Pal2020 nonIceBindingHyperactiveAntifreezeProteino

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

pubs.acs.

org/JPCB Article

Deciphering the Role of the Non-ice-binding Surface in the


Antifreeze Activity of Hyperactive Antifreeze Proteins
Prasun Pal, Sandipan Chakraborty,* and Biman Jana*
Cite This: J. Phys. Chem. B 2020, 124, 4686−4696 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Antifreeze proteins (AFPs) show thermal hysteresis through specific


interaction with the ice crystal. Hyperactive AFPs interact with the ice surface through
Downloaded via UNIV DEGLI STUDI DI PERUGIA on May 22, 2023 at 07:45:35 (UTC).

a threonine-rich motif present at their ice-binding surface (IBS). Ordering of water


around the IBS was extensively investigated. However, the role of non-IBS in ice
growth inhibition is yet to be understood completely. The present study explores the
nature of hydration and its length-scale evaluation around the non-IBS for hyperactive
AFPs. We observed that the hydration layer of non-IBS is liquid-like, even in highly
supercooled conditions, and the nature of hydration is drastically different from the
hydration pattern of non-AFP surfaces. In similar conditions, the hydration layer
around the IBS is ice-like ordered. Non-IBS of the hyperactive AFP exposes toward the
bulk and is able to maintain the liquid-like character of its hydration water up to 15 Å.
We also find that the amino acid compositions and their spatial distribution on the
non-IBS are markedly different from those of the IBS and non-AFP surfaces. These
results elucidate the combined role of IBS and non-IBS in ice-growth inhibition. While
IBS is required to adsorb on ice efficiently, the exposed non-IBS may prevent ice nucleation/growth on top of the bound AFPs.

1. INTRODUCTION with the ice crystal.16−18 However, due to the size limitation,
Understanding the water-ice phase transition and its the IBS of AFP is not suitable for ice nucleation.19 The large
modulation by external perturbation is an active area of size of the IBS of INPs stimulates ice nucleation since it can
research. The process is a central event in many natural form stable ice nuclei that can initiate ice growth. Very
atmospheric phenomena such as snowfall and atmospheric recently, the notion got further credence from the experimental
icing as well as in the cryo-industry. Nature develops a unique report, which showed that the fish type III AFP and beetle
tool to modulate ice growth as a survival strategy for many TmAFP can induce ice nucleation above the homogeneous
organisms living in extremely cold conditions.1,2 These freezing temperature in solution, probed by using a customized
organisms produce antifreeze proteins (AFPs) that can depress microfluidic device that employs nanoliter droplets.19 Ice
the freezing point of water in a noncolligative manner.2 Certain nucleation temperature is strongly correlated with the critical
AFPs are capable of depressing freezing points up to 5−6 °C size of ice nuclei, and a larger protein surface increases the
or even higher, and they are termed as hyperactive AFPs.3,4 temperature where ice nucleation is triggered.19
They are either characterized by β-helical topology,5 a Although the IBSs of both AFPs and INPs17 are suitably
flattened cylindrical structure,6 or a solenoid architecture.7 designed to interact with the ice surface, the specific character
The ice binding surface (IBS) of different hyperactive AFPs is of non-IBS might play a determining role in ice-growth
topologically as well as sequentially similar and commonly inhibition by AFPs. According to the adsorption-inhibition
characterized by the Threonine-X-Threonine (T-X-T) motif model, IBS interacts with the ice plane, and the non-IBS plane
where X can be any amino acid. The degree of repeat is exposes to the bulk solvent. Ice grows in between two bound
different for different hyperactive AFPs, which in turn alters AFPs, not on the non-IBS, which accounts for the generation
the extent of ice binding ability. In general, measurement of of the curved surface within the hysteresis gap.13 Using HD-
thermal hysteresis (TH) activity is an age-old technique used VSFG spectroscopy and molecular dynamics simulation, it has
to quantify the efficacy of AFP and provides evidence of ice
binding. Mutations of residues at the IBS lower the TH activity
and alter the growing ice morphology, evident from ice-etching Received: February 12, 2020
techniques, which infer that the IBS interacts with the ice Revised: May 15, 2020
surface. Recent simulation studies show that the perfect spatial Published: May 19, 2020
arrangements of strategically located amino acid repeats make
the IBS suitable for ice interactions.8−15 In fact, both the IBSs
for AFPs and INPs (ice nucleating proteins) favorably interact

