Heat Transfer in Protein-Water Interfaces
Heat Transfer in Protein-Water Interfaces
Heat Transfer in Protein-Water Interfaces
temperature relaxation process using the solution of the heat diffusion equation we compute the
thermal conductivity and thermal diffusivity of the proteins, as well as the thermal conductance of
the protein–water interface. Our results indicate that the temperature relaxation of the protein can
Downloaded by University of Memphis on 08 October 2012
1610 | Phys. Chem. Chem. Phys., 2010, 12, 1610–1617 This journal is
c the Owner Societies 2010
View Online
organic solvent indicate that a significant amount of dissipation ATPase was the same as that of the vesicle interior. This
can occur through the molecule–solvent interface.7 The time assumption remains unsupported. Hence, in order to further
scale associated with such processes has also been estimated, establish the validity of these assumptions it is necessary to
being of the order of tens of picoseconds in myoglobin.12 This investigate the thermal properties of proteins, and study their
time scale is not far from the one associated to the relaxation dissipation of energy.
through the biomolecular structure, indicating that a vibrational The paper is structured as follows. Firstly we discuss the
coupling between biomolecular and solvent degrees of freedom methodology employed in this work; the transient non-
may take place.13 equilibrium molecular dynamics simulations and the equations
Computer simulation techniques are being used to investigate employed to analyze the temperature relaxation in the proteins.
the transport properties as well as vibrational energy pathways The discussion of our results including the heat capacity of the
of proteins.6,7,10,12,14–28 These studies have reported vibra- proteins as well as the thermal conductivity and thermal
tional energy relaxation times between one to several tens of conductance follows. We finish the paper with the main
picoseconds. The first simulation study of vibrational energy conclusions obtained in this work.
Published on 11 January 2010 on http://pubs.rsc.org | doi:10.1039/B918607G
This journal is
c the Owner Societies 2010 Phys. Chem. Chem. Phys., 2010, 12, 1610–1617 | 1611
View Online
Using the initial condition, T(r, t = 0) = Ti, the solution of estimates for A and B. l1 is obtained from A and t from B. The
eqn (2) is39 thermal conductivity follows from
Tðr;tÞ Tf
Ti Tf rcp R2 3Cp
k¼ ¼ ð9Þ
2 ð3Þ t 4pRt
X
1
sinðln Þ ln cosðln Þ l t R ln r
¼ 4 exp n sin ;
n¼1
2ln sinð2ln Þ t ln r R where R is the radius of gyration of the protein, r is the protein
2 density, which we define as r = m/V, where m is the protein
where t = R /DT, DT is the thermal diffusivity, DT = k/(rcp)
mass and V = 4pR3/3, Cp = mcp (in J K1). Eqn (9) follows
The coefficients, ln, fulfil the following relation,
from DT = R2/t = k/(rcp). Once k is known, the thermal
GR conductance (G) can be estimated using eqn (7). In addition to
1 ln cotðln Þ ¼ ¼ Bi; ð4Þ
k the analysis of the temperature relaxation, the computation
of the thermal conductivity requires the heat capacity, Cp, and
Published on 11 January 2010 on http://pubs.rsc.org | doi:10.1039/B918607G
of G E 102 MW K1 m2, typical of hydrocarbon–water thermal conductance of small alkane nanodroplets in water.30
interfaces,30,37 a protein radius of gyration of E2 nm and a Our results for this quantity when extrapolated to a droplet of
protein thermal conductivity of the order of 0.2 W K1 m1,23 infinite radius were consistent with the estimates obtained for
we get Bi E 1. As we will see below, the thermal conductance planar alkane–water interfaces using the stationary NEMD
of the protein–water interface is similar to that of the approach.37 This lends support to the consistency of our
alkane–water interface. Hence, our estimate of Bi E 1 indicates method.
that the thermal resistance of the protein and the protein– We note that the equations discussed above (3–6) are used
water interface are of the same order, and both the thermal to model heat relaxation in macroscopic objects too. We find
conductivity and the thermal conductance should be included that the relaxation in small objects such as proteins can be
in the diffusion equation to model the temperature relaxation described in the same way as in larger macroscopic objects.
process. One requirement for this to be true is that the velocity
The solution of eqn (3) can be simplified by taking the components of the atoms must follow a normal distribution,
average temperature over the whole volume, V, of the protein. so that we can define a proper microscopic temperature.
