Citation: Tika Ram Lamichhane, Hari Prasad Lamichhane. Heat conduction by thyroid hormone receptors[J]. AIMS Biophysics, 2018, 5(4): 245-256. doi: 10.3934/biophy.2018.4.245
[1] | Tika Ram Lamichhane, Hari Prasad Lamichhane . Structural changes in thyroid hormone receptor-beta by T3 binding and L330S mutational interactions. AIMS Biophysics, 2020, 7(1): 27-40. doi: 10.3934/biophy.2020003 |
[2] | David E. Shoup . Diffusion-controlled reaction rates for clusters of binding sites on a cell. AIMS Biophysics, 2016, 3(4): 522-528. doi: 10.3934/biophy.2016.4.522 |
[3] | Ivo Spicka, Ondrej Krejcar, Robert Frischer, Pavol Kostial, Ali Selamat, Zora Jancikova, Kamil Kuca . Utilization of linearization methods for measuring of thermal properties of materials. AIMS Biophysics, 2018, 5(4): 257-271. doi: 10.3934/biophy.2018.4.257 |
[4] | Haliz Hussein, Hazhmat Ali, Zeki Mohamed, Majeed Mustafa, Khairi Abdullah, Asaad Alasady, Mayada Yalda . Thyroid function and hematological alterations in cardiac catheterization workers: a pre-post observational study on x-ray exposure. AIMS Biophysics, 2025, 12(1): 43-53. doi: 10.3934/biophy.2025004 |
[5] | Panisak Boonamnaj, Pornthep Sompornpisut, Piyarat Nimmanpipug, R.B. Pandey . Thermal denaturation of a coronavirus envelope (CoVE) protein by a coarse-grained Monte Carlo simulation. AIMS Biophysics, 2022, 9(4): 330-340. doi: 10.3934/biophy.2022027 |
[6] | Tomasz Rok, Artur Kacprzyk, Eugeniusz Rokita, Dorota Kantor, Grzegorz Tatoń . Quantitative assessment of thermal effects on the auricle region caused by mobile phones operating in different modes. AIMS Biophysics, 2024, 11(4): 427-444. doi: 10.3934/biophy.2024023 |
[7] | Dina Falah Noori Al-Sabak, Leila Sadeghi, Gholamreza Dehghan . Predicting the correlation between neurological abnormalities and thyroid dysfunction using artificial neural networks. AIMS Biophysics, 2024, 11(4): 403-426. doi: 10.3934/biophy.2024022 |
[8] | Alexander Galashev . Computer implementation of the method for electrolytic production of thin films for biomedical applications: short review. AIMS Biophysics, 2024, 11(1): 39-65. doi: 10.3934/biophy.2024004 |
[9] | Ekaterina Bereznyak, Natalia Gladkovskaya, Evgeniy Dukhopelnykov, Anastasiya Gerus, Anastasiya Lantushenko, Maxim Evstigneev . Thermal analysis of ligand-DNA interaction: determination of binding parameters. AIMS Biophysics, 2015, 2(4): 423-440. doi: 10.3934/biophy.2015.4.423 |
[10] | Karina Kwapiszewska, Kamil Żukowski, Radosław Kwapiszewski, Zbigniew Brzózka . Double casting prototyping with a thermal aging step for fabrication of 3D microstructures in poly(dimethylsiloxane). AIMS Biophysics, 2016, 3(4): 553-562. doi: 10.3934/biophy.2016.4.553 |
Thyroid signaling defects broadly depend on the action of thyroid hormone receptors (THRs), local ligands, thyroid hormone (TH) carrier proteins, coactivators, corepressors, neural development and metabolism [1]. There are two THR subtypes (THR-α and THR-β) consisting of ligand binding domain (LBD) and homogeneous DNA-binding domain (DBD). The LBDs of these subtypes differ only by a single amino acid residue: Ser277 in THR-α and Asn331 in THR-β [2]. THR isoforms differ in chain length by amino as well as carbxy termini. THR-α1 (triiodothyronine: T3-binding splice product of THR-α) is expressed in skeletal muscle, heart and brain. THR-β has three T3-binding splice products: THR-β1 is expressed in wide regions of body tissues; THR-β2 is in retina, brain and inner ear; and THR-β3 is predominant in liver, kidney and lung [1]. THs interact with different signalling pathways and their fuction is based on iodine and nutritional status. The TH-liganded isoforms: THR-β, THR-β1 and THR-α in cartoon view of globular forms are shown in Figure 1. The mutational evidances found in THR isoforms indicate thyroid disorders with resistance to THs (RTH) [1,3]. THR isoform selective agonists interfere with the carboxyterminal helix 12. Helix 12 dynamics is responsible for the interaction of THR with corepressors and coactivators. T3 has higher binding affinity than thyroxine (T4) in THR and the T3 bounded isoforms regulate gene expressions under DNA transcription. The unliganded THR represses basal transcription which is known as transcriptional silencing. The abnormal changes in serum TH concentrations are the indicators of hypothyroidism or hyperthyroidism with subacute thyroiditis which are directly or indirectly associated with T3 regulated gene expressions [4].