© 2020 American Chemical Society https://dx.doi.org/10.1021/acs.jpcb.0c01206


4686 J. Phys. Chem. B 2020, 124, 4686−4696
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

been shown that non-AFP also binds to the ice surface but The present study aims to explore the hydration of non-IBS
with a less affinity than the AFPs.20 Importantly, non-AFPs are to understand how it is different from the hydration of IBS,
incapable to retard ice growth and engulfed within a growing non-AFP, and bulk water. Differential water ordering around
ice crystal as an impurity.20 However, AFPs can retard ice the non-IBS, IBS, non-AFP, and bulk water has been probed to
growth, and there is no ice growth on the non-IBS exposed understand their temperature and length-scale dependencies.
toward the bulk water. Free energy calculation reveals that the Results obtained from the study provide a crucial foundation
binding free energy for the interaction of IBS with the ice to comprehend the understanding on the role of non-IBS in
surface is more favorable than the interaction between non-IBS the antifreeze activities for hyperactive AFPs.
and ice surfaces.9 Liu et al. demonstrated that, for the solid
surface-tethered AFP, the non-IBS exposed to the bulk 2. METHODS
solution impedes ice nucleation, whereas exposed IBS All the simulations were performed using the GROMACS
promotes ice nucleation.21 2016.438,39 package using the AMBER-ILDN40,41 force field
Many molecular dynamics simulations suggested that water with the TIP5P42 water model. The Visual Molecular
is more ordered around the IBS than around the non-IBS, Dynamics (VMD) Package was used for visualization
which rationalizes the preferential ice interaction by IBS. purposes.43
However, it is noteworthy that the presence of a preordered 2.1. Preparation of the Systems. Structures of hyper-
water layer around the IBS is not common to all active insect AFPs and non-AFP were obtained from the
AFPs.8−12,22−24 The presence of “ice-like” water around the protein structural repository, PDB IDs 1EZG6 (TmAFP),
IBS of hyperactive TmAFP was not observed, probed either by 1M8N44 (SbwAFP), and 2BM445 (non-AFP). The non-AFP
the 17O magnetic relaxation dispersion (MRD) method or considered in the study belongs to a similar fold like the
molecular dynamics simulation.25,26 Similar observation was hyperactive AFPs such that the effects of size and geometry are
also reported by Meister et al. for another hyperactive AFP, minimal. The non-AFP is a dimer, and a monomeric unit was
DAFP-1, using vibrational sum-frequency generation spectros- considered in the study. The C-terminal end (residues 120−
copy.27 Duboué-Dijon and Laage, on the other hand, reported 181) does not adopt the β-role topology; thus, it was excluded
IBS induced short-range enhancement of the water structure from the study. All the heteroatoms and crystallographic water
and a greater slowdown of water reorientation dynamics.28 On molecules were also excluded from the simulations. The
the other hand, the presence of ordered water population threonine-rich flat β-sheet of these hyperactive AFPs was
around the IBS for many hyperactive,29 type II,30 and type considered as the IBS, and other residues that were on the
III31 AFPs is crystallographically resolved. Detailed inves- opposite surfaces were considered as non-IBS in our study.
tigations and thermodynamic analysis reveal the role of Details of the residues considered as IBS and non-IBS for the
hydrophobic hydration of IBS residues in the stabilization of two AFPs (TmAFP and SbwAFP) and residues considered for
the ordered water layer in the low-temperature region.10,32−34 the non-AFP are shown in Table S1 (Supporting Information).
The implication of this ordered water layer on ice binding has Mutated TmAFP was constructed by mutating all the
been elucidated at the atomic level for type I,8 II,11 III,12 and threonine residues at the IBS to glycine.
insect AFPs.9,10 While hydration around the IBS is primarily 2.2. Simulation Details. All proteins were initially
explored to interpret its role in ice recognition, the hydration minimized in vacuo using the steepest descent algorithm to
remove any bad contacts. Then, each of them was solvated in a
pattern around the non-IBS is comparatively less explored.
box containing TIP5P water molecules. The dimension of the
Previous computer simulations depicted that the ordering of
simulation box containing TmAFP was 6.24 × 5.92 × 7.99 nm3
water around the non-IBS is significantly lower than that
filled with 9230 water. Two Na+ ions were added to make the
around the IBS.22,23 Kuffel et al. showed that the ordering of
system charge neutral. The dimension of the solvated box for
solvation water around the IBS and non-IBS is distinct. In SbwAFP was 6.81 × 6.99 × 8.39 nm3 with 12,455 water
terms of structural ordering, there are some similarities molecules. Three Cl− ions were added to neutralize the charge
between the solvation water around the IBS and ice.35 On of the system. The box dimension of the solvated non-AFP was
the other hand, structural characterization of hydration water 7.14 × 7.21 × 10.38 nm3 with 16,668 water molecules. Four
probed by the third-order Steinhardt parameter (q3) showed Na+ ions were added to neutralize the charge of the system.
that the order of water around the non-IBS is bulk-like.26 The box dimension of the mutated TmAFP was 6.24 × 5.89 ×
Duboué-Dijon and Laage reported that the average asphericity 7.99 nm3 containing 9175 water molecules. Two Na+ ions were
of hydration water around the non-IBS is almost identical to added to neutralize the charge of that system. The triclinic box
that of the non-AFP.28 However, Kozuch et al. developed a was so chosen that the protein atoms were at least 2 nm apart
neural network-based antifreeze activity prediction platform from the box edges. Long-range electrostatic interactions were
where it has been observed that the short hydrogen bond evaluated by the particle mesh Ewald46 method with a cutoff of
lifetime of hydration water around the non-IBS is an important 1.0 nm and a grid spacing of 0.16 nm. A cutoff of 1.0 nm was
parameter dictating the antifreeze activity.36 Using computer applied to calculate the van der Waals (vdW) interactions.
simulation, it has recently been shown that water density is Each of the solvated systems was initially energy-minimized
higher around the non-IBS plane than around the IBS.37 with 5000 steps by using the steepest descent algorithm. The
However, there is no such measurable difference between the minimized solvated protein was then equilibrated in the
hydration shell thickness and densities between AFPs and non- canonical (NVT) ensemble for 1 ns by employing restraining
AFPs.37 Despite all these efforts, a comprehensive picture of forces with a force constant of 1000 kJ mol−1 nm−1 to the
hydration pattern around the non-IBS, its temperature, and protein (backbone) only, while the water molecules were
length-scale dependency is missing, which is required to allowed to move freely. Final production simulations of 50 ns
understand how the hydration around the non-IBS is different were performed with the equilibrated solvated protein in the
from the hydration pattern around IBS and non-AFP surfaces. NVT ensemble at different temperatures (240, 250, 260, 273,
4687 https://dx.doi.org/10.1021/acs.jpcb.0c01206
J. Phys. Chem. B 2020, 124, 4686−4696
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

and 300 K) where the protein backbone was restrained. The different temperatures (240, 250, 260, 273, and 300 K). The
temperature was controlled by coupling with a thermostat with TIP5P water model has been used in the study since the
a time constant set to 0.1 ps using the v-rescale algorithm.47 To melting point (Tm) of the water model is close to the melting
check the ordering of the bulk, we also performed simulation temperature of ice.52 We have used the AMBER-ILDN force
of a box of 512 TIP5P water molecules with a dimension of 2.5 field for simulation, which is widely used in biomolecular
× 2.5 × 2.5 nm3 at five different temperatures using the same simulation. We have further validated the force field by
simulation parameters mentioned above. Additional simula- assessing its ability to maintain the inward orientations of the
tions for another non-AFP, ubiquitin (PDB ID: 1UBQ), were OH groups of the threonine ladder of the IBS, as observed in
performed at the mentioned five different temperatures using the crystal structure, by calculating the root-mean-square
the same protocol mentioned above. The simulation box deviation from the crystallographic orientation. The observed
dimension was 7.06 × 7.23 × 7.66 nm3, containing 7755 fluctuation is only around 0.04 nm (Figure S1, Supporting
TIP5P water molecules. Information), which indicates that the crystallographic
2.3. Analysis. Structural analysis of water around the orientation of the OH groups is maintained throughout the
protein surface was performed on the last 30 ns of the simulation. We have defined the first hydration shell by
production simulation for each system by calculating the four- considering the water molecules within 0.56 nm from each
body order parameter (F4)48 based on the H−O···O−H surface of the protein. The distance corresponds to the entire
torsional angle of the hydrogen-bonded water dimer and with first peak in the radial distribution function of water with
the modified Steinhardt bond-order parameter (<q6>).49 respect to heavy atoms of the protein. The radial distribution
The nature of ordering of water was probed with the four- function of water around the TmAFP surfaces at 260 K is
body order parameter called the F4 order parameter. One shown in Figure S2 (Supporting Information). Evident from
needs to calculate the H−O···O−H torsional angle (ϕ) of a the figure, the water density is higher around the non-IBS than

÷÷÷÷÷÷÷◊
hydrogen-bonded water dimer. By definition, hydrogen around IBS, in agreement with Nutt and Smith, which
farthest from the partner oxygen atom of the water molecule indicates a more liquid-like water structure around the non-
in the water dimer was used to define the OH vectors to IBS. Notably, ice is more ordered but less dense than liquid
calculate the H−O···O−H torsion angle (ϕ). water.22 Further water structure analysis using O−O−O angle
distribution reveals more ordering of water around the IBS in
F4 = ⟨cos 3ϕ⟩ (1) comparison to the non-IBS (Figure S3, Supporting Informa-
F4 can distinguish between ice, liquid, and clathrate. The 33,48,50 tion), as observed by Kuffel et al.35 These observations validate
F4 values are ∼0.7 for clathrate water, ∼0.0 for liquid water, our simulation methodology, which is then used to explore the
and ∼ −0.3 for hexagonal ice. hydration patterns around different AFP and non-AFP surfaces
Another order parameter used to probe ordering of water in critical detail.
was the modified Steinhardt bond order parameter, <q6>. It is 3.1. Water Ordering around the IBS Is Considerably
a local bond-order parameter that measures the local structure Different from That around the Non-IBS for Hyper-

ÄÅ ÉÑ
around a particle51 and defined as (here, l = 6) active AFPs. The Steinhardt bond-order parameter, <q6>, has
ÅÅ 4π ÑÑ
ÅÅ 2Ñ
|⟨qlm(i)⟩| ÑÑÑ
been used to probe the water ordering around the IBS and
⟨ql(i)⟩ = ÅÅ
ÅÅ 2l + 1 ÑÑ
+l
non-IBS. The parameter is widely used to differentiate between
ÅÇ ÑÖ
∑ liquid water and ice. Upon going from liquid-like water to a
m =−l (2) highly ordered water structure, the <q6> value increases.

