Subscriber access provided by UNIV OF NEW ENGLAND ARMIDALE
Article
How a Lytic Polysaccharide Monooxygenase Binds Crystalline Chitin Bastien Bissaro, Ingvild Isaksen, Gustav Vaaje-Kolstad, Vincent G.H. Eijsink, and Åsmund K. Røhr Biochemistry, Just Accepted Manuscript • DOI: 10.1021/acs.biochem.8b00138 • Publication Date (Web): 02 Mar 2018 Downloaded from http://pubs.acs.org on March 3, 2018
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.
Biochemistry 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 29 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
Biochemistry
1 2 3 4 5
Bastien Bissaro‡, Ingvild Isaksen‡, Gustav Vaaje-Kolstad‡, Vincent G.H. Eijsink‡ and Åsmund K. Røhr‡*
6 7
KEYWORDS; LPMO, ENZYMES, EPR
8 9
ABSTRACT GRAPHIC
How a Lytic Polysaccharide Monooxygenase Binds Crystalline Chitin
CHITIN,
CELLULOSE,
COMPUTATIONAL
CHEMISTRY,
COPPER
10 11 12
ABSTRACT: Lytic polysaccharide monooxygenases (LPMOs) are major players in biomass
13
conversion, both in Nature and in the biorefining industry. How the mono-copper LPMO
14
active site is positioned relative to the crystalline substrate surface to catalyze powerful, but
15
potentially self-destructive, oxidative chemistry is one of the major questions in the field. We
16
have adopted a multi-disciplinary approach, combining biochemical, spectroscopic and
17
molecular modeling methods to study chitin binding by the well-studied LPMO from Serratia
18
marcescens SmAA10A (or CBP21). The orientation of the enzyme on a single chain substrate
19
was determined by analyzing enzyme cutting patterns. Building on this analysis, molecular
20
dynamics (MD) simulations were carried out to study interactions between the LPMO and
21
three different surface topologies of crystalline chitin. The resulting atomistic models showed
22
that most of enzyme-substrate interactions involve the polysaccharide chain that is to be
23
cleaved. The models also revealed a constrained active site geometry as well as a tunnel
24
connecting the bulk solvent to the copper site, through which only small molecules such as
25
H2O, O2 or H2O2 can diffuse. Furthermore, MD simulations, QM/MM calculations and EPR
26
spectroscopy demonstrate that rearrangement of Cu-coordinating water molecules is
27
necessary when binding substrate and also provide a rationale for the experimentally observed
28
C1-oxidative regiospecificity of SmAA10A. This study provides a first, experimentally
29
supported, atomistic view of the interactions between an LPMO and crystalline chitin. The
30
confinement of the catalytic center is likely of crucial importance for controlling the oxidative
31
chemistry carried out by LPMOs and will help guiding future mechanistic studies.
1 ACS Paragon Plus Environment
Biochemistry 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
32
Page 2 of 29
INTRODUCTION
33
Enzymes known as lytic polysaccharide monooxygenases (LPMOs) act on structural,
34
crystalline polysaccharides such as cellulose1–3 or chitin,4 but also various hemicelluloses5,6 or
35
starch.7 This makes LPMOs key players of biomass conversion in Nature, notably during
36
fungal action,8 and has rendered LPMOs crucial for the efficiency of the most recent
37
commercial enzymatic cocktails employed in biorefineries.9 LPMOs may also be involved in
38
microbial pathogenicity.10–12 LPMOs are mono-copper enzymes carrying out hydroxylation of
39
the C1 and/or C4 carbon of the scissile glycosidic bond, leading to bond cleavage via an
40
elimination reaction.13 Importantly, and in contrast to classical hydrolytic enzymes such as
41
cellulases, LPMOs can cleave polysaccharide chains packed in crystalline forms4,14 thus
42
disrupting the crystalline surface.15–17 Substrate hydroxylation is thought to be catalyzed by an
43
activated copper-oxygen species,18–22 and requires a precise assembling of the enzyme-
44
polysaccharide complex, allowing regioselective oxidative chemistry to occur and preventing
45
off-pathway oxidative processes that may damage the enzyme.22
46
Despite recent developments23–25 little is known about how LPMOs interact with their
47
crystalline substrate. An NMR-based approach led to the identification of residues involved in
48
the interaction between the chitin-active LPMO from Serratia marscecens (SmAA10A, also
49
known as CBP21) and crystalline chitin,23 confirming and extending knowledge acquired
50
from a previous mutagenesis study.26 Insights into the interactions between a fungal LPMO
51
and (soluble) cello-oligosaccharides have been obtained from both NMR24 and
52
crystallographic studies.25 The seminal crystallographic work by Frandsen et al.25 revealed
53
atomistic details of the interaction between the catalytic center of the LPMO and a single
54
sugar chain and also provided insights into the effect of substrate-binding on the
55
configuration of the copper site. However, interactions between an LPMO and a soluble
56
substrate likely differ from the interactions with a substrate embedded in a crystalline lattice.
57
Currently, only computational approaches allow simultaneous visualization of an enzyme and
58
a crystalline substrate at the atomic scale.
59
A single study has so far addressed this matter by modeling the interaction of a fungal
60
(AA9) LPMO with cellulose,27 indicating interacting residues and showing high flexibility of
61
two surface loops (not present in SmAA10A). Notably, in this study, the orientation of the
62
LPMO on the cellulose chain was not experimentally probed but set by analogy to previous
63
molecular dynamics studies performed on carbohydrate binding modules.28 The diversity of
64
loops and surface topologies displayed by LPMOs may reflect the diversity of substrate
65
structures (topologies and decorations) but, so far, correlations between surface architecture
66
and substrate specificity remain unknown.29,30 One interesting question is whether substrate
67
interactions primarily (or only) involve the polysaccharide chain that is to be cleaved or also
68
involve contacts with adjacent chains. From that point of view, chitin-LPMO systems are
2 ACS Paragon Plus Environment
Page 3 of 29 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
Biochemistry
69
interesting to study since (i) chitin can be found as α- or β-allomorphs, displaying antiparallel
70
and parallel chain arrangements, respectively, and (ii) some chitin-active LPMOs, such as
71
SmAA10A, are active on both forms but with preferences.31–34
72
In the present study, we aimed at answering fundamental questions pertaining to the
73
geometry of the LPMO-substrate complex, the structural determinants of substrate-binding,
74
and binding-induced effects on the enzyme active site and its access. Using a multi-
75
disciplinary approach combining computational, biophysical and biochemical methods, we
76
have produced the first, experimentally supported, fully atomistic model of an LPMO
77
interacting with a crystalline polysaccharide, here chitin. Enzyme activity assays were used to
78
direct model building and subsequent molecular dynamics simulations, and electron
79
paramagnetic resonance (EPR) spectroscopy was used for experimental confirmation of
80
computational results. All work was performed using a well-studied LPMO, SmAA10A, that
81
can interact with both crystalline α- or β-chitin, allowing thorough comparisons of
82
computational and experimental data. The present findings shed light on how an LPMO
83
orients itself and binds to a crystalline substrate in a mode that allows regioselective catalysis
84
of oxidative polysaccharide cleavage.
85 86 87
MATERIALS AND METHODS
88
Materials. Chemicals were purchased from Sigma-Aldrich. β-chitin extracted from squid
89
pen was purchased from France Chitin (Orange, France). Ascorbic acid (100 mM) stock
90
solutions were prepared in metal-free water (Trace SELECT®, Sigma-Aldrich), aliquoted,
91
stored at -20 °C, and thawed in the dark for 10 min just before use.
92 93
Production and purification of recombinant LPMOs. Recombinant LPMO from Serratia
94
marcescens (SmAA10A) and mutants thereof were expressed and purified according to
95
previously described protocols.26 All LPMOs used in this study, except the wild type used in
96
the EPR experiments (see EPR section), were prepared in sodium phosphate buffer (50 mM,
97
pH 6.0), copper-saturated with Cu(II)SO4 and desalted (PD MidiTrap G-25, GE Healthcare)
98
before use.11
99 100
SmAA10A activity test. Reactions were carried out in 2 mL Eppendorf tubes and the reaction
101
volume was 200 µL (for final time point analysis) or 500 µL (for time-course monitoring).
102
Typical reactions contained the LPMO (10 µM) and substrate (10 g.L-1 for β-chitin or 1 mM
103
for oligosaccharides) in Tris-HCl buffer (pH 8.0, 50 mM) and were pre-incubated during 20
104
min at 40 °C in a Thermomixer (1000 rpm). The reactions were initiated by adding ascorbic
3 ACS Paragon Plus Environment
Biochemistry 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 4 of 29
105
acid (to a final concentration of 1 mM). In control reactions, SmAA10A was replaced by
106
Cu(II)SO4 (10 µM). For time course monitoring, 55 µL samples were taken from the reaction
107
mixtures at regular intervals and soluble fractions were immediately separated from the
108
insoluble substrate by filtration using a 96-well filter plate (Millipore) operated with a
109
vacuum manifold. Immediately after, the filtrate was incubated at 98 °C during 15 min to
110
ensure enzyme inactivation. The heat inactivated samples were then frozen (-20 °C) prior to
111
further analysis.
112
For qualitative product analysis, samples were analyzed by MALDI-ToF MS, as previously
113
described.4 For quantitative analysis, oxidized chito-oligosaccharides were separated by high
114
performance anion exchange chromatography (HPAEC) and monitored by pulsed
115
amperometric detection (PAD) using a Dionex Bio-LC equipped with a CarboPac PA1
116
column as previously described for cello-oligosaccharides.35 All chromatograms were
117
recorded using Chromeleon 7.0 software. In all figures, DPnox refers to (GlcNAc)n-
118
1GlcNAc1A,
119
forms) of a chito-oligosaccharide composed of n glycosyl units. DPnox standards were
120
obtained by oxidation of chito-oligosaccharides with a DP ranging from 1 to 6 by a chito-
121
oligosaccharide oxidase as previously described.11
i.e. the C1-oxidized form (in equilibrium between lactone and aldonic acid
122 123
Electron Paramagnetic Resonance (EPR) Spectroscopy. All samples contained 200 µM
124
SmAA10A in 50 mM MES pH 6.0. Before use, the enzyme was saturated with Cu(II)SO4 and
125
desalted (PD MidiTrap G-25, GE Healthcare), as described above. The SmAA10A-(NAG)6
126
sample was prepared by dissolving 4 mg (NAG)6 (hexa-N-acetyl-chitohexaose, Megazyme)
127
in 200 µL SmAA10A solution. Hydrated chitin particles were obtained by repetitive washing
128
with MES buffer (50 mM, pH 6.0), after which excess buffer was removed using a 0.22 µm
129
centrifugal filter (Millipore), spinning at 10,000 g for 2 minutes. Samples with α-chitin (Sea
130
Garden, Norway, Avaldsnes) or β-chitin (France, Orange), both < 80 mesh-sized particles,
131
were prepared by carefully packing EPR tubes with 300 µL hydrated particles before adding
132
200 µL of a copper saturated 200 µM SmAA10A solution or 200 µL 50 mM MES pH 6.0
133
(control reaction). After 10 minutes, excess solution was removed, resulting in a sample
134
height of ~25 mm, and the samples were frozen in liquid nitrogen. EPR spectra were recorded
135
using a BRUKER EleXsys 560 SuperX instrument equipped with an ER 4122 SHQE SuperX
136
high sensitivity cavity and a cold finger. Spectra were recorded using 1 mW microwave
137
power and 10 G modulation amplitude at 77 K. In presence of substrate the EPR spectrum
138
shows a mixed population corresponding to substrate-bound and unbound fractions of the
139
LPMO. The spectrum corresponding to the unbound LPMO fraction as well as the substrate-
4 ACS Paragon Plus Environment
Page 5 of 29 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
Biochemistry
140
only background were subtracted from all substrate-containing EPR spectra. Spin
141
Hamiltonian parameters were fitted using the EasySpin package.36
142 143
Synchrotron radiation X-ray diffraction (SRXRD). Diffraction data was collected at the
144
Swiss-Norwegian beamline (BM01), ESRF, France using a PILATUS2M detector and a
145
wavelength of 0.7361 Å.37 Hydrated samples of the α- and β-chitin identical to what was used
146
in the EPR experiments were packed in ø 0.8 mm capillaries and diffraction data was
147
collected at room temperature, 60 s exposure, 1.5° s-1. Diffraction patterns were processed
148
using the SNBL Tool BOX software37 and Matlab.
