Chakraborty2019 iceAdsorptionAntifreezeProtein

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

PCCP

View Article Online


PAPER View Journal | View Issue

Ordered hydration layer mediated ice


adsorption of a globular antifreeze protein:
Published on 14 August 2019. Downloaded by University of Perugia on 5/22/2023 8:46:37 AM.

Cite this: Phys. Chem. Chem. Phys.,


2019, 21, 19298 mechanistic insight
Sandipan Chakraborty and Biman Jana *

The ice/water interface recognition mechanism of antifreeze proteins (AFPs) is highly contentious.
Conventionally, protein adsorption on a solid surface is primarily driven by the polar interactions
between the hydrophilic residues of the protein and interfacial water of the solid surface. Ice surface
recognition by a type III AFP is surprising in this context where the ice binding surface (IBS) is
hydrophobic. The present study provides molecular insight into the unusual interface recognition
phenomenon of a type III AFP (QAE isoform) from Macrozoarces americanus. Potential of mean force
calculations show that the type III AFP adsorbs on the ice surface mediated through a layer of ordered
water. Molecular dynamics simulations at lower than ambient temperature reveal that the flat
hydrophobic IBS induces ordering of water. The excellent geometrical synergy between the hydration
water structure around the IBS and water arrangements on the pyramidal surface favours adsorption on
the pyramidal plane. Mutations that interrupt the hydration shell water ordering essentially lead to less
efficient adsorption, which greatly reduces the anti-freezing activity of the AFP. Binding free energy
calculations of the wild-type and several mutant AFPs reveal that the binding affinity is linearly correlated
Received 3rd June 2019, with the experimentally observed thermal hysteresis activity. Therefore, binding to a specific ice plane
Accepted 14th August 2019 with considerable affinity is the dictating factor of the anti-freeze activity for a type III AFP. Mechanistic
DOI: 10.1039/c9cp03135a insights into the ice binding process of the wild-type and different mutant AFPs obtained from this study
pave the way for rational design of type III variants with much improved activity, which possesses ample
rsc.li/pccp industrial applicability, particularly in cryo-preservation.

1. Introduction the last two decades.14–18 Several models have been proposed.16–19
In earlier work of protein adsorption on a polyethylene oxide
Understanding the mechanism of protein and peptide adsorption surface, Jeon et al. proposed diffusion mediated protein adsorp-
on solid/liquid interfaces is a subject of recent interest due tion where van der Waals attraction plays an important role.17
to its immense applications in materials science, medicine, Later, Szleifer et al. proposed a theoretical framework for protein
and technology.1–4 Surface adsorption plays a pivotal role adsorption on surfaces based on generalized single-chain mean
in material design and application, bio-mineralization and field theory where the surface coverage, the concentration of the
biosensor design.3,5–8 The structure and activity of proteins, bulk protein and the nature of the solvent dictate the adsorption
particularly enzymes, on solid surfaces have been studied isotherm.18 Norde et al. on the other hand highlighted that the
extensively to understand the effect of surface adsorption on balance between conformational entropy and specific interactions
the biological function of the proteins.6,9–12 Notably the fate of governs protein adsorption on surfaces.20 All these models explain
nanoparticles in vivo depends strongly on their interactions the surface adsorption at the macroscopic level and the process of
with different body proteins. Thus understanding the mechanism surface recognition of proteins at the molecular level is still poorly
of different interface recognition by proteins emerges as a topic of understood. Recent work of Penna et al.14 and Yu et al.15 put
intense research interest. Although efforts were started in the forward a three-step model where in the first step there is a biased
1960s by Vroman et al.,13 exploration of molecular understanding diffusion of the protein towards the interface, which is followed by
of the process of adsorption has gained tremendous attention in its anchoring in the second hydration shell of the surface using
the hydrophilic surface residues and then popping into the first
School of Chemical Sciences, Indian Association for the Cultivation of Science,
hydration shell of the solid surface. Finally, the protein rearranges
Jadavpur, Kolkata-700032, India. E-mail: pcbj@iacs.res.in; Fax: +91 33 2473 2805; itself to lock down on the solid surface by optimizing the protein–
Tel: +91 33 2473 4971 solid surface interactions. The roles of polar side chains are

19298 | Phys. Chem. Chem. Phys., 2019, 21, 19298--19310 This journal is © the Owner Societies 2019
View Article Online

Paper PCCP

crucial in this three-step model of protein adsorption. However, threonine repeats. Moreover, the identified IBS is hydrophobic
type III antifreeze proteins (AFPs) are a surprise in this context. in nature.21,55,56 The binding plane specificity of type III AFPs is
This class of AFP uses a hydrophobic surface to recognize the another controversial area of research. Initially, prism plane
ice/water interface, which is essentially polar in nature.21 binding was proposed, based on ice etching experiments.55,57
AFPs are evolutionarily selected for ice/water interface recogni- Later, Antson et al. reported modified ice etching data for type
tion, which is required for the survival of living organisms in III AFPs which indicated its binding to different ice planes that
the cold Antarctic winter. These bio-macromolecules have been are parallel or inclined at a small angle to the c-axis.58 Thus type
isolated from Antarctic fish, insects, bacteria, yeast and even III AFPs are thought to be a non-basal multiple plane binder.
plants.22–31 AFPs are commonly classified as fish type I, II, III, A previous simulation study revealed slow water dynamics
and IV, insect AFPs and antifreeze glycoproteins (AFGPs). around the IBS of type III AFP59 and even the presence of
Published on 14 August 2019. Downloaded by University of Perugia on 5/22/2023 8:46:37 AM.