ÅÄÅ ÑÉÑ
where Evident from Figure 1A, ordering of water around the IBS is

ÅÅ ÑÑ
⟨qlm(i)⟩ = ÅÅÅ Å qlm(j)ÑÑÑÑ
Nneighs(i)

ÅÅ Nneighs(i) + 1 ÑÑ
1
ÅÅÇ ÑÑÖ

j=0 (3)
Nneighs(i)
1
qlm(i) = ∑ Ylm(θij , ϕij)
Nneighs(i) j=1 (4)
Ylm (θij, ϕij) is the spherical harmonics, and θij and ϕij are the
polar and azimuthal angles of the bond vector connecting the
ith water molecule to the jth member of its Nneighs(i) measured
in an arbitrary laboratory frame of reference, respectively. The
sum goes over all neighboring oxygen atoms of water Figure 1. (A) Distribution of the <q6> order parameter of water
molecules, Nneighs(i), that are within 3.5 Å from the oxygen within the first solvation shell around the IBS and non-IBS of TmAFP
atom of water(i). The calculation includes also the oxygen upon lowering the temperature from 300 to 240 K. The left inset
atom of that water(i) itself. figure represents changes of the average <q6> values upon lowering
the temperatures. The right inset figure represents the distribution of
3. RESULTS the <q6> order parameter for bulk water at different temperatures. (B)
Distribution of the F4 order parameter of the hydration water around
The ordering of water around the IBS and non-IBS planes of the IBS and non-IBS of TmAFP upon lowering the temperature from
two hyperactive AFPs (TmAFP and SbwAFP) has been probed 300 to 240 K. The left inset figure represents changes of average F4
by using molecular dynamics simulation. For comparison, we values for hydration water around the IBS and non-IBS upon lowering
have also considered a non-AFP with similar topology. the temperatures, and corresponding bulk distribution is shown in the
Simulations of each system have been performed at five right inset figure.

4688 https://dx.doi.org/10.1021/acs.jpcb.0c01206
J. Phys. Chem. B 2020, 124, 4686−4696
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

higher than that around the non-IBS. The <q6> distribution as observed in the case of TmAFP. However, the degree of
around the non-IBS is liquid-like (shown in the inset). change is different. Here, we have also noticed that water
Interestingly, upon lowering the temperature, the hydration around the IBS is more ordered than that around the non-IBS,
water layer around the IBS progressively becomes more evident from the distribution of <q6> (Figure 2A). Upon
ordered as the peak of <q6> distribution shifts from ∼0.52 at lowering the temperature, water around the non-IBS surface
300 K to ∼0.55 at 240 K. However, the non-IBS surface resists remains more liquid-like than the hydration water of IBS and
such ordering of water to some extent upon lowering the bulk, which is evident from both the <q6> and F4 distributions
temperature. The peak of <q6> distribution changes from (Figure 2A,B).
∼0.39 to ∼0.40 upon lowering the temperature from 300 to Similar contrasting behavior of hydration water around the
240 K. This clearly signifies that the hydration water around IBS and non-IBS is evident for both TmAFP and SbwAFP,
the non-IBS retains a liquid-like character even in significant which indicates that the feature is probably common across
supercooled conditions. different hyperactive insect AFPs. The observed results support
On the other hand, the F4 order parameter is used to the observation of Liu et al., where it has been shown that ice
differentiate between different states of water. The F4 values nucleation is facilitated on the exposed IBS surface, whereas
are ∼0.7 for clathrate water, ∼0.0 for liquid water, and −0.3 for exposed non-IBS retards ice nucleation.21 However, a critical
hexagonal ice.53 Previously, F4 was successfully probed to question still remains unexplored. Is this behavior of the
uncover the functional transition from clathrate to ice-like hydration water around the non-IBS similar to the hydration
ordering in water upon lowering the temperature for a pattern of non-AFPs?
hyperactive AFP adsorbed on the ice surface.9 Upon lowering 3.2. Behavior of the Hydration Water around the
the temperature progressively from 300 to 240 K, a gradual Non-IBS of Hyperactive AFP Is Different from That of
shift in the F4 distribution toward the more negative values has Non-AFP. We have considered a non-AFP that possesses
been noticed, which indicates more ice-like ordering of the similar β-roll topology as observed in the case of TmAFP. Also,
hydration water around the IBS (Figure 1B). Marked the sizes of the AFP and non-AFP are very similar, and both of
differences have been observed in the behavior of hydration them contain 7 β-rolls. This consideration minimizes the
water around IBS and non-IBS upon lowering the temperature. effects of geometry and excluded the volume effect on
Lowering the temperature from 300 to 260 K, the F4 observed quantities.
distribution of hydration water around the non-IBS is almost Interestingly, changes of both F4 and <q6> distributions for
unchanged and remains liquid-like (Figure 1B). Further hydration water around the non-AFP surfaces are markedly
lowering the temperature from 260 to 240 K, there is a different from the behavior of water around the non-IBS of
minute shift in the F4 distribution toward the negative values TmAFP upon lowering the temperature. <q6> distribution
for hydration water around the non-IBS. The peak value is ∼ reveals that water around the non-AFP is more ordered than
−0.03 at 240 K for non-IBS, while for IBS, it is −0.112 at the that around the non-IBS at all the studied temperatures for
same temperature. Thus, while non-IBS resists the ice-like TmAFP (Figure 3A). A similar observation has also been
ordering of the hydration water, IBS promotes the ice-like
ordering, which is clearly evident from the average F4 value
upon lowering the temperature (Figure 1B, inset).
We have further evaluated the contrasting behavior of
hydration water around the IBS and non-IBS for another
hyperactive AFP from Spruce budworm (SbwAFP) with the
aid of both <q6> and F4 order parameters to address the
generality of the observations, and results are shown in Figure
2. We have observed a similar pattern of hydration around the
IBS and non-IBS upon lowering the temperatures (Figure 2),

Figure 3. (A) Distribution of the <q6> order parameter of water


within the first solvation shell around the non-IBS of TmAFP and
non-AFP surfaces upon lowering the temperature from 300 to 240 K.
(B) Distribution of the F4 order parameter of hydration water around
the non-IBS of TmAFP and the surfaces of non-AFP upon lowering
the temperature from 300 to 240 K. Corresponding changes in
average <q6> and F4 values are shown in the inset.

noticed for SbwAFP when the temperature-dependent


behavior of its hydration shell has been compared with the
hydration behavior around the non-AFP surfaces (Figure S4,
Figure 2. (A) Distribution of the <q6> order parameter of hydration Supporting Information). Notably, the degree of ordering of
water around the IBS and non-IBS of SbwAFP upon lowering the
temperature from 300 to 240 K. The inset figure represents the
water around the non-AFP is markedly less than that of the IBS
average <q 6> values at different studied temperatures. (B) of both the hyperactive AFPs. To further assess the generality
Distribution of the F4 order parameter of hydration water around of the observation, we choose another globular non-AFP,
the IBS and non-IBS of SbwAFP upon lowering the temperature from namely, ubiquitin, a widely studied model protein. Comparison
300 to 240 K. The inset figure represents the average F4 values of of hydration shell ordering around ubiquitin and non-IBS of
water around the IBS and non-IBS at different temperatures. TmAFP upon lowering the temperature is shown in Figure S5
4689 https://dx.doi.org/10.1021/acs.jpcb.0c01206
J. Phys. Chem. B 2020, 124, 4686−4696
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