This results in Deviations from normality if present are expected to appear
more clearly in very small objects. Hence, to test this point we
hTðr; tÞiV Tf TðtÞ Ti have chosen the smallest protein investigated in this work,
¼
Ti Tf Ti Tf myoglobin. Fig. 1 shows the normalized probability velocity
X1 2
1 ðsinðln Þ ln cosðln ÞÞ2 ln t distribution as well as the quantile–quantile (Q–Q) plot
¼6 3 l sinðl Þ cosðl Þ
exp ; where we compare the velocity distribution generated in the
n¼1 ln n n n t
simulation with a theoretical normal distribution. For this
ð5Þ
analysis we considered the velocities of the protein carbon
which can be further simplified by retaining only the first term atoms. The linearity of the Q–Q plot indicates that our
in the series,39 which represents a good approximation for simulation data are normally distributed. Application of the
t > 0.2 t, Shapiro–Wilk method (using 4000 samples)40,41 did not show
evidence of deviations from normality.
2
TðtÞ Tf 1 ðsinðl1 Þ l1 cosðl1 ÞÞ2 l1 t
¼6 3 exp ; ð6Þ Computer simulations of the proteins
Ti Tf l1 l1 sinðl1 Þ cosðl1 Þ t
We have investigated three protein structures, myoglobin
with
(1MBS entry in the Protein Data Bank),42 green fluorescence
GR protein (1QXT),43 and two different states of the Ca2+-
1 l1 cotðl1 Þ ¼ ¼ Bi: ð7Þ ATPase (1SU44 and 1KJU44). The Ca2+-ATPase enzymes
k
are P-type ion pumps that feature in active transport of
Eqn (6) is more convenient for the analysis of the simulation
Ca2+ across the cellular membrane. The catalytic cycle in-
results and is the one we have employed in this work. From the
volves several conformations. 1SU4 is a conformation where
fitting of the simulation results to eqn (6) we can obtain both
Ca2+ is present (E1 state), whereas 1KJU is a conformation
the l1 parameter appearing in eqn (7) and t. The fitting is
corresponding to the Ca2+-free (E2) state. It has been recently
performed as follows. We rewrite eqn (6) as:
found that ATPases release heat when they are transporting
TðtÞ Tf Ca2+,8 hence knowledge of the thermal transport properties is
ln ¼ A t=B ð8Þ needed to characterize the heat transfer in these enzymes.
Ti Tf
Snapshots of the proteins are given in Fig. 2. It can be seen
where A includes the sin and cos terms and B = t/l21. The that Ca2+-ATPases undergo dramatic conformational
fitting of the temperature relaxation data to eqn (8) provides changes during the enzymatic cycle.
1612 | Phys. Chem. Chem. Phys., 2010, 12, 1610–1617 This journal is
c the Owner Societies 2010
View Online
This journal is
c the Owner Societies 2010 Phys. Chem. Chem. Phys., 2010, 12, 1610–1617 | 1613
View Online
was set to 350 K. The radius of gyration for two of the proteins has capacity which amounts to about 15% of the heat capacity of
been shifted upwards to facilitate the comparison of the results. the proteins in their native states. Table 1 compares our results
with experimental data available in the literature, which were
system was allowed to relax. The temperature of the water bath obtained using calorimetric techniques. We have also included
was maintained constant during the relaxation process using a simulation data from Yu and Leitner,23 who used a computa-
temperature coupling constant of 0.1 ps. We performed several tional approach to estimate the heat capacities of myoglobin
tests to assess the effect of the temperature coupling constant on and GFP. Our results for the heat capacity of myoglobin are in
the protein relaxation process. We found that the results were good agreement with the experimental results of this protein in
insensitive within statistical uncertainty to the value set for the solution. The estimates from Yu and Leitner are slightly below
coupling constant. We performed an additional test using a the experimental result. We note that these simulations were
Nosé-Hoover (NH) thermostat to assess the influence of the performed in anhydrous conditions.
