You are on page 1of 11

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/323671039

QM/MM Approach on the Structural and Stereolectronic Factors governing


Glycosylation by GTF-SI from Streptococcus mutans

Article  in  Organic & Biomolecular Chemistry · March 2018


DOI: 10.1039/C8OB00284C

CITATIONS READS

5 248

7 authors, including:

Gonzalo Jaña Fernanda Mendoza


Universidad Andrés Bello Universidad Andrés Bello
28 PUBLICATIONS   168 CITATIONS    20 PUBLICATIONS   82 CITATIONS   

SEE PROFILE SEE PROFILE

Manuel Osorio Joel Alderete


Universidad Andrés Bello Universidad de Talca
11 PUBLICATIONS   28 CITATIONS    128 PUBLICATIONS   1,626 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

UNVEILING THE FACTORS THAT DETERMINE THE ENZYMATIC C-GLYCOSIDIC BOND FORMATION: A QM/MM APPROACH View project

The structural and thermodynamic factors that determine 3 ', 4' or 6 'transglucosidase activity in the GTF-SI enzyme View project

All content following this page was uploaded by Gonzalo Jaña on 21 March 2018.

The user has requested enhancement of the downloaded file.


Organic &
Biomolecular Chemistry
View Article Online
PAPER View Journal

A QM/MM approach on the structural and


Published on 09 March 2018. Downloaded by Universidad de Concepcion on 20/03/2018 13:42:30.

Cite this: DOI: 10.1039/c8ob00284c


stereoelectronic factors governing glycosylation
by GTF-SI from Streptococcus mutans†
Gonzalo A. Jaña, *a Fernanda Mendoza, a Manuel I. Osorio, a,b Joel B. Alderete, c

Pedro A. Fernandes, d Maria J. Ramos d and Verónica A. Jiménez a

In this work, QM/MM calculations were employed to examine the catalytic mechanism of the retaining
glucosyltransferase GTF-SI enzyme, which participates in the process of caries formation. Our goal was to
characterize, with atomistic details, the mechanism of sucrose hydrolysis and the catalytic factors that
modulate this reaction. Our results suggest a concerted mechanism for sucrose hydrolysis in which the
first event corresponds to the glycosidic bond breakage assisted by Glu515, followed by the nucleophilic
attack of Asp477, leading to the formation of the Covalent Glycosyl Enzyme (CGE) intermediate. A novel
conformational itinerary of the glucosyl moiety along the reaction mechanism was identified: 2H3 → 2H3–
E3 → 4C1, and the calculated energy barrier is 16.4 kcal mol−1, which is in good agreement with experi-
mental evidence showing a major contribution coming from the glycosidic bond breakage. Our calculations
Received 2nd February 2018, also revealed that Arg475 and Asp588 play a critical role as TS-stabilizers by electrostatic and charge transfer
Accepted 9th March 2018
mechanisms, respectively. This is the first report dealing with the specific features of the mechanism and
DOI: 10.1039/c8ob00284c catalytic residues involved in GTF-SI hydrolysis of sucrose, which is a matter of relevance in enzyme catalysis
rsc.li/obc and could be valuable to aid the design of novel and specific inhibitors targeting GTF-SI.

1. Introduction Glucans are synthesized by 1,3-1,6-α-glucansucrases (also


known by the acronyms of GSs or GTFs for glucansucrases and
Dental caries are a major public health problem in most glucosyltransferases, respectively) which are produced by oral
industrialized countries, affecting 60–90% of school-aged chil- bacteria streptococci, like Streptococcus mutans or Streptococcus
dren and the vast majority of adults as a consequence of sobrinus. GTFs are members of the Glycoside Hydrolase (GH)
growing sugar consumption.1–3 This infection relies on the family 70 and catalyze the formation of glycosidic bonds
biosynthesis of an extracellular polymeric film that shelters between glucose units, producing different α-glucans from
microorganisms, and allows their adhesion to biological sur- sucrose through transglucosylation reactions.5,6 GTFs are
faces, thus enabling bacterial growth.4 Glucans are the main classified into GTF-I, GTF-SI and GTF-S. GTF-S is a dextransu-
polymeric components of this biofilm that play a key role in crase that synthesizes predominantly soluble glucans with
the cariogenesis process, for which they have gained impor- α(1 → 6) linkages, whereas GTF-I and GTF-SI are mutansu-
tance in the search for new treatments to prevent dental crases that synthesize mainly insoluble glucans with α(1 → 3)
caries.5 glycosidic linkages (mutans). As the synthesis of mutans aids
Streptococcus mutans (S. mutans) in adhering to the tooth
surface, the inhibition of GTF-SI or GTF-I activities might
a
Departamento de CienciasQuímicas, Facultad de Ciencias Exactas, impair the cariogenic virulence of tooth plaque.7 Recent
Universidad Andres Bello, Sede Concepción, Autopista Concepción-Talcahuano 7100, experimental studies have shown that inactivating two GTF
Talcahuano, Chile. E-mail: gonzalo.jana@unab.cl
b encoding genes of S. mutans markedly reduced the plaque
Doctorado en Fisicoquímica molecular, Universidad Andres Bello, Republica 275,
Santiago, Chile formation on tooth surfaces in vitro, which suggests that mole-
c
Departamento de Química Orgánica, Facultad de Ciencias Químicas, cules that were able to inhibit GTFs could be potential anti-
Universidad de Concepción, Casilla 160-C, Concepción, Chile caries agents. Nowadays, there are natural and synthetic mole-
d
UCIBIO, REQUIMTE, Departamento de Química e Bioquímica, Faculdade de cules whose inhibitory activity results in poor biofilm for-
Ciências, Universidade do Porto, Rua do Campo Alegre, s/n, 4169-007 Porto,
mation; however, it is not known how these molecules target
Portugal
† Electronic supplementary information (ESI) available. See DOI: 10.1039/ GTFs from S. mutans. Although many of the available inhibi-
c8ob00284c tors are highly effective in vitro, it is still unclear whether they

This journal is © The Royal Society of Chemistry 2018 Org. Biomol. Chem.
View Article Online