(Supporting Information). We observed a similar behavior as SbwAFP resist ice-like ordering of hydration water very
noted above. Thus, non-AFPs do not favor ice formation like efficiently up to 260 K. Notably, the Tm of TIP5P water is
the IBS of an AFP. Potentials of mean force calculations also ∼274 K.52 A considerable reduction in F4 values and
reveal that the IBS of AFP possesses a higher binding affinity concomitant increase in <q6> values of hydration water at
with the ice surface than the non-AFPs,20 which is in line with 260 K around the IBS for both the AFPs have been noticed.
the observation obtained from this study. This indicates the ice-like ordering of hydration water around
Changes in the F4 values reveal progressive ordering of the IBS. The hydration water around the non-AFP stays in
hydration water around the non-AFP surfaces with the between non-IBS and IBS of the AFP, which indicates that the
lowering of temperature (Figure 3B, Figures S4B and S5B), non-AFP cannot resist ice-like ordering of its hydration water
which is unlike the behavior of hydration water around the upon lowering the temperature; however, it is also less efficient
non-IBS surface. Thus, the behavior of water around the non- to promote ice-like ordering of the hydration water in
IBS is unique for AFPs, and it resists water ordering into an comparison to the IBS.
ice-like form even in highly supercooled conditions, which Upon lowering the temperature up to 240 K, there is a
might have implications on the ice-growth inhibition ability of marginal ordering in the hydration water around the non-IBS
AFPs. for both the AFPs, evident from the minute decrease in F4 and
3.3. Degree of Differential Water Ordering around increase of <q6> values. However, there is a large increase in
the IBS and Non-IBS: Effect of Temperatures. We have water ordering around the IBS for both the AFPs in the 260−
then probed the effect of temperature on the discriminatory 240 K temperature region, evident from a large decrease in F4
behavior of water ordering induced by the IBS and non-IBS of and an increase in <q6> values. The magnitude of the decrease
AFPs. Also, the changes of water ordering around the non-AFP in F4 values indicates a favorable ice-nucleating condition
surface upon lowering the temperature have been considered templated by the IBS; however, the area of the IBS of the two
for comparison (Figure 4A,B). 300 K temperature has been AFPs used in this study is too small to initiate any significant
heterogeneous ice nucleation. In this low-temperature region,
the hydration shell around the non-IBS of both the AFPs still
retains its liquid-like character.
3.4. Long-Range Effects of Non-IBS-Induced Water
Ordering in Hyperactive AFPs. We have then probed the
long-range effects of the IBS and non-IBS-induced water
ordering and compared with the non-AFP-induced water
ordering using the F4 order parameter. We have shown here
the results obtained at 260 K (Figure 5A). Results obtained at
300 and 273 K are largely similar to that of 260 K and
therefore shown in the Supporting Information (Figure S6).
Figure 4. Changes in average (A) F4 and (B) <q6> order parameters Results obtained at 260 K for SbwAFP are also shown in Figure
of water within the first solvation shell around the IBS, non-IBSs of S7 (Supporting Information), which also show a similar trend.
two AFPs (TmAFP and SbwAFP), and a non-AFP at different We have considered water within 5.6 Å from the protein
temperatures. surface as the first solvation shell, which corresponds to the
entire first peak in the g(r) plot of the protein surface and
considered as the reference temperature, and the degree of water oxygen (Figure S2). Thereafter, we have considered a 2
water ordering has been probed by subtracting the average F4 Å slab to define the next hydration layer. At 260 K, the order of
and <q6> values at 300 K from the respective average values at water around the IBS is maximum (∼ −0.045) and least
a particular temperature. A coherent picture has been evident around the non-IBS (∼ −0.009), and non-AFP (∼ −0.025) is
either probed by F4 or <q6>. The non-IBSs of TmAFP and in between these two extreme cases. The corresponding bulk

Figure 5. (A) Changes in the average F4 order parameter values of different hydration shells around the IBS, non-IBS of TmAFP, and non-AFP at
260 K. (B) Changes of ΔF4 (F4 at 300 K to F4 at 260 K) in different hydration layers around the IBS, non-IBS of TmAFP, and nonAFP and
compared with the bulk values.

4690 https://dx.doi.org/10.1021/acs.jpcb.0c01206
J. Phys. Chem. B 2020, 124, 4686−4696
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

F4 value is highly negative (∼ −0.066) at 260 K. This is the distribution of <q6> shifts toward lower values, and the
expected since it was previously reported that protein distribution of F4 shifts toward zero. These results indicate that
modulates the ordering of water strongly within the first mutations indeed reduce the ice-like water ordering around the
solvation layer;54,55 therefore, these water molecules are less IBS. On the other hand, the distribution of <q6> and F4 order
responsive to external stimuli like temperature. parameters for hydration water around the non-IBS remains
Interestingly, not only the different degree of water ordering essentially unchanged upon mutation at both the temperatures.
modulation by various surfaces is evident in the first hydration This signifies that modulation of water ordering around the
shell, but also it is percolated up to many layers beyond the IBS and non-IBS is probably uncorrelated.
first hydration shell. Figure 5A reveals that the F4 values of 3.6. Comparisons of Amino Acid Composition and Its
hydration water around the IBS approach more closely the Spatial Distribution on the Surfaces of IBS, Non-IBS,
bulk values upon extending the solvation layer length scale in and Non-AFPs. Electrostatic surface potential calculations
comparison to the non-IBS and non-AFP surfaces. Distance- clearly depict that the IBS of both the AFPs is flat and
dependent behavior of hydration water around the non-IBS hydrophobic in comparison to the non-IBS and non-AFP
plane is very unique and retains a more liquid-like character surfaces. Non-IBS of TmAFP is slightly negatively charged.
than other surfaces. At a particular temperature, the degree of There are two non-IBS planes of SbwAFP. One of them is less
ordering in the successive 2 Å hydration layer retains this charged, but the flat and periodic patterns of residues are
differential feature of water ordering around different surfaces; missing.
however, the value decreases continuously upon going further The other non-IBS is more charged. Both positively and
from the different protein surfaces. Effects of protein surfaces negatively charged regions are evident, and the regions are
nullify at ∼17 Å (Figure 5A). However, the degree of changes segregated. On the other hand, there is no such pattern in the
in water ordering in the successive hydration layer is different charge distribution on any of the non-AFP surfaces (Figure 7).
for different protein surfaces, and this has been probed with Regions of the positively and negatively charged regions are
ΔF4, as shown in Figure 5B. Changes of water ordering around less segregated. To understand the minute differences in amino
the IBS saturate within the few layers beyond the first acid compositions of various surfaces of both AFPs and non-
hydration shell. The effect of non-AFP surfaces on protein AFPs, the amino acid distribution pattern has been calculated
hydration also decays sharply within 10 Å as the changes are for six different hyperactive AFPs and 40 representative non-
almost saturating (Figure 5B). Long-range modulation of water homologous structurally uncorrelated non-AFPs from four
ordering is clearly evident around the non-IBS, which extends different classes (All α, All β, α + β, and α/β). The results are
up to 15 Å. Thus, non-IBS is not like the non-AFP surfaces; it shown in Figure 8. Details of the selected AFP and non-AFP
is distinct and unique, which can modulate water ordering and proteins are enlisted in Table S2 (Supporting Information).
dynamics in a different way. Amino acids are grouped into four different classes
3.5. Synergism between the IBS- and Non-IBS- according to Qiao et al.56 Serine, threonine, asparagine,
induced Hydration Water Ordering Is Not Evident in glutamine, cysteine, glycine, proline, and tyrosine have been
Hyperactive AFPs: Observations from the Mutational grouped as polar neutral; arginine, histidine, and lysine have
Studies. The designing principle of IBS is naturally selected been grouped as positively charged; aspartic and glutamic acids
for hyperactive AFPs. There is a T-X-T motif at the IBS; are clubbed as negatively charged; alanine, isoleucine, leucine,
however, the degree of repeat varies in different hyperactive methionine, phenylalanine, tryptophan, and valine have been
AFPs. TmAFP contains seven such repeats (A/T/Q-X-T). We considered as hydrophobic amino acids. Hydrophobic and
have mutated the threonine residues at the IBS to glycine polar neutral amino acids are abundant at the IBS of
residues and then probed water ordering at both the IBS and hyperactive AFPs (Figure 8A). The decomposition of the
non-IBS. Our objective is to understand the effects of distribution at the individual amino acid level reveals that
disruption of water ordering around the IBS on the hydration threonine residues are mostly present among all the amino
pattern of the non-IBS. acids with a polar neutral side chain. It is well known that the
We have considered two temperatures to compare the ice-binding motif of the hyperactive AFPs is T-X-T, where T is
hydration patterns of the wild-type and mutated TmAFP, 273 a threonine residue. It has been shown recently that threonine
and 260 K. Distributions of <q6> (Figure 6A) and F4 (Figure plays a critical role in stabilizing the anchored clathrate water
6B) order parameters for the hydration water at two different array present at the IBS of hyperactive AFPs, which help the
temperatures around the IBS and non-IBS are shown. For IBS, AFP to adsorb on the ice surface.9,10 Hydrophobic amino acids
and the methyl side chain of threonine residues induce water
ordering at low temperatures that mediates ice adsorption
guided by water structure complementarities between the
ordered water layer and interfacial water of a particular ice
surface where the AFP specifically binds.
On the other hand, different polar residues are abundant on
the surface of the non-IBS without any particular preference
toward any particular amino acid. Also, there is a significantly
higher presence of charged amino acids than the IBS, and the
frequencies of hydrophobic amino acids are less on the non-
IBS than on the IBS. However, the number of residues with
either charge is considerably lower on the non-IBS than on the
Figure 6. Distributions of (A) <q6> and (B) F4 order parameters of non-AFP surfaces. Spatial distributions of amino acids from
water molecules within the first solvation shell around the IBS and different groups on the surfaces of TmAFP (Figure 8C) and
non-IBS of wild-type and mutated TmAFP at 273 and 260 K. SbwAFP (Figure S8A) are shown. On the other hand, spatial
4691 https://dx.doi.org/10.1021/acs.jpcb.0c01206
J. Phys. Chem. B 2020, 124, 4686−4696
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