149 150
Development of molecular models. The SmAA10A atomic coordinates were extracted from
151
the PDB entry 2BEM:C.26 The H++ server38 predicted only His74 to have a notably altered
152
pKa value, being positively charged at pH 7.0. Thus, the PDB file was updated accordingly
153
(His74 -> HIP74; i.e. protonated on delta and epsilon nitrogens). The Na+ ion modeled in the
154
copper-binding site in the original crystal structure was replaced by copper. All
155
crystallographic water molecules were retained in the model.
156
No force field parameters exist for the copper sites in LPMOs. Therefore, a minimal
157
model consisting of a truncated histidine brace (N-terminal His28 atom C and His114 atom
158
Cβ capped with H) with Cu(I) or Cu(II) was used to derive partial charges for both
159
oxidization states. During geometry optimization of these complexes, an angle constraint (Nδ-
160
Cu-Nε = 178°) was applied to retain the T-shaped Cu-3N geometry, which is a known, highly
161
conserved structural feature of LPMOs, regardless of their redox state.30 The electrostatic
162
potentials of the Cu(I) and Cu(II) complexes were computed using the program Gaussian 0939
163
with the hybrid functional B3LYP and the 6-311G(d,p) basis set for Cu and 6-31G(d,p) for
164
the remaining atoms. When computing the ESPs, the atomic radius of the Cu atom was set to
165
1.8 Å. Finally, the partial charges were fitted using the RESP method implemented in
166
AmberTools17.40 Bond and angle force constants describing the Cu-histidine brace
167
interaction were computed from the Cu(II) containing minimal model described above to
168
which two water ligands were added, resulting in a near trigonal bipyramidal starting model.
169
Adding the water molecules allowed unrestrained geometry optimization (B3LYP/6-
170
311G(2d,2p)) of the complex, which is a prerequisite for a correct frequency analysis and
171
which resulted in a conserved T-shaped Cu-3N model with an overall distorted square
172
pyramidal molecular geometry. Subsequently to frequency analysis, cartesian force constants
173
were extracted from the Hessian and projected on unit vectors representing bonds and angles,
174
to compute bond and angle force constants encompassing the Cu atom.41 The resulting force
175
field parameters can be found in Table S1.
5 ACS Paragon Plus Environment
Biochemistry 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 6 of 29
176
The most abundant forms of crystalline chitin found in Nature are α-chitin and β-
177
chitin, organized in layers of β-1,4 linked N-acetylglucosamine (NAG) chains arranged in an
178
antiparallel or parallel manner, respectively. For hydrated β-chitin, only the [100] lattice plane
179
exposes H1 atoms that are abstracted during LPMO catalysis (Fig. 2A). This is analogous to
180
what is observed for Iβ-cellulose, the structural organization of which resembles hydrated β-
181
chitin. Regarding α-chitin, the [100] lattice plane has previously been used to represent the
182
polysaccharide surface42 but the [110] lattice plane should also be considered since it exhibits
183
a smoother and more densely packed surface (Fig. 3B). In this study, using crystal structure
184
coordinates of α-chitin43 and hydrated β-chitin44 we built models for β-chitin (β[100]) made of
185
5 layers of 5 parallel (NAG)24 chains, and models for the two α-chitin types (α[100] and
186
α[110]) made of 5 layers of 6 antiparallel (NAG)24 chains.
187
In theory, the LPMO substrate binding patch could interact with the crystalline
188
substrate surface in any orientation around an axis orthogonal to the α- and β-chitin lattice
189
planes (Fig. 3A). The experimental data presented in this paper, demonstrating the orientation
190
of a chitin chain bound to SmAA10A (Fig. 1), substantially reduced the degrees of freedom to
191
be considered when placing the enzyme on the chitin surface. Knowing that the Cu atom must
192
be close to the hydrogen atom to be abstracted during catalysis, the enzyme was placed ~5 Å
193
above the chitin surface such that the Cu ion was hovering directly above the exposed H1 and
194
H4 atoms, but without pre-determined selectivity. In the SmAA10A-β[100] start model the
195
Cu-H1 and Cu-H4 distances were 5.7 and 5.3 Å, respectively. Corresponding distances of 5.5
196
and 5.7 Å were measured for both the SmAA10A-α[100] and α[110] complexes. As a control,
197
we also built start models where the enzyme was rotated 180° around the Cu atom along an
198
axis orthogonal to the chitin surface. All models were solvated in a rectangular box of TIP3P
199
water, with the edges at least 16 Å away from the solute, using chloride ions for
200
neutralization. The solvated models contained in total ~160 000 atoms. The protein ff14SB45
201
and carbohydrate GLYCAM_06-j46 force fields and our own active site parameters (see
202
above) were used to generate input files for AMBER16. Input files for SmAA10A-chitin
203
complexes were prepared using the force field parameters derived for Cu(I) to mimic an
204
active enzyme with weak Cu-water interactions. When setting up the SmAA10A in solution
205
model, the parameters derived for Cu(II) active sites, which increase the affinity for
206
coordinating water molecules, were applied. The difference in water affinity for the Cu(I) and
207
Cu(II) active sites originates from the dissimilar Cu partial charges (see Table S1).
208 209
Molecular dynamics (MD) simulations. The models were subjected to 2500 steps of energy
210
minimization with 5 kcal.mol-1.Å-2 positional restraints on the non-hydrogen atoms of the
211
enzyme and chitin molecules to relieve abnormal molecular contacts. Then, the systems were
6 ACS Paragon Plus Environment
Page 7 of 29 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
Biochemistry
212
heated linearly from 0 K to 300 K for 40 ps at constant volume, with restraints lowered to 1
213
kcal mol-1 Å-2, using the Langevin thermostat with a collision frequency of 1 ps-1. Density
214
equilibrations were run at 300 K for 0.5 ns at a constant pressure of 1 atm using the
215
Berendsen barostat with a pressure relaxation time of 1 ps. The final 50 ns equilibration step
216
was carried out in the NVT ensemble using the weak coupling algorithm and a time constant
217
of 10 ps to regulate the temperature. In this step, 2 kcal.mol-1.Å-2 positional restraints were
218
applied to the C1 atoms of the lowest layer of NAG chains of the chitin models (the “highest”
219
layer being the one interacting with the LPMO). The 300 ns production runs were performed
220
using the same conditions as in the final equilibration step, storing a snapshot every 20 ps,
221
yielding 15,000 snapshots for each trajectory. In all simulations, we used a time step of 2 fs,
222
periodic boundary conditions with a 12 Å cutoff for non-bonded interactions and PME
223
treatment of long range electrostatics, while hydrogen atoms were constrained by the SHAKE
224
algorithm. Simulations were carried out using the CUDA version of PEMEMD included in
225
AMBER16.47 Analysis of production trajectories was performed using the cpptraj module
226
included in AmberTools.48
227 228
Quantum mechanics/molecular mechanics (QM/MM) calculations. The QM/MM model of
229
SmAA10A in absence of chitin (model 1) was taken from a 50 ns MD trajectory (generated
230
using the same equilibration procedure as for the LPMO-chitin complexes), applying a filter
231
that selects frames having an active site geometry (all non-hydrogen atoms of H28, H114,
232
A112, F187, and E60) with RMSD < 1 Å relative to the crystal structure (2BEM:C).
233
The QM/MM model (model 2) of SmAA10A in complex with chitin was taken from the 300
234
ns SmAA10A-β[100] trajectory, applying the same filter as described above, with the addition
235
of a Cu-H1 distance window of 3.8 ± 0.2 Å (the average Cu-H1 distance).
236
QM/MM models 1 and 2 both included Cu(II) and the enzyme residues H28 (link atom C),
237
H114, A112, F187 and E60 (link atom Cα) in the QM region. Model 1 furthermore included
238
two Cu-coordinating water molecules (these two Cu-coordinating water molecules were
239
observed throughout the MD trajectory). Model 2 additionally included one water molecule in
240
a near equatorial position (2.9 Å away from Cu in the MD-snapshot starting structure) and
241
two NAG units in the QM-region. For NAG2, the link atoms were the O1 of the reducing end
242
and the C1 of the NAG unit preceding the non-reducing end.
243
Finally, a plausible18,49,50 copper-oxygen reactive species ([CuO]+ core, singlet state) model,
244
using the geometry optimized model 2 with oxygen replacing the Cu-coordinating water
245
molecule in the equatorial position, was generated.
246
The QM/MM interface in AMBER51 with ORCA352 providing the QM energy and
247
gradient was used to geometry optimize the models described above. The BP86 GGA
7 ACS Paragon Plus Environment
Biochemistry 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 8 of 29
248
functional,53,54 which has previously been shown to yield conform geometries for transition-
249
metal complexes,55 was selected, and dispersion was included through the Grimme’s DFT-D3
250
approach with Becke-Johnson dampening.56,57 The Def2-TZVPP basis set was used for Cu
251
and Def2-SVP for all other atoms,58,59 and the RI-approximation was applied to speed up the
252
calculations.
253
The following 3-stage geometry optimization scheme was applied to the QM/MM models; i)
254
100 steps of unrestrained QM/MM optimization with the TNCG (Truncated Newton linear
255
Conjugate Gradient method with LBFGS preconditioning) algorithm, ii) 10,000 MM steps
256
with the conjugate gradient method (no QM) where all QM atoms used in step i) were frozen
257
using 1000 kcal.mol-1.Å-2 positional restraints, iii) TNCG QM/MM optimization until
258
convergence (drms = 10-4 kcal.mol-1.Å-1). Atoms more than 6 Å away from the QM region
259
were frozen using 1000 kcal.mol-1.Å-2 positional restraints during this step.