Biologically active molecules are responsible for the heat conduction after the sudden increase in temperature by means of chemical or photochemical reactions. Heat energy transfers through the vibrational states and the residues of globular proteins giving rise to anisotropic flow of energy. It is central to understantand the protein function including ligand binding and dissociation [5]. Heat capacity of a protein-hormone system at a particular temperature depends on folding and unfolding states, polar and non-polar groups in solution, hydration states, and protein-ligand interactions [6]. The heat capacity of a native globular protein is close to the value found from the heat capacities of individual amino acid residues in an extended polypeptide chain [7]. Moreover, T3-liganded and unliganded THRs have the different heat capacities. Anharmonicity enhances thermal conductivity of proteins by the energy flow through localized normal modes of vibration [8]. Heat capacity and thermal conductivity being the temperature dependent quantities, thermal diffusivity also depends on temperature. The temperature gradient differs from chemically active cites to LBDs providing microscopic basis for the signal transduction. The energy fluctuation and dissipation or the temperature dependent changes in internal energy during molecular dynamics simulations (MDS) support to calculate heat capacity of the biomolecular system. The mean energy contributing heat capacity of such complex system is calculated from the ensemble averages of individual terms in potential function describing the system [6]. The temperature relaxation across protein-water interface is the basis of finding thermal diffusivity and then thermal conductivity [9,10]. This process relies on the principles of transient non-equilibrium thermodynamics.
Enthalpy or mean energy (E) is expressed in terms of ensemble averages of energy terms (Ui) describing a complex molecular system [6], i.e.
E=〈U〉=∑Ni=1〈Ui〉=∑Ni=1Ei | (1) |
where i refers to the energy related to bond angle, bond length, dihedral, van der Waal, electrostatic, improper and cross-terms giving rise to Hamiltonian or net potential function (U). Each of these terms contributes heat capacity of the system. The terms 〈Ui〉 differ in case of any perturbation, or mutation, or liganded and unliganded THR so that the total energy obtained is used to compute Boltzmann weight. The heat capacity, using the method of energy fluctuation in canonical ensemble [6,11], is defined by:
CV=dEdT=d〈U〉dT=〈δU2〉KBT2=〈U2〉−〈U〉2KBT2=mcV | (2) |
where m is mass and cV is specific heat capacity of the system, KB = 0.00198657 kcal mol−1K−1 is Boltzmann constant, and T is absolute temperature.
The possible sources of change in heat capacity are protein folding and unfolding including polar and apolar hydration contributions, equilibrium effect, electrostatics, vibrational terms, van der Waal interactions, H-bonding and protein conformational entropy. For the protein in solution, the heat capacity changes due to protein-protein interactions as well as hydration effect. After solvating the globular form of protein in a neutral water-ion solution providing the cellular environment, one can determine the heat capacity of the solvated protein by method of mixture or mole fraction [12,13]. Under thermal equilibrium, McV=MPcpV+MWcwV where the suffix “p” refers to protein-ligands and “w” refers to water-ions. Then, the heat capacity of the protein-hormone system in solution is given by:
cpV=McV−MWcwV MP | (3) |
Knowing the relative masses and the specific heat capacities of mixture and solution, we can determine the specific heat capacity of the protein-hormone system under ideal condition. For this purpose, we perform MDS approach separately for the solvated system or mixture and for the water-ion solution at constant temperature.