thermostat. We found that both the Berendsen and NH Our results show a significant increase in the heat capacity
reproduced the same relaxation behavior for the temperature. in moving from myoglobin to GFP and to the Ca2+-ATPases,
The relaxation process was investigated for about 100 ps. The both containing a larger number of carbon atoms, E103 and
equipartition principle expression was employed to compute the E5 103, respectively, as compared with myoglobin
temperature of the protein as a function of time. The temperature E8 102. The heat capacity is an extensive property for
was then fitted to eqn (6) using the Levenberg–Marquadt macroscopic systems, but for small systems consisting of
algorithm.51 The results reported below were obtained from an a few atoms the heat capacity is non extensive.55 Hence our
analysis performed over three independent runs. computations of the heat capacity of single proteins can
The computation of the thermal conductance requires provide a test to assess the extensivity of this property in
previous knowledge of the effective radius of the proteins proteins of varying molecular weight. A representation of the
(radius of gyration) and the heat capacity (see eqn (9)). We experimental heat capacity of the protein as a function of
found that the radius of gyration of the protein fluctuates the number of carbon atoms offers an approach to test the
around a well defined average during the 100 ps duration of consistency of the simulated heat capacities reported above
the relaxation process (see Fig. 4). (see Fig. 5). We note that more involved approaches are
The heat capacity of the proteins was estimated using possible. Gomez et al.53 considered the molecular weight and
constant temperature runs of the protein immersed in water the accessible surface of the proteins, in order to fit heat
using in this case periodic boundary conditions. The heat capacities of proteins in the non-native state. Fig. 5 clearly
capacity was calculated from direct numerical differentiation shows that the experimental results follow a linear dependence
of the internal energy of the protein using temperatures with the number of carbon atoms, as would be expected from
around the temperature of interest, T = 300 K. Typically the extensive character of the heat capacity. Our simulation
we used 5 temperatures, 310, 305, 300, 295 and 290 K, to results conform to this linear behavior and in general are in
calculate the derivative of the internal energy. Each simulation good agreement with the experimental measurements. To the
was 1.2 ns long. All simulations were performed using the best of our knowledge there are no experimental results of the
GROMACS simulation package.52 heat capacity of Ca2+-ATPases. We find that the heat capacity
of these proteins is of the order of 180 kJ mol1 K1. This
result agrees well with the heat capacity obtained from the
Results extrapolation of the experimental results for smaller proteins
(cf. Fig. 5). Within the statistical accuracy of our computations
Heat capacity of proteins
the heat capacities of the two Ca2+ATPase conformations,
The heat capacity plays an important role in determining 1SU4 and 1KJU, are the same, indicating the heat capacity is
the energetics associated to protein folding processes. This not sensitive to the structural detail of the proteins, and
1614 | Phys. Chem. Chem. Phys., 2010, 12, 1610–1617 This journal is
c the Owner Societies 2010
View Online
Published on 11 January 2010 on http://pubs.rsc.org | doi:10.1039/B918607G
number of carbon atoms in the protein. Experiments 298 K, simula- (bottom-left) and ATPase-1KJU (bottom-right). Ti and Tf represent
tions 300 K. The experimental values (full circles) are taken from the initial and solvent temperatures, 350 and 250 K, respectively. The
ref. 53. The labels denote different proteins; bovine pancreatic trypsin dashed line represents the fitting of the temperature relaxation to
inhibitor (BPTI), ubiquitin (UB), rybonuclease T1 (RT1), lysozyme eqn (6).
