat. Chem. Biol. 2017, 13, 1123-1128. 32. Bissaro, B.; Várnai, A.; Røhr, A. K.; Eijsink, V. G. H. Oxidoreductases and Reactive Oxygen Species in Conversion of Lignocellulosic Biomass. Microbiol. Mol. Biol. Rev. 2018, 82, e00029-18. 33. Hangasky, J. A.; Iavarone, A. T.; Marletta, M. A. Reactivity of O2 Versus H2O2 with Polysaccharide Monooxygenases. Proc. Natl. Acad. Sci. U. S. A. 2018, 115, 4915-4920. 34. Breslmayr, E.; Hanzek, M.; Hanrahan, A.; Leitner, C.; Kittl, R.; Santek, B.; Oostenbrink, C.; Ludwig, R. A Fast and Sensitive Activity Assay for Lytic Polysaccharide Monooxygenase. Biotechnol. Biofuels 2018, 11, 79. 35. Kuusk, S.; Bissaro, B.; Kuusk, P.; Forsberg, Z.; Eijsink, V. G. H.; Sørlie, M.; Väljamäe, P. Kinetics of H2O2-driven Degradation of Chitin by a Bacterial Lytic Polysaccharide Monooxygenase. J. Biol. Chem. 2018, 293, 523-531. 36. Kuusk, S.; Kont, R.; Kuusk, P.; Heering, A.; Sørlie, M.; Bissaro, B.; Eijsink, V. G. H.; Väljamäe, P. Kinetic Insights into the Role of the Reductant in H2O2-driven Degradation of Chitin by a Bacterial Lytic Polysaccharide Monooxygenase. J. Biol. Chem. 2019, 294, 15161528. 37. Müller, G.; Chylenski, P.; Bissaro, B.; Eijsink, V. G. H.; Horn, S. J. The Impact of Hydrogen Peroxide Supply on LPMO Activity and Overall Saccharification Efficiency of a Commercial Cellulase Cocktail. Biotechnol. Biofuels 2018, 11, 209. 38. Wang, D. M.; Li, J.; Wong, A. C. Y.; Aachmann, F. L.; Hsieh, Y. S. Y. A Colorimetric Assay to Rapidly Determine the Activities of Lytic Polysaccharide Monooxygenases. Biotechnol. Biofuels 2018, 11, 215. 48 ACS Paragon Plus Environment

Page 48 of 59

Page 49 of 59 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 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

1204 1205 1206 1207 1208 1209 1210 1211 1212 1213 1214 1215 1216 1217 1218 1219 1220 1221 1222 1223 1224 1225 1226 1227 1228 1229 1230 1231 1232 1233 1234 1235 1236 1237 1238 1239 1240 1241 1242 1243 1244 1245 1246 1247 1248

ACS Catalysis

39. Bissaro, B.; Isaksen, I.; Vaaje-Kolstad, G.; Eijsink, V. G. H.; Rohr, A. K. How a Lytic Polysaccharide Monooxygenase Binds Crystalline Chitin. Biochemistry 2018, 57, 1893-1906. 40. Horn, S. J.; Vaaje-Kolstad, G.; Westereng, B.; Eijsink, V. G. H. Novel Enzymes for the Degradation of Cellulose. Biotechnol. Biofuels 2012, 5, 45. 41. Eibinger, M.; Ganner, T.; Bubner, P.; Rosker, S.; Kracher, D.; Haltrich, D.; Ludwig, R.; Plank, H.; Nidetzky, B. Cellulose Surface Degradation by a Lytic Polysaccharide Monooxygenase and Its Effect on Cellulase Hydrolytic Efficiency. J. Biol. Chem. 2014, 289, 35929-35938. 42. Eibinger, M.; Sattelkow, J.; Ganner, T.; Plank, H.; Nidetzky, B. Single-molecule Study of Oxidative Enzymatic Deconstruction of Cellulose. Nat. Commun. 2017, 8, 894. 43. Vermaas, J. V.; Crowley, M. F.; Beckham, G. T.; Payne, C. M. Effects of Lytic Polysaccharide Monooxygenase Oxidation on Cellulose Structure and Binding of Oxidized Cellulose Oligomers to Cellulases. J. Phys. Chem. B 2015, 119, 6129-6143. 44. Harris, P. V.; Xu, F.; Kreel, N. E.; Kang, C.; Fukuyama, S. New Enzyme Insights Drive Advances in Commercial Ethanol Production. Curr. Opin. Chem. Biol. 2014, 19, 162-170. 45. Johansen, K. S. Discovery and Industrial Applications of Lytic Polysaccharide Monooxygenases. Biochem. Soc. Trans. 2016, 44, 143-149. 46. Cannella, D.; Hsieh, C. W. C.; Felby, C.; Jørgensen, H. Production and Effect of Aldonic Acids During Enzymatic Hydrolysis of Lignocellulose at High Dry Matter Content. Biotechnol. Biofuels 2012, 5, 26. 47. Chylenski, P.; Petrović, D. M.; Müller, G.; Dahlström, M.; Bengtsson, O.; Lersch, M.; Siika-aho, M.; Horn, S. J.; Eijsink, V. G. H. Enzymatic Degradation of Sulfite-pulped Softwoods and the Role of LPMOs. Biotechnol. Biofuels 2017, 10, 177. 48. Hu, J. G.; Arantes, V.; Pribowo, A.; Gourlay, K.; Saddler, J. N. Substrate Factors That Influence the Synergistic Interaction of AA9 and Cellulases During the Enzymatic Hydrolysis of Biomass. Energy Environ. Sci. 2014, 7, 2308-2315. 49. Hu, J. G.; Chandra, R.; Arantes, V.; Gourlay, K.; van Dyk, J. S.; Saddler, J. N. The Addition of Accessory Enzymes Enhances the Hydrolytic Performance of Cellulase Enzymes at High Solid Loadings. Bioresour. Technol. 2015, 186, 149-153. 50. Müller, G.; Várnai, A.; Johansen, K. S.; Eijsink, V. G. H.; Horn, S. J. Harnessing the Potential of LPMO-containing Cellulase Cocktails Poses New Demands on Processing Conditions. Biotechnol. Biofuels 2015, 8, 187. 51. Johansen, K. S. Lytic Polysaccharide Monooxygenases: the Microbial Power Tool for Lignocellulose Degradation. Trends Plant Sci. 2016, 21, 926-936. 52. Span, E. A.; Marietta, M. A. The Framework of Polysaccharide Monooxygenase Structure and Chemistry. Curr. Opin. Struct. Biol. 2015, 35, 93-99. 53. Frandsen, K. E. H.; Lo Leggio, L. Lytic Polysaccharide Monooxygenases: a Crystallographer's View on a New Class of Biomass-degrading Enzymes. IUCrJ 2016, 3, 448467. 54. Vaaje-Kolstad, G.; Forsberg, Z.; Loose, J. S.; Bissaro, B.; Eijsink, V. G. Structural Diversity of Lytic Polysaccharide Monooxygenases. Curr. Opin. Struct. Biol. 2017, 44, 67-76. 55. Meier, K. K.; Jones, S. M.; Kaper, T.; Hansson, H.; Koetsier, M. J.; Karkehabadi, S.; Solomon, E. I.; Sandgren, M.; Kelemen, B. Oxygen Activation by Cu LPMOs in Recalcitrant Carbohydrate Polysaccharide Conversion to Monomer Sugars. Chem. Rev. 2018, 118, 25932635.