Paper Organic & Biomolecular Chemistry

are suitable for in vivo use, and their mechanisms remain addition, there are no computational studies considering
largely unknown. Furthermore, the consequences for the sucrose acting as the donor substrate, and therefore a detailed
human metabolic GHs (like α-amylases) or the health-associ- understanding of the mechanistic and catalytic factors that
ated microbiota are also unclear.7 rule the mutan biosynthesis performed by either of these
Even though the catalytic mechanism is known for several enzymes is a key factor to identify specific inhibitors for GTF-I
members from different GH families,8–10 the specific details or GTF-SI.
about how GTF-SI from Streptococcus mutans catalyzes the for- The recent elucidation of the crystal structure of the GTF-SI
mation of α-glucans have never been addressed at a molecular enzyme (PDB ID: 3AIE) from Streptococcus mutans15 represents
atomic level due to the lack of a crystalline structure. Thus, the a great opportunity for understanding the catalytic mechanism
Published on 09 March 2018. Downloaded by Universidad de Concepcion on 20/03/2018 13:42:30.

catalytic mechanism of glucan synthesis has been proposed by involved in mutan synthesis, which could allow the identifi-
direct analogy with GH13 (α-amylases) and GH77 (amylomal- cation of structural and electronic changes that are relevant
tases) families (all of them belong to GH-H clan).11,12 These during the catalysis. The X-ray structure possesses an appropri-
enzymes, and therefore the GTFs,12–15 follow an α-retaining ate resolution (2.10 Å) to conduct the computational modeling
double displacement mechanism that starts with the attack of of the catalytic conversion from sucrose to mutan. From this
a nucleophilic residue (Asp) to the anomeric carbon of the glu- crystal structure, three protein residues have been proposed to
cosyl moiety of sucrose concerted with a glutamic acid proto- play a relevant catalytic role, namely Asp477 as a nucleophile,
nation of the glycosidic oxygen of the fructosyl leaving group. Glu515 as an acid/base, and Asp588 as a transition state stabil-
In this way, the covalent β-glucosyl enzyme intermediate (CGE) izer.15 Homologous catalytic residues have been found in
is formed via an oxocarbenium ion-like transition state (TS), another GTF crystal structure from Lactobacillus reuteri 180.12
this step being rate-limiting. In the second step, the leaving Thus, the 3AIE crystal structure offers a suitable model to
fructose is exchanged by a new acceptor sucrose (R-OH, Fig. 1), conduct exhaustive computational studies on GTF-SI, aimed at
which gets deprotonated by a glutamate residue acting as a elucidating the fundamental aspects regarding the mecha-
base. The R-OH attacks the anomeric carbon of Asp-sucrose nism, such as the key amino acid residues stabilizing tran-
(CGE) while the leaving group Asp departs. Thus, the final sition state structures, and the catalytic steps involved in the
α-glycosidic product is formed in a concerted step via another reaction path.
oxocarbenium ion-like TS. Pinto et al.16 studied the catalytic Herein we carried out Quantum Mechanics (QM) and
mechanism of human pancreatic α-amylase (HPA) from the Quantum Mechanics/Molecular Mechanics (QM/MM) calcu-
GH13 family by means of QM/MM calculations. These calcu- lations starting from the structure of sucrose in complex with
lations enabled the analysis of atomic structural details related the crystalline structure of the GTF-SI enzyme. In a first stage,
to the transfer reaction, which led to the characterization of a we built a reduced cluster model of the active site including
mechanism that turned out to be in good agreement with the the main catalytic residues Asp477 and Glu515. Then, the
empirically proposed mechanism for α-amylase’s enzymes. enzymatic environment was incorporated in order to fully
Although the general mechanism is already understood, characterize the catalytic strategies of this enzyme by means of
homologous catalytic residues within the GH13 (α-amylase) the QM/MM method.
and GH70 (GTF-SI) families do not necessarily imply identical
catalytic functions17 nor similar mechanisms.18 Additionally,
it is known that different GH-family enzymes catalyzing the 2. Methods
same kind of mechanism can show different conformational
itineraries of the glucosyl ring.19 Thus, the role of several 2.1 Models
active site amino acids in the reaction catalyzed by GTFs from The crystal structure of GTF-SI (2.10 Å resolution) corresponds
Streptococcus mutans has been speculative and further evidence to an octamer composed of equal chains in structure and
must be provided in order to characterize its catalytic mecha- sequence, with no sucrose in the active site (apoenzyme).
nism and the conformational itinerary of the glucosyl ring. In Thus, to model the GTF-SI·sucrose binary complex the chain C

Fig. 1 General double displacement mechanism proposed for the synthesis of α-glucans by GTFs.

Org. Biomol. Chem. This journal is © The Royal Society of Chemistry 2018
View Article Online

Organic & Biomolecular Chemistry Paper

until obtaining a viable TS structure using the Powell/


Murtagh–Sargent scheme.28,29 The stationary points were
characterized as minima or transition state (TS) structures by
means of frequency calculations. Finally, single point energy cal-
culations were performed with the Hybrid-Generalized Gradient
Approximation (H-GGA)ωB97X-V30 functional with nonlocal cor-
relation, using the 6-311++G(d,p) basis set. This functional has
been proved to describe very well different molecular pro-
perties.31 Natural charges and Second Order Perturbation ana-
Published on 09 March 2018. Downloaded by Universidad de Concepcion on 20/03/2018 13:42:30.

lyses were carried out through NBO calculations using the NBO
v5.0 program32 as implemented in Q-Chem.