260 261
Computation of EPR spin Hamiltonian parameters. EPR parameters were computed from
262
the QM part of the converged QM/MM geometries using ORCA352 with the hybrid functional
263
PBE060 and the chain-of-spheres approximation. The (NAG)2 part of the model was omitted
264
to reduce system size. This was considered safe considering that for model 2 only 2.5 % of
265
the spin density was located at (NAG)2. Scalar relativistic effects were included with ZORA.
266
The basis sets Def2-TZVPP and Def2-TZVP were decontracted and used for Cu and all other
267
atoms, respectively. The integration grid (Grid6) was increased for Cu (SpecialGridIntAcc 7).
268
The following contributions were included when calculating the hyperfine couplings: 1) the
269
isotropic Fermi contact term, 2) the magnetic spin dipole interaction between the nucleus spin
270
and electron spin and 3) the second order contribution form spin-orbit coupling.
271 272
RESULTS AND DISCUSSION
273
Overview. The catalytic core of the LPMO, a Cu-N3 complex hosting a reactive oxygen
274
species, must be carefully maneuvered close to the inert C-H bond to be functionalized
275
leading in fine to polysaccharide chain cleavage. LPMOs acting on the β(1,4) glycosidic
276
bonds of chitin or cellulose often display strict regio-selectivity, meaning that they
277
exclusively hydroxylate the C1 or C4 carbon, which are separated by 2.5 Å in the
278
polysaccharide chain. Notably, there is no known inherent property of the active site itself that
279
could determine this regioselectivity. LPMO action on chitin leads to oxidation at C1. The
280
LPMO active site is part of a relatively flat, solvent-exposed surface displaying conserved
281
amino acid side chains that have been shown by experiments23,26 to be crucial for enzyme
282
activity. There are no available crystallographic data or models describing interactions
283
between an LPMO and chitin or (soluble) chito-oligosaccharides that would allow
8 ACS Paragon Plus Environment
Page 9 of 29 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
Biochemistry
284
rationalization of the mode of action of these enzymes. Therefore, to fill this gap and provide
285
sound fundaments for future mechanistic studies, we have built atomistic models of enzyme-
286
substrate complexes, the validity of which is supported by experimental data.
287 288
Determining the relative orientation LPMO-substrate. In 2005, a site-directed mutagenesis
289
study conducted on SmAA10A (then called CBP21) revealed that several surface exposed
290
amino acid side chains (Tyr54, Glu55, Glu60, His114, Asp182 and Asn185) are important for
291
binding to chitin.26 Complementary to this study, and subsequent to the discovery of LPMO
292
activity in 2010,4 HSQC NMR experiments demonstrated that β-chitin shields the backbone
293
amides of specific residues of SmAA10A strongly (Tyr54, Glu55, Gln57, Ser58 and Thr111)
294
or moderately (Gln53, Leu110, Ala112, His114 and Thr116) from deuterium exchange with
295
bulk solvent (Fig. S1).23 Most of these residues are highly conserved within the chitin-active
296
AA10 subfamily. The strongly affected residues form an elongated patch on the SmAA10A
297
surface (Fig. S1), stretching from Tyr54, through the Cu-site, to Asp182 (Fig. 1). Knowing
298
that this patch and the Cu-site must face the surface of the crystalline substrate limits the
299
number of possible enzyme orientations. However, the enzyme can still be rotated 360º on the
300
substrate surface, leaving a substantial number of conformations to be explored.
301
We hypothesized that a chitin oligomer, (N-acetyl-glucosamine)6, or (NAG)6, could
302
cover this area with conserved residues and possibly be cleaved, although SmAA10A has only
303
been reported to be active on insoluble substrates.4 Using high enzyme loadings we indeed
304
detected SmAA10A activity on (NAG)6 (Figs. 1 and S2), but not on shorter oligomers (Fig.
305
S3). Importantly, the dominating oxidized product had a degree of polymerization of 4
306
(DP4ox), clearly indicating that there is one preferred binding orientation that, if we assume
307
that the substrate indeed aligns with the stretch of conserved residues, places the reducing end
308
of the hexamer near Asp182 and the non-reducing end near Tyr54 (Fig. 1 & S4A). In the case
309
of the opposite orientation, DP2ox would have been dominating (Fig. S4B), which is not the
310
case (Fig. S2A). The determination of this relative orientation allows introduction of a
311
putative subsite numbering, according to the CAZyme nomenclature,61 with subsites spanning
312
from -4 to +2, involving a stretch of exposed residues running from Tyr54 to Asp182,
313
respectively (Fig. 1). Figs. 1 and S2 also show production of DP5ox and, to a lesser extent,
314
DP3ox from NAG6, meaning that other productive binding modes are possible (Fig. S4E-H).
315
The production of considerable amount of DP5ox supports the notion that the substrate is
316
oriented with its non-reducing end pointing towards Tyr54 (Fig. S4E).
9 ACS Paragon Plus Environment
Biochemistry 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 29
317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337
Figure. 1. Probing the binding orientation of (NAG)6 in SmAA10A and the role of Tyr54. Panel (A) shows HPAEC-PAD chromatograms of C1-oxidized chito-oligosaccharides released from β-chitin (10 g.L-1) or from (NAG)6 (1 mM) by SmAA10A-WT or its Y54A mutant (10 µM) after 60 min of reaction. Reactions were initiated by adding AscA (1 mM) and carried out in Tris-HCl buffer (50 mM, pH 8.0), with incubation at 40 °C in a thermomixer (1000 rpm). In control reactions SmAA10A was replaced by Cu(II)SO4 (10 µΜ). Panel (B) shows time-courses for the release of DP4ox in reactions described in panel A. Note: the enzyme is no longer active at the first time point, indicating fast inactivation, a selfdestructive reaction that was notably shown to occur in absence of substrate.22,62 Such timecourse profile with (NAG)6 is thus probably diagnostic of weak binding to this substrate. The error bars show ± s.d. (n = 3, independent experiments). Panel (C) shows a model (extracted from MD simulations performed on β-chitin; see Fig. 3) of SmAA10A interacting with (NAG)6 (see Fig. S1 for a 90°-rotated top view). The histidine brace, composed of His28 and His114, and bound Cu are shown in orange. The side chains of Tyr54 and Asp182, in subsites -4 and +2, respectively, are also shown (grey sticks). The C1 of the NAG unit located in subsite -1 is the carbon subject to hydroxylation during catalysis by SmAA10A. Abbreviations: DP, degree of polymerization; NAG, N-acetylglucosamine; NR, non-reducing end; R, reducing end.
338
To confirm the positioning of the oligosaccharide in the active site we tested the
339
effect of the Y54A mutation, in the proposed subsite -4, on the activity towards (NAG)6.
340
Expectedly, this mutation decreased the activity of SmAA10A on (NAG)6 (Fig. 1B).
341
Furthermore, the production of DP5ox was much more affected than that of DP3ox (Fig. 1A),
342
which is consistent with the notion that Tyr54 in subsite -4 is involved in the binding mode
10 ACS Paragon Plus Environment
Page 11 of 29 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
Biochemistry
343
leading to formation of DP5ox (Fig. S4E). Interestingly, the effect of the Y54A mutation on
344
the activity towards β-chitin was minimal (Fig. 1). The crystalline structure of β-chitin,
345
possibly provides additional anchoring points for the LPMO (see MD simulations below) and
346
this may dampen the effect of the Y54A mutation.
347 348
Molecular dynamics simulations of SmAA10A-chitin interactions. The experimentally
349
determined orientation of the substrate relative to the LPMO was used as input to generate
350
large scale MD start models of SmAA10A-Cu(I) interacting with α- and β-chitin. Figure 2,
351
based on analysis of transmission electron microscopy images63,64 and crystal structures43,44 of
352
α- and β-chitin, shows microfibril cross-sections and lattice planes. Any putative interaction
353
of SmAA10A with the solvent exposed surfaces [120] and [12ത0] of the microfibril would not
354
allow the enzyme active site to get close to the glycosidic bond for catalysis to occur. For the
355
[010] surfaces, the glycosidic bonds are not exposed to the solvent at all. In the case of β-
356
chitin, alike Iβ-cellulose, the surface with exposed glycosidic bonds is defined by the lattice
357
plane [100], called β[100]. In the case of α-chitin the situation is less straightforward since
358
two lattice planes could be considered, namely α[100] and α[110] (Fig. 2). The MD results
359
presented here (see below) show that SmAA10A can bind in a productive manner to all these
360
three chitin surface topologies. It is interesting to note that the mature β-chitin microfibril
361
from the diatom Thalassiosira weissflogi displays a SmAA10A accessible surface (Fig. 2A)
362
while the β-chitin microfibril from the tubeworm Lamellibrachia satsuma would need peeling
363
before a [100] surface appears (Fig. 2B). The mature α-chitin microfibril from the alga
364
Phaeocystis globose also needs trimming before a large [100] or [110] surface emerges (Fig.
365
2C). Therefore, LPMO efficiency likely depends on the source of the polysaccharide, as has
366
indeed been observed.34
367
11 ACS Paragon Plus Environment
Biochemistry 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
368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396
Page 12 of 29
Figure. 2. Possible binding sites for SmAA10A on crystalline chitin fibers. The figures show scaled cross-sections of a β-chitin microfibril from T. weissflogi (A),64 a β-chitin microfibril from L. satsuma (B)64 and an α-chitin microfibril from P. globose (C)63 derived from electron micrographs and crystal structures. Each microfibril is composed of a multitude of individual chitin chains (shown as black jagged lines with an endon view, i.e. oriented perpendicular to the cross-section plane) arranged in a parallel (for A and B) or anti-parallel manner (for C). The green lines indicate the lattice plane surfaces on the crystalline chitin fibers that allow productive LPMO binding (relatively flat surfaces where H1-atoms are exposed towards the enzyme active site). The surfaces indicated by red lines are very rough and display deep grooves that are incompatible with productive LPMO binding because steric clashes between the chitin chains and the enzyme will not allow the necessary proximity between the copper center and the H1-atoms to be abstracted.
12 ACS Paragon Plus Environment
Page 13 of 29 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
Biochemistry
397
A global view of SmAA10A interacting with β[100] is shown in Fig. 3A and a movie
398
illustrating the enzyme docking on the chitin surface is available as supporting information
399
(Movie S1). End-on views (i.e. looking along the polysaccharide chains) of the β[100],
400
α[100], and α[110] crystalline chitin models reveal very different surface topologies (Fig.
