Abstract
Free full text
Covalent Bonds versus van der Waals Forces: A Picture in Thermal Conduction of Organic Materials
Abstract
We present a direct comparison of the heat transport properties between the state in which the constituent molecules are assembled by intermolecular forces and the one in which they are covalently bonded, in a molecular system with identical constituent elements and masses, as well as a nearly identical structure and density. This comparison leading to an essential understanding of thermal conduction in organic materials is made possible by the unique compound found by Wudl et al., which exhibits a single-crystal-to-single-crystal topochemical polymerization with a yield of >99%, in combination with microtemperature wave analysis (μTWA), which allows accurate measurements of the thermal diffusivity of small single crystals. At room temperature, the thermal conductivity of monomer and polymer single crystals is not significantly different. For both crystals, the thermal conductivity increases monotonically with decreasing temperature. However, below the Debye temperature, the thermal conductivity of the polymer single crystal increases exponentially, giving much larger values than those of the monomer single crystal. Based on physical quantities related to the behavior of phonons, derived from the specific heat analysis, we discuss the differences in heat transport properties in the two states and provide guidelines for achieving high thermal conductivity in organic materials.
Introduction
Phonons are responsible for heat transport in materials where free electrons are absent.1 Therefore, unlike inorganic and carbon materials in which atoms are infinitely linked by ionic and covalent bonds, in organic materials, which are substances where the constituent molecules are assembled through weak intermolecular forces, heat transport by phonons does not efficiently occur. Furthermore, since molecules themselves are substances, in which multiple elements are linked in various bonding modes, and have complex inherent vibrational modes, the heat energy stored as intramolecular vibrations is not necessarily involved in heat transport.2 Reflecting these facts, it is commonly believed that organic materials do not intrinsically exhibit high thermal conduction.3−9
Now, what would be the thermal conductivity characteristics if molecules were infinitely linked by covalent bonds? To examine this idea, it is necessary to find an organic material that meets three requirements: (1) it must be of sufficient size to enable experimental thermal conductivity measurements; (2) it must have two structurally characterizable states, one covalently bonded and one not, to allow for comparative verification; (3) other factors related to the constituent molecules, including mass, elements present, and assembly structure, must be almost identical in these two states. To the best of our knowledge, only one system that satisfies these strict requirements has been reported thus far, namely, [2,2′-bi-1H-indene]-3,3′-dihydroxy-1,1′-heptanoate (1, Figure Figure11a), discovered by Wudl et al. in 2014.10 This molecule undergoes quantitative (>99%) single-crystal-to-single-crystal topochemical polymerization into poly-1 (Figure Figure11b) upon exposure to visible light (λ = 350–550 nm) and depolymerization upon heating at 195 °C, to recover monomer 1 while maintaining its single crystalline form.
Here, we describe the heat transport properties and their temperature dependence below room temperature for single crystals of monomer 1 and poly-1. We obtained the thermal diffusivity of 1 and poly-1 at each temperature using a microtemperature wave analysis (μTWA) technique,11,12 which allows for the accurate evaluation of thermal diffusivity and its anisotropy even for micrometer-sized small single crystals.13 Using the experimentally obtained thermal diffusivity, specific heat, and density, the thermal conductivity of 1 and poly-1 was determined. Interestingly, while the difference in thermal conductivity between 1 and poly-1 is small at high temperatures, below their Debye temperatures, a significant increase in thermal conductivity was observed only for poly-1. The origin of this behavior is discussed.