49 ACS Paragon Plus Environment

ACS Catalysis 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 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

1249 1250 1251 1252 1253 1254 1255 1256 1257 1258 1259 1260 1261 1262 1263 1264 1265 1266 1267 1268 1269 1270 1271 1272 1273 1274 1275 1276 1277 1278 1279 1280 1281 1282 1283 1284 1285 1286 1287 1288 1289 1290 1291 1292 1293

56. Tandrup, T.; Frandsen, K. E. H.; Johansen, K. S.; Berrin, J. G.; Lo Leggio, L. Recent Insights into Lytic Polysaccharide Monooxygenases (LPMOs). Biochem. Soc. Trans. 2018, 46, 1431-1447. 57. Frandsen, K. E. H.; Simmons, T. J.; Dupree, P.; Poulsen, J. C. N.; Hemsworth, G. R.; Ciano, L.; Johnston, E. M.; Tovborg, M.; Johansen, K. S.; von Freiesleben, P.; Marmuse, L.; Fort, S.; Cottaz, S.; Driguez, H.; Henrissat, B.; Lenfant, N.; Tuna, F.; Baldansuren, A.; Davies, G. J.; Lo Leggio, L.; Walton, P. H. The Molecular Basis of Polysaccharide Cleavage by Lytic Polysaccharide Monooxygenases. Nat. Chem. Biol. 2016, 12, 298-303. 58. Zhang, L.; Koay, M.; Mahert, M. J.; Xiao, Z.; Wedd, A. G. Intermolecular Transfer of Copper Ions from the CopC Protein of Pseudomonas syringae. Crystal Structures of Fully Loaded CuICuII Forms. J. Am. Chem. Soc. 2006, 128, 5834-5850. 59. Forsberg, Z.; Mackenzie, A. K.; Sørlie, M.; Røhr, A. K.; Helland, R.; Arvai, A. S.; VaajeKolstad, G.; Eijsink, V. G. H. Structural and Functional Characterization of a Conserved Pair of Bacterial Cellulose-oxidizing Lytic Polysaccharide Monooxygenases. Proc. Natl. Acad. Sci. U. S. A. 2014, 111, 8446-8451. 60. Petrović, D. M.; Bissaro, B.; Chylenski, P.; Skaugen, M.; Sørlie, M.; Jensen, M. S.; Aachmann, F. L.; Courtade, G.; Várnai, A.; Eijsink, V. G. H. Methylation of the N-terminal Histidine Protects a Lytic Polysaccharide Monooxygenase from Auto-oxidative Inactivation. Protein Sci. 2018, 27, 1636-1650. 61. Hemsworth, G. R.; Taylor, E. J.; Kim, R. Q.; Gregory, R. C.; Lewis, S. J.; Turkenburg, J. P.; Parkin, A.; Davies, G. J.; Walton, P. H. The Copper Active Site of CBM33 Polysaccharide Oxygenases. J. Am. Chem. Soc. 2013, 135, 6069-6077. 62. Wu, M.; Beckham, G. T.; Larsson, A. M.; Ishida, T.; Kim, S.; Payne, C. M.; Himmel, M. E.; Crowley, M. F.; Horn, S. J.; Westereng, B.; Igarashi, K.; Samejima, M.; Stahlberg, J.; Eijsink, V. G. H.; Sandgren, M. Crystal Structure and Computational Characterization of the Lytic Polysaccharide Monooxygenase GH61D from the Basidiomycota Fungus Phanerochaete chrysosporium. J. Biol. Chem. 2013, 288, 12828-12839. 63. Vu, V. V.; Beeson, W. T.; Phillips, C. M.; Cate, J. H. D.; Marletta, M. A. Determinants of Regioselective Hydroxylation in the Fungal Polysaccharide Monooxygenases. J. Am. Chem. Soc. 2014, 136, 562-565. 64. Danneels, B.; Tanghe, M.; Joosten, H. J.; Gundinger, T.; Spadiut, O.; Stals, I.; Desmet, T. A Quantitative Indicator Diagram for Lytic Polysaccharide Monooxygenases Reveals the Role of Aromatic Surface Residues in HjLPMO9A Regioselectivity. PLoS One 2017, 12, e0178446. 65. Danneels, B.; Tanghe, M.; Desmet, T. Structural Features on the Substrate-binding Surface of Fungal Lytic Polysaccharide Monooxygenases Determine Their Oxidative Regioselectivity. Biotechnol. J. 2019, 14, e1800211. 66. Forsberg, Z.; Bissaro, B.; Gullesen, J.; Dalhus, B.; Vaaje-Kolstad, G.; Eijsink, V. G. H. Structural Determinants of Bacterial Lytic Polysaccharide Monooxygenase Functionality. J. Biol. Chem. 2018, 293, 1397-1412. 67. Book, A. J.; Yennamalli, R. M.; Takasuka, T. E.; Currie, C. R.; Phillips, G. N.; Fox, B. G. Evolution of Substrate Specificity in Bacterial AA10 Lytic Polysaccharide Monooxygenases. Biotechnol. Biofuels 2014, 7, 109. 68. Li, X.; Beeson, W. T.; Phillips, C. M.; Marletta, M. A.; Cate, J. H. D. Structural Basis for Substrate Targeting and Catalysis by Fungal Polysaccharide Monooxygenases. Structure 2012, 20, 1051-1061.