Different classes of AFPs bear no structural similarities and ice-like ordered hydration water within the first hydration shell
they are evolutionary independent from each other.32 was also reported.60,61 The presence of such an ordered water
The mechanism of ice recognition by AFPs is highly population was also evident from vibrational sum frequency
controversial. The ice binding surface (IBS) of type I and insect generation spectroscopy.62 Grabowska et al.63 and Xu et al.64
AFPs contains an array of polar residues. Their repeat distances hinted that both long range solvent perturbation and short
match the oxygen atom repeat distances of the particular ice range local interactions determine the hysteresis activity.
plane where they specifically bind.33,34 This observation is the Mahatabuddin et al. recently reported the crystal structure of
origin of the first proposed mechanism, known as the ‘‘hydrogen an improved activity variant of a type III AFP from notched-fin
bond matching hypothesis’’.33 However, the hypothesis was later eelpout (A20I mutant) where a polypentagonal ice-like water
rejected when mutational data on type I AFPs were available. network is clearly evident on the IBS of the AFP, which binds to
Substitution of equidistant Thr residues at the IBS with Ser the pyramidal binding plane.65 However, Sun et al. using fusion
(similar hydrogen bonding ability) leads to complete loss of protein crystallography showed that the water layer around the
activity; however the AFP still remains B30% active upon IBS of a type III AFP which binds to the pyramidal plane is
replacement of Thr with Val, which does not possess hydrogen sparse and does not form a robust network.66 Notably, crystal
bonding side chains.22 Later, a surface complementarity matching water molecules are strongly influenced by crystal packing
hypothesis has been proposed motivated by the observation artifacts and the conditions of the experiments (temperature,
that many AFPs possess a flat IBS and substitution of the IBS buffer, salt concentrations). Also, a pentagonal ice-like water
residues by amino acids with bulkier side chains diminishes structure has been reported in many other non-antifreeze proteins
the activity.21,35,36 However, this mechanism is also highly too, like crambin,67 bacterial malate dehydrogenase,68 etc. There-
criticized since the ice/water interface is not a rigid surface to fore, the prospects of hydration water mediated ice recognition for
provide enough strength for AFP adsorption, which is known to type III AFPs remain controversial, particularly due to the fact that
be irreversible in nature.37 the structure of type III AFPs bound to an ice surface in the
Recently, AFPs have come into prominence due to their presence of an ice/water interface is yet to be characterized. Thus
unusual ability to control hydration water dynamics, particularly molecular-scale insight into the process of adsorption of a type III
around the IBS.38–41 Initially, it has been argued that an AFP AFP on an ice surface still remains elusive.
induces long range perturbations to water dynamics which alter The present work for the first time provides the atomistic
the local ice growth kinetics at the ice/water interface40,42 and details of the ice binding mechanism of a type III AFP on
even induce ice melting.43,44 However, microfluidic experiments different ice planes using equilibrium simulations and free
and fluorescence microscopy clearly showed that the AFPs energy calculations. The structure and thermodynamic aspects
adhere to the ice surface10,37,45–47 and the binding is nearly of type III AFP adsorption on the ice surface have been critically
irreversible in nature.45 Later it has been demonstrated that explored. Our objectives are to explore in critical detail how the
both short-range interactions and long-range protein induced hydrophobic protein surface adsorbs on the ice/water interface
water dynamics dictate the antifreeze activity, at least for the and the role of hydrophobic hydration in the process of
insect AFPs.41 However, atomistic insight into the process of adsorption. We have further applied our proposed adsorption
adsorption is largely lacking. Very recently, we have proposed a model to explain the experimentally observed activity data of
clathrate mediated ice adsorption mechanism for type I48 and several type III AFP mutants.
insect AFPs.49–53 According to the mechanism, the IBS first
adheres to the ice/water interface mediated through the ordered
hydration layer and there is formation of ordered water cages 2. Methods
around the IBS. In insect AFPs, the presence of threonine
repeats at the IBS stabilizes a part of the water cages, known All the molecular dynamics simulations and free energy calcula-
as ‘‘anchored clathrate water’’ (ACW),54 which forms dual tions were performed using the GROMACS package69–71 in
hydrogen bonds involving both the AFP and ice surface, thus combination with the OPLS/AA72–74 force field and SPC/E water
anchoring the AFP on the ice surface. model.75 Validations of the water model and force field in low
Type III AFPs belong to the category of globular proteins. temperature simulations of the AFPs were discussed in detail in
Unlike the type I and insect AFPs, type III AFPs do not have any our recent publications.48,49 Previously, Bryk et al. demonstrated

This journal is © the Owner Societies 2019 Phys. Chem. Chem. Phys., 2019, 21, 19298--19310 | 19299
View Article Online

PCCP Paper

that the basal and prism plane of ice interfaces can be appropriately translational movement of the protein was restricted. The
mimicked by SPC/E water at a temperature of 225  5 K. Within trajectory was saved at every 1 ps. In the case of the four
this range, the density, translational, orientational and dynamic different mutants, similar protocols were followed during
order parameters smoothly and continuously change from the the 50 ns simulation for each system in the NPT ensemble.
crystal to the liquid across the interface.76 All the visualizations All the simulations for the mutant AFPs were performed only at
were carried out using the Visual Molecular Dynamics (VMD) a single temperature (225 K) at which the PMF calculations
package.77 were carried out.
All the analyses were performed using the available analysis
2.1. Preparation of the systems tools in GROMACS. For RMSD calculations, structural alignment
The structure of the type III AFP from Macrozoarces americanus was carried out on Ca atoms and the RMSD was calculated on all
Published on 14 August 2019. Downloaded by University of Perugia on 5/22/2023 8:46:37 AM.

was obtained from the protein data bank (PDB ID: 1MSI).55 heavy atoms of the protein from the simulation. The RMSF was
Residue protonation states were assigned according to neutral calculated on all heavy atoms of the protein from the simulation.
pH. All the heteroatoms and crystallographic water molecules
were not considered in the simulation. Different AFP mutants 2.3. Molecular docking of AFPs with ice planes
(A16M, L19A//V41A, N14S, and T18N) were constructed by The energy minimized wild-type AFP was initially docked to
mutating a selected residue of the wild-type protein to the three different ice planes (basal, prism and pyramidal planes)
required residue at the particular position. An appropriate side using a rigid body docking procedure implemented in the
chain rotamer was selected from the rotamer library where PatchDock webserver.80 Docking generates many possible
minimum steric clashes were observed. docked AFP–ice complexes. The possible docked complex where
the experimentally known binding sites oriented towards the
2.2. Equilibrium simulations of wild-type and mutant type III ice surface was chosen. In the case of mutant AFPs, adsorption
AFPs was considered only on the pyramidal plane. The wild-type
Wild-type and different mutant AFPs were energy minimized AFP–pyramidal ice complex obtained from molecular docking
in vacuo using the steepest descent algorithm. Each system was was selectively mutated at the particular IBS residue to generate
then immersed in a cubic box containing SPC/E water in such a mutant AFP–pyramidal ice complexes with the aid of mutagenesis
way that the minimum distance between the protein atoms and tools implemented in the VMD package.77
box edges was greater than 10 Å. Minimizations in aqueous
environments were then performed for all the systems using 2.4. Calculations of potential of mean force (PMF) for
another 500 steps of energy minimization using the steepest wild-type and mutant AFP adsorption on ice planes
descent algorithm. Then a 1 ns position restraint simulation For the wild-type AFP, the free energy of binding was calculated
was performed for each AFP in the isothermal–isobaric (NPT) for basal, prism and pyramidal plane adsorption. While for the
ensemble. Protein backbone atoms were restrained using a mutant systems, the free energy of binding was calculated for
harmonic potential with a force constant of 1000 kJ mol1 nm2 the pyramidal plane only. AFP–ice docked complexes were
in all three dimensions, however, water molecules were allowed to considered initially. Each docked complex was prepared and
move freely. All the simulations were performed under periodic equilibrated. Briefly, each system was aligned in such a way
boundary conditions. that the ice slab was on the X–Y plane and the width of the slab
To study the conformational dynamics of the wild-type and was along the Z-axis. Then the system was solvated in a
mutant AFPs, we performed equilibrium simulations at 298 K. rectangular box containing SPC/E water. The box dimension
The temperature was kept constant by coupling to a thermostat was chosen in such a way that in the X–Y plane it did not collide
using the V-rescale algorithm78 with a time constant for with its image and along the Z-axis the box distance was greater
coupling set to 0.1 ps. The pressure was kept constant at 1 bar than double the final pull distance. For wild type AFP adsorp-
by employing the isotropic Parrinello–Rahman barostat with a tion on basal and prism planes the box dimensions were
time-constant of 2 ps. Electrostatic interactions were calculated 80  80  100 Å3 and for wild-type and mutant AFP adsorption
using the particle mesh Ewald summation (PME) method79 with on the pyramidal plane the box dimensions were chosen as
default values for the grid spacing and a 10 Å distance cut-off 100  80  100 Å3. All the simulations were performed under
was used to cut short-range electrostatic interactions. Finally, a periodic boundary conditions. All the systems were initially
500 ns production simulation was carried out for each system energy minimized using the steepest descent algorithm followed
in the NPT ensemble using the same parameters and the by 1 ns position restraint dynamics in the NVT ensemble at
coordinates of the simulated systems were saved every 20 ps. 225 K. During the simulation the ice slab was frozen. The
For hydration study, the energy minimized solvated struc- temperature was kept constant by coupling to a thermostat
ture of wild-type AFP was subjected to 1 ns position restrained using the V-rescale algorithm with a time constant for coupling
equilibrations at four different temperatures (210 K, 225 K, set to 0.1 ps. Electrostatic interactions were calculated using the
250 K, and 298 K) in the NPT ensemble. The same parameters PME method. Then finally a 3 ns NVT simulation was carried out
for temperature and pressure coupling were used. Final 50 ns for each system at 225 K where both the proteins and water
production runs were carried out in the NPT ensemble at molecules were allowed to move freely, while the ice slab was
four different temperatures. During the production run, the kept frozen and the simulation parameters remain unaltered.