Results and Discussion
Monomer 1 and its single crystals were prepared by following the reported procedures.10 When orange-colored single crystals of 1 were exposed to a white LED light (OLYMPUS, SZ-LW61), topochemical polymerization took place to give colorless single crystals of poly-1 (Figure Figure11c,d). The quantitative conversion of 1 into poly-1 was confirmed by diffuse reflectance spectroscopy as well as extraction of the polymer crystals using dichloromethane as reported previously.10
By means of μTWA, we evaluated thermal diffusivities for single crystals of 1 and poly-1 in the directions parallel (αM// and αP//) and perpendicular (αM and αP) to the polymer chains (i.e., the molecular stacking direction for 1). The values of αM//, αP//, αM, and αP at 300 K were determined to be 1.47 ± 0.14 × 10–7, 2.01 ± 0.18 × 10–7, 1.16 ± 0.14 × 10–7, and 1.17 ± 0.16 × 10–7 m2 s–1, respectively (Figure Figure11e). We originally predicted that the poly-1 crystal in which the monomer units are covalently bonded should have a much higher thermal diffusivity than the monomer 1 crystal in which only intermolecular forces operate. However, no significant difference in thermal diffusivity was observed before and after polymerization or with changes in the polymer chain direction, although the value of αP// is higher than in other cases. Thus, covalent bonding between the monomer units leads to some improvement in the heat transport properties in the crystalline state, but not to any great extent at room temperature, giving a thermal diffusivity change of αP///αM// = 1.4.
Notably, when the temperature is decreased, a large difference in the thermal diffusivity before and after polymerization is observed. As shown in Figure Figure22a, the values of αM// (red) and αP// (blue) increase monotonically in proportion to T–0.84 (at 80–300 K) and T–0.95 (at 100–300 K), respectively, with decreasing temperature. This behavior is due to an increase in the scattering time of thermal carriers, and both exponents are close to −1, indicating that phonons are scattered by the Umklapp process.14−16 Note that the change in thermal diffusivity of monomer 1 and poly-1 with temperature becomes abrupt below a certain point. While the temperature dependence of αM// below 80 K appears to be roughly in proportion to T–1.73, that of αP// below 100 K does not follow a power law. Obviously, the influence of the covalent bonding between the monomers on the behavior of phonon scattering is pronounced in the low temperature region.
To investigate the behavior of thermal conductivity, we also measured the temperature dependence of specific heat at constant pressure (Cp) and density (ρ) of monomer 1 (red) and poly-1 (blue) (Figure Figure22b,c). The Cp values are almost identical to each other, meaning that the monomer and polymer in their crystalline states have the same number of degrees of freedom, with no significant difference in the intra- and intermolecular energies corresponding to each degree of freedom. The ρ values of 1 and poly-1 at selected temperatures were determined by single-crystal X-ray crystallography. As expected, after topochemical polymerization, the density increases.
Using the experimentally obtained values of thermal diffusivity (αM// and αP//), Cp, and ρ, we obtained the temperature dependence of thermal conductivity (κ = αCpρ) in the corresponding directions (κM// and κP//). As shown in Figure Figure22d, at 294 K, κM// (red) and κP// (blue) are 0.24 and 0.33 W m–1 K–1, respectively, giving a ratio of κP///κM// = 1.4. In the temperature range of 100–300 K, the values of κM// and κP// are not significantly different. However, below 100 K, a remarkable change in κM// and κP// was observed. For example, at 43 K, the values of κM// and κP// were determined to be 0.29 and 2.10 W m–1 K–1, respectively. Thus, the thermal conductivity of poly-1 becomes 7.2 times larger compared with that at 294 K. Importantly, the temperature, at which κP///κM// begins to abruptly increase, roughly corresponds to the Debye temperatures (θD) of 1 (67.2 K) and poly-1 (83.3 K), obtained from the analysis of the temperature dependence of their Cp values as described below.
The Debye temperature is a physical quantity unique to a material, and by determining the Debye temperatures for monomer 1 and poly-1, their phonon group velocity and mean free paths can be evaluated. For molecular crystals, the specific heat at constant volume (Cv) can be represented by both Debye and Einstein specific heat models.17 The former describes Cv, originating from the three translational and three rotational degrees of freedom when a molecule is considered as a rigid body,18 and is expressed as
where CvD, NA, kB, and θD are the Debye specific heat at constant volume, the Avogadro constant, the Boltzmann constant, and the Debye temperature, respectively.19 The latter describes Cv originating from intramolecular degrees of freedom. The Einstein specific heat model is expressed as
where CvE and θE are the Einstein specific heat at constant volume and the Einstein temperature, respectively.20
Using these two models, we obtained the Debye temperatures of 1 and poly-1, both of which consist of 72 atoms, and thus, there is a total of 216 degrees of freedom. For simplicity, we treat each of the three translational (trans) and three rotational (rot) degrees of freedom with one Debye specific heat model (CvD_trans and CvD_rot), and four Einstein specific heat models (CvEi, CvEj, CvEk, and CvEl; i, j, k, and l = the number of intramolecular vibrational modes) are applied to the remaining 210 intramolecular degrees of freedom.2 Accordingly, the total specific heat (Cv) can be expressed as
where i + j + k + l = 210.