2.3 QM/MM calculations


The initial GTF-SI·sucrose complex was used as a starting
structure for QM/MM calculations. The system containing
13 893 atoms was partitioned into a Quantum Mechanics part
Fig. 2 Active site view of the superimposed crystal structures GTF-SI
(green) and GTF180ΔN (cyan). (QM) defined by 61 atoms (link atoms not included): 45 atoms
of sucrose, the side chain of the nucleophile Asp477 (6 atoms),
and the side chain of the acid Glu515 (10 atoms). Two link
was chosen and aligned with the GTF180-ΔN crystal structure, atoms were used, each one of them located along the Cα-QM
which is in complex with sucrose (PDB code 3HZ3). The struc- of Asp477 and Glu515. The total charge of the QM subsystem
tural alignment showed a QH20 value of 0.800, which indicates was −1. The geometry of the QM subsystem was described at
a high structural and sequence similarity between 3AIE and the BP86/6-31+G(d) level of theory. The Molecular Mechanics
3HZ3 structures. Thus, the matching procedure enabled locat- (MM) part was described using the Charmm27-all33 atoms
ing the sucrose unit within the active site of GTF-SI (Fig. 2), force field, and an electronic embedding scheme was used for
forming this way the GTF-SI·sucrose complex. This model the QM/MM treatment. CHARMM34,35 and Q-CHEM packages
already features the most relevant interactions for the catalytic were employed to carry out the calculations. Initially, the QM
mechanism and was employed to conduct further QM and part was frozen and the rest of the system was optimized using
QM/MM calculations. a gradient criterion of 0.005 kcal mol−1 Å−1. Then, we defined
The PROPKA method21 was applied to determine the proto- an active region of 25 Å radius centered at the glycosidic
nation states of all ionizable residues at pH = 6.0 in which the oxygen. Every residue within the active region was set free,
GTF-SI enzyme exhibits its maximum activity,22 using the while the remaining residues were frozen during the optimi-
PDB2PQR2.1.1 server.23 According to PROPKA, Glu515 was zations. This partial freezing method was recently found to be
predicted to exist in its protonated state. All other residues appropriate for the treatment of large systems.36 The pro-
were assigned standard protonation states. This complex was cedure afforded an optimized reactant structure that was used
solvated using a water sphere of a 25 Å radius centered at the as a starting point for the initial reaction path calculation.
glycosidic oxygen, forming a model of 13 893 atoms. This initial guess was built using a Reaction Coordinate (RC)
that allowed the conversion of the reactant to the CGE inter-
mediate (Fig. 3).
2.2 Quantum mechanical calculations The initial path was used as a guess to compute the
The initial model was truncated to a smaller system composed Minimum Energy Path (MEP), which was calculated using the
of sucrose, Asp477, Glu515, and three crystalline water mole- Conjugate Peak Refinement (CPR) algorithm37 as
cules (residue numbers 55, 127 and 128) that interact with the implemented in the TREK module of the CHARMM program.
glucosyl and fructosyl moieties of sucrose. Geometry optimi- The CPR algorithm rebuilds the whole path that connects the
zations in the gas phase were computed at the BP86/6-31+G(d) desired reactant and product structures, and locates the
level of theory. This functional has been successfully tested for correct transition state structure, ensuring that the energy
carbohydrate related enzymes showing a good compromise barrier is not affected by a biasing description of the
between computational efficiency and accuracy.24 All calcu- process.38 This method allows determination of the mecha-
lations were carried out using the Q-Chem program nism of complex reactions in proteins and has been success-
package.25,26 The Freezing String Method27 (FSM) was used to fully used in conjunction with QM/MM energy functions to
locate the TS structure. Briefly, this algorithm consists of an study the hydrolysis of chondroitin disaccharide catalyzed by
iterative process in which growing pairs of molecular struc- chondroitin lyase39 and other complex enzymatic
tures along the reaction path, starting from the reactant and reactions.40–42 The CPR criterion for deciding whether a tran-
the product, complete the “string” generating a guess TS struc- sition state has been found was that a gradient of less than
ture. Later, local optimizations of the guess structure were 10−2 kcal mol−1 Å−1 is attained. CPR identifies these points
performed using an exact initial Hessian that was updated along the path where the energy is highest (the peaks) and

This journal is © The Royal Society of Chemistry 2018 Org. Biomol. Chem.
View Article Online

Paper Organic & Biomolecular Chemistry

Glu515: 3O′–HE2 = 1.75 Å, as displayed in Fig. 4. Besides, a


water molecule is interacting with the 4H and 3O of glucose at
1.84 Å and 1.75 Å, respectively. The conformation of glucose at
the reactant is highly distorted exhibiting ϕ = 155° and θ = 82°,
which is consistent with a skew-boat 2SO type, according to the
Cremer–Pople parameters.43 Additionally, the putative nucleo-
phile Asp477 seems to be well positioned for the nucleophilic
attack, with the OD1 atom located at 3.83 Å from the anomeric
carbon of glucose (C1 in Fig. 4).
Published on 09 March 2018. Downloaded by Universidad de Concepcion on 20/03/2018 13:42:30.

The CGE structure was obtained from a flexible scan along


the bond between the glycosidic oxygen and the acid proton of
Glu515 using the optimized reactant structure as a starting
point. Then, the FSM algorithm was applied to localize the TS
structure using the optimized structures of the reactant and
the CGE intermediate. The TS structure has an oxocarbenium
character demonstrated by the change of the C1–O5 bond dis-
tance, which has decreased from 1.43 Å in the reactant to
Fig. 3 QM (black)/MM (gray) partition used in this work. Red lines indi-
1.30 Å in the TS, showing a significant double bond character.
cate the location of the link atoms, and the β-face of the glucosyl The values of ϕ and θ are 205° and 45°, respectively, indicating
moiety is highlighted (opposite face of fructosyl). The labels of several a conformation close to the half-chair 4H3, similar to that
atoms are shown and will be referenced throughout the results section. observed for the retaining α-galactosidases (GH27 family).44
The Reaction Coordinate (RC) has been defined as RC = d(C1–OG) −
Normal mode analysis revealed a unique imaginary frequency
d(C1–OD1) − d(HE2–OG).
of 125.5i cm−1, confirming the nature of a saddle point con-
necting the reactant and the CGE structures. The vibrational
mode was associated with stretching of the C1–OG, C1–
iteratively moves these points closer to the MEP by controlled
OD1Asp477 and the OG–HE2Glu515 bonds, in agreement with a
conjugated-gradient minimization. The transition states were
concerted mechanism for the CGE formation in which the
further validated by means of normal mode analyses of the
proton transfer and the nucleophilic attack occur after the gly-
QM part.
cosidic bond breakage, as confirmed by IRC calculations.
Finally, single point energy calculations were carried out for
Additionally, Natural Population Analysis (NPA) revealed that
the stationary points using the H-GGA ωB97X-V functional
the TS charge, including the atomic charges of the anomeric
together with the 6-311++G(d,p) basis set. The ωB97X-V/6-
center (O5–C1–H1), was 0.30 a.u. higher than the charge of the
311++G(d,p) method was also used for the NBO and electro-
reactant, thus revealing a positive charge development, in
static calculations.
agreement with a oxocarbenium ion-like TS. Natural Bond
Orbital (NBO) analysis confirmed the double bond character of
the C1–O5 bond at the TS and also confirmed that the glycosi-
3. Results and discussion dic bond is already broken, with a distance of 2.15 Å. The
interaction between Glu515 and the glycosidic oxygen (OG)
This work describes the results of QM and QM/MM calcu-
is still present in the TS (HE2–OG = 1.65 Å), showing that the
lations dealing with the mechanism of sucrose hydrolysis by
proton transfer has not occurred yet. The nucleophilic oxygen
GTF-SI. The characterization and analysis of the mechanism
of Asp477 is located at 2.64 Å from the anomeric carbon,
were carried out considering the structural and stereoelectro-
1.19 Å closer than the distance observed at the reactant, pro-
nic factors that could modulate the nature of the TS and the
viding stabilization to the oxocarbenium ion. A new interaction
conformational itinerary of the glucosyl moiety during the
between the 2-hydroxyl group (2OH) of glucose and the OG
reaction, which is an undetermined issue for GTF-SI and its
(2H–OG = 1.65 Å) was observed at the TS, which was attributed
whole GH70 family.
to the increase of negative charge on the OG (0.25 a.u.) upon
glycosidic bond breakage. This suggests that the 2H–OG inter-
3.1 CGE formation using a QM cluster model action has an important TS-stabilizing effect. The nature of the
The cluster model employed in the present study includes 2H–OG interaction was further analyzed using a Second Order
sucrose, Asp477, Glu515 and three water molecules. After geo- Perturbation analysis of the Fock matrix in Natural Bond Orbital
metry optimization, a reactant-state system was obtained (NBO) basis, which detailed important interactions between
showing a series of stabilizing interactions between the donor Lewis-type and acceptor non-Lewis-type NBOs describing
sucrose and the amino acid residues. For example, the hydro- the quantum region (Table 1). The initial geometrical con-
gen bonds between the glucose that will be transferred and clusion about the relevance of the 2H–OG interaction on the
Asp477: (i) 6H–OD2 = 1.77 Å and (ii) 2H–OD2 = 1.52 Å; and stabilization of the oxocarbenium was thus confirmed through
one interaction between the leaving fructosyl group and this analysis. The results indicate a nOG ! σ*2OH interaction,