50 ACS Paragon Plus Environment

Page 50 of 59

Page 51 of 59 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 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

1294 1295 1296 1297 1298 1299 1300 1301 1302 1303 1304 1305 1306 1307 1308 1309 1310 1311 1312 1313 1314 1315 1316 1317 1318 1319 1320 1321 1322 1323 1324 1325 1326 1327 1328 1329 1330 1331 1332 1333 1334 1335 1336 1337 1338

ACS Catalysis

69. Beeson, W. T.; Phillips, C. M.; Cate, J. H. D.; Marletta, M. A. Oxidative Cleavage of Cellulose by Fungal Copper-dependent Polysaccharide Monooxygenases. J. Am. Chem. Soc. 2012, 134, 890-892. 70. Kim, S.; Stahlberg, J.; Sandgren, M.; Paton, R. S.; Beckham, G. T. Quantum Mechanical Calculations Suggest That Lytic Polysaccharide Monooxygenases Use a Copper-oxyl, Oxygenrebound Mechanism. Proc. Natl. Acad. Sci. U. S. A. 2014, 111, 149-154. 71. O'Dell, W. B.; Agarwal, P. K.; Meilleur, F. Oxygen Activation at the Active Site of a Fungal Lytic Polysaccharide Monooxygenase. Angew. Chem. Int. Ed. 2017, 56, 767-770. 72. Bacik, J. P.; Mekasha, S.; Forsberg, Z.; Kovalevsky, A. Y.; Vaaje-Kolstad, G.; Eijsink, V. G. H.; Nix, J. C.; Coates, L.; Cuneo, M. J.; Unkefer, C. J.; Chen, J. C. H. Neutron and Atomic Resolution X-ray Structures of a Lytic Polysaccharide Monooxygenase Reveal Copper-mediated Dioxygen Binding and Evidence for N-terminal Deprotonation. Biochemistry 2017, 56, 25292532. 73. Hedegård, E. D.; Ryde, U. Targeting the Reactive Intermediate in Polysaccharide Monooxygenases. J. Biol. Inorg. Chem. 2017, 22, 1029-1037. 74. Hedegård, E. D.; Ryde, U, Molecular Mechanism of Lytic Polysaccharide Monooxygenases. Chem. Sci. 2018, 9, 3866-3880. 75. Hemsworth, G. R.; Davies, G. J.; Walton, P. H. Recent Insights into Copper-containing Lytic Polysaccharide Mono-oxygenases. Curr. Opin. Struct. Biol. 2013, 23, 660-668. 76. Donoghue, P. J.; Tehranchi, J.; Cramer, C. J.; Sarangi, R.; Solomon, E. I.; Tolman, W. B. Rapid C-H Bond Activation by a Monocopper(III)-hydroxide Complex. J. Am. Chem. Soc. 2011, 133, 17602-17605. 77. Himes, R. A.; Karlin, K. D. Copper-dioxygen Complex Mediated C-H Bond Oxygenation: Relevance for Particulate Methane Monooxygenase (pMMO). Curr. Opin. Chem. Biol. 2009, 13, 119-131. 78. Bertini, L.; Breglia, R.; Lambrughi, M.; Fantucci, P.; De Gioia, L.; Borsari, M.; Sola, M.; Bortolotti, C. A.; Bruschi, M. Catalytic Mechanism of Fungal Lytic Polysaccharide Monooxygenases Investigated by First-principles Calculations. Inorg. Chem. 2018, 57, 86-97. 79. Wang, B. J.; Johnston, E. M.; Li, P. F.; Shaik, S.; Davies, G. J.; Walton, P. H.; Rovira, C. QM/MM Studies into the H2O2-dependent Activity of Lytic Polysaccharide Monooxygenases: Evidence for the Formation of a Caged Hydroxyl Radical Intermediate. ACS Catal. 2018, 8, 1346-1351. 80. Kjaergaard, C. H.; Qayyum, M. F.; Wong, S. D.; Xu, F.; Hemsworth, G. R.; Walton, D. J.; Young, N. A.; Davies, G. J.; Walton, P. H.; Johansen, K. S.; Hodgson, K. O.; Hedman, B.; Solomon, E. I. Spectroscopic and Computational Insight into the Activation of O2 by the Mononuclear Cu Center in Polysaccharide Monooxygenases. Proc. Natl. Acad. Sci. U. S. A. 2014, 111, 8797-8802. 81. Borisova, A. S.; Isaksen, T.; Dimarogona, M.; Kognole, A. A.; Mathiesen, G.; Várnai, A.; Røhr, A. K.; Payne, C. M.; Sørlie, M.; Sandgren, M.; Eijsink, V. G. H. Structural and Functional Characterization of a Lytic Polysaccharide Monooxygenase with Broad Substrate Specificity. J. Biol. Chem. 2015, 290, 22955-22969. 82. Hangasky, J. A.; Marletta, M. A. A Random-sequential Kinetic Mechanism for Polysaccharide Monooxygenases. Biochemistry 2018, 57, 3191-3199. 83. Cragg, S. M.; Beckham, G. T.; Bruce, N. C.; Bugg, T. D. H.; Distel, D. L.; Dupree, P.; Etxabe, A. G.; Goodell, B. S.; Jellison, J.; McGeehan, J. E.; McQueen-Mason, S. J.; Schnorr, K.;

51 ACS Paragon Plus Environment

ACS Catalysis 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 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

1339 1340 1341 1342 1343 1344 1345 1346 1347 1348 1349 1350 1351 1352 1353 1354 1355 1356 1357 1358 1359 1360 1361 1362 1363 1364 1365 1366 1367 1368 1369 1370 1371 1372 1373 1374 1375 1376 1377 1378 1379 1380 1381 1382 1383