The properties of the biological macromolecules explain their functional impact on the living organisms in response to an external perturbation such as thermal gradient. The linear response theory describes the small perturbation relating non-equilibrium transport to equilibrium thermodynamic properties of the system. The complex biomolecular system is solvated under non-periodic boundary conditions forming a spherical droplet. The local temperature T(r,t) of the system depends on position (r) and time (t) which is governed by the heat diffusion equation:
∂T(r,t)∂t=D∇2T(r,t) | (4) |
where D is thermal diffusivity. For the radius R of the solvated protein-ligand in a globular form, the boundary conditions are: T(r,0)=Tifor r<R,and T(r,t)=Tffor r>R where Ti is initial temperature of the system and Tf is final temperature of the surrounding fluid-shell of thickness Δr. Using these boundary conditions, solution of the
〈T(t)〉=Tf+6Ti−Tfπ2∑∞n=11n2exp[−(nπR)2Dt] | (5) |
Knowing the thermal diffusivity (D) by fitting the simulated data with
k=ρcVD | (6) |
Crystal structures of three THR isoforms consisting of TH-ligands (T3 and/or T4): THRT3-β, THRT3-β1 and THRT3T4-α were obtained from the protein data bank (pdb) codes 3GWS [15], 1XZX [16] and 4LNX [17], respectively. CHARMM-force field topologies and parameters [18,19,20] were implemented to prepare the simulation packages for solvation, energy minimization, equilibration and production runs by using visual molecular dynamics (VMD) [21] and nanoscale molecular dynamics (NAMD) [22] interfaces. The ligands T3 and T4 were parameterized suitably [20] for the NAMD runs.
The separate THR isoforms: THR-β, THRT3-β, THRT3-β1 and THRT3T4-α were solvated fully in an explicit solvent of neutral water-ions providing the cellular environment. Here, THRT3T4-α was truncated in size by omitting initial residue IDs from 145 to 156 of almost free strand. At first, THR-β as well as THRT3-β was solvated under periodic boundary conditions in TIP3P-water box of cell basis vectors 50.36, 60.92 and 74.78 Å, neutrallized by adding Na+ and Cl− ions in the concentration of 0.15 mol/L, performed energy minimization up to 3000 conjugate gradient (CG) steps and then separate equilibration runs were performed up to 2 ns at five different constant temperatures 300, 305, 310, 315 and 320 K. The Langevin constant temperature and pressure (1 atm) controls were used with damping coefficient of 1 ps−1 and integrator papameter of 2 fs/step by keeping rigid bonds of H-atoms. The force field related switching, cut-off and pair list distances were 10, 12 and 14 Å, repectively with 1-4 scaling 1.0. The mean energy, 〈U〉 of THR-β and that of THRT3-β at the five different temperatures were noted down from the last 0.5 ns runs. The slope of straight line fit in 〈U〉 vs. T plot provides the heat capacity (CV) of a mixture of solvating T3-liganded or unliganded THR at the body scale temperature 310 K. The same procedure with identical terms and conditions was implemented to a neutral water-ion box of cell basis vectors 30, 45 and 60 Å to find the heat capacity (CwV) of the solvent only. It is to be noted that CwV found for the solvent (water) by keeping rigid bonds of H-atoms with 2 fs/step is contributed by the non-bond (electrostatic and van der Waal) energies only. Then, from
Each of the THR isoforms was solvated in a water-sphere (TIP3P water model) neutralized with Na+ and Cl− ions in the concentration of 0.15 mol/L to study their heat conduction properties such as heat capacity (CpV), thermal diffusivity (D) and thermal conductivity (k). During the solvation under non-periodic boundary conditions, the radii of THRT3-β, THRT3-β1 and THRT3T4-α related water-droplets were 37.48, 36.99 and 31.29 Å, respectively. The force-filed and integrator parameters, and Langevin dynamics for constant temperature control were fixed as mentioned in the periodic boundary conditions. Each of the systems was geometrically optimized up to the sufficient (2000-5000) CG steps and equilibrated up to 20 ns. Conformational stability of the system was monitored by plotting graphs for root mean square deviation (RMSD) and/or radius of gyration (RG) during equilibration as well as production runs. The systems were also subjected to the additional production runs of 2 ns each in NVT ensemble at 310 K to find the heat capacity (CV) by using energy fluctuations (
The final coordinates of each equilibrated droplet consisting of a THR isoform were taken after its 20 ns run at 310 K. The inner sphere (r < R) was kept at initial temperature, Ti = 310 K and the outer shell (r > R) was fixed at the final temperature, Tf = 200 K where R = 33 Å and Δr = 4.48 Å for THRT3-β; R = 33 Å and Δr = 3.99 Å for THRT3-β; and R = 27 Å and Δr = 4.29 Å for THRT3T4-α. Then, NAMD simulations of temperature relaxation were performed without any constraint imposed into the atomic vibrations to monitor the change in system's temperature. The cooling process was done up to 20 ps each. By fitting the temperature data with the theoretical expression in
The separately performed 20 ns equilibration runs at 310 K of THR-isoforms solvated in water-droplets result mean ± SD of protein's RMSD: 2.13 ± 0.38 Å for THRT3-β, 1.98 ± 0.21 Å for THRT3-β1, and 2.24 ± 0.31 Å for THRT3T4-α; and the related RGs are 18.74 ± 0.07, 18.84 ± 0.11, and 18.41 ± 0.06 Å, respectively. In case of unliganded THR-β, RG is 18.70 ± 0.09 Å and RMSD is 2.61 ± 0.50 Å during its 20 ns long equilibration. Almost constant values of RG with small SD indicate that the protein-hormone systems are conformationally stable during their equilibration runs as plotted as shown in Figure 2a. The small fluctuations seen in RMSD plots (Figure 2b) are due to formation and breaking of H-bonds among the more stable residues, α-helix and β-sheets of the protein systems while searching for their conformational stability during the course of simulations. The additional production runs up to 2 ns in canonical (NVT) ensemble yield more small values of RMSD: 1.23 ± 0.11, 1.41 ± 0.18 and 1.13 ± 0.11 Å for THRT3-β, THRT3-β1 and THRT3T4-α proteins, respectively.