Org. Biomol. Chem. This journal is © The Royal Society of Chemistry 2018
View Article Online

Organic & Biomolecular Chemistry Paper


Published on 09 March 2018. Downloaded by Universidad de Concepcion on 20/03/2018 13:42:30.

Fig. 4 Snapshots of stationary points along the reaction path for the cluster model.

Table 1 Stabilization energy (ΔE) due to the interaction between the ΔΔE (51.5 kcal mol−1) corresponding to nOG ! σ*ðC1O5Þ high-
donor and acceptor orbitals obtained from Second Order Perturbation lights the relevance of stereoelectronic interactions on the
analysis of the QM cluster model. The ΔE values for reactant and TS
oxocarbenium stabilization. This interaction gains relevance at
structures are given in kcal mol−1. The ΔΔE is the difference between
the ΔE at the TS compared to that for the reactant
the TS even though the glycosidic bond (C1–OG) is already
broken (2.15 Å). It is interesting that although the mechanism
Reactant TS ΔΔE was found to be a concerted one, the proton transfer has not
occurred yet (HE2–OG = 1.65 Å). The reason underlying this
n(OE1)–σ*(1OH′) 10.7 29.3 18.6
n(OD1)–σ*(C1–O5) Missing 8.7 8.7 finding could be understood by means of 2nd-order pertur-
n(OD2)–σ*(6OH) 10.5 21.7 11.2 bation analysis, which shows the important stabilizing effect
σ(C2–H2)–σ*(C1–O5) Missing 15.5 15.5 of nOG ! σ*ðHE2OE2Þ (19.8 kcal mol−1), explaining that even
n(O5)–σ*(C1–O2) 3.9 8.9 5.0
n(OG)–σ*(OE2–HE2) 1.6 21.4 19.8 though the proton transfer has not occurred, there is an
n(OG)–σ*(2OH) Missing 16.0 16.0 important favorable effect of this interaction.
n(OG)–σ*(C1–O5) 8.6 60.1 51.5 The CGE intermediate shows a structure in which the
leaving fructosyl group has already been protonated by Glu515,
and Asp477 has attacked the glucosyl moiety. In addition, the
suggesting a strong hydrogen bond. The presumable stabilizing Cremer–Pople parameters (ϕ = 240° and θ = 79°) indicate that
effect of the hydrogen bonds formed by Asp477 and Glu515 glucose reaches an envelope ∼4E conformation with an energy
with sucrose (Fig. 4) is supported by the high calculated ΔΔE, reaction of −7.77 kcal mol−1, which is in agreement with other
for example, for nOD1 ! σ*ð6OHÞ or nOE1 ! σ *ð1OH′ Þ , which provide theoretical predictions.16 Thus both the proton transfer and
some of the strongest stabilizations for the TS of 11.2 and the nucleophilic attack seem to occur without an apparent
18.6 kcal mol−1, respectively. The important effect of the hydro- energy barrier, indicating that the energy barrier is mostly
gen bond between the OD1Asp477 and 6H of glucose is in agree- associated with the glycosidic bond breakage.
ment with what is known about the relevance of this interaction The calculated activation energy for the reaction at the
in other similar enzymes, which usually accounts for a stabiliz- BP86/6-31+G(d) level is 15.6 kcal mol−1, which increases to
ing effect at the TS of ∼5 kcal mol−1.16 Additionally, the highest 26.4 kcal mol−1 when single point energy calculations are

This journal is © The Royal Society of Chemistry 2018 Org. Biomol. Chem.
View Article Online

Paper Organic & Biomolecular Chemistry

carried out at the ωB97X-V/6-311++G(d,p) level. The first an antisymmetric stretching of C1–OG, HE2Glu515–OG and
method has been widely used due to its good geometrical per- OD1Asp477–C1 bonds, related to the breakage of the glycosidic
formance and computational efficiency; however it has been bond along with the proton transfer and the nucleophilic
found that it underestimates the energy barriers related to attack.
carbohydrate mechanisms.24,45 Thus, all the energy values dis- The results obtained for the full enzyme model revealed
cussed from now on are at the ωB97X-V/6-311++G(d,p) level, that the presence of the enzymatic environment decreases the
which has shown fewer errors regarding both reaction and acti- energy of the TS in about 9 kcal mol−1, compared to the
vation energies.31 The energy barrier obtained with this func- cluster model (Fig. 5). In addition, QM/MM calculations indi-
tional, 26.4 kcal mol−1, is rather high compared to experi- cate a concerted mechanism in which the glycosidic bond is
Published on 09 March 2018. Downloaded by Universidad de Concepcion on 20/03/2018 13:42:30.