(LYS), staphylococcal nuclease (SN), myoglobin (MY), bovine
chymotrypsinogen A (BCA), green fluorescence protein (GFP) and
conductivity and the thermal conductance of the protein–
Ca2+-ATP-ases. The squares represent the simulation results obtained
in this work and the line is a linear fitting to the experimental data.
water interface.
Fig. 7 shows the temperature relaxation results along with
the corresponding fitting using eqn (6). Our model (eqn (6)),
suggesting that the number of degrees of freedom for both which takes into account the different thermal conductivities
conformations is very similar. of the solvent and the protein, provides a good representation
of the temperature relaxation. This is particularly clear for the
Thermal conductance and thermal conductivity of proteins larger proteins (ATPase) where the temperature fluctuations
Fig. 6 shows a representative temperature relaxation curve are significantly reduced as compared to the smaller proteins
for one of the proteins investigated in this work (Ca2+- (cf. Fig. 7).
ATPase-1SU4). All the relaxation processes investigated here Table 2 contains numerical data of the heat transport
involved a time span of about 100 ps. We find that the time properties of the proteins for (Ti, Ts) = (350, 250) K. To test
scale for temperature relaxation is tens of picoseconds. the sensitivity of our results to the simulation conditions we
Assuming, as a first approximation, an exponential decay of performed simulations with different initial and solvent
the temperature with time, exp(t/t) we can estimate a temperatures, (400, 300) K and (350, 250) K. We obtained
characteristic relaxation time, t, for the cooling process. This the same results within the statistical accuracy of our method.
is of the order of 10–20 ps for the proteins investigated in this Hence we report in Table 2 the results for (350, 250) K.
work. We have fitted the simulated temperature relaxation The thermal diffusivities obtained in this work are of the
data to eqn (6) to quantify the thermal diffusivity, thermal order expected for biological tissue56 and they are also in the
range of the thermal diffusivities of small alkane droplets in
water 5–12 Å2 ps1. Our results for the smaller proteins,
myoglobin and GFP, E5 Å2 ps1 are similar to the thermal
diffusivity of Rhodospseudomonas viridis, 7 Å2 ps1, obtained
by different authors14 using a transient non-equilibrium
approach similar to ours. Yu and Leitner23 have employed
more recently a different approach, based on the scaling of the
Protein D k G R lK
Myoglobin 41 0.13 0.02 100 5 1.529 0.003 1.3
Fig. 6 Temperature relaxation of Ca2+-ATPase (1SU4) as a function GFP 31 0.12 0.01 270 20 1.667 0.002 0.4
of time. The dashed line represents the fitting of the temperature ATPase-1SU4 17 2 0.22 0.02 260 70 3.821 0.004 0.9
ATPase-1KJU 16 2 0.23 0.03 210 40 3.69 0.01 1
relaxation to eqn (6).
This journal is
c the Owner Societies 2010 Phys. Chem. Chem. Phys., 2010, 12, 1610–1617 | 1615
View Online
energy diffusion coefficient with the vibrational mode frequency advanced in the discussion of the Biot number (cf. eqn (4)).
of the protein, to estimate the thermal diffusivity of myoglobin The Biot number can be rewritten in terms of the Kapitza
and GFP. These authors equilibrated the proteins in water length as, Bi = R/lK. Hence, for Bi E 1 ) lK E R, and both
before computing the transport properties in vacuo. Our the thermal conductance and the thermal conductivity must be
results lie below the ones reported by these authors, 21.1 included in the description of the thermal relaxation.
and 18.7 Å2 ps1 for GFP and myoglobin, respectively,
although they are comparable in magnitude.