Walton, P. H.; Watts, J. E. M.; Zimmer, M. Lignocellulose Degradation Mechanisms Across the Tree of Life. Curr. Opin. Chem. Biol. 2015, 29, 108-119. 84. Loose, J. S. M.; Forsberg, Z.; Fraaije, M. W.; Eijsink, V. G. H.; Vaaje-Kolstad, G. A Rapid Quantitative Activity Assay Shows That the Vibrio cholerae Colonization Factor GbpA Is an Active Lytic Polysaccharide Monooxygenase. FEBS Lett. 2014, 588, 3435-3440. 85. Hofrichter, M.; Ullrich, R. Oxidations Catalyzed by Fungal Peroxygenases. Curr. Opin. Chem. Biol. 2014, 19, 116-125. 86. Matthews, S.; Belcher, J. D.; Tee, K. L.; Girvan, H. M.; McLean, K. J.; Rigby, S. E. J.; Levy, C. W.; Leys, D.; Parker, D. A.; Blankley, R. T.; Munro, A. W. Catalytic Determinants of Alkene Production by the Cytochrome P450 Peroxygenase OleTJE. J. Biol. Chem. 2017, 292, 5128-5143. 87. Shoji, O.; Watanabe, Y. Peroxygenase Reactions Catalyzed by Cytochromes P450. J Biol Inorg. Chem. 2014, 19, 529-539. 88. Hrycay, E. G.; Bandiera, S. M. Monooxygenase, Peroxidase and Peroxygenase Properties and Reaction Mechanisms of Cytochrome P450 Enzymes. In Monooxygenase, Peroxidase and Peroxygenase Properties and Mechanisms of Cytochrome P450, Hrycay, E. G.; Bandiera, S. M., Eds.; Springer International Publishing: Cham, 2015; Vol. 851, pp 1-61. 89. Krauser, J. A.; Guengerich, F. P. Cytochrome P450 3A4-catalyzed Testosterone 6βHydroxylation Stereochemistry, Kinetic Deuterium Isotope Effects, and Rate-limiting Steps. J. Biol. Chem. 2005, 280, 19496-19506. 90. Fukami, T.; Katoh, M.; Yamazaki, H.; Yokoi, T.; Nakajima, M. Human Cytochrome P450 2A13 Efficiently Metabolizes Chemicals in Air Pollutants: Naphthalene, Styrene, and Toluene. Chem. Res. Toxicol. 2008, 21, 720-725. 91. Kumar, S.; Liu, H.; Halpert, J. R. Engineering of Cytochrome P450 3A4 for Enhanced Peroxide-mediated Substrate Oxidation Using Directed Evolution and Site-directed Mutagenesis. Drug Metab. Dispos. 2006, 34, 1958-1965. 92. Joo, H.; Lin, Z. L.; Arnold, F. H. Laboratory Evolution of Peroxide-mediated Cytochrome P450 Hydroxylation. Nature 1999, 399, 670-673. 93. Girhard, M.; Kunigk, E.; Tihovsky, S.; Shumyantseva, V. V.; Urlacher, V. B. Lightdriven Biocatalysis with Cytochrome P450 Peroxygenases. Biotechnol. Appl. Biochem. 2013, 60, 111-118. 94. Shoji, O.; Fujishiro, T.; Nakajima, H.; Kim, M.; Nagano, S.; Shiro, Y.; Watanabe, Y. Hydrogen Peroxide Dependent Monooxygenations by Tricking the Substrate Recognition of Cytochrome P450BSβ. Angew. Chem. Int. Ed. 2007, 46, 3656-3659. 95. Girhard, M.; Schuster, S.; Dietrich, M.; Dürre, P.; Urlacher, V. B. Cytochrome P450 Monooxygenase from Clostridium acetobutylicum: a New α-Fatty Acid Hydroxylase. Biochem. Biophys. Res. Commun. 2007, 362, 114-119. 96. Nordblom, G. D.; White, R. E.; Coon, M. J. Studies on Hydroperoxide-dependent Substrate Hydroxylation by Purified Liver Microsomal Cytochrome P-450. Arch. Biochem. Biophys. 1976, 175, 524-533. 97. Li, Q.-S.; Ogawa, J.; Shimizu, S. Critical Role of the Residue Size at Position 87 in H2O2-dependent Substrate Hydroxylation Activity and H2O2 Inactivation of Cytochrome P450BM-3. Biochem. Biophys. Res. Commun. 2001, 280, 1258-1261. 98. Cirino, P. C.; Arnold, F. H. A Self-sufficient Peroxide-driven Hydroxylation Biocatalyst. Angew. Chem. Int .Ed. 2003, 42, 3299-3301.

52 ACS Paragon Plus Environment

Page 52 of 59

Page 53 of 59 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 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

1384 1385 1386 1387 1388 1389 1390 1391 1392 1393 1394 1395 1396 1397 1398 1399 1400 1401 1402 1403 1404 1405 1406 1407 1408 1409 1410 1411 1412 1413 1414 1415 1416 1417 1418 1419 1420 1421 1422 1423 1424 1425 1426 1427 1428