19300 | Phys. Chem. Chem. Phys., 2019, 21, 19298--19310 This journal is © the Owner Societies 2019
View Article Online

Paper PCCP

The umbrella sampling technique was applied to construct


the potential of mean force (PMF) curve for wild-type and
mutant AFP adsorption to the ice surface. The AFP and ice slab
center of mass distance along the Z-direction was considered as
the reaction coordinate. The AFP was then pulled from the
respective ice plane along the Z-direction using the umbrella
sampling technique. The interval between two successive
windows was 1 Å along the reaction coordinate. In each
window, a force constant of 2000 kJ mol1 nm2 was applied
to maintain the required reaction coordinate value using an
Published on 14 August 2019. Downloaded by University of Perugia on 5/22/2023 8:46:37 AM.

umbrella potential. In each umbrella window, a 1 ns equili-


bration simulation and a 5 ns production run were performed
in the NVT ensemble. All the simulation parameters remained
the same as mentioned in the equilibration process. Convergence
of the simulation in each window was confirmed by performing Fig. 1 Structure of the type III AFP from Macrozoarces americanus shown
block averaging. The weighted histogram analysis method as a cyan cartoon (A). Regions of the IBS are shown in a blue transparent
surface representation. The rest of the protein is considered as the non-
(g_wham) was then used to construct the PMF profiles.81
IBS. (B) The electrostatic surface potential of the type III AFP is shown.
Sufficient overlaps among the distributions of reaction coordinates Regions of positive and negative potentials are shown in blue and red
were confirmed by histogram analysis. colours, respectively. The IBS with sparse negative electrostatic potential is
shown within the circle.
2.5. Equilibrium simulation of wild-type and mutant AFP–ice
complexes
Structures of the AFP–ice complexes corresponding to the PMF around the AFP. The melting point of the SPC/E water model is
minima were further simulated in the NVT ensemble at 225 K B215 K. Therefore, we have simulated the AFP at four different
for 30 ns to characterize the adsorbed state. All the simulation temperatures (210 K, 225 K, 250 K, and 298 K) and analysed the
parameters remained unaltered as mentioned above. For all the water ordering within the first hydration shell (Fig. 2).
simulations the ice slab was kept frozen in all three dimensions. As is evident from the figure, there is a preferential accumu-
lation of water around the IBS compared to the non-IBS plane at
all temperatures. The peak appears at B3.7 Å, which is larger
3. Results and discussion than the hydrogen bond donor–acceptor distance limit (3 Å),
which reinforces the fact that these water molecules are not
In this work, we have studied the molecular mechanism of ice hydrogen bonded to the IBS planes. We have further analysed
recognition of a type III AFP (QAE isoform) from Macrozoarces the ordering of these water molecules using the O–O–O angle
americanus. We have considered the ice binding surface (IBS) distribution. This order parameter is a highly sensitive one that
based on previously reported experimental mutation data and can differentiate between interstitial and tetrahedrally oriented
initial docking studies.21,55 Garnham et al. also showed that water.83,84 The distribution is characterized by two peaks. The
there are two adjacent ice recognition sites aligned with an low angle peak B601 is related to the interstitial water and the
angle B1501.82 Here, we have considered all the residues of peak at B100–1041 signifies tetrahedrally ordered hydrogen
both recognition sites to define the IBS. Residues 9, 12–16, 18– bonded water. The most common arrangement of hydrogen
20, 41, 44 and 51 are considered to be involved in the IBS.21,55,82 bonded water molecules around an oxygen atom of a central
Other residues of the protein are considered as the non-IBS water is nearly tetrahedral. Due to the dynamic nature of liquid
plane. A pictorial representation of the IBS and non-IBS is water, hydrogen bonds vary in distance as well as in symmetry,
shown in Fig. 1A. Unlike other classes of AFP, the IBS of type which resulted in the formation of transient irregular tetrahedral
III AFPs is relatively flat and electrostatic surface potential arrangements. A representative tetrahedral arrangement of water
calculations reveal that the IBS is hydrophobic in comparison is shown in the inset of Fig. 2. However, in the case of ice, the
to the non-IBS plane (Fig. 1B). There is a patch of weak perfect ordering of water molecules in the crystal lattice makes
negatively charged surface at the tip of the IBS indicated by the hydrogen bonds highly symmetrical, which results in the
the circle in the figure. This is primarily contributed by Asn14, formation of a perfect tetrahedron. Room temperature liquid
Thr15, Thr18, and Gln44. water exhibits the peak of the O–O–O angle distribution of
tetrahedral water at B991. A shift of the peak maximum towards
3.1. Type III AFP has the ability to induce water ordering higher angle signifies more tetrahedral ordering and an increase
around the IBS in magnitude signifies an increase in the population of such
We have investigated the water ordering around both the IBS ordered water. We have noticed that the higher angle peak
and non-IBS and its variation upon lowering the temperature. maximum of the O–O–O angle distribution of the first hydration
AFPs are active around the freezing temperature of water. shell water around the IBS progressively shifts toward a higher
We have used the SPC/E water model to mimic the hydration value and its magnitude increases with a decrease in temperature

This journal is © the Owner Societies 2019 Phys. Chem. Chem. Phys., 2019, 21, 19298--19310 | 19301
View Article Online

PCCP Paper
Published on 14 August 2019. Downloaded by University of Perugia on 5/22/2023 8:46:37 AM.

Fig. 2 Analysis of hydration water at four different temperatures around the IBS and non-IBS (nIBS) planes of the type III AFP. IBS and non-IBS planes are
represented in black and red colours respectively. Radial distributions of water (oxygen) molecules around different planes of the AFP at four different
temperatures are shown and the corresponding O–O–O angle distributions among water molecules within the first hydration shell (within 5.5 Å) around
each surface of the protein are also shown in the inset. A representative tetrahedral arrangement of water is shown in the inset.

compared to the non-IBS plane. This trend clearly suggests that


upon lowering the temperature water molecules around the IBS
adopt a more ordered structure. The population of tetrahedrally
ordered water is higher around the IBS compared to the non-IBS
plane at all the temperatures, which is particularly pronounced in
low temperature regions. Noticeably, the higher angle peak is
shifted by 21 in the case of IBS compared to the non-IBS plane,
which is more pronounced at low temperatures and is related
to the hydrophobic hydration of the IBS in the low temperature
region.
The water structure around the IBS obtained at 30 ns of
equilibrium simulation at 210 K is shown in Fig. 3. The water
molecules within the first hydration shell of the IBS are highly
ordered and there is the formation of water cages stabilized
through a highly cooperative hydrogen bond network (Fig. 3).
Notably, Meister et al. also noticed water ordering in the
aqueous solution of a type III AFP, even above the freezing
temperature, by using femtosecond sum frequency generation
spectroscopy.62

3.2. Type III AFP preferentially adsorbs on the pyramidal Fig. 3 Structure of the water shell around the IBS of the type III AFP.
plane mediated through a layer of water The IBS is shown as a blue transparent surface and the protein is rendered
as a cyan cartoon. Oxygen atoms of the hydration layer water are rendered
We have performed potential of mean force calculations using in a red vdW mode and the hydrogen bond between them is shown as a
umbrella sampling techniques to delineate the binding plane red line.