401
3B). β[100] chitin is nearly planar, whereas α[100] displays deep grooves between the
402
polysaccharide chains. In the α[110] model, the polysaccharide chains overlay slightly, and
403
the grooves are shallow compared to α[100]. The MD simulations of the three SmAA10A-
404
chitin models showed that the alignments of the copper ion and the histidine brace with the
405
to-be-cleaved glycosidic bond, are similar (see Fig. S5A), with, notably, identical average
406
Cu-H1 distances (Fig. 3E). Also, the SmAA10A active site geometry is not altered
407
significantly upon substrate binding (Fig. 3C and Fig. S5B). The Cu-H1 distances are 3.8 Å
408
on average and are always more than 0.8 Å shorter than the corresponding Cu-H4 distances
409
(Fig. 3E). The closer proximity to C1 is consistent with the established C1 oxidative
410
regiospecificity of SmAA10A4 (see discussion below and Fig. 5C).
411
Both the LPMO and the chitin displayed very stable packing during all simulations
412
with average RMSDs of ~ 0.5 Å for chitin-C1 atoms and ~ 0.8 Å for SmAA10A-Cα atoms
413
(Fig. 3E). Comparison of SmAA10A-Cα atom positions relative to chitin-C1 atom positions
414
showed that the SmAA10A-chitin complexes are also relatively stable (Fig. 3E) although
415
much wider distributions are observed (see Fig. S6 for time-resolved data). SmAA10A-β[100]
416
and SmAA10A-α[110] showed considerably lower RMSD values (1.16 and 0.96 Å) compared
417
to SmAA10A-α[100] (1.49 Å) (Fig. 3E), indicating that SmAA10A can move more freely on
418
the α[100] surface, probably due to the difference in surface topology (Fig. 3B).
419
To visualize the origin of these LPMO-chitin complex RMSD distributions, a
420
principal component analysis (PCA) was applied to identify the most significant concerted
421
motions, referred to as normal modes #1 and #2, with mode #1 having the highest eigenvalue.
422
The PCA shows that the largest motion mode for SmAA10A-β[100] and SmAA10A-α[110] is
423
a rotating motion (mode #1, Fig. S7) around an axis orthogonal to the chitin surface and
424
centered close to the histidine brace. The second most significant modes show mixed
425
translational/rocking and rotational/rocking modes for SmAA10A-β[100] and SmAA10A-
426
α[110], respectively (mode #2, Fig. S7). The α[100] model, with its much more rugged
427
surface (Fig. 3B) shows an opposite arrangement of mode #1 and #2, with more
428
rocking/translational movement in mode #1 and a rotating motion for mode #2 (Fig. S7).
429
The intermolecular interactions between SmAA10A and the chitin models were
430
dissected by close contact and hydrogen bond analysis of the MD-trajectories, enabling
431
detailed examination at the residue level. Fig. 3D shows heat-maps for close contacts of
432
SmAA10A with crystalline chitin for each of the three models. The close contact analysis
433
reveals enzyme residues with atoms within a distance of 6 Å from any NAG atom in the
13 ACS Paragon Plus Environment
Biochemistry 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 14 of 29
434
trajectory (Fig. 3F). A general observation is that the heat-maps correspond well with the
435
patch of surface-exposed, conserved amino acid residues thought to be involved in substrate
436
binding by SmAA10A (see Fig. S1). There are, however, differences between the three chitin
437
forms, likely because of their distinct surface topologies. The interactions between SmAA10A
438
and β[100] or α[110], which have relatively flat surfaces, involve a larger fraction of the
439
surface compared to α[100] (Fig. 3D). Fig. 3F shows that 71%, 70% and 66% of all measured
440
close contacts occur between the enzyme and the to-be-cleaved polysaccharide chain
441
(subsites -5 to +3) for β[100], α[100] and α[110], respectively. This is consistent with
442
biochemical data presented above but also shows that the contribution of adjacent chains is
443
not negligible in a crystalline context. The close contact analysis also reveals that the
444
aromatic ring of Tyr54 is close to the NAG unit in subsite -4 during the simulation,
445
supporting the idea that solvent-exposed aromatics in LPMOs, of which there often are only
446
few, are key factors for docking onto the polysaccharide. The hydrogen bond analysis
447
allowed us to see that residues Glu55, Thr111, His114, Gln57 and Asp182 form well defined
448
hydrogen bonds to chitin (Fig. S8B). For an in-detail view of amino acids interacting with
449
each subsite see Fig. S9.
450
Additionally, the models presented here allow to investigate the structural differences
451
that may be involved in the preference for one chitin allomorph over the other, a matter on
452
which sound biochemical data are scarce owing to substrate heterogeneity and lack of
453
systematic comparison. Experiments have shown that SmAA10A displays a binding
454
preference for β-chitin over α-chitin.23,65 Here, the close contact (Fig. 3) and, especially, the
455
hydrogen bond (Fig. S8) analyses pointed notably at Arg113 as prominent in subsite c when
456
interacting with β[100], but not interacting with α[100]. Arg113, not considered or detected
457
in previous studies, is a non-conserved residue found in the middle of the motif “TAXH”,
458
totally conserved in the chitin-active AA10 sub-family, where H is His114 of the histidine
459
brace and X = A, M, P, Q or R. Our models suggest that mapping out phenotypic sub-clades
460
(e.g. allomorphs preferences or degrees of chitin acetylation) among chitin-active AA10s may
461
require to consider non-conserved surface-exposed residues.
462
The previous mutagenesis26 and NMR23 studies aiming at mapping interactions
463
between SmAA10A and β-chitin relied on different principles and gave slightly different
464
results (Fig. S1), which now can be explained. For instance, Glu60, Asp182, and Asn185
465
were not detected in the NMR approach but found to be important in the mutagenesis study.
466
The MD simulations reveal that Glu60 and Asp182 to a large extent, and Asn185 to a lesser
467
extent, interact with β[100] through bridging water molecules (Fig. S8C). This could explain
468
why these residues are important for binding but were not detected by the NMR approach,
14 ACS Paragon Plus Environment
Page 15 of 29 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
Biochemistry
469
which was based on detecting protection of the backbone amides from exchange with solvent
470
by chitin binding.
471
Overall, there is a striking compliance between the previous mutagenesis and NMR
472
data and the MD-simulations presented here (Fig. S1). The present experimentally supported
473
models complete the picture by revealing details of the interaction of each amino acid with
474
crystalline chitin. As a control experiment, we also investigated the ability of the enzyme to
475
bind to the polysaccharide surface in the opposite direction by running control simulations for
476
all the model substrates where SmAA10A was rotated 180° around an axis orthogonal to the
477
chitin surface and centered on the copper atom. With α[100] or α[110]-chitins, the enzyme
478
drifted away from the substrate (results not shown). With β[100]-chitin a complex was
479
formed, which, however, did not look plausible with a O-H1 distance of 4.2 Å and a Cu-O-H1
480
angle of 86° (see Fig. 5C for comparison) and few intermolecular interactions involving
481
conserved residues were observed (not shown). Taken together, the simulations and
482
experimental data strongly indicate a substrate orientation where the non-reducing end of the
483
scissile polysaccharide chain points towards Tyr54. Surprisingly, this conclusion contrasts
484
with the crystal structure of a fungal, cellulose-active, C4-specific AA9 LPMO in complex
485
with cello-oligomers, which shows an opposite substrate/LPMO relative orientation (Fig.
486
S10).25 Despite this striking difference, in both cases, the N-terminal histidine is located in the
487
subsite hosting the sugar unit to be modified, i.e. in the +1 subsite for C4-oxidation (e.g.
488
LsAA9A) or in the -1 subsite for C1-oxidation (e.g. SmAA10A). Also, in both systems, a Tyr
489
residue is located in the negative subsites interacting with the non-reducing end of the
490
substrate, three subsites away from the subsite where catalysis occurs (Fig. S10). Altogether,
491
this suggests that the evolutionary shift between bacterial strict C1-oxidizers (belonging to
492
AA10s) and fungal strict C4-oxidizers (belonging to AA9s) may not originate from a simple
493
“small” translation of the active site over the glycosidic bond to be cleaved. Instead, an
494
appropriate and significantly different network of interactions must have been evolved so that
495
the reactive intermediate is properly poised over the target C-H bond for catalysis to occur.
496
On the contrary, it is more likely that the C1 vs C4 oxidative regioselectivity divergence
497
occurring within common phylogenetic groups (e.g. within fungal AA9s or bacterial AA10s)
498
is the result of more subtle interactions re-arrangements, as suggested by recent mutagenesis
499
studies.66,67 This inversed substrate-LPMO orientation is intriguing and its impact, on
500
substrate specificity and/or oxidative regioselectivity, deserves further investigations.
501 502 503
15 ACS Paragon Plus Environment
Biochemistry 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
504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526
Page 16 of 29
Figure 3. Interactions of SmAA10A with chitin probed by MD simulations. (A) Global view of SmAA10A interacting with β-chitin (SmAA10A-β[100]); explicit solvent excluded). (B) End-on view, looking along the polysaccharide chains, of the crystalline chitin models β[100], α[100], and α[110]. The black solid lines indicate the lattice planes corresponding to the LPMO binding surfaces. (C) Averaged structures of SmAA10A in solution (brown) and bound to β-chitin (green) from 300 ns trajectories show that the copper ion coordinating amino acid side chains retain their conformation upon substrate binding (see Fig. S5 for more details). (D) Heat-mapped contact surfaces of SmAA10A when interacting with crystalline chitin. The color gradient, from blue to red, indicates the frequency of contacts from low to high, respectively. For the sake of clarity, only the NAG units responsible for most of the contacts are represented as filled rings, identified by numbers (-5 to +3) or letters (a, b, …) when belonging to the main chain or adjacent chains, respectively. (E) The left panels show the distribution of RMSD values for chitin C1 atoms (green), enzyme Cα atoms (yellow) and chitin C1 atoms versus enzyme Cα atoms (purple), whereas the right panels show the Cu-H1 (red) and Cu-H4 (blue) distances, as derived from 300 ns simulations, fitted with the kernel probability density function (the corresponding time-resolved data are shown in Fig. S7). (F) Number of close contacts between NAG units of the polysaccharide, numbered according to the binding sites indicated in panel D, and residues of SmAA10A. Close contacts with less than 20 contacts were not included in the analysis. Contacts involving sugars bound to subsites -5 to +3 dominate and their fraction of the total contacts is indicated for each chitin model (66 - 71 %). Blue bars indicate conserved residues and green bars non-conserved residues (see Fig S1 for conservation scores). See Fig. S9 for visualization of the interactions
16 ACS Paragon Plus Environment
Page 17 of 29 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
Biochemistry
527 528 529
and Fig. S8 for analysis of hydrogen bond and water bridges. Abbreviations: NAG, N-acetylglucosamine.