ACS Catalysis

99. Fujishiro, T.; Shoji, O.; Nagano, S.; Sugimoto, H.; Shiro, Y.; Watanabe, Y. Crystal Structure of H2O2-dependent Cytochrome P450SPα with Its Bound Fatty Acid Substrate: Insight into the Regioselective Hydroxylation of Fatty Acids at the Alpha Position. J. Biol. Chem. 2011, 286, 29941-29950. 100. Girvan, H. M.; Poddar, H.; McLean, K. J.; Nelson, D. R.; Hollywood, K. A.; Levy, C. W.; Leys, D.; Munro, A. W. Structural and Catalytic Properties of the Peroxygenase P450 Enzyme CYP152K6 from Bacillus methanolicus. J. Inorg. Biochem. 2018, 188, 18-28. 101. Matsunaga, I.; Yamada, M.; Kusunose, E.; Miki, T.; Ichihara, K. Further Characterization of Hydrogen Peroxide-dependent Fatty Acid α-Hydroxylase from Sphingomonas paucimobilis. J. Biochem. 1998, 124, 105-110. 102. Lee, D.-S.; Yamada, A.; Sugimoto, H.; Matsunaga, I.; Ogura, H.; Ichihara, K.; Adachi, S.-i.; Park, S.-Y.; Shiro, Y. Substrate Recognition and Molecular Mechanism of Fatty Acid Hydroxylation by Cytochrome P450 from Bacillus subtilis: Crystallographic, Spectroscopic, and Mutational Studies. J. Biol. Chem. 2003, 278, 9761-9767. 103. Matsunaga, I.; Ueda, A.; Sumimoto, T.; Ichihara, K.; Ayata, M.; Ogura, H. Site-directed Mutagenesis of the Putative Distal Helix of Peroxygenase Cytochrome P450. Arch. Biochem. Biophys. 2001, 394, 45-53. 104. Span, E. A.; Suess, D. L. M.; Deller, M. C.; Britt, R. D.; Marletta, M. A. The Role of the Secondary Coordination Sphere in a Fungal Polysaccharide Monooxygenase. ACS Chem. Biol. 2017, 12, 1095-1103. 105. Dimarogona, M.; Topakas, E.; Olsson, L.; Christakopoulos, P. Lignin Boosts the Cellulase Performance of a GH-61 Enzyme from Sporotrichum thermophile. Bioresour. Technol. 2012, 110, 480-487. 106. Westereng, B.; Cannella, D.; Agger, J. W.; Jørgensen, H.; Andersen, M. L.; Eijsink, V. G. H.; Felby, C. Enzymatic Cellulose Oxidation is Linked to Lignin by Long-range Electron Transfer. Sci. Rep. 2015, 5, 18561. 107. Muraleedharan, M. N.; Zouraris, D.; Karantonis, A.; Topakas, E.; Sandgren, M.; Rova, U.; Christakopoulos, P.; Karnaouri, A. Effect of Lignin Fractions Isolated from Different Biomass Sources on Cellulose Oxidation by Fungal Lytic Polysaccharide Monooxygenases. Biotechnol. Biofuels 2018, 11, 296. 108. Brenelli, L.; Squina, F. M.; Felby, C.; Cannella, D. Laccase-derived Lignin Compounds Boost Cellulose Oxidative Enzymes AA9. Biotechnol. Biofuels 2018, 11, 10. 109. Várnai, A.; Umezawa, K.; Yoshida, M.; Eijsink, V. G. H. The Pyrroloquinoline-quinonedependent Pyranose Dehydrogenase from Coprinopsis cinerea Drives Lytic Polysaccharide Monooxygenase Action. Appl. Environ. Microbiol. 2018, 84, e00156-18. 110. Tan, T. C.; Kracher, D.; Gandini, R.; Sygmund, C.; Kittl, R.; Haltrich, D.; Hallberg, B. M.; Ludwig, R.; Divne, C. Structural Basis for Cellobiose Dehydrogenase Action During Oxidative Cellulose Degradation. Nat. Commun. 2015, 6, 7542. 111. Courtade, G.; Wimmer, R.; Røhr, A. K.; Preims, M.; Felice, A. K. G.; Dimarogona, M.; Vaaje-Kolstad, G.; Sørlie, M.; Sandgren, M.; Ludwig, R.; Eijsink, V. G. H.; Aachmann, F. L. Interactions of a Fungal Lytic Polysaccharide Monooxygenase with β-Glucan Substrates and Cellobiose Dehydrogenase. Proc. Natl. Acad. Sci. U. S. A. 2016, 113, 5922-5927. 112. Loose, J. S. M.; Forsberg, Z.; Kracher, D.; Scheiblbrandner, S.; Ludwig, R.; Eijsink, V. G. H.; Vaaje-Kolstad, G. Activation of Bacterial Lytic Polysaccharide Monooxygenases with Cellobiose Dehydrogenase. Protein Sci. 2016, 25, 2175-2186.

53 ACS Paragon Plus Environment

ACS Catalysis 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 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

1429 1430 1431 1432 1433 1434 1435 1436 1437 1438 1439 1440 1441 1442 1443 1444 1445 1446 1447 1448 1449 1450 1451 1452 1453 1454 1455 1456 1457 1458 1459 1460 1461 1462 1463 1464 1465 1466 1467 1468 1469 1470 1471 1472 1473