mental activation energy for the related GTF-I enzyme46 being simultaneously broken as the proton transfer occurs, fol-
(∼16 kcal mol−1). Despite this high discrepancy, the minimal lowed by the nucleophilic attack, unlike cluster model results.
cluster model is very useful because it separates the pure The main structural and stereoelectronic factors that might be
chemistry of the process from the electrostatic stabilization of relevant for the stabilization of the TS and the conformations
the enzyme, something that would start mixing if we had used adopted by the glucosyl moiety during the reaction will be dis-
a larger model of the enzyme. The results of the full enzyme cussed next.
model will be discussed in the next section. The interactions established by sucrose in the reactant
structure (Fig. S2†) give rise to a conformation of glucose in
3.2 CGE formation using a QM/MM model which the calculated puckering parameters (ϕ = 150° and θ =
36°) indicate a moderately distorted structure that resembles
The minimal cluster model provided a first indication of the
the 2H3 conformation. Interestingly, this kind of sugar confor-
mechanism, revealing that the two catalytic residues that
mation was previously proposed for the glucose in GTF180ΔN
directly participate in the reaction, Asp477 and Glu515, are
and therefore, the finding of the 2H3 conformation in GTF-SI
also very important for stabilizing the fructosyl and glucosyl
comes to support this proposal.5
groups during the reaction. To further analyze the catalytic
The mechanistic calculations show that the reaction
strategies of GTF-SI, the reaction was studied using a full
starts with the lengthening of the C1–OG bond while Asp477
enzyme model. The reaction path obtained at the QM (BP86/6-
and Glu515 are approaching the glucose, which is in part
31+G(d))/MM(CHARMM27) (Fig. S1†), using the reaction co-
caused by the growing conformational distortion of the
ordinated (RC) previously defined, shows only one reaction
glucose. In the TS structure the C1–OG is 2.16 Å, the C1–
event for the formation of the CGE intermediate, which is
OD1Asp477 is 2.43 Å and the HE2–OG is 1.11 Å, which demon-
obtained after overcoming an energy barrier of ∼15 kcal mol−1.
strates that the glycosidic bond breakage and the proton trans-
Thus, based on this guess reaction path, the Minimum Energy
fer occur in a concerted way (Table 2 and Fig. 6). The confor-
Path (MEP) was built using the CPR algorithm. The energies
mation of glucose at the TS is closer to both the E3 and the
associated with the TS and the CGE structures were corrected 2
H3 (ϕ = 173° and θ = 43°). The oxocarbenium nature of the TS
at the ωB97X-V/6-311++(d,p) level (Fig. 5), obtaining an energy
is confirmed by the distance C1–O5, which has decreased from
barrier of 17.9 kcal mol−1, which is in very good agreement
1.38 Å to 1.29 Å (Table 2), evidencing a strong double-bond
with the experimental evidence.46
character; additionally, the NPA charges on the O5–C1–
Although the CPR algorithm guarantees the location of the
H1group of atoms are 0.26 a.u. higher than the calculated
true TS,37,38 this was further characterized by means of normal
charges for the reactant, which shows a decrease of the elec-
mode analysis showing one imaginary frequency of 188.9i cm−1
tron density on the anomeric carbon as reactants evolve to the
with an intensity of 245 kcal mol−1, which was associated with
TS structure.
The formation of a partial positive charge on the anomeric
center was confirmed by NBO analysis, which revealed a
decrease in the p-character of C1 and the formation of a new

Table 2 Selected distances d (Å), conformations, and change of posi-


tive charge at the anomeric center (Δq in a.u.) for the reactant, TS and
CGE intermediate

Reactant TS CGE

d(C1–OG) 1.51 2.16 2.82


d(C1–OD1Asp477) 3.09 2.43 1.49
d (C1–O5) 1.38 1.29 1.40
d(HE2–OG) 1.78 1.11 1.05
d(HE2–OE2) 1.01 1.37 1.55
Fig. 5 Potential energy profiles (kcal mol−1) that schematically describe Conformation ∼2H3 2
H3–E3 ∼4C1
the CGE formation for the cluster and the QM/MM models. Δq(O5–C1–H1) 0.00 0.26 −0.02

Org. Biomol. Chem. This journal is © The Royal Society of Chemistry 2018
View Article Online

Organic & Biomolecular Chemistry Paper


Published on 09 March 2018. Downloaded by Universidad de Concepcion on 20/03/2018 13:42:30.

Fig. 6 Snapshots of stationary points along the reaction path for the full enzyme model.

π bonding interaction between C1 and O5, in agreement with Table 3 Stabilization energy (ΔE) due to the interaction between the
the expected Lewis structure of an oxocarbenium ion charac- donor and acceptor orbitals obtained from Second Order Perturbation
analysis of the full enzyme model. The ΔE values for reactant and TS
terized by a rehybridization of C1 from sp3 to sp2.
structures are given in kcal mol−1. The ΔΔE corresponds to the differ-
In order to get deeper insight into the nature of the main ence between the ΔE at the TS compared to that for the reactant
interactions that contribute to the stabilization of the TS from
a molecular orbital point of view, a second-order perturbation Reactant TS ΔΔE
analysis was carried out as displayed in Table 3.
n(OE1)–σ*(1OH′) 7.1 8.9 1.8
Our results revealed that the highest ΔΔE is related to n(OG)–σ*(OE2–HE2) 18.9 Missing −18.9
nOE2 ! σ*ðOG HE2Þ interaction (108 kcal mol−1), revealing that n(OE2)–σ*(OG–HE2) Missing 108.0 108.0
Glu515 acts as a catalytic acid but also as a strong TS stabilizer. n(OD1)–σ*(C1–O5) 0.1 17.6 17.5
n(OD2)–σ*(6OH) 11.2 27.3 16.1
This interaction was not observed in the cluster model. The σ(C5–H5)–σ*(C1–O5) 1.2 14.8 13.6
great contribution of the nOE2 ! σ*ðOG HE2Þ . interaction in the n(O5)–σ*(C1–O2) 4.9 8.9 4.0
QM/MM model can be related to the large positive charge on n(OG)–σ*(C1–O5) 4.6 50.1 45.5
σ(C–OE2)–σ*(1OH′) 4.2 8.2 4.0
the HE2 atom (0.52 a.u.), which is located between two n(OG)–σ*(C1–C2) 8.4 9.8 1.4
oxygens with high electron density: the OE2 (−0.78 a.u.) and n(OG)–σ*(C4–C5) 7.5 10.2 2.7
the OG (−0.82 a.u.) atoms (Fig. 7B). These atoms are separated n(OG)–σ*(C5–C6) 7.8 9.6 1.8
by a distance of 2.48 Å (Fig. 7A), which is characteristic of
special short-strong hydrogen bonds.47 This kind of strong
interaction has not been described for other related GH- transferred proton as compared with GTF-SI and therefore, it
enzymes, such as members of GH13 or GH77 families. The is not able to establish such a short-strong hydrogen bond.16
only QM/MM study carried out with a similar enzyme of the Additionally, the positive charge of the Arg475 residue in
GH13 family, the human pancreatic α-amylase, showed that the active site can assist the proton transfer from Glu515 to
the acid residue homologue to Glu515 is farther from the the fructosyl leaving group by stabilizing the growing negative