530 531
MD simulations reveal a gated, main access to the active site. An important aspect of LPMO
532
catalysis pertains to the accessibility of the copper site when the enzyme is bound to the
533
polysaccharide. The idea that the copper site would be confined and shielded from bulk
534
solvent, is appealing since it would allow good control of the ongoing redox chemistry, but
535
also raises questions for example because the copper site would be shielded from potential
536
reducing agents that are needed to deliver electrons. Due to the latter conundrum, the
537
existence of an intramolecular electron transfer pathway has been proposed.20,21 The existence
538
of such an electron transfer pathway has never been demonstrated and is in fact not necessary
539
when considering H2O2 as a co-substrate of LPMO catalysis.22 In a H2O2-based mechanism,
540
electron delivery to the enzyme-substrate complex is not needed, provided that the enzyme
541
was reduced prior to binding. A recent kinetic study of SmAA10A supports the involvement
542
of H2O262 and it has also been shown that reduction of the copper promotes substrate
543
binding.68
544
Our models reveal the existence of a solvent tunnel formed between SmAA10A and
545
the chitin surface upon substrate binding. The ~12 Å long tunnel, predicted using Caver,69
546
connects the bulk solvent to the copper site and consists of a narrow monolayer of water
547
molecules (Fig. 4A). The Glu60 residue, associated with water bridged substrate interactions
548
(see above), displays three different rotamers (R1, R2 and R3) (Fig. 4B), influencing access
549
to the active site (Fig. 4C). Other residues lining the tunnel, e.g. Asn185, display limited
550
fluctuations compared to Glu60. For each configuration, a priority parameter (0 ≤ Pt ≤ 1) is
551
calculated on the basis of how functional a tunnel is predicted to be with respect to molecule
552
transport and how often the tunnel is observed in the ensemble. A high Pt value indicate a
553
functional tunnel observed frequently in the ensemble. In the β-chitin model, the R1 state
554
corresponds mainly to a “closed” tunnel (with a low Pt of 0.01) (Fig. 4D). In contrast, in the
555
less populated states R2 (Pt = 0.49) and R3 (Pt = 0.59) the Glu60 side chain is rotated away
556
from the crystalline surface, increasing access to the active site. The different surface
557
topologies of α-chitin, α[100] and α[110], seem to reduce the importance of Glu60 as an
558
gatekeeper (see Table S2 for details). Importantly, for all three chitin models the tunnel
559
maximum bottleneck radius is ~1.6 Å, meaning that the tunnel is too narrow to allow large
560
molecules such as ascorbate or diphenolic compounds (which are known LPMO reductants)
561
to reach the active site. Our data are therefore in favor of a scenario where copper reduction
562
occurs prior to polysaccharide binding, as supported by a recent study showing better
563
substrate binding by the LPMO-Cu(I) versus LPMO-Cu(II).68 However, the tunnel would
17 ACS Paragon Plus Environment
Biochemistry 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 18 of 29
564
allow molecules such as O2, O2•-, H2O2 or H2O to diffuse in and out of the active site. Given
565
the open and closed conformations observed for Glu60 we speculate that this residue is
566
involved in regulation of the access of reagents (e.g. H2O2) to the copper center and may thus
567
be crucial for LPMO activity but also for the enzyme’s operational stability. How such a
568
gatekeeping function would be controlled remains to be investigated. Of note, residues that
569
could be functionally analogous to Glu60 (mostly Glu or Gln) seems to appear in all
570
LPMOs29 and mutation of Glu60 in SmAA10A drastically reduced its boosting effect on
571
chitinase action26.
572
573 574 575 576 577 578 579 580 581 582 583 584 585 586 587
Figure 4. Accessibility of the active site in the SmAA10A-β[100] complex. “(A) An active site access tunnel (shown as beige surface), formed upon binding of SmAA10A to the polysaccharide, connects the active site copper (shown as orange sphere) with the bulk solvent. (B) MD-simulations show that the side chain of Glu60 displays three main rotamer populations indicated by R1, R2 and R3, which shape the morphology of the tunnel at proximity of the active site (see panel D). Asn185, which is also involved in the definition of the tunnel shape, displays only restricted fluctuations, not influencing the tunnel shape. (C) Snapshots representative of the different rotamer populations. (D) Approximately 500 snapshots were analyzed for each population of R1 (left panel), R2 (middle panel) and R3 (right panel) using Caver.69 The green lines indicate the resulting tunnels identified for the different populations, and the blocking effect of Glu60 is apparent for the R1 population. Parameters describing the tunnels are given in Table S2.
588 589
Insights into active-site properties upon substrate binding by QM/MM and EPR.
590
Interestingly, the MD simulations did not reveal substantial structural alterations of the active
591
site amino acids due to substrate-binding (Fig. 3C and S5B), although EPR experiments (see
592
below) clearly show that the SmAA10A copper environment is altered upon binding (Fig. 5).
18 ACS Paragon Plus Environment
Page 19 of 29 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
Biochemistry
593
Previously, it has been speculated that rearrangement of Cu-coordinating waters would cause
594
the change in the EPR spectrum when AA9s bind substrate.70 In a subsequent study, a similar
595
substrate binding-induced effect on the EPR signal was observed, but only in presence of
596
chloride, leading to the suggestion that binding of co-substrate plays a role in modulating the
597
EPR spectrum (see below).25
598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624
Figure 5. X-band EPR spectra of SmAA10A recorded in the absence and presence of substrate. The figures show EPR spectra (black line) of SmAA10A-Cu(II) (200 µM) in the absence of substrate (A) and in the presence of (NAG)6 (a soluble substrate) (B) or two forms of crystalline chitin, β- or αchitin (solid substrates) (C). Simulated spectra are also shown (sim, grey line); in panel C, only the simulation of the EPR spectrum recorded with β-chitin is displayed. Panel (D) shows the superhyperfine structure of the gx,yregion of the spectrum recorded for SmAA10A-Cu(II) in the presence of β-chitin. The simulation of this spectrum is based on the g-values and AzCu values determined from the experimental spectrum and Ax,yCu and AN values computed by DFT (see main text). All samples were prepared in MES buffer (50 mM, pH 6.0). All spectra were recorded at 77 K, 1 mW microwave power and 10 Gauss modulation amplitude. Simulated and computed spin Hamiltonian parameters are listed in Table 1.
625 626 627 628 629 630 631 632 633 634 635 636
To connect the computational results with spectroscopic data, two models of
637
SmAA10A-Cu(II), namely 1 (in solution, Fig. 6A) and 2 (with crystalline β-chitin, Fig. 6B),
19 ACS Paragon Plus Environment
Biochemistry 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 20 of 29
638
were subjected to QM/MM calculations. As should be expected for an accurate Cu(II)-active
639
site model, the geometry optimized model 1 displays a water coordination mode and active
640
site geometry closely resembling the low X-ray dose structure of the chitin-active AA10A
641
from Enterococcus faecalis (EfAA10A, PDB ID 4ALC) (Fig. S11), which is homologous to
642
SmAA10A (49 % sequence identity).
643
QM/MM calculations show that model 1 displays a distorted intermediate of the
644
trigonal bipyramid and square pyramid geometries (Fig. 6A), in agreement with the
645
experimentally observed rhombic EPR spectrum (d(x2-y2) ground state) of SmAA10A (Fig.
646
5A, Table 1). Superposition of model 1 and model 2 shows that one of the Cu-coordinating
647
water molecules of model 1 would clash with the bound chitin (Fig. S12). Consequently,
648
simultaneous presence of two Cu(II)-coordinating water molecules and substrate is not
649
feasible. Thus, rearrangement of at least one of the water molecules is expected when
650
SmAA10A binds chitin. On that note, Frandsen et al. have shown that upon formation of a
651
complex between a fungal LPMO and a cello-oligomer the copper-coordinating water in the
652
axial position is displaced by the C6 hydroxymethyl group of the +1 glucose and the
653
equatorial water molecule is replaced by a chloride ion.25 Here, the QM/MM geometry
654
optimized model 2 show that the single water molecule binding to Cu(II) lies in the equatorial
655
position, resulting in a distorted square planar geometry (Fig. 6B). This change in copper site
656
geometry agrees with spectroscopic experiments, since the SmAA10A-substrate EPR
657
envelopes indicate a shift towards axial symmetry compared to the free enzyme (Fig. 5B&C,
658
Table 1).
659
The spin Hamiltonian parameters for model 1 and 2 were computed and are
660
compared to the experimentally obtained values in Table 1. Mulliken charges and spin
661
populations are listed in Table S3. The experimental spin Hamiltonian parameters were
662
acquired by simulation of the spectra (Fig. 5). The experimentally determined gx, gy, AxCu and
663
AyCu values should only be considered as estimates while gz and AzCu could be accurately
664
determined. Though the computed gz and AzCu values deviate from the experimentally
665
observed values, a well-known situation in studies on Cu-complexes by DFT71, the trends of
666
effects induced by binding to crystalline substrate were reproduced: gz values decrease and
667
|AzCu| values increase. Notably, binding of (NAG)6 had slightly different spectral features
668
compared to binding of β-chitin, where binding of (NAG)6 resulted in a weaker shift of AzCu
669
(from 346 to 560, compared to 610 MHz for β-chitin). This is not surprising, considering the
670
higher flexibility of (NAG)6 compared to crystalline β-chitin, which will lead to a larger
671
distribution of binding geometries. In the presence of (NAG)6, a weak superhyperfine
672
splitting pattern appears in the gx,y region of the spectrum (Fig. 5B). This phenomenon is
673
considerably more apparent in the presence of α-chitin or β-chitin, which seem to alter the
674
SmAA10A active site in a similar manner (Fig. 5C), despite the obvious differences in the
20 ACS Paragon Plus Environment
Page 21 of 29 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
Biochemistry
675
molecular organization and crystalline properties of the chitin materials (Fig. S13). In the
676
crystallographic study by Frandsen et al., the changes in EPR signature upon substrate
677
binding (and in presence of chloride) were proposed to originate from equatorial coordination
678
of the copper by a chloride ion and from changes in the hydrogen bonding pattern between
679
the N-terminal His and a non-Cu-coordinating water molecule.25 In contrast to this previous
680
study, where chloride spin was included when analyzing the EPR superhyperfine pattern and
681
where ligand nitrogens were described by isotropic hyperfine coupling constants,25 we here
682
applied the anisotropic values derived from DFT, without chloride being present. The use of
683
AxCu, AyCu, AN, ANδ and ANε values computed from the QM/MM model 2 enabled us to
684
reproduce the superhyperfine splitting features observed in the experimental SmAA10A-β-
685
chitin EPR spectrum with high accuracy (Fig. 5C and D). This shows that QM/MM model 2
686
(Fig. 6B) is a plausible example of the physical SmAA10A-β-chitin complex with one water
687
coordinating in equatorial position to the Cu(II) ion.