19302 | Phys. Chem. Chem. Phys., 2019, 21, 19298--19310 This journal is © the Owner Societies 2019
View Article Online

Paper PCCP

specificities of this class of AFP. We have considered three correlates with the binding affinity data obtained from the PMF
different ice planes in this study, namely basal, prism and calculations. We have also calculated the number of direct
pyramidal planes. hydrogen bonds between the AFP and the ice surface in three
Fig. 4A reveals that the type III AFP preferentially binds to different AFP–ice complexes from the simulation trajectories and
the pyramidal plane and the order of binding affinities towards the results are shown in Fig. 4C. As is evident from the figure, the
different ice planes is pyramidal plane 4 prism plane 4 basal basal plane has the highest ability to form direct hydrogen bonds
plane. Thus the type III AFP is essentially a non-basal ice plane with the IBS and the pyramidal plane has the least affinity to form
binder, which is in line with experimental evidence.58 We have direct hydrogen bonds with the IBS. The observed order of the
further characterized the nature of the AFP–ice complexes by number of hydrogen bonds between the IBS and three different
performing equilibrium simulations of the complexes corres- ice planes is exactly opposite to the binding affinity order obtained
Published on 14 August 2019. Downloaded by University of Perugia on 5/22/2023 8:46:37 AM.

ponding to the PMF minima. Recent work on the binding of a from the PMF calculations. Thus direct hydrogen bond formation
type I AFP and an insect AFP on the ice surface revealed that is not the factor that drives the AFP adsorption to a particular ice
AFPs essentially adsorb on the ice plane through a layer of plane, rather the number of interfacial water molecules that
ordered water.48–50 The functional relevance of the water layer anchor the AFP towards the ice surface dictates the binding
has been critically analysed.49 Here, we have analysed the specificities. Hydrogen bonds may facilitate the complex forma-
nature of binding of the AFP on the ice surface, i.e., is it tion but are not the deciding factor. Therefore, a type III AFP is
adsorption of the AFP mediated through a layer of ordered primarily adsorbed on the pyramidal plane through a layer of
water or an AFP–ice hydrogen bonded complex? Distributions ordered water. This mode of binding also has been reported for
of the number of water molecules sandwiched between the IBS type I AFPs and insect AFPs.48–50,85 However, the structure of a
and the ice surface during the recognition of three different ice type III AFP is distinctively different from those classes of AFPs.
planes are shown in Fig. 4B. The number of such water Topologically and evolutionarily, a type III AFP does not bear any
molecules anchoring the IBS to the pyramidal plane is the structural resemblance to type I and insect AFPs.86
highest, followed by the prism plane, and it is least in the case Further insight into the structure of the adsorbed complex
of basal plane binding (Fig. 4B). Interestingly, the order strongly reveals that the IBS of the type III AFP is engulfed by a layer of

Fig. 4 (A) Potential of mean force (PMF) calculations for type III AFP adsorption on three different ice planes: pyramidal plane (black); prism plane (green);
and basal plane (pink). The results are represented as mean  S.D. (n = 4). Closer insights into the minima of the PMF profiles are shown in the inset. (B) The
distribution of interfacial water molecules between the IBS and ice surface during different ice plane adsorptions is shown. Water molecules are considered
to be interfacial when the oxygen atom of water is within 4.5 Å from the oxygen atoms of the ice surface and the IBS plane. (C) The number of hydrogen
bonds between the AFP and three different ice surfaces obtained from the equilibrium simulation of the adsorbed complexes are shown.

This journal is © the Owner Societies 2019 Phys. Chem. Chem. Phys., 2019, 21, 19298--19310 | 19303
View Article Online

PCCP Paper

resulted in less effective adsorption complex formation on the


basal surface. Hudait et al. recently showed that the hydration
water around the IBS for a hyperactive insect AFP in solution
does not adopt a clathrate like structure, however, upon binding,
the hydration layer around the IBS adopts clathrate like
ordering.52 We have also observed that water cage formation
around the IBS is particularly pronounced in the adsorbed state
of the type III AFP on the pyramidal surface.

3.3. Effect of mutations on the conformational dynamics of


Published on 14 August 2019. Downloaded by University of Perugia on 5/22/2023 8:46:37 AM.

the type III AFP


Mutagenesis was often used to identify the ice binding hotspot
residues and also used to infer the mechanism of binding for type
III AFPs.21,55,87 Graether et al. summarised the activity of all the
previously reported type III AFP mutants and newly characterized
mutants in their work to provide a comprehensive understanding
of hotspot residues.88 From the TH data, it has been observed
Fig. 5 3-D structure of the type III AFP–ice complex adsorbed on the
that the T18N mutation severely affects the antifreeze activity
pyramidal plane. The AFP is shown in a pickle green cartoon representa-
tion and the IBS residues are represented in vdW mode. The ice layer is
(90% loss in TH activity).87 Another substitution N14S causes a
shown in a stick representation. The interfacial water layer that anchors the 75% loss in activity.87 These mutations were assumed to cause
IBS on the ice surface is shown as a red transparent surface. Detailed steric interference during ice plane binding, which results in a
insight into the structure is shown in the inset (Side view: the water layer loss of hydrogen bonding between IBS and the ice surface.55 As a
that anchors the AFP on the ice surface is shown as red balls and the
control we have also considered the A16M mutation, which only
hydrogen bonds between them are shown as a red dotted line; upper
view: the IBS of the type III AFP is shown as a pale green surface and the reduces the activity by 15%.35,89 Baardsnes et al. evaluated the role
hydration water around the IBS is shown as blue circles, the top surface of of steric interactions in ice recognition by performing alanine
the pyramidal plane adjacent to the IBS is also shown in stick mode). scanning mutation of several residues close to the proposed IBS.21
They found that single mutants L19A, V20A, and V41A showed
only a 20% loss of activity, however, double mutants showed more
water which enables the adherence of the AFP on the ice than a 50% loss of activity. Among them, L19A/V41A showed a
surface (Fig. 5). Evident from the figure, most of the hydro- 75% loss of activity. Thus we have considered this particular
phobic residues face the ice surface. Very few polar groups mutant in this study.21 We have critically evaluated the effects of
orient towards the ice plane, which signifies fewer hydrogen these mutations on the structure and binding ability of the type III
bonding possibilities between the IBS and the ice surface. AFP. Notably, the chosen type III mutants showed a wide change
Further analysis of the water layer that anchors the AFP on of TH activity and the nature of substitution is different from
the ice surface reveals that there is a formation of water cages each other. Choosing these widely different mutants with
around the IBS. Hydrophobic residues at the IBS induce water varied activity profile provides a more robust validation of our
ordering. This ordered water layer effectively adsorbs AFP on proposed adsorption model.
the ice surface. In the final adsorbed state, the ice/water inter- Initially, we have evaluated the effect of those mutations
face and the ice surface complete the cage formation around on the conformational landscape of the protein and the IBS
the IBS, which anchors the AFP to the pyramidal plane. The dynamics using 500 ns of equilibrium simulation for each
water structure around the IBS closely resembles the surface system. The results are shown in Fig. 6.
water arrangement of the pyramidal plane (Fig. 5, inset upper RMSD is a global parameter that signifies structural deviations
view), as hydration water molecules (blue circles) are often of the protein during the simulation from the crystallographic
closely superimposable with the pyramidal surface water conformation and Rg is correlated with the size and shape of the
arrangements (stick representation). This structural synergy protein. 2-D scatter plots of the RMSD and Rg of wild-type and
implies that upon adsorption a few intermediate water mole- mutant AFPs are shown in Fig. 6A. Here, each dot in the plot
cules sandwiched between the IBS and the ice surface are represents a conformation of the protein sampled during the
released to the bulk and as a result the ice surface becomes simulation. Conformations of wild-type AFPs are confined in 2-D
part of the water cages framing the IBS. Close structural space close to the crystallographic native form with RMSD
matches between the hydration water around the IBS and the deviations of 1–2 Å. The T18N and N14S mutants show similar
pyramidal surface water arrangements rationalize the highest distributions to the wild-type AFP. However, RMSF analysis
binding affinity of the type III AFP for the pyramidal plane. The reveals that the fluctuations of the IBS are highest in the case
basal surface is hexagonal and therefore there is a mismatch in of the N14S mutant (Fig. 6B). Therefore, the observed loss of
the water arrangement of the basal surface and the hydration activity of the N14S mutant can be partly attributed to the loss of
layer water structure of the type III AFP. Thus the spontaneity in structural rigidity of the IBS. However, the IBS is highly stable for
water cage formation has been reduced significantly, which the T18N mutant. Therefore, the loss of activity of the T18N mutant