This journal is © The Royal Society of Chemistry 2018 Org. Biomol. Chem.
View Article Online

Paper Organic & Biomolecular Chemistry


Published on 09 March 2018. Downloaded by Universidad de Concepcion on 20/03/2018 13:42:30.

Fig. 7 Active site view of the TS indicating: A, the interactions and residues that are important for TS stabilization and B, a short-strong hydrogen
bond depicted as a (red) high and (green) low electron density isosurface for the nOE2 ! σ*ðOG HE2Þ interaction.

charge on the glutamate moiety by means of long-range stabilizer; however this amino acid residue has also been pro-
electrostatic interactions (Fig. 7). Actually, the distance posed as a CGE stabilizer.4,12,15 Thus, the role of this residue
between Arg475 and Glu515 decreases from 4.32 to 4.12 Å as has not been confirmed yet. As can be seen in Fig. 7, Asp588 is
evolving from Reactant to TS. Thus, we propose that Arg475 interacting with sucrose through its glucosyl and fructosyl
plays a very important role to properly describe the concerted moiety along the whole reaction path and thus, it is expected
nature of this mechanism. to exert an influence during the mechanism. It is worth noting
Another important interaction for TS stabilization corres- that the interaction between glucose and Asp588 (2OH–
ponds to nOG ! σ*ðC1O5Þ (45.5 kcal mol−1), which was also OD2Asp588) is very strong showing an average distance of 1.55 Å
observed in the cluster model, thus highlighting the impor- over the reactant, TS and CGE structures. If Asp588 were not
tance of this stereoelectronic interaction in the oxocarbenium present in the active site, it is likely that the 2OH group will
stabilization. The hydrogen bonds formed between sucrose reorient to form another similar interaction. Actually, in the
and Asp477 and Glu515 are also relevant for the stabilization reactant of the cluster model the 2OH group was interacting
of the TS, which is supported by the ΔΔE related to the with Asp477 and then this group was reoriented to the OG
nOD1Asp477 ! σ*6ðHOÞ interaction (16.1 kcal mol−1), which is (opposite face of glucose) at the TS and CGE structures. This
5.0 kcal mol−1 stronger than the interaction observed in the suggests that Asp588 is important for maintaining the 2OH
cluster model. A relevant interaction observed in both models group correctly oriented during the reaction mechanism.
corresponds to nOD2Asp477 ! σ*C1O5 , which shows a two-fold Moreover, the closeness between Asp588 and the Glu515 (∼3 Å)
increase in the QM/MM model compared to the cluster could be especially important at the reactant state because it
system. To further analyze the role of Asp477 in the stabiliza- might increase the pKa of Glu515, which is necessary for cataly-
tion of oxocarbenium, this amino acid was removed from the sis. In order to examine whether Asp588 exerts a transfer charge
QM part of the QM/MM system, and single point calculations effect during the mechanism, Asp588 was included into the QM
were carried out for reactant and TS structures. The results part of the QM/MM system and the energy barrier and the reac-
showed an energy barrier increase in about 22 kcal mol−1, tion energy were recalculated.
which clearly indicates the key role of Asp477 in the stabiliza- Thus, the calculated energy barrier and reaction energy are
tion of the oxocarbenium ion-like TS. 16.4 and −0.88 kcal mol−1, respectively. This shows a stabiliz-
Interestingly, the σC5H5 ! σ*C1O5 interaction has a ΔΔE of ing effect of 1.5 and 3.1 kcal mol−1 for the TS and CGE,
13.6 kcal mol−1 and is not present in the cluster model. respectively, suggesting that transfer charge from Asp588 to
However, an analogue interaction between σC2H2 ! σ*C1O5 is sucrose could be important for both the TS and CGE stabiliz-
found in the reduced model (15.5 kcal mol−1). In both cases, ation. Finally, when Asp477 attacks the anomeric carbon the
the acceptor antibonding orbital is the σ*C1O5 , which suggests CGE intermediate is formed showing a distance of 1.49 Å for
that these two interactions could favor a slightly different dis- the C1–OD1Asp477 (Fig. 6). In the intermediate, the glucosyl
torted conformation of the glucosyl moiety, as is revealed by the moiety adopts a chair conformation (4C1), similarly to that
different conformational itineraries obtained from each model. observed for several GHs.44
The interactions established by Asp588 could be crucial for It is worth noting that the conformational itineraries of
the catalytic activity of GTF-SI since mutagenesis experiments glucose along the reaction path are not clear for members of
in which the homologue of Asp588 (Asp1136) in GTF180-ΔN GH13 neither for GH70 families;19,48 therefore, this is the first
was mutated resulted in a drastic decrease in activity.12 Due to study that reveals conformational features of the glucose during
this, both Asp1136 and Asp588 were proposed as a putative TS the catalytic process: ∼2H3 → 2H3–E3 → 4C1. This new knowl-

Org. Biomol. Chem. This journal is © The Royal Society of Chemistry 2018
View Article Online

Organic & Biomolecular Chemistry Paper

edge could be very valuable for the design of specific inhibitors Acknowledgements
of GTF-SI that could be employed for dental caries treatment.
The authors acknowledge financial support from Grant
FONDECYT no. 1150704.
4. Conclusion
Despite the number of mechanistic studies regarding the References
Glycosyl Hydrolase (GH) family, a number of these enzymes
have not been comprehensively addressed in terms of the 1 P. E. Petersen, S. Estupinan-Day and C. Ndiaye, Bull.
Published on 09 March 2018. Downloaded by Universidad de Concepcion on 20/03/2018 13:42:30.