Five NVT simulations up to 2 ns each at temperature difference of 5 K about the body scale temperature 310 K result the values of <U> for liganded and unliganded THR-β under the periodic boundary conditions. The slope of linear fit in <U> vs. T plot (Figure 3a) obtains the heat capacity, CV of the entire THRT3-β solvated water-box which is 84.59 ± 0.74 kcal mol−1K−1 at about 310 K. Since the total mass of THRT3-β solvated water-box is 131765 amu, the related specific heat, cV is 2688.52 ± 23.52 Jkg−1K−1. Similarly, CV of THR-β solvated water-box (mass = 131186 amu) is 85.59 ± 1.12 kcal mol−1K−1 and the related cV is 2732.31 ± 35.75 Jkg−1K−1. Upon implementing the same procedure with identical terms and conditions for the simulations of a neutral water-ion box with 2 fs/step and rigid bonds of H-atoms, its heat capacity, CwV contributed by electrostatic and van der Waal energy terms is 31.72 ± 0.43 kcal mol−1K−1. Since mass of the modelled water-ion box is 45357 amu, the related specific heat, cwV = 2928.75 ± 39.70 Jkg−1K−1. Figure 3b shows the best linear fit of <U> vs. T for the water-box where the slope of straight line gives CwV. The mass of THR-β is 27640 amu and that of THRT3-β is 28291 amu. By using
The heat capacities of the systems are also determined from the mean energy fluctuations in their canonical ensembles (
Molecule in water sphere | No. of atoms in mixture | Total mass, M (amu) | 〈U2〉 (k cal mol−1)2 |
〈U〉2 (k cal mol−1)2 |
CV=〈U2〉−〈U〉2KBT2 (k cal mol−1K−1) |
cV (J kg−1K−1) |
THR-β | 21157 | 132213 | 4393896241.65 | 4393879748.25 | 86.39 ± 0.28 |
2736.42 ± 8.87 |
THRT3-β | 21192 | 132864 | 4384867565.23 | 4384850985.47 | 86.85 ± 0.28 | 2737.51 ± 8.82 |
THRT3-β1 | 20266 | 127468 | 3925226341.90 | 3925210205.54 | 84.52 ± 0.27 | 2776.85 ± 8.87 |
THRT3T4-a | 12249 | 79395.7 | 1225218557.93 | 1225208975.47 | 50.19 ± 0.16 | 2647.36 ± 8.44 |
@ The error has been calculated from CV=CV(U,T),i.e.ΔCV=CV√(ΔUU)2+(ΔTT)2 where ΔU and ΔT are standard errors in potential energy and temperature, respectively. |
Molecule | No. of atoms | Mass, mp(amu) | cPV (Jkg−1K−1) |
D (m2/s) |
k (Wm−1K−1) |
THR-β | 3895 | 27640 | 2008.75 ± 107.77 | (1.56 ± 0.02) × 10−7 | 0.30 ± 0.023 |
THRT3-β | 3930 | 28291 | 2030.62 ± 105.32 | (1.57 ± 0.03) × 10−7 | 0.30 ± 0.023 |
THRT3-β1 | 4014 | 29069 | 2244.53 ± 95.73 | (1.35 ± 0.05) × 10−7 | 0.29 ± 0.022 |
THRT3T4-α | 3986 | 28853 | 2154.44 ± 46.32 | (1.28 ± 0.04) × 10−7 | 0.26 ± 0.017 |
@@ The error has been calculated from k=k(CV,D,ρ),i.e.Δk=k√(ΔCVCV)2+(ΔDD)2+(Δρρ)2 where ΔCV, ΔD and Δρ are standard errors in specific heat, thermal diffusivity and density, respectively. a From the calorimetric measurements by Privalov et al. 1986 [25], cPV of native globular proteins ranges from 1200 to 2300 Jkg−1K−1 at 300 K. b D (water) = 14.6 Å2/ps, D (protein) = 21.1 Å2/ps for green fluoroscent protein (GFP), 18.7 Å2/ps for myoglobin [8]; and D (protein) = 4-18 Å2/ps at the temperature relaxation from 350 to 250 K [9]. c k (water) = 0.64 Wm−1K−1 & k (protein) = 0.13-0.28 Wm−1K−1 at 300 K [8,9]. |
Thermal diffusivity of THRs ranges from 1.28 × 10−7 to 1.57 × 10−7 m2/s at the temperature relaxation from 310 to 200 K that is calculated from the best fitting of theoretical expression (