19304 | Phys. Chem. Chem. Phys., 2019, 21, 19298--19310 This journal is © the Owner Societies 2019
View Article Online

Paper PCCP
Published on 14 August 2019. Downloaded by University of Perugia on 5/22/2023 8:46:37 AM.

Fig. 6 Effects of several deleterious mutations on the dynamics of the type III AFP studied by using equilibrium simulations. (A) 2-D scatter plots of root
mean square deviations (RMSD) and radius of gyration (Rg) for wild type and various mutant type III AFPs obtained from equilibrium simulations are
shown. (B) Root mean square fluctuations (RMSF) of the ice binding plane residues in wild-type and mutant AFPs are also shown. (C) Cartoon
representations of the simulation average structure of type III and mutant AFPs coloured on the basis of their secondary structure (helix: red; b-sheet:
yellow; turn/coil: green) are shown. The IBS is shown within the red circle. Mutated residues are shown in a stick representation. The crystal structures of
the A16M (PDB id: 2MSI35) and T18N (PDB id: 9MSI88) mutants are shown in the inset.

is not due to the structural alteration of the protein due to sampling techniques. The potential of mean force for pyramidal
mutation. 2-D scatter plots for the A16M and L19A//V41A mutants plane adsorption has been calculated for the wild-type and
reveal that both the proteins sample conformations during the different mutants and the results are shown in Fig. 7. Notably,
simulation that are structurally distinct from the wild-type AFP we have only considered the pyramidal plane binding affinity
(Fig. 6A). However, the IBS is highly stable during the simulation since it has been shown before that the type III AFP binds to the
for the A16M mutant (Fig. 6B), which implies that the mutation pyramidal plane with the highest affinity (Fig. 4A).
might alter the dynamics of the non-IBS regions rather than the Mutations indeed reduce the binding affinity towards the
IBS. However, the double mutant shows slightly higher IBS fluctua- pyramidal plane and very interestingly the loss of activity is
tions during the simulation (Fig. 6B). The average simulation strongly correlated with the loss of binding affinity (Fig. 7B),
structure reveals that a region of the IBS appears as b-bridge except for the double mutant. Thus a mutation that reduces
(Fig. 6C, indicated by the circle) for the A16M and N14S mutants the ice plane binding affinity essentially reduces the antifreeze
rather than b-sheet, as exists in the wild-type AFP. Thus the activity of the AFP, which indicates that appropriate ice plane
alteration in the structure of the protein induced by mutagenesis binding with considerable affinity is a critical event in the anti-
is not the principal cause of the observed loss of activity for most of freezing activity of an AFP, at least for type III AFPs.
the AFP mutants. However, high IBS fluctuations and a mild loss of As we have shown that the wild-type AFP adsorbs on the
structural rigidity of the IBS for the N14S mutant might play a role pyramidal plane with a layer of ordered water and hydrogen
in the activity loss of the mutant to some extent. bonds between the IBS and the ice surface further stabilize
the complex, so we have then separately analysed how each
3.4. Type III AFP mutants bind with the pyramidal plane with of the mutations changes the ordering of the water layer
reduced affinity that anchors AFP to the ice surface and also the number
We have further investigated the role of these mutations of IBS–ice surface hydrogen bonds. The results are shown in
in the binding affinity with the ice plane using umbrella Fig. 8.

This journal is © the Owner Societies 2019 Phys. Chem. Chem. Phys., 2019, 21, 19298--19310 | 19305
View Article Online

PCCP Paper
Published on 14 August 2019. Downloaded by University of Perugia on 5/22/2023 8:46:37 AM.

Fig. 7 (A) Potential of mean force (PMF) profiles for pyramidal plane adsorption for the wild type and different type III AFP mutants are shown. Results are
represented as mean  S.D. (n = 4). (B) The correlation between the calculated binding free energy and experimentally observed thermal hysteresis
data21,35,55,87,89 (scaled with respect to the wild-type AFP activity) is shown.

Fig. 8 (A) The average number of water molecules anchoring the IBS with the pyramidal surface (black square) for the wild-type and different mutants
and the associated number fluctuations (pink square) are shown. Water molecules are considered as anchored when the oxygen atom of water is within
3.5 Å from the oxygen atoms of the ice surface and the IBS plane. (B) Distributions of IBS–ice surface hydrogen bonds for wild-type and mutant type III
AFP adsorption on the pyramidal plane are shown.

From the equilibrium simulations of the bound complexes time frame and hNIWi is the average number of such water
of the wild-type and mutant AFPs adsorbed on the pyramidal molecules calculated from the entire simulation trajectory. Low
plane, we have calculated the average number of interfacial fluctuations implicate the high stability of the water network
water molecules sandwiched between the IBS and the ice surface. around the IBS.
We have also characterized the average number fluctuations hwi of The wild-type AFP adsorbs on the pyramidal plane with the
those water molecules anchoring the wild-type/mutant AFPs on aid of a water layer constituted by 18–19 highly ordered water
the ice surface using the following expression: molecules that form cages around the IBS. The water cage is
  highly stable and associated with very little number fluctua-
1 X ðjNIW  hNIW ijÞ
hwi ¼ (1) tions (Fig. 8A). For the A16M mutant, the number of anchoring
n hNIW i
water molecules decreases, while their associated fluctuations
hwi is the average number fluctuation, which is calculated by increase marginally, which rationalizes the observed marginal
averaging the instantaneous number fluctuations (wi) over loss of binding affinity obtained from PMF calculations. Notably,
the entire simulation timescale. NIW is the number of water the A16M mutant is B85% active with respect to the wild-type.
molecules in between the IBS and the ice surface in a particular In the case of the double mutant, there is a significant drop in

19306 | Phys. Chem. Chem. Phys., 2019, 21, 19298--19310 This journal is © the Owner Societies 2019
View Article Online

Paper PCCP

the number of molecules of the anchoring water layer that


adsorbs the AFP on the ice surface and also its associated
fluctuation increases. In this mutant, bulkier side-chains are
mutated with small methyl groups that enable AFP to interact
more closely with the ice surface. It is clearly evident from the
PMF plot (Fig. 7A) that the double mutant interacts with the
pyramidal plane more closely in comparison to the wild-type.
Therefore, the number of interfacial water molecules sand-
wiched between the IBS and the ice surface decreases and
reduced complementarity increases the fluctuations of the water
Published on 14 August 2019. Downloaded by University of Perugia on 5/22/2023 8:46:37 AM.