specific details of their catalytic mechanisms. Among them is W. H. O., 2005, 83, 642–642.
GTF-SI, which is involved in the development of dental caries, 2 P. E. Petersen, D. Bourgeois, H. Ogawa, S. Estupinan-Day
and has been identified as a target for preventive treatments. and C. Ndiaye, Bull. W. H. O., 2005, 83, 661–669.
Thus, understanding the atomistic details of GTF-SI hydrolysis 3 R. A. Giacaman, P. M. Reyes and V. B. Leon, Braz. Oral Res.,
of sucrose is relevant for drug design. 2013, 27, 7–13.
In this work QM and QM/MM calculations were carried out 4 X. F. Meng, J. Gangoiti, Y. X. Bai, T. Pijning, S. S. Van
using a cluster and a full enzyme model built from the crystal Leeuwen and L. Dijkhuizen, Cell. Mol. Life Sci., 2016, 73,
structure of GTF-SI in a complex with sucrose. The TS was 2681–2706.
found to be strongly stabilized by the hydrogen bond 5 H. Leemhuis, T. Pijning, J. M. Dobruchowska, S. S. van
OG⋯HE2⋯OE2, which according to the distance OG⋯OE2, Leeuwen, S. Kralj, B. W. Dijkstra and L. Dijkhuizen,
charges and 2nd-order perturbation analysis can be considered J. Biotechnol., 2013, 163, 250–272.
as a short-strong hydrogen bond. This strong stabilizing inter- 6 H. Leemhuis, T. Pijning, J. M. Dobruchowska,
action seems to be a new insight into the catalytic strategies B. W. Dijkstra and L. Dijkhuizen, Biocatal. Biotransform.,
used for this enzyme since it has not been observed in other 2012, 30, 366–376.
members of related-GH families. 7 Z. Ren, L. L. Chen, J. Y. Li and Y. Q. Li, J. Oral Microbiol.,
As for other GH enzymes, the nucleophile (Asp477 in 2016, 8, 31095.
GTF-SI) is able to provide important stabilization for the TS; in 8 L. Petersen, A. Ardevol, C. Rovira and P. J. Reilly, J. Am.
this case the interaction is established between Asp477 and Chem. Soc., 2010, 132, 8291–8300.
the 6OH group of the glucosyl moiety. Besides, Arg475 likely 9 X. Biarnes, A. Ardevol, J. Iglesias-Fernandez, A. Planas and
assists the proton transfer stabilizing the nascent negative C. Rovira, J. Am. Chem. Soc., 2011, 133, 20301–20309.
charge on the glutamate moiety. Another important active site 10 L. Raich, V. Borodkin, W. X. Fang, J. Castro-Lopez,
residue is Asp588, which interacts with sucrose through its glu- D. M. F. van Aalten, R. Hurtado-Guerrero and C. Rovira,
cosyl (2OH group) and fructosyl moiety (1OH′) along the whole J. Am. Chem. Soc., 2016, 138, 3325–3332.
reaction path. Asp588 seems to be crucial for maintaining the 11 B. Henrissat and G. Davies, Curr. Opin. Struct. Biol., 1997,
2OH group suitably oriented during the reaction mechanism, 7, 637–644.
exerting a stabilizer role not only in the TS, as has been pro- 12 A. Vujicic-Zagar, T. Pijning, S. Kralj, C. A. Lopez,
posed, but also in the CGE intermediate. Actually, the W. Eeuwema, L. Dijkhuizen and B. W. Dijkstra, Proc. Natl.
inclusion of Asp588 in the QM region results in a decrease of Acad. Sci. U. S. A., 2010, 107, 21406–21411.
the activation and reaction energies down to 16.4 and 13 V. Monchois, R. M. Willemot and P. Monsan, FEMS
−0.88 kcal mol−1, respectively. Microbiol. Rev., 1999, 23, 131–151.
We consider that this new information provides relevant 14 R. Ohrlein, Biocatalysis - from Discovery to Application, 1999,
new knowledge about specific enzymatic catalysis aspects vol. 200, pp. 227–254.
with implications not only for GTF-SI but also for the whole 15 K. Ito, S. Ito, T. Shimamura, S. Weyand, Y. Kawarasaki,
GH70 family. Thus, we propose for the first time a confor- T. Misaka, K. Abe, T. Kobayashi, A. D. Cameron and
mational itinerary for the glucosyl moiety of sucrose: ∼2H3 → S. Iwata, J. Mol. Biol., 2011, 408, 177–186.
2
H3–E3 → 4C1. 16 G. P. Pinto, N. F. Bras, M. A. S. Perez, P. A. Fernandes,
We hope that these key structural and stereoelectronic fea- N. Russo, M. J. Ramos and M. Toscano, J. Chem. Theory
tures of the TS might be considered in the future design of Comput., 2015, 11, 2508–2516.
specific inhibitors for GTF-SI. Upcoming efforts will be 17 R. I. Sadreyev, B. H. Kim and N. V. Grishin, Curr. Opin.
focused on unraveling the factors that modulate the specificity Struct. Biol., 2009, 19, 321–328.
(α-1,3 or α-1,6 linkage) of the new glycosidic bond that is 18 M. Kitamura, M. Okuyama, F. Tanzawa, H. Mori, Y. Kitago,
synthesized by GTF-SI. N. Watanabe, A. Kimura, I. Tanaka and M. Yao, J. Biol.
Chem., 2008, 283, 36328–36337.
19 G. J. Davies, A. Planas and C. Rovira, Acc. Chem. Res., 2012,
Conflicts of interest 45, 308–316.
20 E. Roberts, J. Eargle, D. Wright and Z. Luthey-Schulten,
There are no conflicts to declare. BMC Bioinf., 2006, 7, 382.