Thermal conductivity of THR proteins, as depicted in Table 2, ranges from 0.26-0.30 Wm−1K−1 provided with the standard error of about 0.02 Wm−1K−1. This result shows that the protein-hormone systems are bad conductors of heat in comparision to water (kw = 0.64 Wm−1K−1). As reported by Yu et al. 2005 [8], thermal conductivity of both GFP and myoglobin is kp = 0.27 Wm−1K−1. The value of kp ranges from 0.13-0.23 Wm−1K−1 for the different proteins according to Lervik et al. 2010 [9]. The slightly higher value of kp has been observed for THRT3-β than that for THRT3-β1 and THRT3T4-α. Such change in heat conduction property is related to the effect of protein surface curvature, hydrophilic and hydrophobic patches and H-bonding between protein residues and protein-water interface [9,30]. According to the theory of non-equilibrium thermodynamics, the development of temperature gradient across THRs-water interface and the related thermal conduction are important biophysical phenomena linked with the functions of the protein-hormone systems towards the regulation of body temperature. The result of lower values of thermal conductivity of proteins than that of water is supporting the principle of body temperature regulation.
We have performed molecular dynamics simulations to find heat transfer coefficients of THRs. Heat capcity has been determined by using formula of mole fraction to the constituents of THR solvated periodic-box as well as nano-droplet of water. Numerical differentiation or slope of the best fitted straight line in mean potential function vs. temperature of the periodic box yields the same heat capacity as obtained from the method of mean energy fluctuations of the nano-droplet in canonical ensemble. The heat capacity of THR isoforms in solution ranges from 2000 to 2200 Jkg−1K−1 with the standard error limit of about 100 Jkg−1K−1 at 310 K. This result lies within the range of 1200 to 2300 Jkg−1K−1 as reported by Privalov et al. 1986 and a bit higher than 1500 Jkg−1K−1 as given by Yu et al. 2005 for the native globular proteins at 300 K. Our results for the specific heats of the protein-hormone systems are also in the close agreement with that of antifreeze protein in the hydration limit ranging from 2065 Jkg−1K−1 for dry state to 2533 Jkg−1K−1 for fully hydrated protein at about 300 K as studied by Pandey et al. 2017. In overt hypothyroid disorder, THRs become free of T3. The heat capaity values of T3-liganded and unliganded THR-β are different in some extent due to the change in degrees of freedom, protein-ligand interactions, H-bonding, and hydrophobic, vibrational and conformational changes in presence and absence of T3-hormone in LBD of the receptor. Thermal properties of THRs are directly associated with gene expressions and regulation of body temperature. Thermal diffusivity of THRs has been found to be ranging from 1.28 × 10−7 to 1.57 × 10−7 m2/s near to 1.46 × 10−7 m2/s of water at body scale temperature 310 K. A little bit higher value of thermal diffusion coefficient for THRT3-β than that of THRT3-β1 and THRT3T4-α is associated with the effect of inter-residue and protein-water H-bonding, thermal transport channels and boundary conductance in the protein-water interface. Thermal conductivity of THRs ranges from 0.26-0.30 Wm−1K−1 with the higher value for THRT3-β in the standard error limit of about 0.02 Wm−1K−1 which is about half the value, 0.64 Wm−1K−1 for water at 310 K. Our result for the thermal conductivity is also consistent with 0.27 Wm−1K−1 as reported by Yu et al. 2005 for GFP and myoglobin. The slightly different values of the heat transfer coefficients for THR-isoforms are related to H-bonding in protein-water interface, surface curvature, and hydrophobic and hydrophillic patches. In conclusion, specific heat, thermal diffusivity and thermal conductivity of THRs are consistent with that of other proteins as published in the previous literatures supporting the theory and the methodology implemented for their calculations.