688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707
Figure 6. QM/MM geometry optimized active sites. Panel (A) shows the QM region of SmAA10A in absence of chitin with two water molecules coordinating to Cu(II) in a distorted intermediate of the trigonal bipyramid and square pyramid geometries. When bound to chitin, only one water molecule is sterically allowed to interact with Cu(II) and the copper site adopts a distorted square planar geometry (B). For both (A) and (B) the delocalized spin densities (yellow surfaces) indicate the effect of the interaction between Cu(II) and the ligands. Details of the active site geometries are shown on the right side in panel (A) and (B). The active site of SmAA10A modeled by QM/MM with an [CuO]+ core shows how the reactive oxygen species is directed to the H1 atom that is abstracted during catalysis, with a distance O-H1 of 2.08 Å and a Cu-O-H1 angle of 146° (C). The link atoms for NAG2 are indicated by “L” and shown as black stick and ball (see experimental section).
708
21 ACS Paragon Plus Environment
Biochemistry 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
709
Finally,
we
12,18,21,22,49,50,72
geometry
optimized
a
Page 22 of 29
QM/MM
model
representing
a
+
710
plausible
711
site (Fig. 6C). In the light of the recent finding showing that H2O2 is a co-substrate for
712
LPMOs,22 and by analogy to calculations performed on cellulose-active AA9s,18,49,50 a Cu-oxo
713
or Cu-oxyl is suggested to be the species responsible for H-atom abstraction in the reaction
714
leading to hydroxylation of C1 in chitin (scheme 1). Thus, it was of interest to examine if a
715
[CuO]+ species in our SmAA10A-β[100] model would be compatible with catalysis. The Cu
716
bound oxygen atom aligns in a plane with Cu and the Cu-coordinating nitrogen atoms from
717
the histidine-brace (Fig. 6C). For this singlet species, our calculations predict a Cu-O distance
718
of 1.78 Å, while the O-H1 and O-H4 distances are 2.08 and 3.21 Å, respectively. The shorter
719
O-H1 distance, and also the Cu-O-H1 angle of 145.5°, are indeed what could be expected for
720
a pre-catalytic complex when comparing with data obtained with the analogues [FeO]2+
721
core.73 Moreover, the longer O-H4 distance indicate that H4 abstraction is energetically
722
unfavorable compared to H1 abstraction, in line with the observed C1 oxidizing activity of
723
the enzyme.
pre-catalytic complex with a [CuO] species in the SmAA10A active
724
725 726 727 728 729 730 731 732 733 734 735 736
Scheme 1. SmAA10A-catalyzed reaction mechanism of chitin oxidation. The resting state of SmAA10A (only copper-coordinating histidines are shown) is the Cu(II) state (shown as orange sphere) which undergoes a priming reduction yielding the active Cu(I) state (blue sphere). As proposed by Bissaro et al.22, following binding to chitin (represented by a chitobiose fragment), the reaction of the co-substrate H2O2 with Cu(I) leads to the release of a water molecule and formation of a copper-oxyl intermediate. As shown in the present study, the [CuO]+ core is properly poised to allow regioselective hydrogen atom abstraction (HAA) and subsequent hydroxylation at the C1 carbon of the -1 NAG unit, via a copper-oxyl, oxygen rebound mechanism.18 The resulting hydroxylated polysaccharide is then subject to a spontaneous elimination reaction leading to glycosidic bond cleavage.13
737 738
Table 1. Spin Hamiltonian parameters for experimental spectra and QM/MM modelsa Spin Hamiltonian parameters
SmAA10A
SmAA10A
SmAA10A
Model 1
+ (NAG)6
+ β-chitin
(SmAA10A)
Model 2 (SmAA10A + β-chitin)
22 ACS Paragon Plus Environment
Page 23 of 29 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
Biochemistry
g – tensor
ACu – tensor (MHz)
gx
2.018
2.025
2.023
2.034
2.048
gy gz
2.116
2.066
2.064
2.095
2.059
2.259
2.212
2.216
2.198
2.183
Ax
218
3
36
190
36
Ay
131
25
-25
-61
-25
Az
346
560
-610b
-484
-593
Ax, Ay, Az
n.a.c
n.a.
34, 34, 50
32, 32, 47
34, 34, 50
Ax, Ay, Az
n.a.
n.a.
39, 28, 29
42, 31, 32
39, 28, 29
Ax, Ay, Az
n.a.
n.a.
29, 38, 30
38, 30, 29
29, 38, 30
AN – tensor (MHz) ANδ – tensor (MHz) ANε – tensor (MHz)
739 740 741 742 743 744 745 746
a
Parameters in normal font are derived from simulation of experimental EPR spectra (assuming collinear g- and Atensors) and parameters in italics are computed by DFT (PBE0 hybrid functional and Def2-TZVPP for Cu and Def2-TZVP for other atoms, A-tensors are Euler rotated into g-tensor frame). Notably, parameters in italics found in the column “SmAA10A + β-chitin” are computed values obtained from model 2. b The negative sign suggested by DFT was kept while using the experimental determined AzCu when simulating the spectrum in Fig. 5C and D. c Not applicable.
23 ACS Paragon Plus Environment
Biochemistry 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
747
Page 24 of 29
CONCLUDING REMARKS
748
We have demonstrated how biochemical analysis of LPMO catalytic action can be
749
used to assist the building of reliable enzyme-substrate models, by reducing the number of
750
possible starting molecular conformations. After extensive molecular dynamics simulations,
751
we were able to analyze interactions between the LPMO and its crystalline substrate,
752
revealing how evolutionary conserved amino acids secure tight binding. We could also
753
demonstrate that the conserved amino acids on the LPMO surface mainly interact with the
754
polysaccharide chain that ends up being cleaved by the enzyme. Our biochemical data and
755
models indicate that SmAA10A can cleave chitin chains at crystalline edges with very narrow
756
substrate surfaces as well as amorphous chitin substrates.
757
The LPMO-chitin complexes showed a well-defined LPMO-substrate association,
758
which was reproducible for all three models, adding confidence to the data. The positioning
759
of the LPMO on the crystalline surface brings the activated oxygen-species close to the
760
hydrogen atom on the C1 atom of the scissile bond, which is compatible with the
761
experimentally observed C1 specificity of SmAA10A. EPR spectra predicted on the basis of
762
our models and experimentally determined EPR spectra for SmAA10A interacting with
763
substrate were similar. It has been shown for several LPMOs that substrate binding alters the
764
Cu(II) EPR signal and several possible causes have been proposed. The present data
765
demonstrate that reorganization of Cu-coordinating water molecules is the most likely cause
766
of the changed EPR spectrum.
767
Importantly, the models showed that the mono-copper active site of SmAA10A is
768
connected to the bulk solvent through a narrow tunnel that would allow diffusion of small
769
molecules such as H2O, O2 and H2O2. This tunnel appears to be gated by Glu60, which is
770
highly conserved and known to be catalytically important from site-directed mutagenesis
771
studies. The accurate description of the LPMO-substrate complex presented here shows a
772
constrained molecular geometry of the catalytic center that likely is of crucial importance for
773
controlling the powerful oxidative chemistry carried out by LPMOs and that will help guiding
774
future mechanistic studies.
775 776
ASSOCIATED CONTENT
777
Supporting information. The supporting information contains supplemental figures 1 to 13,
778
tables S1-S3 and movie S1.
779 780
AUTHORS INFORMATION
781 782
Corresponding author *
[email protected] 24 ACS Paragon Plus Environment
Page 25 of 29 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
Biochemistry
783 784 785 786 787 788
ORCID Bastien Bissaro: 0000-0001-8354-3892 Vincent G.H. Eijsink: 0000-0002-9220-8743 Gustav Vaaje-Kolstad: 0000-0002-3077-8003 Åsmund K. Røhr: 0000-0002-4956-4865
789
Notes
790
The authors declare no competing financial interest
791 792
Present address
793 794 795
‡
Faculty of Chemistry, Biotechnology and Food Science, Norwegian University of Life Sciences, N-1432 Ås, Norway
796 797
ACKNOWLEDGEMENTS
798 799 800 801 802 803 804 805 806 807 808 809 810
This work was supported by the Research Council of Norway grants 240967, 249865 and 262853. Computational work was performed on the Abel Cluster, owned by the University of Oslo, and the Norwegian metacenter for High Performance Computing (NOTUR) and the Extreme Science and Engineering Discovery Environment (XSEDE), which is supported by National Science Foundation grant number ACI-1548562, through allocation TGMCB090159. We would also like to acknowledge the Swiss-Norwegian Beamlines team at the European Synchrotron Radiation Facility, Grenoble, France, for assistance with powder diffraction experiments. We would also like to thank Prof. Einar Sagstuen and Efim Brondz for their assistance at the EPR laboratory, University of Oslo, Norway. Finally, we want to thank Dr. Jennifer Loose for supplying samples of the Y54A mutant of SmAA10A and of the chito-oligosaccharide oxidase.
811 812 813 814 815 816 817 818
AscA, ascorbic acid; CHOS, chito-oligosaccharides; DFT, density functional theory; DP, degree of polymerization; EfAA10A, AA10A from Enterococcus faecalis; EPR, electron paramagnetic resonance; LsAA9A, AA9A from Lentinus similis; LPMO: Lytic polysaccharide monooxygenase; MD, molecular dynamics; NAG, N-acetyl glucosamine; NR, non-reducing end; PCA, principal component analysis; QM/MM, quantum mechanics/molecular mechanics; R, reducing end; SmAA10A or CBP21: AA10A from Serratia marcescens
819 820 821 822 823 824 825 826 827 828 829 830 831 832
REFERENCES
ABBREVIATIONS
(1) Forsberg, Z., Vaaje-kolstad, G., Westereng, B., Bunsæ, A. C., Stenstrøm, Y., Mackenzie, A., Sørlie, M., Horn, S. J., and Eijsink, V. G. H. (2011) Cleavage of cellulose by a CBM33 protein. Protein Sci. 20, 1479–1483. (2) Quinlan, R. J., Sweeney, M. D., Lo Leggio, L., Otten, H., Poulsen, J.-C. N., Johansen, K. S., Krogh, K. B. R. M., Jorgensen, C. I., Tovborg, M., Anthonsen, A., Tryfona, T., Walter, C. P., Dupree, P., Xu, F., Davies, G. J., and Walton, P. H. (2011) Insights into the oxidative degradation of cellulose by a copper metalloenzyme that exploits biomass components. Proc. Natl. Acad. Sci. 108, 15079–15084. (3) Phillips, C. M., Beeson, W. T., Cate, J. H., and Marletta, M. A. (2011) Cellobiose dehydrogenase and a copper-dependent polysaccharide monooxygenase potentiate cellulose degradation by Neurospora crassa. ACS Chem. Biol. 6, 1399–1406. (4) Vaaje-Kolstad, G., Westereng, B., Horn, S. J., Liu, Z., Zhai, H., Sørlie, M., and Eijsink, V. G. H. (2010) An oxidative enzyme boosting the enzymatic conversion of recalcitrant polysaccharides. Science 330, 219–222.