cage. A significant disruption of the hydration water shell upon


ice adsorption can rationalize the high loss (B75%) in the
anti-freezing activity for the double mutant. However, since the
AFP mutant approaches closely to the ice surface, therefore
the number of IBS–ice surface hydrogen bonds increases
(Fig. 8B), which helps the mutant to gain in binding affinity
during adsorption. This explains the fact that L19A//V41A Fig. 9 O–O–O angle distributions of first hydration shell water molecules
exists as an outlier in the binding affinity and TH activity (within 5.5 Å) around the ice binding surface of the wild-type and different
correlation curve (Fig. 7B). Interestingly, for the N14S mutant AFP mutants are shown. The peak related to the tetrahedral water is
zoomed in on in the inset.
AFP, the number of anchoring water molecules is high and the
water network is also associated with much less fluctuation.
Therefore, the disruption of the hydration water shell around the hydration water around the IBS (Fig. 9). Interestingly, the
the IBS is not the reason for the loss of binding affinity. order of reduction in tetrahedrality of the hydration water in
However, there is a significant loss of hydrogen bonding the wild-type and AFP mutants (wild-type 4 A16M 4 L19A//
between the IBS and the ice surface (Fig. 8B). Notably, we V41A 4 N14S 4 T18N) is strongly correlated with the order of
have demonstrated that for the N14S mutant, the IBS lost the binding affinity obtained from the PMF calculations.
its rigidity. Thus structural alterations play a significant
role in the loss of binding affinity as well as its activity. Very 3.5. Proposed mechanism
fascinating results have been observed for the T18N mutant. Based on the detailed investigations on the mechanism of ice
The mutant AFP forms more direct hydrogen bonds with the plane adsorption of the type III AFP and its different mutants,
ice surface than the wild-type (Fig. 8B), yet the mutant has structured hydration water layer mediated ice recognition by
the least binding affinity. The loss of binding affinity is the type III AFP is proposed. The mechanism is schematically
completely associated with the loss of water cage formation shown in Fig. 10.
during pyramidal plane adsorption. The number of anchoring
water molecules decreases drastically and also its associated
fluctuations are very high, which signifies the least stability of
the water cages that anchor the T18N mutant AFP to the ice
surface (Fig. 8A). Our findings rationalize the fact that the hydration
water around the thr18 residue of the wild type AFP is highly
tetrahedral, evident from sum frequency generation spectroscopy.59
The result also clearly signifies hydrogen bond formation with the
ice surface is not the driving force for binding, rather ordering of
the hydration water around the IBS and proper complementarity
between the hydration water structure and ice/water interface
structure drives the AFP–ice adsorption complex formation. Hydro-
gen bonds that complement and stabilize the water network
around the IBS favour the complex formation further.
In fact, the loss of stability of the water cages around the
IBS during adsorption to the pyramidal surface is strongly
correlated with the inherent loss of hydration shell ordering
in different AFP mutants. Comparisons of the O–O–O angle
distributions around the IBS among wild-type and mutant
Fig. 10 Schematic outline of the adsorption mechanism of the type III
AFPs show that the frequency of the higher angle peak corres-
AFP on the ice surface. The AFP and the ice layer are rendered in cartoon
ponding to the tetrahedrally ordered water is lowest in the case and stick representations, respectively. The IBS is shown in a surface
of the T18N mutant. Also, the peak is shifted to a lower angle representation and the ordered hydration water cages around the IBS
value, which indicates the lowering of tetrahedral ordering in are shown as red sticks.

This journal is © the Owner Societies 2019 Phys. Chem. Chem. Phys., 2019, 21, 19298--19310 | 19307
View Article Online

PCCP Paper

The ice binding surface (IBS) of the type III AFP is mostly flat layer of the solid surface plays a central role. A type III AFP
and hydrophobic. Hydrophobic hydration at low temperature recognizes ice using its flat hydrophobic surface. Our study
induces the ordering of water molecules in the hydration shell shows that hydrophobic hydration is the primary driving force
around the IBS. There is even the formation of water cage for the AFP–ice complex formation, not the direct hydrogen
structures BTm. The close synergy between the water arrangements bond formation between the IBS and the ice surface. Mutations
in the water cages around the IBS and the surface water that interrupt the hydration shell water ordering essentially
arrangements of the pyramidal plane drives the AFP to adhere lead to less efficient adsorption on the ice surface, which in
on the pyramidal ice surface. In the adsorbed complex, the ice turn greatly reduces the activity.
surface becomes part of the water cages, which effectively Comparisons of our results with the ice binding mechanism
adsorbs the AFP on the ice surface. A specific mutation of the of other classes of AFPs, type I, type II90 and insect AFPs,48,49,53
Published on 14 August 2019. Downloaded by University of Perugia on 5/22/2023 8:46:37 AM.

IBS residues that disrupts the ordering of the hydration layer indicate a universal mechanism of ice recognition among
around the IBS leads to less effective adsorption. Disruption of different classes of AFPs: ordered hydration layer mediated
the water ordering around the IBS results in a large reduction ice recognition. However, the origin of such water ordering
in the binding affinity of the mutant AFPs which can’t be around the IBS is completely different in different types of AFPs
compensated for by direct hydrogen bond formation with the and strongly depends on the sequence and topologies of the IBS.
ice surface. The principal driving force for this hydration water The evolution of different classes of AFPs was independent,32 which
mediated ice recognition is hydrophobic hydration, which essentially leads to such structural diversities in the IBS among
becomes highly significant in the low temperature regime. different AFPs.86 As a result the ice binding affinities and activities
It is noteworthy that the working temperature of AFPs is around of different classes of AFPs greatly vary from weak, moderate to
freezing temperature. A strong linear correlation between the hyperactive. However, the core recognition mechanism remains
binding affinity and activity data in several AFP mutants preserved through the process of natural selection.
suggests that the binding to a specific ice plane dictates the
observed thermal hysteresis (TH) activity.
Conflicts of interest
There are no conflicts to declare.
4. Conclusions
The present study explores the mechanism of ice/water inter-
face recognition by a type III AFP. Previous studies on type I and Acknowledgements
hyperactive insect AFPs suggested an anchored clathrate water The authors sincerely acknowledge financial assistance from
mediated ice adsorption mechanism.48–52 The IBSs of both the Department of Science and Technology, Government of India
hyperactive insect AFPs and type I AFPs contain equally spaced (SERB grant EMR/2016/001333). The authors also gratefully
periodic threonine residues, known as Thr ladders. Methyl acknowledge the central supercomputing facility (CRAY) at the
side-chain and hydroxyl groups play a crucial role in stabilizing Indian Association for the Cultivation of Science, Kolkata.
the anchored clathrate water array which anchors the AFP to
the ice surface by forming co-operative dual hydrogen bonds.49
However, type III AFPs are globular proteins and structurally References
and topologically a type III AFP is completely different from
type I and insect AFPs. Importantly, the IBS does not contain 1 M. M. Orosco, C. Pacholski and M. J. Sailor, Nat. Nanotech-
any repeat of threonine or any other related amino acids. nol., 2009, 4, 255.
Rather, the IBS is comparatively flat and more hydrophobic 2 T. Fischer, A. Agarwal and H. Hess, Nat. Nanotechnol., 2009,
compared to the insect AFPs. Here, we have observed an IBS 4, 162.
induced ordered hydration water mediated ice/water interface 3 H. Im, X.-J. Huang, B. Gu and Y.-K. Choi, Nat. Nanotechnol.,
recognition mechanism. The flat hydrophobic IBS induces 2007, 2, 430.
water structure formation in the freezing temperature regime 4 A. E. Nel, L. Mädler, D. Velegol, T. Xia, E. M. V. Hoek,
which drives the AFP to adhere to the ice/water interface of the P. Somasundaran, F. Klaessig, V. Castranova and M. Thompson,
pyramidal plane and to some extent to the prism plane too. Nat. Mater., 2009, 8, 543.
Structural synergies between the water cages formed around 5 C. Wu, M. Chen, A. A. Skelton, P. T. Cummings and
the IBS and the interfacial water of the pyramidal plane highly T. Zheng, ACS Appl. Mater. Interfaces, 2013, 5, 2567–2579.
assist the adsorption. This observation justifies the fact that 6 Y. Cheng, L.-D. Koh, D. Li, B. Ji, Y. Zhang, J. Yeo, G. Guan,
type III AFPs are essentially non-basal binders supported by the M.-Y. Han and Y.-W. Zhang, ACS Appl. Mater. Interfaces,
ice-etching study.58 2015, 7, 21787–21796.
The ice/water interface recognition mechanism by type III 7 K. Yao, P. Tan, Y. Luo, L. Feng, L. Xu, Z. Liu, Y. Li and
AFPs is surprising in the context of the currently accepted three R. Peng, ACS Appl. Mater. Interfaces, 2015, 7, 12270–12277.
step model for protein adsorption on solid surfaces where 8 Q. Q. Hoang, F. Sicheri, A. J. Howard and D. S. C. Yang,
anchoring by the hydrophilic residues to the interfacial water Nature, 2003, 425, 977.