Figure 7. Cartoon representations of the (A) TmAFP, (B) SbwAFP, and (C) non-AFP are shown. (D) The electrostatic surface potentials of
different planes of TmAFP, SbwAFP, and non-AFP are shown. Blue and red colors are used to indicate the positively and negatively charged
regions.

hydrophobic amino acids, and amino acids of different natures


are randomly distributed on the solvent-exposed surfaces.
There can be the presence of small microdomains, but clearly,
different types of amino acids are not segregated. In contrast,
the non-IBS surfaces of both TmAFP and SbwAFP are
characterized by a segregated pattern of different domains of
polar and charged amino acids. Charged amino acids are
absent on the exposed non-IBS surface, which faces the bulk
when the IBS of AFP is bound to the ice surface; rather, they
exist at the interface between the IBS and non-IBS. Clearly, the
amino acid composition and its spatial distribution of non-IBS
are markedly different from those of the IBS. Also, it is unique
and not comparable to the surfaces of other non-AFPs with
varied topology. Notably, there is enormous structural diversity
among non-AFPs. We have considered four structurally
distinct representative non-AFPs and observed a unanimous
trend among them; however, further consideration of many
other non-AFPs is needed to consolidate the observations.
Figure 8. (A) Normalized distributions of different types of amino
acids present at the IBS and non-IBS of six AFPs and 40 non-AFPs are 4. DISCUSSION
shown. (B) Normalized distribution of each of the 20 amino acids
present at the IBS and non-IBS of six AFPs and 40 non-AFPs is also
Ice growth inhibition by AFP is a multidimensional problem
depicted. The front view of the spatial distributions of different types and yet to be understood completely. Most of the studies focus
of amino acids on the (C) TmAFP and (D) a non-AFP with similar on the role of IBS on ice binding. A highly ordered water
topology, a pentapeptide repeat protein from M. tuberculosis (PDB ID: population around the IBS was previously probed with the aid
2BM4), is shown. The orange color signifies a polar uncharged amino of both experiments and computer simulations.8−12,20,57
acid, the blue color represents a positively charged amino acid, the red Recent free energy simulations further elucidated the role of
color represents a negatively charged amino acid, and the violet color these ordered water layers in ice binding for different classes of
signifies a hydrophobic amino acid side chain. Proline, cysteine, and AFPs.8−12 However, this preordering of hydration water is an
glycine are rendered in green color. essential criterion for ice recognition or not still remains highly
debatable.8−12,22−24 Recently, the presence of an ordered water
distributions of amino acids on four structurally uncorrelated layer during the adsorption of a hyperactive AFP on the ice
non-AFPs are shown in Figure 8D and Figure S8B−D. In surface has been probed using HD-VSFG spectroscopy.20 This
addition to the pentapeptide repeat protein from Mycobacte- study also showed that the non-AFP protein can adsorb/bind
rium tuberculosis, which shows topological similarities with on the ice surface but with a less affinity than the AFP.20 Apart
TmAFP, three other model globular non-AFPs, namely, from this distinction, AFP and non-AFP showed different
ubiquitin, lysozyme, and barnase, have been considered. In behaviors in terms of ice nucleation and growth upon lowering
the case of non-AFPs, the protein core is composed of the temperature. It has been suggested that, upon lowering the
4692 https://dx.doi.org/10.1021/acs.jpcb.0c01206
J. Phys. Chem. B 2020, 124, 4686−4696
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