The authors are thankful to Prof. Dr. Raju Khanal, Central Department of Physics, Tribhuvan University, Kathmandu, Nepal for providing the computing facilities. This work has been partially supported by Nepal Academy of Science and Technology (NAST) through a grant of PhD fellowship to the first author.
The authors declare that they have no conflict of interest.
[1] |
Brent GA (2012) Mechanisms of thyroid hormone action. J Clin Invest 122: 3035–3043. doi: 10.1172/JCI60047
![]() |
[2] |
Souza PCT, Puhl AC, Martínez L, et al. (2014) Identification of a new hormone-binding site on the surface of thyroid hormone receptor. Mol Endocrinol 28: 534–545. doi: 10.1210/me.2013-1359
![]() |
[3] |
Bochukova E, Schoenmakers N, Agostini M, et al. (2012) A mutation in the thyroid hormone receptor alpha gene. New Eng J Med 366: 243–249. doi: 10.1056/NEJMoa1110296
![]() |
[4] |
Brent GA (1994) The molecular basis of thyroid hormone action. Eng J Med 331: 847–853. doi: 10.1056/NEJM199409293311306
![]() |
[5] | Leitner DM (2008) Energy flow in proteins. Annu Rev Phys Chem 59: 233–259. |
[6] |
Prabhu NV, Sharp KA (2005) Heat capacity in proteins. Annu Rev Phys Chem 56: 521–548. doi: 10.1146/annurev.physchem.56.092503.141202
![]() |
[7] |
Privalov PL, Tiktopulo EI, Venyaminov SY, et al. (1989) Heat capacity and conformation of proteins in the denatured state. J Mol Biol 205: 737–750. doi: 10.1016/0022-2836(89)90318-5
![]() |
[8] |
Yu X, Leitner DM (2005) Heat flow in proteins: Computation of thermal transport coefficients. J Chem Phys 122: 054902. doi: 10.1063/1.1830431
![]() |
[9] |
Lervik A, Bresme F, Kjelstrup S, et al. (2010) Heat transfer in protein-water interfaces. Phys Chem Chem Phys 12: 1610–1617. doi: 10.1039/b918607g
![]() |
[10] | Helbing J, Devereux M, Nienhaus K, et al. (2011) Temperature dependence of the heat diffusivity of proteins. J Phys Chem A 116: 2620–2628. |
[11] |
Nosé S (1984) A molecular dynamics method for simulations in the canonical ensemble. Mol Phys 52: 255–268. doi: 10.1080/00268978400101201
![]() |
[12] | Suurkuusk J (1974) Specific heat measurements on lysozyme, chymotrypsinogen, and ovalbumin in aqueous solution and in solid state. Acta Chem Scand B 28: 409–417. |
[13] | Yang PH, Rupley JA (1979) Protein-water interactions. Heat capacity of the lysozyme-water system. Biochemistry 18: 2654–2661. |
[14] |
Lervik A, Bresme F, Kjelstrup S (2009) Heat transfer in soft nanoscale interfaces: The influence of interface curvature. Soft Matter 5: 2407–2414. doi: 10.1039/b817666c
![]() |
[15] |
Nascimento AS, Dia SMG, Nunes FM, et al. (2006) Structural rearrangements in the thyroid hormone receptor hinge domain and their putative role in the receptor function. J Mol Biol 360: 586–598. doi: 10.1016/j.jmb.2006.05.008
![]() |
[16] |
Sandler B, Webb P, Apriletti JW, et al. (2004) Thyroxine-thyroid hormone receptor interactions. J Biol Chem 279: 55801–55808. doi: 10.1074/jbc.M410124200
![]() |
[17] |
Souza PCT, Puhl AC, Martínez L, et al. (2014) Identification of a new hormone-binding site on the surface of thyroid hormone receptor. Mol Endocrinol 28: 534–545. doi: 10.1210/me.2013-1359
![]() |