113. Zamocky, M.; Ludwig, R.; Peterbauer, C.; Hallberg, B. M.; Divne, C.; Nicholls, P.; Haltrich, D. Cellobiose Dehydrogenase - a Flavocytochrome from Wood-degrading, Phytopathogenic and Saprotropic Fungi. Curr. Protein Pept. Sci. 2006, 7, 255-280. 114. Harada, H.; Onoda, A.; Uchihashi, T.; Watanabe, H.; Sunagawa, N.; Samejima, M.; Igarashi, K.; Hayashi, T. Interdomain Flip-flop Motion Visualized in Flavocytochrome Cellobiose Dehydrogenase Using High-speed Atomic Force Microscopy During Catalysis. Chem. Sci. 2017, 8, 6561-6565. 115. Henriksson, G.; Johansson, G.; Pettersson, G. A Critical Review of Cellobiose Dehydrogenases. J. Biotechnol. 2000, 78, 93-113. 116. Igarashi, K.; Momohara, I.; Nishino, T.; Samejima, M. Kinetics of Inter-domain Electron Transfer in Flavocytochrome Cellobiose Dehydrogenase from the White-rot Fungus Phanerochaete chrysosporium. Biochem. J. 2002, 365, 521-526. 117. Mason, M. G.; Nicholls, P.; Wilson, M. T. Rotting by Radicals - the Role of Cellobiose Oxidoreductase? Biochem. Soc. Trans. 2003, 31, 1335-1336. 118. Sygmund, C.; Santner, P.; Krondorfer, I.; Peterbauer, C. K.; Alcalde, M.; Nyanhongo, G. S.; Guebitz, G. M.; Ludwig, R. Semi-rational Engineering of Cellobiose Dehydrogenase for Improved Hydrogen Peroxide Production. Microb. Cell Fact. 2013, 12, 38. 119. Matsumura, H.; Umezawa, K.; Takeda, K.; Sugimoto, N.; Ishida, T.; Samejima, M.; Ohno, H.; Yoshida, M.; Igarashi, K.; Nakamura, N. Discovery of a Eukaryotic Pyrroloquinoline Quinone-dependent Oxidoreductase Belonging to a New Auxiliary Activity Family in the Database of Carbohydrate-active Enzymes. PLoS One 2014, 9, e104851. 120. Cannella, D.; Möllers, K. B.; Frigaard, N. U.; Jensen, P. E.; Bjerrum, M. J.; Johansen, K. S.; Felby, C. Light-driven Oxidation of Polysaccharides by Photosynthetic Pigments and a Metalloenzyme. Nat. Commun. 2016, 7, 11134. 121. Bissaro, B.; Forsberg, Z.; Ni, Y.; Hollmann, F.; Vaaje-Kolstad, G.; Eijsink, V. G. H. Fueling Biomass-degrading Oxidative Enzymes by Light-driven Water Oxidation. Green Chem. 2016, 18, 5357-5366. 122. Möllers, K. B.; Mikkelsen, H.; Simonsen, T. I.; Cannella, D.; Johansen, K. S.; Bjerrum, M. J.; Felby, C. On the Formation and Role of Reactive Oxygen Species in Light-driven LPMO Oxidation of Phosphoric Acid Swollen Cellulose. Carbohydr. Res. 2017, 448, 182-186. 123. Frommhagen, M.; Westphal, A. H.; Hilgers, R.; Koetsier, M. J.; Hinz, S. W. A.; Visser, J.; Gruppen, H.; van Berkel, W. J. H.; Kabel, M. A. Quantification of the Catalytic Performance of C1-cellulose-specific Lytic Polysaccharide Monooxygenases. Appl. Microbiol. Biotechnol. 2018, 102, 1281-1295. 124. Gusakov, A. V.; Bulakhov, A. G.; Demin, I. N.; Sinitsyn, A. P. Monitoring of Reactions Catalyzed by Lytic Polysaccharide Monooxygenases Using Highly-sensitive Fluorimetric Assay of the Oxygen Consumption Rate. Carbohydr. Res. 2017, 452, 156-161. 125. Hegnar, O. A.; Petrović, D. M.; Bissaro, B.; Alfredsen, G.; Várnai, A.; Eijsink, V. G. H. pH-Dependent Relationship between Catalytic Activity and Hydrogen Peroxide Production Shown via Characterization of a Lytic Polysaccharide Monooxygenase from Gloeophyllum trabeum. Appl. Environ. Microbiol. 2019, 85, e02612-18. 126. Simmons, T. J.; Frandsen, K. E. H.; Ciano, L.; Tryfona, T.; Lenfant, N.; Poulsen, J. C.; Wilson, L. F. L.; Tandrup, T.; Tovborg, M.; Schnorr, K.; Johansen, K. S.; Henrissat, B.; Walton, P. H.; Lo Leggio, L.; Dupree, P. Structural and Electronic Determinants of Lytic Polysaccharide Monooxygenase Reactivity on Polysaccharide Substrates. Nat. Commun. 2017, 8, 1064.

54 ACS Paragon Plus Environment

Page 54 of 59

Page 55 of 59 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 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

1474 1475 1476 1477 1478 1479 1480 1481 1482 1483 1484 1485 1486 1487 1488 1489 1490 1491 1492 1493 1494 1495 1496 1497 1498 1499 1500 1501 1502 1503 1504 1505 1506 1507 1508 1509 1510 1511 1512 1513 1514 1515 1516 1517 1518

ACS Catalysis

127. Kracher, D.; Andlar, M.; Furtmuller, P. G.; Ludwig, R. Active-site Copper Reduction Promotes Substrate Binding of Fungal Lytic Polysaccharide Monooxygenase and Reduces Stability. J. Biol. Chem. 2018, 293, 1676-1687. 128. Loose, J. S. M.; Arntzen, M. O.; Bissaro, B.; Ludwig, R.; Eijsink, V. G. H.; VaajeKolstad, G. Multipoint Precision Binding of Substrate Protects Lytic Polysaccharide Monooxygenases from Self-destructive Off-pathway Processes. Biochemistry 2018, 57, 41144124. 129. Mutahir, Z.; Mekasha, S.; Loose, J. S. M.; Abbas, F.; Vaaje-Kolstad, G.; Eijsink, V. G. H.; Forsberg, Z. Characterization and Synergistic Action of a Tetra-modular Lytic Polysaccharide Monooxygenase from Bacillus cereus. FEBS Lett. 2018, 592, 2562-2571. 130. Courtade, G.; Forsberg, Z.; Heggset, E. B.; Eijsink, V. G. H.; Aachmann, F. L. The Carbohydrate-binding Module and Linker of a Modular Lytic Polysaccharide Monooxygenase Promote Localized Cellulose Oxidation. J. Biol. Chem. 2018, 293, 13006-13015. 131. Liu, P. H.; Murakami, K.; Seki, T.; He, X. M.; Yeung, S. M.; Kuzuyama, T.; Seto, H.; Liu, H. W. Protein Purification and Function Assignment of the Epoxidase Catalyzing the Formation of Fosfomycin. J. Am. Chem. Soc. 2001, 123, 4619-4620. 132. Wang, C.; Chang, W. C.; Guo, Y. S.; Huang, H.; Peck, S. C.; Pandelia, M. E.; Lin, G. M.; Liu, H. W.; Krebs, C.; Bollinger, J. M. Evidence That the Fosfomycin-producing Epoxidase, HppE, Is a Non-heme-iron Peroxidase. Science 2013, 342, 991-995. 133. Kittl, R.; Kracher, D.; Burgstaller, D.; Haltrich, D.; Ludwig, R. Production of Four Neurospora crassa Lytic Polysaccharide Monooxygenases in Pichia pastoris Monitored by a Fluorimetric Assay. Biotechnol. Biofuels 2012, 5, 79. 134. Galbe, M.; Zacchi, G. Pretreatment: the Key to Efficient Utilization of Lignocellulosic Materials. Biomass Bioenergy 2012, 46, 70-78. 135. Teeri, T. T.; Koivula, A.; Linder, M.; Wohlfahrt, G.; Divne, C.; Jones, T. A. Trichoderma reesei Cellobiohydrolases: Why So Efficient on Crystalline Cellulose? Biochem. Soc. Trans. 1998, 26, 173-178. 136. Penttila, P. A.; Várnai, A.; Pere, J.; Tammelin, T.; Salmen, L.; Siika-aho, M.; Viikari, L.; Serimaa, R. Xylan as Limiting Factor in Enzymatic Hydrolysis of Nanocellulose. Bioresour. Technol. 2013, 129, 135-141. 137. Yu, L.; Lyczakowski, J. J.; Pereira, C. S.; Kotake, T.; Yu, X. L.; Li, A.; Mogelsvang, S.; Skaf, M. S.; Dupree, P. The Patterned Structure of Galactoglucomannan Suggests It May Bind to Cellulose in Seed Mucilage. Plant Physiol. 2018, 178, 1011-1026. 138. Busse-Wicher, M.; Gomes, T. C. F.; Tryfona, T.; Nikolovski, N.; Stott, K.; Grantham, N. J.; Bolam, D. N.; Skaf, M. S.; Dupree, P. The Pattern of Xylan Acetylation Suggests Xylan May Interact with Cellulose Microfibrils as a Twofold Helical Screw in the Secondary Plant Cell Wall of Arabidopsis thaliana. Plant J. 2014, 79, 492-506. 139. Busse-Wicher, M.; Grantham, N. J.; Lyczakowski, J. J.; Nikolovski, N.; Dupree, P. Xylan Decoration Patterns and the Plant Secondary Cell Wall Molecular Architecture. Biochem. Soc. Trans. 2016, 44, 74-78. 140. Berlin, A.; Maximenko, V.; Gilkes, N.; Saddler, J. Optimization of Enzyme Complexes for Lignocellulose Hydrolysis. Biotechnol. Bioeng. 2007, 97, 287-96. 141. Selig, M. J.; Knoshaug, E. P.; Adney, W. S.; Himmel, M. E.; Decker, S. R. Synergistic Enhancement of Cellobiohydrolase Performance on Pretreated Corn Stover by Addition of Xylanase and Esterase Activities. Bioresour. Technol. 2008, 99, 4997-5005.