temperature, ice grows on top of the non-AFP, and it retains as amino acid compositions and its spatial distributions, which is
an impurity in the growing ice slab.20 However, AFP retards markedly different. Previously, it was shown that different types
ice growth on top of it within the thermal hysteresis limit, of amino acids modulate the water orientation differently, and
which is in accordance with the adsorption-inhibition the length scale of such an influence widely varied among
model.13,20 These observations imply a different nature and different classes of amino acids.56 Negatively charged amino
degree of water ordering around the non-IBS of AFP. acids orient 98% hydration water dipole toward the protein
However, the hydration pattern of non-IBS is very little surface, and such effects percolated up to 16 Å, and positively
understood. The order of the hydration shell around the non- charged amino acids orient 94% water up to 12 Å.56 Water
IBS and its temperature and length-scale dependence are reorientation abilities of neutral polar and hydrophobic amino
mostly unexplored. Using a combined approach of molecular acids were also evident.56 Thus, different residues alter water
dynamics simulations at varied temperature range and orientations differently. Notably, the nature of water
sequence analysis, we have evaluated the nature of water reorientation ability and its length scale may change in the
ordering around the IBS and non-IBS and compared it to the low-temperature region, but the details are completely
hydration of non-AFPs. We have observed that water is highly unexplored. The unique composition and spatial distribution
ordered around the IBS of AFP compared to non-IBS and non- of neutral polar and charged residues at the non-IBS might be
AFP surfaces at all temperatures. This is in accordance with responsible for the observed temperature-resistant liquid-like
many previous studies.8−12,22,23 The water around the IBS is hydration pattern around the surface. A notable observation is
ice-like ordered, and the population of ice-like water increases the long-range effects of non-IBS-induced water reorganiza-
upon lowering the temperature for both of the hyperactive tions. At least for hyperactive AFPs, we have observed that
AFPs studied here. Temperature-dependent studies demon- non-IBS retains a significant population of liquid-like water in
strate that the non-IBS surface resists water ordering up to highly supercooled conditions up to 15 Å, while the effect of
significantly low-temperature regions. The water around the IBS and non-AFP surface induced water ordering decays
non-IBS remains liquid-like up to 260 K, and after that, there is around 10 Å. Notably, bulk water also adopts a more ice-like
a very minute increase in ordering of water upon further character upon lowering the temperature, but the degree of
lowering of temperature. Notably, we have used TIP5P water ordering is lower than that of the hydration water around the
in our study. The Tm of the water model is 274 K.52 Thus, non- IBS, as evident from the <q6> values at different temperatures.
IBS retains the liquid-like nature of its hydration water even in It is pertinent to mention that the perturbations to the F4 and
significantly supercooled conditions. This feature is markedly <q6> values upon lowering temperature may be different on
different from the hydration around the IBS. Also, the nature protein surfaces compared to the bulk. These two parameters
of the hydration shell structure is not similar to the hydration move in the opposite direction with the increment in ice-like
water around the non-AFP surfaces where we found an water population: the F4 value decreases, while the <q6> values
increment in the ordering of water upon lowering the increase. However, how these two parameters behave upon
temperature. However, the degree of ice-like ordering in the different conditions and perturbations is worthy to investigate
hydration shell of non-AFP is significantly lower than that of in more details. Notably, the value of an order parameter for
the IBS. Differences in the degree of hydration shell ordering hydration water around different surfaces at a particular
around the different surfaces upon lowering the temperature temperature reveals that the ice-like ordering is found to be
are summarized in Figure 9. Here, we represent the ordering in still significantly higher in the hydration shell around the IBS
terms of two order parameters, F4 and <q6>. than around other surfaces. Notably, Duboué-Dijon and
The differential pattern of hydration shell ordering around Laage28 studied the hydration layer ordering around a
the IBS and non-IBS might stem from the difference in their hyperactive antifreeze protein using the TIP4P/2005 water
model at 300, 260, and 235 K and noticed that the water
dynamics around the IBS is slower than the entire AFP
hydration shell, probed by the reorientation dynamics
retardation factor. This observation implies that the non-IBS
hydration shell shows different temperature sensitivity from the
IBS. However, mean asphericity calculation did not reveal such
temperature-dependent behavior. Thus, the choice of appro-
priate order parameter is crucial to elucidate subtle changes in
the hydration behavior. Kuffel et al.58 also showed the
difference in hydration shell behavior around the IBS and
non-IBS planes in a wide range of temperatures, 240−300 K,
using the translational and rotational diffusion coefficients as
order parameters. Using the surface pair correlation function,
Nutt and Smith22 showed a significant increase in water
ordering around the IBS upon lowering the temperature from
300 to 220 K, but the changes around the non-IBS is
insignificant.
Our study clearly highlights the role of non-IBS in the
Figure 9. 3D plot of average F4 and <q6> values of the first hydration antifreeze activity of hyperactive AFPs. IBS is designed to
shell water around the different protein surfaces (IBS, non-IBS, and interact with the ice surface in a strong and specific manner,
non-AFP) upon lowering the temperatures. Bulk values at the while the non-IBS-induced liquid-like ordering of its hydration
respective temperature are also shown. shell may prevent the ice nucleation and subsequent growth on
4693 https://dx.doi.org/10.1021/acs.jpcb.0c01206
J. Phys. Chem. B 2020, 124, 4686−4696
The Journal of Physical Chemistry B


pubs.acs.org/JPCB Article

top of the bound AFPs. Thus, the combined effect of IBS and AUTHOR INFORMATION
non-IBS might be operative to inhibit ice growth.
Corresponding Authors
5. CONCLUSIONS Sandipan Chakraborty − Amity Institute of Biotechnology,
Amity University, Kolkata 700135, India;
We have shown that the IBS and non-IBS of the hyperactive Email: sandipanchakraborty.13@gmail.com
AFPs are unique in terms of amino acid compositions and their Biman Jana − School of Chemical Sciences, Indian Association
spatial distributions, which enable the non-IBS plane to control for the Cultivation of Science, Kolkata 700032, India;
the organization of its hydration layer in a liquid-like orcid.org/0000-0001-7684-1963; Phone: +91 33 2473
arrangement. Non-IBS is able to maintain the liquid-like 4971; Email: pcbj@iacs.res.in; Fax: +91 33 2473 2805
character of its hydration layer even in highly supercooled
conditions, and the effects are percolated to many layers Author
beyond the first solvation shell, up to 15 Å. The behavior of Prasun Pal − School of Chemical Sciences, Indian Association for
hydration shell ordering around the non-IBS is also markedly the Cultivation of Science, Kolkata 700032, India
different from the non-AFP surfaces both in terms of
temperature and length-scale behavior. Non-AFP surfaces do Complete contact information is available at:
not possess the ability to retain the liquid-like character of its https://pubs.acs.org/10.1021/acs.jpcb.0c01206
hydration shell as efficiently as the non-IBS. Thus, combined
effects of IBS and non-IBS might orchestrate the ice-growth Notes
inhibition ability of hyperactive AFPs. The present study The authors declare no competing financial interest.
elucidates the intrinsic differences of the hydration layer
ordering around different AFP and non-AFP surfaces and their
temperature and length-scale dependency. Therefore, the study
■ ACKNOWLEDGMENTS
This research is supported by the Department of Science and
provides a direct correlation between the nature of surfaces
Technology SERB grant EMR/2016/001333 for funding. P.P.
and observed ordering in the hydration layer. The observations
acknowledges Mr. Sridip Parui for his help during the
obtained from the present study lay the foundation to interpret
development of the order parameter code. P.P. also thanks
the hydration behavior in a more complex scenario where AFP
Department of Science and Technology (DST), India for
is adsorbed on a heterogeneous ice/water interface during ice-
fellowship.


growth inhibition. The role of amino acids spatial distributions
in controlling the water structure in highly supercooled
conditions paves the way for designing highly effective, REFERENCES
nontoxic AFP mimetics with potential industrial applications. (1) Jana, B.; Chakraborty, S. Antifreeze Proteins: An Unusual Tale of