25 ACS Paragon Plus Environment
Biochemistry 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
833 834 835 836 837 838 839 840 841 842 843 844 845 846 847 848 849 850 851 852 853 854 855 856 857 858 859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891 892
Page 26 of 29
(5) Agger, J. W., Isaksen, T., Varnai, A., Vidal-Melgosa, S., Willats, W. G. T., Ludwig, R., Horn, S. J., Eijsink, V. G. H., and Westereng, B. (2014) Discovery of LPMO activity on hemicelluloses shows the importance of oxidative processes in plant cell wall degradation. Proc. Natl. Acad. Sci. 111, 6287– 6292. (6) Frommhagen, M., Sforza, S., Westphal, A. H., Visser, J., Hinz, S. W. A., Koetsier, M. J., Van Berkel, W. J. H., Gruppen, H., and Kabel, M. A. (2015) Discovery of the combined oxidative cleavage of plant xylan and cellulose by a new fungal polysaccharide monooxygenase. Biotechnol. Biofuels 8, 101. (7) Vu, V. V., Beeson, W. T., Span, E. A., Farquhar, E. R., and Marletta, M. A. (2014) A family of starch-active polysaccharide monooxygenases. Proc. Natl. Acad. Sci. 111, 13822–13827. (8) Floudas, D., Binder, M., Riley, R., Barry, K., Blanchette, R. A., Henrissat, B., Martínez, A. T., Otillar, R., Spatafora, J. W., Yadav, J. S., Aerts, A., Benoit, I., Boyd, A., Carlson, A., Copeland, A., Coutinho, P. M., De Vries, R. P., Ferreira, P., Findley, K., Foster, B., Gaskell, J., Glotzer, D., Górecki, P., Heitman, J., Hesse, C., Hori, C., Igarashi, K., Jurgens, J. A., Kallen, N., Kersten, P., Kohler, A., Kües, U., Kumar, T. K. A., Kuo, A., LaButti, K., Larrondo, L. F., Lindquist, E., Ling, A., Lombard, V., Lucas, S., Lundell, T., Martin, R., McLaughlin, D. J., Morgenstern, I., Morin, E., Murat, C., Nagy, L. G., Nolan, M., Ohm, R. A., Patyshakuliyeva, A., Rokas, A., Ruiz-Dueñas, F. J., Sabat, G., Salamov, A., Samejima, M., Schmutz, J., Slot, J. C., John, F. S., Stenlid, J., Sun, H., Sun, S., Syed, K., Tsang, A., Wiebenga, A., Young, D., Pisabarro, A., Eastwood, D. C., Martin, F., Cullen, D., Grigoriev, I. V., and Hibbett, D. S. (2012) The paleozoic origin of enzymatic lignin decomposition reconstructed from 31 fungal genomes. Science 336, 1715–1719. (9) Johansen, K. S. (2016) Discovery and industrial applications of lytic polysaccharide monooxygenases. Biochem. Soc. Trans. 44, 143–149. (10) 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., and van Aalten, D. M. F. (2012) The Vibrio cholerae colonization factor GbpA possesses a modular structure that governs binding to different host surfaces. PLoS Pathog. (Ghosh, P., Ed.) 8, e1002373. (11) Loose, J. S. M., Forsberg, Z., Fraaije, M. W., Eijsink, V. G. H., and Vaaje-Kolstad, G. (2014) A rapid quantitative activity assay shows that the Vibrio cholerae colonization factor GbpA is an active lytic polysaccharide monooxygenase. FEBS Lett. 588, 3435–3440. (12) Agostoni, M., Hangasky, J. A., and Marletta, M. A. (2017) Physiological and Molecular Understanding of Bacterial Polysaccharide Monooxygenases. Microbiol. Mol. Biol. Rev. 81, e0001517. (13) Beeson, W. T., Phillips, C. M., Cate, J. H. D., and Marletta, M. A. (2012) Oxidative cleavage of cellulose by fungal copper-dependent polysaccharide monooxygenases. J. Am. Chem. Soc. 134, 890– 892. (14) Eibinger, M., Ganner, T., Bubner, P., Rošker, S., Kracher, D., Haltrich, D., Ludwig, R., Plank, H., and Nidetzky, B. (2014) Cellulose surface degradation by a lytic polysaccharide monooxygenase and its effect on cellulase hydrolytic efficiency. J. Biol. Chem. 289, 35929–35938. (15) Vermaas, J. V., Crowley, M. F., Beckham, G. T., and Payne, C. M. (2015) Effects of Lytic Polysaccharide Monooxygenase Oxidation on Cellulose Structure and Binding of Oxidized Cellulose Oligomers to Cellulases. J. Phys. Chem. B 119, 6129–6143. (16) Villares, A., Moreau, C., Bennati-Granier, C., Garajova, S., Foucat, L., Falourd, X., Saake, B., Berrin, J. G., and Cathala, B. (2017) Lytic polysaccharide monooxygenases disrupt the cellulose fibers structure. Sci. Rep. 7, 40262. (17) Eibinger, M., Sattelkow, J., Ganner, T., Plank, H., and Nidetzky, B. (2017) Single-molecule study of oxidative enzymatic deconstruction of cellulose. Nat. Commun. 8, 894. (18) Kim, S., Stahlberg, J., Sandgren, M., Paton, R. S., and Beckham, G. T. (2014) Quantum mechanical calculations suggest that lytic polysaccharide monooxygenases use a copper-oxyl, oxygenrebound mechanism. Proc. Natl. Acad. Sci. 111, 149–154. (19) 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., and Solomon, E. I. (2014) Spectroscopic and computational insight into the activation of O2 by the mononuclear Cu center in polysaccharide monooxygenases. Proc. Natl. Acad. Sci. 111, 8797–8802. (20) Beeson, W. T., Vu, V. V., Span, E. A., Phillips, C. M., and Marletta, M. A. (2015) Cellulose Degradation by Polysaccharide Monooxygenases. Annu. Rev. Biochem. 84, 923–946. (21) Walton, P. H., and Davies, G. J. (2016) On the catalytic mechanisms of lytic polysaccharide monooxygenases. Curr. Opin. Chem. Biol. 31, 195–207. (22) Bissaro, B., Røhr, Å. K., Müller, G., Chylenski, P., Skaugen, M., Forsberg, Z., Horn, S. J., VaajeKolstad, G., and Eijsink, V. G. H. (2017) Oxidative cleavage of polysaccharides by monocopper
26 ACS Paragon Plus Environment
Page 27 of 29 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
Biochemistry
893 894 895 896 897 898 899 900 901 902 903 904 905 906 907 908 909 910 911 912 913 914 915 916 917 918 919 920 921 922 923 924 925 926 927 928 929 930 931 932 933 934 935 936 937 938 939 940 941 942 943 944 945 946 947 948 949 950 951 952
enzymes depends on H2O2. Nat. Chem. Biol. 13, 1123–1128. (23) Aachmann, F. L., Sorlie, M., Skjak-Braek, G., Eijsink, V. G. H., and Vaaje-Kolstad, G. (2012) NMR structure of a lytic polysaccharide monooxygenase provides insight into copper binding, protein dynamics, and substrate interactions. Proc. Natl. Acad. Sci. 109, 18779–18784. (24) Courtade, G., Wimmer, R., Røhr, Å. K., Preims, M., Felice, A. K. G., Dimarogona, M., VaajeKolstad, G., Sørlie, M., Sandgren, M., Ludwig, R., Eijsink, V. G. H., and Aachmann, F. L. (2016) Interactions of a fungal lytic polysaccharide monooxygenase with β-glucan substrates and cellobiose dehydrogenase. Proc. Natl. Acad. Sci. 113, 5922–5927. (25) 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., and Walton, P. H. (2016) The molecular basis of polysaccharide cleavage by lytic polysaccharide monooxygenases. Nat. Chem. Biol. 12, 298–303. (26) Vaaje-Kolstad, G., Houston, D. R., Riemen, A. H. K., Eijsink, V. G. H., and Van Aalten, D. M. F. (2005) Crystal structure and binding properties of the Serratia marcescens chitin-binding protein CBP21. J. Biol. Chem. 280, 11313–11319. (27) 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., Ståhlberg, J., Eijsink, V. G. H., and Sandgren, M. (2013) Crystal structure and computational characterization of the lytic polysaccharide monooxygenase GH61D from the basidiomycota fungus Phanerochaete chrysosporium. J. Biol. Chem. 288, 12828–12839. (28) Nimlos, M. R., Beckham, G. T., Matthews, J. F., Bu, L., Himmel, M. E., and Crowley, M. F. (2012) Binding preferences, surface attachment, diffusivity, and orientation of a family 1 carbohydratebinding module on cellulose. J. Biol. Chem. 287, 20603–20612. (29) Vaaje-Kolstad, G., Forsberg, Z., Loose, J. S., Bissaro, B., and Eijsink, V. G. (2017) Structural diversity of lytic polysaccharide monooxygenases. Curr. Opin. Struct. Biol. 44, 67–76. (30) Frandsen, K. E. H., and Lo Leggio, L. (2016) Lytic polysaccharide monooxygenases: A crystallographer’s view on a new class of biomass-degrading enzymes. IUCrJ 3, 448–467. (31) Vaaje-Kolstad, G., Bøhle, L. A., Gåseidnes, S., Dalhus, B., Bjørås, M., Mathiesen, G., and Eijsink, V. G. H. (2012) Characterization of the chitinolytic machinery of enterococcus faecalis V583 and highresolution structure of its oxidative CBM33 enzyme. J. Mol. Biol. 416, 239–254. (32) Nakagawa, Y. S., Eijsink, V. G. H., Totani, K., and Vaaje-Kolstad, G. (2013) Conversion of αchitin substrates with varying particle size and crystallinity reveals substrate preferences of the chitinases and lytic polysaccharide monooxygenase of Serratia marcescens. J. Agric. Food Chem. 61, 11061–11066. (33) Nakagawa, Y. S., Kudo, M., Loose, J. S. M., Ishikawa, T., Totani, K., Eijsink, V. G. H., and Vaaje-Kolstad, G. (2015) A small lytic polysaccharide monooxygenase from Streptomyces griseus targeting α- And β-chitin. FEBS J. 282, 1065–1079. (34) Forsberg, Z., Nelson, C. E., Dalhus, B., Mekasha, S., Loose, J. S. M., Crouch, L. I., Røhr, Å. K., Gardner, J. G., Eijsink, V. G. H., and Vaaje-Kolstad, G. (2016) Structural and functional analysis of a lytic polysaccharide monooxygenase important for efficient utilization of chitin in Cellvibrio japonicus. J. Biol. Chem. 291, 7300–7312. (35) Westereng, B., Agger, J. W., Horn, S. J., Vaaje-Kolstad, G., Aachmann, F. L., Stenstrøm, Y. H., and Eijsink, V. G. H. (2013) Efficient separation of oxidized cello-oligosaccharides generated by cellulose degrading lytic polysaccharide monooxygenases. J. Chromatogr. A 1271, 144–152. (36) Stoll, S., and Schweiger, A. (2006) EasySpin, a comprehensive software package for spectral simulation and analysis in EPR. J. Magn. Reson. 