[18] | Mackerell AD, Feig M, Brooks CL (2004) Extending the treatment of backbone energetics in protein force fields: Limitations of gas-phase quantum mechanics in reproducing protein conformational distributions in molecular dynamics simulations. J Comp Chem 25: 1400–1415. |
[19] |
Mackerell JAD, Bashford D, Bellott MLDR, et al. (1998) All-atom empirical potential for molecular modeling and dynamics studies of proteins. J Phys Chem B 102: 3586–3616. doi: 10.1021/jp973084f
![]() |
[20] |
Zoete V, Cuendet MA, Grosdidier A, et al. (2011) SwissParam: A fast force field generation tool for small organic molecules. J Comp Chem 32: 2359–2368. doi: 10.1002/jcc.21816
![]() |
[21] |
Humphrey W, Dalke A, Schulten K (1996) VMD: Visual molecular dynamics. J Mol Graph 14: 33–38. doi: 10.1016/0263-7855(96)00018-5
![]() |
[22] | Phillips JC, Braun R, Wang W, et al. (2005) Scalable molecular dynamics with NAMD. J Comp Chem 26: 1781–1802. |
[23] |
Eftink MR, Anusiem AC, Biltonen RL (1983) Enthalpy-entropy compensation and heat capacity changes for protein-ligand interactions: General thermodynamic models and data for the binding of nucleotides to ribonuclease A. Biochemistry 22: 3884–3896. doi: 10.1021/bi00285a025
![]() |
[24] |
Sturtevant JM (1977) Heat capacity and entropy changes in processes involving proteins. Proc Natl Acad Sci 74: 2236–2240. doi: 10.1073/pnas.74.6.2236
![]() |
[25] |
Privalov PL, Potekhin SA (1986) Scanning microcalorimetry in studying temperature-induced changes in proteins. Method enzymol 131: 4–51. doi: 10.1016/0076-6879(86)31033-4
![]() |
[26] | Privalov PL, Dragan AI (2007) Microcalorimetry of biological macromolecules. Biophys Chem 126: 16–24. |
[27] |
Gomez J, Hilser VJ, Xie D, et al. (1995) The heat capacity of proteins. Proteins: Struct, Funct, Bioinf 22: 404–412. doi: 10.1002/prot.340220410
![]() |
[28] |
Cooper A (2000) Heat capacity of hydrogen-bonded networks: An alternative view of protein folding thermodynamics. Biophys Chem 85: 25–39. doi: 10.1016/S0301-4622(00)00136-8
![]() |
[29] |
Pandey HD, Leitner DM (2017) Thermodynamics of hydration water around an antifreeze protein: A molecular simulation study. J Phys Chem B 121: 9498–9507. doi: 10.1021/acs.jpcb.7b05892
![]() |
[30] | Xu Y, Leitner DM (2014) Vibrational energy flow through the green fluorescent protein-water interface: Communication maps and thermal boundary conductance. J Phys Chem B 118: 7818–7826. |
1. | Tika Ram Lamichhane, Hari Prasad Lamichhane, Structural changes in thyroid hormone receptor-beta by T3 binding and L330S mutational interactions, 2020, 7, 2377-9098, 27, 10.3934/biophy.2020003 | |
2. | Tika Ram Lamichhane, Hari Prasad Lamichhane, 2021, 2319, 0094-243X, 050002, 10.1063/5.0037068 | |
3. | Tika Ram Lamichhane, Madhav Prasad Ghimire, Evaluation of SARS-CoV-2 main protease and inhibitor interactions using dihedral angle distributions and radial distribution function, 2021, 7, 24058440, e08220, 10.1016/j.heliyon.2021.e08220 | |
4. | Ikuo Kurisaki, Seiya Tanaka, Ichiro Mori, Toshihito Umegaki, Yoshiharu Mori, Shigenori Tanaka, Thermal conductivity and conductance of protein in aqueous solution: Effects of geometrical shape, 2023, 44, 0192-8651, 857, 10.1002/jcc.27048 |
Molecule in water sphere | No. of atoms in mixture | Total mass, M (amu) | 〈U2〉 (k cal mol−1)2 |
〈U〉2 (k cal mol−1)2 |
CV=〈U2〉−〈U〉2KBT2 (k cal mol−1K−1) |
cV (J kg−1K−1) |