*
ASSOCIATED CONTENT
sı Supporting Information
Structural Evolution, Hydration and Function. Proc. Indian Natl. Sci.
Acad. 2019, 100, 169−187.
(2) Jia, Z.; Davies, P. L. Antifreeze proteins: an unusual receptor−
ligand interaction. Trends Biochem. Sci. 2002, 27, 101−106.
The Supporting Information is available free of charge at (3) Scotter, A. J.; Marshall, C. B.; Graham, L. A.; Gilbert, J. A.;
https://pubs.acs.org/doi/10.1021/acs.jpcb.0c01206. Garnham, C. P.; Davies, P. L. The basis for hyperactivity of antifreeze
Residues considered in the study for IBS and non-IBS proteins. Cryobiology 2006, 53, 229−239.
(4) Hanada, Y.; Nishimiya, Y.; Miura, A.; Tsuda, S.; Kondo, H.
for two different AFPs and non-AFP, details of the AFPs
Hyperactive antifreeze protein from an Antarctic sea ice bacterium
and non-AFPs considered for residue analysis, root- Colwellia sp. has a compound ice-binding site without repetitive
mean-square deviation of threonine OH groups at the sequences. FEBS J. 2014, 281, 3576−3590.
IBS of TmAFP with respect to the crystallographic (5) Graether, S. P.; Kuiper, M. J.; Gagné, S. M.; Walker, V. K.; Jia,
orientation throughout the simulation timescale, radial Z.; Sykes, B. D.; Davies, P. L. β-Helix structure and ice-binding
distribution function of water (oxygen) around the two properties of a hyperactive antifreeze protein from an insect. Nature
surfaces (IBS and non-IBS) of TmAFP at 260 K, 2000, 406, 325−328.
distribution of the O−O−O angle of water molecules (6) Liou, Y.-C.; Tocilj, A.; Davies, P. L.; Jia, Z. Mimicry of ice
within the first solvation shell around the IBS and non- structure by surface hydroxyls and water of a β-helix antifreeze
IBS of SbwAFP at 273 K, distributions of the <q6> and protein. Nature 2000, 406, 322−324.
(7) Hakim, A.; Nguyen, J. B.; Basu, K.; Zhu, D. F.; Thakral, D.;
F4 order parameters of water within the first solvation
Davies, P. L.; Isaacs, F. J.; Modis, Y.; Meng, W. Crystal Structure of an
shell around the non-IBS of SbwAFP and a non-AFP Insect Antifreeze Protein and Its Implications for Ice Binding. J. Biol.
upon lowering the temperature from 300 to 240 K, Chem. 2013, 288, 12295−12304.
distributions of the <q6> and F4 order parameters of (8) Chakraborty, S.; Jana, B. Conformational and hydration
water within the first solvation shell of the non-IBS of properties modulate ice recognition by type I antifreeze protein and
TmAFP and a non-AFP (ubiquitin) upon lowering the its mutants. Phys. Chem. Chem. Phys. 2017, 19, 11678−11689.
temperature from 300 to 240 K, changes in the average (9) Chakraborty, S.; Jana, B. Molecular Insight into the Adsorption
F4 order parameter values of hydration shells with of Spruce Budworm Antifreeze Protein to an Ice Surface: A Clathrate-
different thicknesses around the IBS, non-IBS of Mediated Recognition Mechanism. Langmuir 2017, 33, 7202−7214.
TmAFP, and non-AFP at 273 and 300 K, changes in (10) Chakraborty, S.; Jana, B. Optimum Number of Anchored
Clathrate Water and Its Instantaneous Fluctuations Dictate Ice Plane
the average F4 order parameter values of different
Recognition Specificities of Insect Antifreeze Protein. J. Phys. Chem. B
hydration shells around the IBS, non-IBS of SbwAFP, 2018, 122, 3056−3067.
and non-AFP at 260 K, spatial distribution of different (11) Chakraborty, S.; Jana, B. Calcium ion implicitly modulates the
types of amino acids on the SbwAFP and three non- adsorption ability of ion-dependent type II antifreeze proteins on an
AFPs (PDF) ice/water interface: a structural insight. Metallomics 2019, 1387.

4694 https://dx.doi.org/10.1021/acs.jpcb.0c01206
J. Phys. Chem. B 2020, 124, 4686−4696
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