19308 | Phys. Chem. Chem. Phys., 2019, 21, 19298--19310 This journal is © the Owner Societies 2019
View Article Online

Paper PCCP

9 X. Sun, Z. Feng, T. Hou and Y. Li, ACS Appl. Mater. Interfaces, 39 K. Meister, S. Ebbinghaus, Y. Xu, J. G. Duman, A. DeVries,
2014, 6, 7153–7163. M. Gruebele, D. M. Leitner and M. Havenith, Proc. Natl.
10 Y. Yu, H. Sun, K. Gilmore, T. Hou, S. Wang and Y. Li, Acad. Sci. U. S. A., 2013, 110, 1617–1622.
ACS Appl. Mater. Interfaces, 2017, 9, 32452–32462. 40 S. Ebbinghaus, K. Meister, B. Born, A. L. DeVries, M. Gruebele
11 A. M. Sultan, Z. C. Westcott, Z. E. Hughes, J. P. Palafox- and M. Havenith, J. Am. Chem. Soc., 2010, 132, 12210–12211.
Hernandez, T. Giesa, V. Puddu, M. J. Buehler, C. C. Perry and 41 S. Ebbinghaus, K. Meister, M. B. Prigozhin, A. L. DeVries,
T. R. Walsh, ACS Appl. Mater. Interfaces, 2016, 8, 18620–18630. M. Havenith, J. Dzubiella and M. Gruebele, Biophys. J., 2012,
12 A. A. Skelton, T. Liang and T. R. Walsh, ACS Appl. Mater. 103, L20–L22.
Interfaces, 2009, 1, 1482–1491. 42 S. S. Mallajosyula, K. Vanommeslaeghe and A. D. MacKerell,
13 L. Vroman, Nature, 1962, 196, 476. J. Phys. Chem. B, 2014, 118, 11696–11706.
Published on 14 August 2019. Downloaded by University of Perugia on 5/22/2023 8:46:37 AM.

14 M. J. Penna, M. Mijajlovic and M. J. Biggs, J. Am. Chem. Soc., 43 G. Todde, C. Whitman, S. Hovmöller and A. Laaksonen,
2014, 136, 5323–5331. J. Phys. Chem. B, 2014, 118, 13527–13534.
15 J. Yu, M. L. Becker and G. A. Carri, Langmuir, 2012, 28, 44 G. Todde, S. Hovmöller and A. Laaksonen, J. Phys. Chem. B,
1408–1417. 2015, 119, 3407–3413.
16 P. Harder, M. Grunze, R. Dahint, G. M. Whitesides and 45 N. Pertaya, C. B. Marshall, C. L. DiPrinzio, L. Wilen, E. S.
P. E. Laibinis, J. Phys. Chem. B, 1998, 102, 426–436. Thomson, J. S. Wettlaufer, P. L. Davies and I. Braslavsky,
17 S. I. Jeon, J. H. Lee, J. D. Andrade and P. G. De Gennes, Biophys. J., 2007, 92, 3663–3673.
J. Colloid Interface Sci., 1991, 142, 149–158. 46 N. Pertaya, C. B. Marshall, Y. Celik, P. L. Davies and
18 I. Szleifer, Biophys. J., 1997, 72, 595–612. I. Braslavsky, Biophys. J., 2008, 95, 333–341.
19 L. Li, S. Chen, J. Zheng, B. D. Ratner and S. Jiang, J. Phys. 47 R. Drori, Y. Celik, P. L. Davies and I. Braslavsky, J. R. Soc.,
Chem. B, 2005, 109, 2934–2941. Interface, 2014, 11, 20140526.
20 W. Norde, Colloids Surf., B, 2008, 61, 1–9. 48 S. Chakraborty and B. Jana, Phys. Chem. Chem. Phys., 2017,
21 J. Baardsnes and P. L. Davies, Biochim. Biophys. Acta, Proteins 19, 11678–11689.
Proteomics, 2002, 1601, 49–54. 49 S. Chakraborty and B. Jana, Langmuir, 2017, 33, 7202–7214.
22 W. Zhang and R. A. Laursen, J. Biol. Chem., 1998, 273, 50 S. Chakraborty and B. Jana, J. Phys. Chem. B, 2018, 122,
34806–34812. 3056–3067.
23 Q. Ye, E. Leinala and Z. Jia, Acta Crystallogr., Sect. D: Biol. 51 A. Hudait, N. Odendahl, Y. Qiu, F. Paesani and V. Molinero,
Crystallogr., 1998, 54, 700–702. J. Am. Chem. Soc., 2018, 140, 4905–4912.
24 J. H. Kim, H. J. Lee, B. Y. Hur, W. C. Lee, S.-H. Park and 52 A. Hudait, D. R. Moberg, Y. Qiu, N. Odendahl, F. Paesani and
B.-W. Koo, Mar. Drugs, 2017, 15, 27. V. Molinero, Proc. Natl. Acad. Sci. U. S. A., 2018, 115, 8266.
25 W. Gronwald, M. C. Loewen, B. Lix, A. J. Daugulis, 53 K. Meister, C. J. Moll, S. Chakraborty, B. Jana, A. L. DeVries,
F. D. Sönnichsen, P. L. Davies and B. D. Sykes, Biochemistry, H. Ramløv and H. J. Bakker, J. Chem. Phys., 2019, 150, 131101.
1998, 37, 4712–4721. 54 C. P. Garnham, R. L. Campbell and P. L. Davies, Proc. Natl.
26 N. Xiao, K. Suzuki, Y. Nishimiya, H. Kondo, A. Miura, Acad. Sci. U. S. A., 2011, 108, 7363–7367.
S. Tsuda and T. Hoshino, FEBS J., 2009, 277, 394–403. 55 Z. Jia, C. I. DeLuca, H. Chao and P. L. Davies, Nature, 1996,
27 S. P. Graether, M. J. Kuiper, S. M. Gagne, V. K. Walker, Z. Jia, 384, 285–288.
B. D. Sykes and P. L. Davies, Nature, 2000, 406, 325–328. 56 D. Verreault, S. Alamdari, S. J. Roeters, R. Pandey,
28 M. M. Tomczak, C. B. Marshall, J. A. Gilbert and P. L. Davies, J. Pfaendtner and T. Weidner, Phys. Chem. Chem. Phys.,
Biochem. Biophys. Res. Commun., 2003, 311, 1041–1046. 2018, 20, 26926–26933.
29 S. Y. Gauthier, A. J. Scotter, F.-H. Lin, J. Baardsnes, G. L. 57 C. C. Cheng and A. L. DeVries, The role of antifreeze
Fletcher and P. L. Davies, Cryobiology, 2008, 57, 292–296. glycopeptides and peptides in the freezing avoidance of cold-
30 Z. Jia and P. L. Davies, Trends Biochem. Sci., 2002, 27, 101–106. water fish, Berlin, Heidelberg, 1991, pp. 1–14.
31 A. L. Devries, Science, 1971, 172, 1152–1155. 58 A. A. Antson, D. J. Smith, D. I. Roper, S. Lewis, L. S. D. Caves,
32 C.-H. C. Cheng, Curr. Opin. Genet. Dev., 1998, 8, 715–720. C. S. Verma, S. L. Buckley, P. J. Lillford and R. E. Hubbard,
33 A. L. Devries and Y. Lin, Biochim. Biophys. Acta, Protein J. Mol. Biol., 2001, 305, 875–889.
Struct., 1977, 495, 388–392. 59 Z. F. Brotzakis, I. K. Voets, H. J. Bakker and P. G. Bolhuis,
34 D. Wen and R. A. Laursen, Biophys. J., 1992, 63, 1659–1662. Phys. Chem. Chem. Phys., 2018, 20, 6996–7006.
35 C. I. DeLuca, P. L. Davies, Q. Ye and Z. Jia, J. Mol. Biol., 1998, 60 N. Smolin and V. Daggett, J. Phys. Chem. B, 2008, 112, 6193–6202.
275, 515–525. 61 C. Yang and K. A. Sharp, Biophys. Chem., 2004, 109, 137–148.
36 D. S. C. Yang, W.-C. Hon, S. Bubanko, Y. Xue, J. Seetharaman, 62 K. Meister, S. Strazdaite, A. L. DeVries, S. Lotze, L. L. C.
C. L. Hew and F. Sicheri, Biophys. J., 1998, 74, 2142–2151. Olijve, I. K. Voets and H. J. Bakker, Proc. Natl. Acad. Sci.
37 Y. Celik, R. Drori, N. Pertaya-Braun, A. Altan, T. Barton, U. S. A., 2014, 111, 17732–17736.
M. Bar-Dolev, A. Groisman, P. L. Davies and I. Braslavsky, 63 J. Grabowska, A. Kuffel and J. Zielkiewicz, J. Chem. Phys.,
Proc. Natl. Acad. Sci. U. S. A., 2013, 110, 1309–1314. 2016, 145, 075101.
38 D. R. Nutt and J. C. Smith, J. Am. Chem. Soc., 2008, 130, 64 Y. Xu, A. Bäumer, K. Meister, C. G. Bischak, A. L. DeVries,
13066–13073. D. M. Leitner and M. Havenith, Chem. Phys. Lett., 2016, 647, 1–6.