This journal is © The Royal Society of Chemistry 2018 Org. Biomol. Chem.
View Article Online

Paper Organic & Biomolecular Chemistry

21 H. Li, A. D. Robertson and J. H. Jensen, Proteins: Struct., S. Ha, D. Joseph-McCarthy, L. Kuchnir, K. Kuczera,
Funct., Bioinf., 2005, 61, 704–721. F. T. K. Lau, C. Mattos, S. Michnick, T. Ngo, D. T. Nguyen,
22 J. S. Chia, C. S. Yang and J. Y. Chen, Infect. Immun., 1998, B. Prodhom, W. E. Reiher, B. Roux, M. Schlenkrich,
66, 4797–4803. J. C. Smith, R. Stote, J. Straub, M. Watanabe,
23 T. J. Dolinsky, J. E. Nielsen, J. A. McCammon and J. Wiorkiewicz-Kuczera, D. Yin and M. Karplus, J. Phys.
N. A. Baker, Nucleic Acids Res., 2004, 32, W665–W667. Chem. B, 1998, 102, 3586–3616.
24 A. T. Pereira, A. J. M. Ribeiro, P. A. Fernandes and 34 M. Karplus, J. Comput. Chem., 1983, 4, 187–217.
M. J. Ramos, Int. J. Quantum Chem., 2017, 117, e25409. 35 B. R. Brooks, C. L. Brooks, A. D. Mackerell, L. Nilsson,
25 Y. Shao, L. F. Molnar, Y. Jung, J. Kussmann, C. Ochsenfeld, R. J. Petrella, B. Roux, Y. Won, G. Archontis, C. Bartels,
Published on 09 March 2018. Downloaded by Universidad de Concepcion on 20/03/2018 13:42:30.

S. T. Brown, A. T. B. Gilbert, L. V. Slipchenko, S. Boresch, A. Caflisch, L. Caves, Q. Cui, A. R. Dinner,


S. V. Levchenko, D. P. O’Neill, R. A. DiStasio Jr., M. Feig, S. Fischer, J. Gao, M. Hodoscek, W. Im,
R. C. Lochan, T. Wang, G. J. O. Beran, N. A. Besley, K. Kuczera, T. Lazaridis, J. Ma, V. Ovchinnikov, E. Paci,
J. M. Herbert, C. Yeh Lin, T. Van Voorhis, S. Hung Chien, R. W. Pastor, C. B. Post, J. Z. Pu, M. Schaefer, B. Tidor,
A. Sodt, R. P. Steele, V. A. Rassolov, P. E. Maslen, R. M. Venable, H. L. Woodcock, X. Wu, W. Yang,
P. P. Korambath, R. D. Adamson, B. Austin, J. Baker, D. M. York and M. Karplus, J. Comput. Chem., 2009, 30,
E. F. C. Byrd, H. Dachsel, R. J. Doerksen, A. Dreuw, 1545–1614.
B. D. Dunietz, A. D. Dutoi, T. R. Furlani, S. R. Gwaltney, 36 A. R. Calixto, M. J. Ramos and P. A. Fernandes, J. Chem.
A. Heyden, S. Hirata, C.-P. Hsu, G. Kedziora, Theory Comput., 2017, 13, 5486–5495.
R. Z. Khalliulin, P. Klunzinger, A. M. Lee, M. S. Lee, 37 S. Fischer and M. Karplus, Chem. Phys. Lett., 1992, 194,
W. Liang, I. Lotan, N. Nair, B. Peters, E. I. Proynov, 252–261.
P. A. Pieniazek, Y. Min Rhee, J. Ritchie, E. Rosta, C. David 38 F. A. Kiani and S. Fischer, Proc. Natl. Acad. Sci. U. S. A.,
Sherrill, A. C. Simmonett, J. E. Subotnik, H. Lee Woodcock 2014, 111, E2947–E2956.
Iii, W. Zhang, A. T. Bell, A. K. Chakraborty, D. M. Chipman, 39 J. Xiong and D. G. Xu, J. Phys. Chem. B, 2017, 121, 931–941.
F. J. Keil, A. Warshel, W. J. Hehre, H. F. Schaefer Iii, 40 H. Batebi and P. Imhof, Theor. Chem. Acc., 2016, 135, 262.
J. Kong, A. I. Krylov, P. M. W. Gill and M. Head-Gordon, 41 M. Culka, F. J. Gisdon and G. M. Ullmann, in Structural
Phys. Chem. Chem. Phys., 2006, 8, 3172–3191. and Mechanistic Enzymology, ed. T. KarabenchevaChristova,
26 A. I. Krylov and P. M. W. Gill, Wiley Interdiscip. Rev.: Elsevier Academic Press Inc, San Diego, 2017, vol. 109, pp.
Comput. Mol. Sci., 2013, 3, 317–326. 77–112.
27 A. Behn, P. M. Zimmerman, A. T. Bell and M. Head- 42 C. Diederich, M. Leypold, M. Culka, H. Weber,
Gordon, J. Chem. Phys., 2011, 135, 224108. R. Breinbauer, G. M. Ullmann and W. Blankenfeldt, Sci.
28 B. A. Murtagh and R. W. H. Sargent, Comput. J., 1970, 13, Rep., 2017, 7, 6272.
185–194. 43 D. Cremer and J. A. Pople, J. Am. Chem. Soc., 1975, 97,
29 M. J. D. Powell, Math. Program., 1971, 1, 26–57. 1354–1358.
30 N. Mardirossian and M. Head-Gordon, Phys. Chem. Chem. 44 A. Ardèvol and C. Rovira, J. Am. Chem. Soc., 2015, 137,
Phys., 2014, 16, 9904–9924. 7528–7547.
31 S. Dasgupta and J. M. Herbert, J. Comput. Chem., 2017, 38, 45 J. Kóňa and I. Tvaroška, Chem. Pap., 2009, 63, 598–607.
869–882. 46 K. S. Devulpalle, S. D. Goodman, Q. Gao, A. Hemsley and
32 E. D. Glendening, J. K. Badenhoop, A. E. Reed, G. Mooser, Protein Sci., 1997, 6, 2489–2493.
J. E. Carpenter, J. A. Bohmann, C. M. Morales and 47 E. V. Anslyn and D. A. Dougherty, in Modern Physical
F. Weinhold, NBO 5.0, Theoretical Chemistry Institute, Organic Chemistry, ed. J. Murdzek, 2004, ch. 3.2, binding
W. Madison, 2001; http://www.chem.wisc.edu/~nbo5. forces, p. 177.
33 A. D. MacKerell, D. Bashford, M. Bellott, R. L. Dunbrack, 48 A. Ardèvol, J. Iglesias-Fernandez, V. Rojas-Cervellera and
J. D. Evanseck, M. J. Field, S. Fischer, J. Gao, H. Guo, C. Rovira, Biochem. Soc. Trans., 2016, 44, 51–60.

Org. Biomol. Chem. This journal is © The Royal Society of Chemistry 2018

View publication stats

You might also like