Conclusions and final remarks
The thermal conductivities of the different proteins
are comparable. We obtain values of the order of We have performed an investigation of the thermal properties
0.15–0.2 W K1 m1. These values are similar to the ones of three structurally different proteins; myoglobin, GFP and
obtained in small hydrocarbon droplets,30 and again they are Ca2+-ATPase in water. Myoglobin and ATPases control
comparable to the thermal conductivity reported by Yu and different biological processes, oxygen transport, and active
Leitner, 0.27 W K1 m1 using a different simulation transport of Ca2+ across the cellular membrane, respectively.
Published on 11 January 2010 on http://pubs.rsc.org | doi:10.1039/B918607G
approach. Our results show that all the proteins investi- These proteins are dramatically different in structure and size,
gated in this work are worse heat conductors than water 17951 vs. 109706 Da. In order to obtain information on the
(k E 0.6 W K1 m1) and consequently they should develop transport properties of these biomolecules we have performed
larger temperature gradients as compared with bulk water. transient non-equilibrium molecular dynamics simulations,
Downloaded by University of Memphis on 08 October 2012
The analysis of the temperature relaxation using the whereby we model the temperature relaxation of the protein
solution of the heat diffusion equation, (cf. eqn (6)) provides towards the temperature of the surrounding bath. The analysis
a route to estimate the thermal conductance of the protein– of the temperature relaxation with time offers an approach to
water interface. We note that the protein surface is hetero- extract the thermal conductivity and thermal diffusivity
geneous in chemical composition, and irregular in shape, of the protein as well as the thermal conductance of the
consisting of regions of different curvature. Because the protein–water interface. With this purpose we have considered
complex surface topography of the proteins, consisting of a solution of the diffusion equation that incorporates both the
curved and roughly planar areas, our results for the thermal thermal conductivity of the protein and the thermal conductance
conductance have to be interpreted as illustrative of the as fitting parameters.
average or effective thermal conductance of the protein–water We find that the thermal conductivity of the proteins is low,
interface. Our thermal conductances of the protein–water 0.1–0.2 W K1 m1 as compared with the thermal conductivity
interfaces 100–250 MW K1 m2, are similar to the ones of water, 0.6 W K1 m1, hence proteins should be able to
reported by us for hydrocarbon nanodroplets immersed in sustain large thermal gradients across their structure. This may
water, which are higher than the thermal conductances of have implications on biological processes involving molecular
planar alkane–water interfaces, 65 MW K1 m2.37 In motors such as Ca2+-ATPase. Following our estimates of the
particular, our estimate of the thermal conductance of GFP thermal conductivity, it seems feasible that these proteins are
is in excellent agreement with recent results reported for this capable of maintaining temperature gradients under stationary
protein using a different simulation approach, (i.e. stationary conditions. According to the non-equilibrium thermo-
non-equilibrium simulations), and a different protein force- dynamics theory31–33 the development of such temperature
field (i.e. full atom model).28 These thermal conductances gradient could then, due to reciprocity of phenomena, trigger
are of the order of the ones observed in water–hydrophilic uptake of Ca2+ as well as ATP-hydrolysis, pointing to a
interfaces,57 indicating a strong thermal coupling between possible role of the Ca2+-ATPase as a temperature regulator,
water and the protein surface. We have recently suggested as proposed by de Meis and coworkers.58
that the thermal conductances of curved interfaces can be Within the statistical uncertainty of our computations we do
higher than those of planar interfaces.30 This is a physical not find significant differences in the thermal conductivities of
effect that goes beyond the chemical nature of the surface. the three proteins investigated in this work. The thermal