This journal is © the Owner Societies 2019 Phys. Chem. Chem. Phys., 2019, 21, 19298--19310 | 19309
View Article Online

PCCP Paper

65 S. Mahatabuddin, D. Fukami, T. Arai, Y. Nishimiya, 77 W. Humphrey, A. Dalke and K. Schulten, J. Mol. Graphics,
R. Shimizu, C. Shibazaki, H. Kondo, M. Adachi and 1996, 14, 33–38.
S. Tsuda, Proc. Natl. Acad. Sci. U. S. A., 2018, 115, 5456. 78 G. Bussi, D. Donadio and M. Parrinello, J. Chem. Phys., 2007,
66 T. Sun, S. Y. Gauthier, R. L. Campbell and P. L. Davies, 126, 014101.
J. Phys. Chem. B, 2015, 119, 12808–12815. 79 T. Darden, D. York and L. Pedersen, J. Chem. Phys., 1993, 98,
67 M. M. Teeter, Proc. Natl. Acad. Sci. U. S. A., 1984, 81, 6014. 10089–10092.
68 R. Talon, N. Coquelle, D. Madern and E. Girard, Front. 80 D. Schneidman-Duhovny, Y. Inbar, R. Nussinov and
Microbiol., 2014, 5, 66. H. J. Wolfson, Nucleic Acids Res., 2005, 33, W363–W367.
69 D. Van Der Spoel, E. Lindahl, B. Hess, G. Groenhof, A. E. 81 J. S. Hub, B. L. de Groot and D. van der Spoel, J. Chem.
Mark and H. J. C. Berendsen, J. Comput. Chem., 2005, 26, Theory Comput., 2010, 6, 3713–3720.
Published on 14 August 2019. Downloaded by University of Perugia on 5/22/2023 8:46:37 AM.

1701–1718. 82 C. P. Garnham, A. Natarajan, A. J. Middleton, M. J. Kuiper,


70 E. Lindahl, B. Hess and D. van der Spoel, J. Mol. Model., I. Braslavsky and P. L. Davies, Biochemistry, 2010, 49, 9063–9071.
2001, 7, 306–317. 83 S. Parui and B. Jana, J. Phys. Chem. B, 2017, 121, 7016–7026.
71 H. J. C. Berendsen, D. van der Spoel and R. van Drunen, 84 S. Parui and B. Jana, J. Phys. Chem. B, 2018, 122, 9827–9839.
Comput. Phys. Commun., 1995, 91, 43–56. 85 J. Grabowska, A. Kuffel and J. Zielkiewicz, Phys. Chem. Chem.
72 G. A. Kaminski, R. A. Friesner, J. Tirado-Rives and W. L. Phys., 2018, 20, 25365–25376.
Jorgensen, J. Phys. Chem. B, 2001, 105, 6474–6487. 86 S. Chakraborty and B. Jana, Proc. Indian Natl. Sci. Acad.,
73 W. L. Jorgensen and J. Tirado-Rives, J. Am. Chem. Soc., 1988, 2019, 85, 169–187.
110, 1657–1666. 87 H. Chao, C. I. DeLuca, P. L. Davies, B. D. Sykes and
74 W. L. Jorgensen, D. S. Maxwell and J. Tirado-Rives, J. Am. F. D. Sönnichsen, Protein Sci., 1994, 3, 1760–1769.
Chem. Soc., 1996, 118, 11225–11236. 88 S. P. Graether, C. I. DeLuca, J. Baardsnes, G. A. Hill, P. L.
75 H. J. C. Berendsen, J. R. Grigera and T. P. Straatsma, J. Phys. Davies and Z. Jia, J. Biol. Chem., 1999, 274, 11842–11847.
Chem., 1987, 91, 6269–6271. 89 C. I. DeLuca, H. Chao, F. D. Sönnichsen, B. D. Sykes and
76 T. Bryk and A. D. J. Haymet, J. Chem. Phys., 2002, 117, P. L. Davies, Biophys. J., 1996, 71, 2346–2355.
10258–10268. 90 S. Chakraborty and B. Jana, Metallomics, 2019, 11, 1387–1400.

19310 | Phys. Chem. Chem. Phys., 2019, 21, 19298--19310 This journal is © the Owner Societies 2019

You might also like