(12) Chakraborty, S.; Jana, B. Ordered hydration layer mediated ice (32) Parui, S.; Jana, B. Pairwise Hydrophobicity at Low Temper-
adsorption of a globular antifreeze protein: mechanistic insight. Phys. ature: Appearance of a Stable Second Solvent-Separated Minimum
Chem. Chem. Phys. 2019, 19298. with Possible Implication in Cold Denaturation. J. Phys. Chem. B
(13) Kuiper, M. J.; Morton, C. J.; Abraham, S. E.; Gray-Weale, A. 2017, 121, 7016−7026.
The biological function of an insect antifreeze protein simulated by (33) Parui, S.; Jana, B. Molecular Insights into the Unusual Structure
molecular dynamics. eLife 2015, 4, No. e05142. of an Antifreeze Protein with a Hydrated Core. J. Phys. Chem. B 2018,
(14) Mochizuki, K.; Molinero, V. Antifreeze Glycoproteins Bind 122, 9827−9839.
Reversibly to Ice via Hydrophobic Groups. J. Am. Chem. Soc. 2018, (34) Pandey, H. D.; Leitner, D. M. Thermodynamics of Hydration
140, 4803−4811. Water around an Antifreeze Protein: A Molecular Simulation Study. J.
(15) Kondo, H.; Mochizuki, K.; Bayer-Giraldi, M. Multiple binding Phys. Chem. B 2017, 121, 9498−9507.
modes of a moderate ice-binding protein from a polar microalga. Phys. (35) Kuffel, A.; Czapiewski, D.; Zielkiewicz, J. Unusual structural
Chem. Chem. Phys. 2018, 20, 25295−25303. properties of water within the hydration shell of hyperactive antifreeze
(16) Hudait, A.; Qiu, Y.; Odendahl, N.; Molinero, V. Hydrogen- protein. J. Chem. Phys. 2014, 141, No. 055103.
Bonding and Hydrophobic Groups Contribute Equally to the Binding (36) Kozuch, D. J.; Stillinger, F. H.; Debenedetti, P. G. Combined
of Hyperactive Antifreeze and Ice-Nucleating Proteins to Ice. J. Am. molecular dynamics and neural network method for predicting
Chem. Soc. 2019, 141, 7887−7898. protein antifreeze activity. Proc. Natl. Acad. Sci. U. S. A. 2018, 115,
(17) Hudait, A.; Odendahl, N.; Qiu, Y.; Paesani, F.; Molinero, V. 13252.
Ice-Nucleating and Antifreeze Proteins Recognize Ice through a (37) Zanetti-Polzi, L.; Biswas, A. D.; Del Galdo, S.; Barone, V.;
Diversity of Anchored Clathrate and Ice-like Motifs. J. Am. Chem. Soc. Daidone, I. Hydration Shell of Antifreeze Proteins: Unveiling the Role
2018, 140, 4905−4912. of Non-Ice-Binding Surfaces. J. Phys. Chem. B 2019, 123, 6474−6480.
(18) Pandey, R.; Usui, K.; Livingstone, R. A.; Fischer, S. A.; (38) Berendsen, H. J. C.; van der Spoel, D.; van Drunen, R.
Pfaendtner, J.; Backus, E. H. G.; Nagata, Y.; Fröhlich-Nowoisky, J.; GROMACS: A message-passing parallel molecular dynamics
Schmüser, L.; Mauri, S.; Scheel, J. F.; Knopf, D. A.; Pöschl, U.; Bonn, implementation. Comput. Phys. Commun. 1995, 91, 43−56.
M.; Weidner, T. Ice-nucleating bacteria control the order and (39) Van Der Spoel, D.; Lindahl, E.; Hess, B.; Groenhof, G.; Mark,
dynamics of interfacial water. Sci. Adv. 2016, 2, No. e1501630. A. E.; Berendsen, H. J. C. GROMACS: Fast, flexible, and free. J.
(19) Eickhoff, L.; Dreischmeier, K.; Zipori, A.; Sirotinskaya, V.; Comput. Chem. 2005, 26, 1701−1718.
Adar, C.; Reicher, N.; Braslavsky, I.; Rudich, Y.; Koop, T. Contrasting (40) Hornak, V.; Abel, R.; Okur, A.; Strockbine, B.; Roitberg, A.;
Behavior of Antifreeze Proteins: Ice Growth Inhibitors and Ice Simmerling, C. Comparison of multiple Amber force fields and
Nucleation Promoters. J. Phys. Chem. Lett. 2019, 10, 966−972. development of improved protein backbone parameters. Proteins
(20) Meister, K.; Moll, C. J.; Chakraborty, S.; Jana, B.; DeVries, A. 2006, 65, 712−725.
L.; Ramløv, H.; Bakker, H. J. Molecular structure of a hyperactive (41) Lindorff-Larsen, K.; Piana, S.; Palmo, K.; Maragakis, P.; Klepeis,
antifreeze protein adsorbed to ice. J. Chem. Phys. 2019, 150, 131101. J. L.; Dror, R. O.; Shaw, D. E. Improved side-chain torsion potentials
(21) Liu, K.; Wang, C.; Ma, J.; Shi, G.; Yao, X.; Fang, H.; Song, Y.; for the Amber ff99SB protein force field. Proteins 2010, 78, 1950−
Wang, J. Janus effect of antifreeze proteins on ice nucleation. Proc. 1958.
Natl. Acad. Sci. U. S. A. 2016, 113, 14739−14744. (42) Mahoney, M. W.; Jorgensen, W. L. A five-site model for liquid
(22) Nutt, D. R.; Smith, J. C. Dual Function of the Hydration Layer water and the reproduction of the density anomaly by rigid,
around an Antifreeze Protein Revealed by Atomistic Molecular nonpolarizable potential functions. J. Chem. Phys. 2000, 112, 8910−
Dynamics Simulations. J. Am. Chem. Soc. 2008, 130, 13066−13073. 8922.
(23) Midya, U. S.; Bandyopadhyay, S. Hydration Behavior at the Ice- (43) Humphrey, W.; Dalke, A.; Schulten, K. VMD: Visual molecular
Binding Surface of the Tenebrio molitor Antifreeze Protein. J. Phys. dynamics. J. Mol. Graphics 1996, 14, 33−38.
Chem. B 2014, 118, 4743−4752. (44) Leinala, E. K.; Davies, P. L.; Jia, Z. Crystal Structure of β-
(24) Brotzakis, Z. F.; Voets, I. K.; Bakker, H. J.; Bolhuis, P. G. Water Helical Antifreeze Protein Points to a General Ice Binding Model.
structure and dynamics in the hydration layer of a type III anti-freeze Structure 2002, 10, 619−627.
protein. Phys. Chem. Chem. Phys. 2018, 20, 6996−7006. (45) Hegde, S. S.; Vetting, M. W.; Roderick, S. L.; Mitchenall, L. A.;
(25) Modig, K.; Qvist, J.; Marshall, C. B.; Davies, P. L.; Halle, B. Maxwell, A.; Takiff, H. E.; Blanchard, J. S. A Fluoroquinolone
High water mobility on the ice-binding surface of a hyperactive Resistance Protein from Mycobacterium tuberculosis That Mimics
antifreeze protein. Phys. Chem. Chem. Phys. 2010, 12, 10189−10197. DNA. Science 2005, 308, 1480.
(26) Hudait, A.; Moberg, D. R.; Qiu, Y.; Odendahl, N.; Paesani, F.; (46) Darden, T.; York, D.; Pedersen, L. Particle mesh Ewald: An N·
Molinero, V. Preordering of water is not needed for ice recognition by log(N) method for Ewald sums in large systems. J. Chem. Phys. 1993,
hyperactive antifreeze proteins. Proc. Natl. Acad. Sci. U. S. A. 2018, 98, 10089.
115, 8266. (47) Bussi, G.; Donadio, D.; Parrinello, M. Canonical sampling
(27) Meister, K.; Lotze, S.; Olijve, L. L. C.; DeVries, A. L.; Duman, J. through velocity rescaling. J. Chem. Phys. 2007, 126, No. 014101.
G.; Voets, I. K.; Bakker, H. J. Investigation of the Ice-Binding Site of (48) Rodger, P. M.; Forester, T. R.; Smith, W. Simulations of the
an Insect Antifreeze Protein Using Sum-Frequency Generation methane hydrate/methane gas interface near hydrate forming
Spectroscopy. J. Phys. Chem. Lett. 2015, 6, 1162−1167. conditions conditions. Fluid Phase Equilib. 1996, 116, 326−332.
(28) Duboué-Dijon, E.; Laage, D. Comparative study of hydration (49) Reinhardt, A.; Doye, J. P. K.; Noya, E. G.; Vega, C. Local order
shell dynamics around a hyperactive antifreeze protein and around parameters for use in driving homogeneous ice nucleation with all-
ubiquitin. J. Chem. Phys. 2014, 141, 22D529. atom models of water. J. Chem. Phys. 2012, 137, 194504.
(29) Garnham, C. P.; Campbell, R. L.; Davies, P. L. Anchored (50) Parui, S.; Jana, B. Factors Promoting the Formation of
clathrate waters bind antifreeze proteins to ice. Proc. Natl. Acad. Sci. U. Clathrate-Like Ordering of Water in Biomolecular Structure at
S. A. 2011, 108, 7363−7367. Ambient Temperature and Pressure. J. Phys. Chem. B 2019, 123, 811−
(30) Arai, T.; Nishimiya, Y.; Ohyama, Y.; Kondo, H.; Tsuda, S. 824.
Calcium-Binding Generates the Semi-Clathrate Waters on a Type II (51) Steinhardt, P. J.; Nelson, D. R.; Ronchetti, M. Bond-
Antifreeze Protein to Adsorb onto an Ice Crystal Surface. Biomolecules orientational order in liquids and glasses. Phys. Rev. B 1983, 28,
2019, 9, 162. 784−805.
(31) Mahatabuddin, S.; Fukami, D.; Arai, T.; Nishimiya, Y.; Shimizu, (52) García Fernández, R.; Abascal, J. L. F.; Vega, C. The melting
R.; Shibazaki, C.; Kondo, H.; Adachi, M.; Tsuda, S. Polypentagonal point of ice Ih for common water models calculated from direct
ice-like water networks emerge solely in an activity-improved variant coexistence of the solid-liquid interface. J. Chem. Phys. 2006, 124,
of ice-binding protein. Proc. Natl. Acad. Sci. U. S. A. 2018, 115, 5456. 144506.

4695 https://dx.doi.org/10.1021/acs.jpcb.0c01206
J. Phys. Chem. B 2020, 124, 4686−4696
The Journal of Physical Chemistry B pubs.acs.org/JPCB Article

(53) Fidler, J.; Rodger, P. M. Solvation Structure around Aqueous


Alcohols. J. Phys. Chem. B 1999, 103, 7695−7703.
(54) Nandi, N.; Bhattacharyya, K.; Bagchi, B. Dielectric Relaxation
and Solvation Dynamics of Water in Complex Chemical and
Biological Systems. Chem. Rev. 2000, 100, 2013−2046.
(55) Nandi, N.; Bagchi, B. Dielectric Relaxation of Biological Water.
J. Phys. Chem. B 1997, 101, 10954−10961.
(56) Qiao, B.; Jiménez-Á ngeles, F.; Nguyen, T. D.; Olvera de la
Cruz, M. Water follows polar and nonpolar protein surface domains.
Proc. Natl. Acad. Sci. U. S. A. 2019, 116, 19274.
(57) Meister, K.; Strazdaite, S.; DeVries, A. L.; Lotze, S.; Olijve, L. L.
C.; Voets, I. K.; Bakker, H. J. Observation of ice-like water layers at an
aqueous protein surface. Proc. Natl. Acad. Sci. U. S. A. 2014, 111,
17732−17736.
(58) Kuffel, A.; Czapiewski, D.; Zielkiewicz, J. Unusual dynamic
properties of water near the ice-binding plane of hyperactive
antifreeze protein. J. Chem. Phys. 2015, 143, 135102.

4696 https://dx.doi.org/10.1021/acs.jpcb.0c01206
J. Phys. Chem. B 2020, 124, 4686−4696

You might also like