178, 42–55. (37) Dyadkin, V., Pattison, P., Dmitriev, V., and Chernyshov, D. (2016) A new multipurpose diffractometer PILATUS@SNBL. J. Synchrotron Radiat. 23, 825–829. (38) Anandakrishnan, R., Aguilar, B., and Onufriev, A. V. (2012) H++ 3.0: Automating pK prediction and the preparation of biomolecular structures for atomistic molecular modeling and simulations. Nucleic Acids Res. 40, W537–W541. (39) Frisch, M. J., Trucks, G. W., Schlegel, H. B., Scuseria, G. E., Robb, M. A., Cheeseman, J. R., Scalmani, G., Barone, V., Mennucci, B., Petersson, G. A., Nakatsuji, H., Caricato, M., Li, X., Hratchian, H. P., Izmaylov, A. F., Bloino, J., Zheng, G., Sonnenberg, J. L., Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., Kitao, O., Nakai, H., Vreven, T., Montgomery Jr., J. A., Peralta, J. E., Ogliaro, F., Bearpark, M., Heyd, J. J., Brothers, E., Kudin, K. N., Staroverov, V. N., Kobayashi, R., Normand, J., Raghavachari, K., Rendell, A., Burant, J. C., Iyengar, S. S., Tomasi, J., Cossi, M., Rega, N., Millam, J. M., Klene, M., Knox, J. E., Cross, J. B., Bakken, V., Adamo, C., Jaramillo, J., Gomperts, R., Stratmann, R. E., Yazyev, O., Austin, A. J.,
27 ACS Paragon Plus Environment
Biochemistry 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
953 954 955 956 957 958 959 960 961 962 963 964 965 966 967 968 969 970 971 972 973 974 975 976 977 978 979 980 981 982 983 984 985 986 987 988 989 990 991 992 993 994 995 996 997 998 999 1000 1001 1002 1003 1004 1005 1006 1007 1008 1009 1010 1011 1012
Page 28 of 29
Cammi, R., Pomelli, C., Ochterski, J. W., Martin, R. L., Morokuma, K., Zakrzewski, V. G., Voth, G. A., Salvador, P., Dannenberg, J. J., Dapprich, S., Daniels, A. D., Farkas, Ö., Foresman, J. B., Ortiz, J. V, Cioslowski, J., and Fox, D. J. (2009) Gaussian 09, Revision D.01. Gaussian Inc. Gaussian, Inc., Wallingford CT. (40) D.A. Case, D.S. Cerutti, T.E. Cheatham, III, T.A. Darden, R.E. Duke, T.J. Giese, H. Gohlke, A.W. Goetz, D. Greene, N. Homeyer, S. Izadi, A. Kovalenko, T.S. Lee, S. LeGrand, P. Li, C. Lin, J. Liu, T. Luchko, R. Luo, D. Mermelstein, K.M. Merz, G. Monard, H., D. M. Y. and P. A. K. (2017) AMBER 2017, University of California, San Francisco. (41) Seminario, J. M. (1996) Calculation of intramolecular force fields from second-derivative tensors. Int. J. Quantum Chem. 60, 1271–1277. (42) Beckham, G. T., and Crowley, M. F. (2011) Examination of the α-chitin structure and decrystallization thermodynamics at the nanoscale. J. Phys. Chem. B 115, 4516–4522. (43) Sikorski, P., Hori, R., and Wada, M. (2009) Revisit of α-chitin crystal structure using high resolution X-ray diffraction data. Biomacromolecules 10, 1100–1105. (44) Sawada, D., Nishiyama, Y., Langan, P., Forsyth, V. T., Kimura, S., and Wada, M. (2012) Water in crystalline fibers of dihydrate β-chitin results in unexpected absence of intramolecular hydrogen bonding. PLoS One (Gasset, M., Ed.) 7, e39376. (45) Maier, J. A., Martinez, C., Kasavajhala, K., Wickstrom, L., Hauser, K. E., and Simmerling, C. (2015) ff14SB: Improving the Accuracy of Protein Side Chain and Backbone Parameters from ff99SB. J. Chem. Theory Comput. 11, 3696–3713. (46) Kirschner, K. N., Yongye, A. B., Tschampel, S. M., González-Outeiriño, J., Daniels, C. R., Foley, B. L., and Woods, R. J. (2008) GLYCAM06: A generalizable biomolecular force field. carbohydrates. J. Comput. Chem. 29, 622–655. (47) Salomon-Ferrer, R., Götz, A. W., Poole, D., Le Grand, S., and Walker, R. C. (2013) Routine microsecond molecular dynamics simulations with AMBER on GPUs. 2. Explicit solvent particle mesh ewald. J. Chem. Theory Comput. 9, 3878–3888. (48) Roe, D. R., and Cheatham, T. E. (2013) PTRAJ and CPPTRAJ: Software for processing and analysis of molecular dynamics trajectory data. J. Chem. Theory Comput. 9, 3084–3095. (49) Hedegård, E. D., and Ryde, U. (2017) Targeting the reactive intermediate in polysaccharide monooxygenases. J. Biol. Inorg. Chem. 22, 1029–1037. (50) Bertini, L., Breglia, R., Lambrughi, M., Fantucci, P., De Gioia, L., Borsari, M., Sola, M., Bortolotti, C. A., and Bruschi, M. (2017) Catalytic Mechanism of Fungal Lytic Polysaccharide Monooxygenases Investigated by First-Principles Calculations. Inorg. Chem. acs.inorgchem.7b02005. (51) Götz, A. W., Clark, M. A., and Walker, R. C. (2014) An extensible interface for QM/MM molecular dynamics simulations with AMBER. J. Comput. Chem. 35, 95–108. (52) Neese, F. (2012) The ORCA program system. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2, 73–78. (53) Becke, A. D. (1988) Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 38, 3098–3100. (54) Perdew, J. P. (1986) Density-functional approximation for the correlation energy of the inhomogeneous electron gas. Phys. Rev. B 33, 8822–8824. (55) Bühl, M., and Kabrede, H. (2006) Geometries of transition-metal complexes from densityfunctional theory. J. Chem. Theory Comput. 2, 1282–1290. (56) Grimme, S., Antony, J., Ehrlich, S., and Krieg, H. (2010) A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 132, 154104. (57) Grimme, S., Ehrlich, S., and Goerigk, L. (2011) Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 32, 1456–1465. (58) Weigend, F., and Ahlrichs, R. (2005) Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 7, 3297. (59) Weigend, F. (2006) Accurate Coulomb-fitting basis sets for H to Rn. Phys. Chem. Chem. Phys. 8, 1057. (60) Adamo, C., and Barone, V. (1999) Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 110, 6158–6170. (61) Sunna, A., Moracci, M., Rossi, M., and Antranikian, G. (1997) Glycosyl hydrolases from hyperthermophiles. Extremophiles 1, 2–13. (62) Kuusk, S., Bissaro, B., Kuusk, P., Forsberg, Z., Eijsink, V. G. H., Sørlie, M., and Väljamäe, P. (2017) Kinetics of H2O2-driven degradation of chitin by a bacterial lytic polysaccharide monooxygenase. J. Biol. Chem. jbc.M117.817593. (63) Ogawa, Y., Kimura, S., Wada, M., and Kuga, S. (2010) Crystal analysis and high-resolution
28 ACS Paragon Plus Environment
Page 29 of 29 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
Biochemistry
1013 1014 1015 1016 1017 1018 1019 1020 1021 1022 1023 1024 1025 1026 1027 1028 1029 1030 1031 1032 1033 1034 1035 1036 1037 1038 1039 1040 1041 1042 1043
imaging of microfibrillar α-chitin from Phaeocystis. J. Struct. Biol. 171, 111–116. (64) Ogawa, Y., Kimura, S., and Wada, M. (2011) Electron diffraction and high-resolution imaging on highly-crystalline β-chitin microfibril. J. Struct. Biol. 176, 83–90. (65) Suzuki, K., Suzuki, M., Taiyoji, M., Nikaidou, N., and Watanabe, T. (1998) Chitin Binding Protein (CBP21) in the Culture Supernatant of Serratia marcescens 2170. Biosci. Biotechnol. Biochem. 62, 128–135. (66) Danneels, B., Tanghe, M., Joosten, H. J., Gundinger, T., Spadiut, O., Stals, I., and Desmet, T. (2017) A quantitative indicator diagram for lytic polysaccharide monooxygenases reveals the role of aromatic surface residues in HjLPMO9A regioselectivity. PLoS One 12, e0178446. (67) Forsberg, Z., Bissaro, B., Gullesen, J., Dalhus, B., Vaaje-Kolstad, G., and Eijsink, V. G. H. (2018) Structural determinants of bacterial lytic polysaccharide monooxygenase functionality. J. Biol. Chem. 293, 1397–1412. (68) Kracher, D., Andlar, M., Furtmüller, P. G., and Ludwig, R. (2017) Active-site copper reduction promotes substrate binding of fungal lytic polysaccharide monooxygenase and reduces stability. J. Biol. Chem. 293, 1676-1687. (69) Chovancova, E., Pavelka, A., Benes, P., Strnad, O., Brezovsky, J., Kozlikova, B., Gora, A., Sustr, V., Klvana, M., Medek, P., Biedermannova, L., Sochor, J., and Damborsky, J. (2012) CAVER 3.0: A Tool for the Analysis of Transport Pathways in Dynamic Protein Structures. PLoS Comput. Biol. 8. (70) Borisova, A. S., Isaksen, T., Dimarogona, M., Kognole, A. A., Mathiesen, G., Várnai, A., Røhr, Å. K., Payne, C. M., Sørlie, M., Sandgren, M., and Eijsink, V. G. H. (2015) Structural and functional characterization of a lytic polysaccharide monooxygenase with broad substrate specificity. J. Biol. Chem. 290, 22955–22969. (71) de Visser, S. P., Quesne, M. G., Martin, B., Comba, P., and Ryde, U. (2014) Computational modelling of oxygenation processes in enzymes and biomimetic model complexes. Chem. Commun. 50, 262–282. (72) Meier, K. K., Jones, S. M., Kaper, T., Hansson, H., Koetsier, M. J., Karkehabadi, S., Solomon, E. I., Sandgren, M., and Kelemen, B. (2017) Oxygen Activation by Cu LPMOs in Recalcitrant Carbohydrate Polysaccharide Conversion to Monomer Sugars. Chem. Rev. acs.chemrev.7b00421. (73) Ye, S., and Neese, F. (2011) Nonheme oxo-iron ( IV ) intermediates form an oxyl radical upon approaching the C – H bond activation transition state. Proc. Natl. Acad. Sci. 108, 1228–1233.
29 ACS Paragon Plus Environment