55 ACS Paragon Plus Environment

ACS Catalysis 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 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

1519 1520 1521 1522 1523 1524 1525 1526 1527 1528 1529 1530 1531 1532 1533 1534 1535 1536 1537 1538 1539 1540 1541 1542 1543 1544 1545 1546 1547 1548 1549 1550 1551 1552 1553 1554 1555 1556 1557 1558 1559 1560 1561 1562 1563

142. Várnai, A.; Huikko, L.; Pere, J.; Siika-Aho, M.; Viikari, L. Synergistic Action of Xylanase and Mannanase Improves the Total Hydrolysis of Softwood. Bioresour. Technol. 2011, 102, 9096-104. 143. Hu, J. G.; Arantes, V.; Saddler, J. N. The Enhancement of Enzymatic Hydrolysis of Lignocellulosic Substrates by the Addition of Accessory Enzymes Such as Xylanase: Is It an Additive or Synergistic Effect? Biotechnol. Biofuels 2011, 4, 36. 144. Chylenski, P.; Forsberg, Z.; Stahlberg, J.; Várnai, A.; Lersch, M.; Bengtsson, O.; Saebo, S.; Horn, S. J.; Eijsink, V. G. Development of Minimal Enzyme Cocktails for Hydrolysis of Sulfite-pulped Lignocellulosic Biomass. J. Biotechnol. 2017, 246, 16-23. 145. Hu, J. G.; Arantes, V.; Pribowo, A.; Saddler, J. N. The Synergistic Action of Accessory Enzymes Enhances the Hydrolytic Potential of a "Cellulase Mixture" But Is Highly Substrate Specific. Biotechnol. Biofuels 2013, 6, 112. 146. Kojima, Y.; Várnai, A.; Ishida, T.; Sunagawa, N.; Petrovic, D. M.; Igarashi, K.; Jellison, J.; Goodell, B.; Alfredsen, G.; Westereng, B.; Eijsink, V. G. H.; Yoshida, M. A Lytic Polysaccharide Monooxygenase with Broad Xyloglucan Specificity from the Brown-rot Fungus Gloeophyllum trabeum and Its Action on Cellulose-xyloglucan Complexes. Appl. Environ. Microbiol. 2016, 82, 6557-6572. 147. Grantham, N. J.; Wurman-Rodrich, J.; Terrett, O. M.; Lyczakowski, J. J.; Stott, K.; Iuga, D.; Simmons, T. J.; Durand-Tardif, M.; Brown, S. P.; Dupree, R.; Busse-Wicher, M.; Dupree, P. An Even Pattern of Xylan Substitution Is Critical for Interaction with Cellulose in Plant Cell Walls. Nat. Plants 2017, 3, 859-865. 148. Humbird, D.; Davis, R.; Tao, L.; Kinchin, C.; Hsu, D.; Aden, A.; Schoen, P.; Lukas, J.; Olthof, B.; Worley, M.; Sexton, D.; Dudgeon, D. Process Design and Economics for Biochemical Conversion of Lignocellulosic Biomass to Ethanol. Dilute-acid Pretreatment and Enzymatic Hydrolysis of Corn Stover; Technical Report for National Renewable Energy Laboratory: Golden, CO, May 2011:NREL/TP-5100-47764. 149. Adav, S. S.; Chao, L. T.; Sze, S. K. Quantitative Secretomic Analysis of Trichoderma reesei Strains Reveals Enzymatic Composition for Lignocellulosic Biomass Degradation. Mol. Cell. Proteomics 2012, 11, M111.012419. 150. Martinez, D.; Berka, R. M.; Henrissat, B.; Saloheimo, M.; Arvas, M.; Baker, S. E.; Chapman, J.; Chertkov, O.; Coutinho, P. M.; Cullen, D.; Danchin, E. G. J.; Grigoriev, I. V.; Harris, P.; Jackson, M.; Kubicek, C. P.; Han, C. S.; Ho, I.; Larrondo, L. F.; de Leon, A. L.; Magnuson, J. K.; Merino, S.; Misra, M.; Nelson, B.; Putnam, N.; Robbertse, B.; Salamov, A. A.; Schmoll, M.; Terry, A.; Thayer, N.; Westerholm-Parvinen, A.; Schoch, C. L.; Yao, J.; Barbote, R.; Nelson, M. A.; Detter, C.; Bruce, D.; Kuske, C. R.; Xie, G.; Richardson, P.; Rokhsar, D. S.; Lucas, S. M.; Rubin, E. M.; Dunn-Coleman, N.; Ward, M.; Brettin, T. S. Genome Sequencing and Analysis of the Biomass-degrading Fungus Trichoderma reesei (syn. Hypocrea jecorina). Nat. Biotechnol. 2008, 26, 553-560. 151. Scott, B. R.; Huang, H. Z.; Frickman, J.; Halvorsen, R.; Johansen, K. S. Catalase Improves Saccharification of Lignocellulose by Reducing Lytic Polysaccharide Monooxygenaseassociated Enzyme Inactivation. Biotechnol. Lett. 2016, 38, 425-434. 152. Du, J.; Song, W. X.; Zhang, X.; Zhao, J.; Liu, G. D.; Qu, Y. B. Differential Reinforcement of Enzymatic Hydrolysis by Adding Chemicals and Accessory Proteins to High Solid Loading Substrates with Different Pretreatments. Bioprocess Biosyst. Eng. 2018, 41, 11531163.