Although it is not possible to disentangle surface topographic diffusivity seems to be more sensitive to the protein nature
(curvature) and chemical (hydrophilic and hydrophobic though, with an increase in this property of about three times
patches) effects easily, we expect that similar curvature effects in going from the smaller proteins (myoglobin and GFP) to
may apply to the protein–water interface, and hence, the the larger ones (Ca2+-ATPase). The thermal conductances of
corresponding thermal conductance may be higher than that the three proteins vary from 100 to 270 MW K1 m2,
of the corresponding planar interface. myoglobin showing the lower thermal conductance of all the
Finally, using our simulation results we have estimated the systems investigated. Our results for the heat capacity show
Kaptiza length, which represents the thickness of a layer with good agreement with available experimental data of proteins
thermal conductivity, k, which has thermal conductance, G, obtained from calorimetric measurements. In addition to the
i.e., the same as the protein–water interface. The typical values proteins listed above we performed simulations of the heat
for hydrophobic interfaces are of the order of 10 nm, as capacity of lysozyme, rybonuclease T1, and bovine chymo-
compared with 1 nm for hydrophilic interfaces.57 We find trypsinogen (BCA). All are in good agreement with the
Kapitza lengths of the order of E1 nm, i.e., of the order of the experiment, with BCA showing the largest deviations from
radius of gyration of the proteins (see Table 2), showing that the experimental data. The heat capacity correlates linearly with
the interfacial conductance is important for defining the the number of carbon atoms of the proteins, as expected for an
thermal transport of the biomolecules. This idea has been extensive property. Using this result and the experimental data of
1616 | Phys. Chem. Chem. Phys., 2010, 12, 1610–1617 This journal is
c the Owner Societies 2010
View Online
small proteins, we have estimated the heat capacity of the Ca2+- 18 D. M. Leitner, Phys. Rev. Lett., 2001, 87, 188102.
ATPase. The extrapolated result is in excellent agreement with 19 D. E. Sagnella, J. E. Straub and D. Thirumalai, J. Chem. Phys.,
2000, 113, 7702–7711.
our computational estimate of 180 kJ mol1 K1. To the best of 20 D. E. Sagnella and J. E. Straub, J. Phys. Chem. B, 2001, 105,
our knowledge this is the first computational estimate of the heat 7057–7063.
capacity of this important protein. 21 K. Moritsugu, O. Miyashita and A. Kidera, J. Phys. Chem. B,
Our work, in conjunction with previous investigations, 2003, 107, 3309–3317.
22 X. Yu and D. M. Leitner, J. Phys. Chem. B, 2003, 107, 1698–1707.
offers a consistent image of the heat transport in biomolecules. 23 X. Yu and D. M. Leitner, J. Chem. Phys., 2005, 122, 054902.
The thermal conductance of the water–protein interface 24 N. Ota and D. A. Agard, J. Mol. Biol., 2005, 351, 345–354.
should play a relevant role in controlling the energy relaxation 25 T. Ishikura and T. Yamato, Chem. Phys. Lett., 2006, 432, 533–537.
26 K. Sharp and J. J. Skinner, Proteins: Struct., Funct., Bioinf., 2006,
of the protein. Overall, our results show that the water–protein 65, 347–361.
interface has a high conductance, similar to the one measured 27 M. Takayanagi, H. Okumura and M. Nagaoka, J. Phys. Chem. B,
in water adsorbed at hydrophilic surfaces.57 We suggest that in 2007, 111, 864–869.
addition to the heterogeneous chemical structure of the 28 N. Shenogina, P. Keblinski and S. Garde, J. Chem. Phys., 2008,
129, 155105.
Published on 11 January 2010 on http://pubs.rsc.org | doi:10.1039/B918607G
proteins, with hydrophobic and hydrophilic patches, other 29 M. J. Assael, E. Charitidou, C. A. N. de Castro and
factors such as the surface topography, characterized by highly W. A. Wakeham, Int. J. Thermophys., 1987, 8, 663–670.
curved interfaces, may be contributing to the enhancement of the 30 A. Lervik, F. Bresme and S. Kjelstrup, Soft Matter, 2009, 5,
thermal conductance of these nanoscale interfaces. We believe 2407–2414.
Downloaded by University of Memphis on 08 October 2012
This journal is
c the Owner Societies 2010 Phys. Chem. Chem. Phys., 2010, 12, 1610–1617 | 1617