THR-β | 21157 | 132213 | 4393896241.65 | 4393879748.25 | 86.39 ± 0.28 |
2736.42 ± 8.87 |
THRT3-β | 21192 | 132864 | 4384867565.23 | 4384850985.47 | 86.85 ± 0.28 | 2737.51 ± 8.82 |
THRT3-β1 | 20266 | 127468 | 3925226341.90 | 3925210205.54 | 84.52 ± 0.27 | 2776.85 ± 8.87 |
THRT3T4-a | 12249 | 79395.7 | 1225218557.93 | 1225208975.47 | 50.19 ± 0.16 | 2647.36 ± 8.44 |
@ The error has been calculated from CV=CV(U,T),i.e.ΔCV=CV√(ΔUU)2+(ΔTT)2 where ΔU and ΔT are standard errors in potential energy and temperature, respectively. |
Molecule | No. of atoms | Mass, mp(amu) | cPV (Jkg−1K−1) |
D (m2/s) |
k (Wm−1K−1) |
THR-β | 3895 | 27640 | 2008.75 ± 107.77 | (1.56 ± 0.02) × 10−7 | 0.30 ± 0.023 |
THRT3-β | 3930 | 28291 | 2030.62 ± 105.32 | (1.57 ± 0.03) × 10−7 | 0.30 ± 0.023 |
THRT3-β1 | 4014 | 29069 | 2244.53 ± 95.73 | (1.35 ± 0.05) × 10−7 | 0.29 ± 0.022 |
THRT3T4-α | 3986 | 28853 | 2154.44 ± 46.32 | (1.28 ± 0.04) × 10−7 | 0.26 ± 0.017 |
@@ The error has been calculated from k=k(CV,D,ρ),i.e.Δk=k√(ΔCVCV)2+(ΔDD)2+(Δρρ)2 where ΔCV, ΔD and Δρ are standard errors in specific heat, thermal diffusivity and density, respectively. a From the calorimetric measurements by Privalov et al. 1986 [25], cPV of native globular proteins ranges from 1200 to 2300 Jkg−1K−1 at 300 K. b D (water) = 14.6 Å2/ps, D (protein) = 21.1 Å2/ps for green fluoroscent protein (GFP), 18.7 Å2/ps for myoglobin [8]; and D (protein) = 4-18 Å2/ps at the temperature relaxation from 350 to 250 K [9]. c k (water) = 0.64 Wm−1K−1 & k (protein) = 0.13-0.28 Wm−1K−1 at 300 K [8,9]. |
Molecule in water sphere | No. of atoms in mixture | Total mass, M (amu) | 〈U2〉 (k cal mol−1)2 |
〈U〉2 (k cal mol−1)2 |
CV=〈U2〉−〈U〉2KBT2 (k cal mol−1K−1) |
cV (J kg−1K−1) |
THR-β | 21157 | 132213 | 4393896241.65 | 4393879748.25 | 86.39 ± 0.28 |
2736.42 ± 8.87 |
THRT3-β | 21192 | 132864 | 4384867565.23 | 4384850985.47 | 86.85 ± 0.28 | 2737.51 ± 8.82 |
THRT3-β1 | 20266 | 127468 | 3925226341.90 | 3925210205.54 | 84.52 ± 0.27 | 2776.85 ± 8.87 |
THRT3T4-a | 12249 | 79395.7 | 1225218557.93 | 1225208975.47 | 50.19 ± 0.16 | 2647.36 ± 8.44 |
@ The error has been calculated from CV=CV(U,T),i.e.ΔCV=CV√(ΔUU)2+(ΔTT)2 where ΔU and ΔT are standard errors in potential energy and temperature, respectively. |
Molecule | No. of atoms | Mass, mp(amu) | cPV (Jkg−1K−1) |
D (m2/s) |
k (Wm−1K−1) |
THR-β | 3895 | 27640 | 2008.75 ± 107.77 | (1.56 ± 0.02) × 10−7 | 0.30 ± 0.023 |
THRT3-β | 3930 | 28291 | 2030.62 ± 105.32 | (1.57 ± 0.03) × 10−7 | 0.30 ± 0.023 |
THRT3-β1 | 4014 | 29069 | 2244.53 ± 95.73 | (1.35 ± 0.05) × 10−7 | 0.29 ± 0.022 |
THRT3T4-α | 3986 | 28853 | 2154.44 ± 46.32 | (1.28 ± 0.04) × 10−7 | 0.26 ± 0.017 |
@@ The error has been calculated from k=k(CV,D,ρ),i.e.Δk=k√(ΔCVCV)2+(ΔDD)2+(Δρρ)2 where ΔCV, ΔD and Δρ are standard errors in specific heat, thermal diffusivity and density, respectively. a From the calorimetric measurements by Privalov et al. 1986 [25], cPV of native globular proteins ranges from 1200 to 2300 Jkg−1K−1 at 300 K. b D (water) = 14.6 Å2/ps, D (protein) = 21.1 Å2/ps for green fluoroscent protein (GFP), 18.7 Å2/ps for myoglobin [8]; and D (protein) = 4-18 Å2/ps at the temperature relaxation from 350 to 250 K [9]. c k (water) = 0.64 Wm−1K−1 & k (protein) = 0.13-0.28 Wm−1K−1 at 300 K [8,9]. |