56 ACS Paragon Plus Environment

Page 56 of 59

Page 57 of 59 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 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

1564 1565 1566 1567 1568 1569 1570 1571 1572 1573 1574 1575 1576 1577 1578 1579 1580 1581 1582 1583 1584 1585 1586 1587 1588 1589 1590 1591 1592 1593 1594 1595 1596 1597 1598 1599 1600 1601 1602 1603 1604 1605 1606 1607 1608

ACS Catalysis

153. Kim, I. J.; Youn, H. J.; Kim, K. H, Synergism of an Auxiliary Activity 9 (AA9) from Chaetomium globosum with Xylanase on the Hydrolysis of Xylan and Lignocellulose. Process Biochem. 2016, 51, 1445-1451. 154. Sanhueza, C.; Carvajal, G.; Soto-Aguilar, J.; Lienqueo, M. E.; Salazar, O. The Effect of a Lytic Polysaccharide Monooxygenase and a Xylanase from Gloeophyllum trabeum on the Enzymatic Hydrolysis of Lignocellulosic Residues Using a Commercial Cellulase. Enzyme Microb. Tech. 2018, 113, 75-82. 155. Peciulyte, A.; Samuelsson, L.; Olsson, L.; McFarland, K. C.; Frickmann, J.; Østergård, L.; Halvorsen, R.; Scott, B. R.; Johansen, K. S. Redox Processes Acidify and Decarboxylate Steam-pretreated Lignocellulosic Biomass and Are Modulated by LPMO and Catalase. Biotechnol. Biofuels 2018, 11, 165. 156. Garcia-Ochoa, F.; Gomez, E.; Santos, V. E.; Merchuk, J. C. Oxygen Uptake Rate in Microbial Processes: an Overview. Biochem. Eng. J. 2010, 49, 289-307. 157. Sheldon, R. A. The Road to Biorenewables: Carbohydrates to Commodity Chemicals. ACS Sustain. Chem. Eng. 2018, 6, 4464-4480. 158. Wyman, C. E. Ethanol Production from Lignocellulosic Biomass: Overview. In Handbook on Bioethanol, Wyman, C. E., Ed.; Taylor & Francis: Washington, DC, 1996; pp 118. 159. Müller, G.; Kalyani, D. C.; Horn, S. J. LPMOs in Cellulase Mixtures Affect Fermentation Strategies for Lactic Acid Production from Lignocellulosic Biomass. Biotechnol. Bioeng. 2017, 114, 552-559. 160. de Oliveira, R. A.; Komesu, A.; Rossell, C. E. V.; Maciel, R. Challenges and Opportunities in Lactic Acid Bioprocess Design – from Economic to Production Aspects. Biochem. Eng. J. 2018, 133, 219-239. 161. Wingren, A.; Galbe, M.; Zacchi, G. Techno-economic Evaluation of Producing Ethanol from Softwood: Comparison of SSF and SHF and Identification of Bottlenecks. Biotechnol. Prog. 2003, 19, 1109-1117. 162. Cannella, D.; Jørgensen, H. Do New Cellulolytic Enzyme Preparations Affect the Industrial Strategies for High Solids Lignocellulosic Ethanol Production? Biotechnol. Bioeng. 2014, 111, 59-68. 163. Berlemont, R. Distribution and Diversity of Enzymes for Polysaccharide Degradation in Fungi. Sci. Rep. 2017, 7, 222. 164. Lenfant, N.; Hainaut, M.; Terrapon, N.; Drula, E.; Lombard, V.; Henrissat, B. A Bioinformatics Analysis of 3400 Lytic Polysaccharide Oxidases from Family AA9. Carbohydr. Res. 2017, 448, 166-174. 165. Wong, E.; Vaaje-Kolstad, G.; Ghosh, A.; Hurtado-Guerrero, R.; Konarev, P. V.; Ibrahim, A. F. M.; Svergun, D. I.; Eijsink, V. G. H.; Chatterjee, N. S.; van Aalten, D. M. F. The Vibrio cholerae Colonization Factor GbpA Possesses a Modular Structure That Governs Binding to Different Host Surfaces. PLoS Pathog. 2012, 8, e1002373. 166. Paspaliari, D. K.; Loose, J. S. M.; Larsen, M. H.; Vaaje-Kolstad, G. Listeria monocytogenes Has a Functional Chitinolytic System and an Active Lytic Polysaccharide Monooxygenase. FEBS J. 2015, 282, 921-936. 167. Agostoni, M.; Hangasky, J. A.; Marletta, M. A. Physiological and Molecular Understanding of Bacterial Polysaccharide Monooxygenases. Microbiol. Mol. Biol. Rev. 2017, 81, e00015-17.

57 ACS Paragon Plus Environment

ACS Catalysis 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 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

1609 1610 1611 1612

168. Reese, E. T.; Siu, R. G. H.; Levinson, H. S. The Biological Degradation of Soluble Cellulose Derivatives and Its Relationship to the Mechanism of Cellulose Hydrolysis. J. Bacteriol. 1950, 59, 485-497.

58 ACS Paragon Plus Environment

Page 58 of 59

Page 59 of 59 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 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

1613 1614 1615

ACS Catalysis

TOC Graphics

1616 1617 1618

59 ACS Paragon Plus Environment