Since Cv is difficult to obtain experimentally and the difference between Cp and Cv in a real system is considered to be small,2 we analyzed the experimentally obtained Cp curve (Figure Figure22e,f) using the three equations, where θD, θE, i, j, k, and l were used as parameters. The best fit was obtained when the values of CvD_trans and CvD_rot were equal, giving Debye temperatures (θD) of 67.2 and 83.3 K for monomer 1 and poly-1, respectively. These temperatures roughly coincide with those at which the thermal diffusivity changes abruptly (Figure Figure22a). Regarding the four Einstein temperatures (θE), the Cp curve of 1 (Figure Figure22e) was best fitted with 16.7, 132.3, 341.4, and 1378.4 K. On the other hand, the Cp curve of poly-1 (Figure Figure22f) was better reproduced when fitting with three parameters (i, j, and k) than with four (i, j, k, and l), and thus, the three values of θE were successfully determined to be 141.0, 352.5, and 1370.0 K. The peak observed around 30 K for 1 (Figure Figure22e) is due to a vibrational mode with an energy of 16.7 K, while the corresponding peak is absent for poly-1 (Figure Figure22f). This suggests that the covalent bonding between the monomeric 1 induces a shift of the intramolecular vibrational modes to a higher energy region.
Assuming isotropic phonon dispersions, we estimated phonon group velocity (vph) using the equation
where kB is the Boltzmann constant, is the Dirac constant, and n is the number density of molecules. Based on this equation, the values of vph were calculated to be 1970 and 2420 m s–1 for 1 and poly-1, respectively. The ratio of vph before and after polymerization is 1.2, which is in good agreement with the experimental observation (κP///κM// = 1.4 at 294 K, Figure Figure22d).21 Thus, the large difference in thermal conductivity between 1 and poly-1 in the lower temperature region (<100 K) cannot be explained in terms of phonon group velocity. We considered that this difference is due to changes in the mean free path (l) and scattering time (τ) of phonons, rather than the phonon group velocity.
Using the equations22
the values of l and τ for 1 and poly-1 at 300 and 43 K were calculated (Table 1). The calculated mean free path and scattering time of phonons are similar before and after polymerization around room temperature. It is noteworthy, however, that the difference in these values between 1 and poly-1 increases with decreasing temperature. At 43 K, poly-1 (83.8 Å) has a 5.8 times larger l value than 1 (14.4 Å), and the value of τ for poly-1 (34.6 × 10–13 s) is 4.7 times larger compared with 1 (7.31 × 10–13 s). Clearly, the difference in heat transport properties depending on whether the monomer units are assembled by intermolecular forces or linked by covalent bonds becomes apparent only in the region below the Debye temperature (θD = 67.2 and 83.3 K for monomer 1 and poly-1, respectively).
Table 1
300 K | 43 K | ||||
---|---|---|---|---|---|
vph (m s–1) | l (Å) | τ (s) | l (Å) | τ (s) | |
1 | 1970 | 2.23 | 1.13 × 10–13 | 14.4 | 7.31 × 10–13 |
poly-1 | 2420 | 2.49 | 1.03 × 10–13 | 83.8 | 34.6 × 10–13 |
When the temperature drops below θD, phonon excitation is suppressed, resulting in increases in both l and τ. It is well-known that at high temperatures, l and τ follow a power law with temperature, while below θD, they behave in an exponential manner [exp(cθD/T), c = constant].15,23 For thermal conduction to take on finite values, in other words, for thermal resistance to occur, phonons must undergo Umklapp scattering, in which crystal momentum including the reciprocal lattice vector must be conserved. When Umklapp scattering takes place below θD, at least one phonon with energy as large as θD should be involved in the scattering process. Below θD, the existence probability of such a phonon is proportional to exp(−cθD/T). This means that the lower the temperature, the fewer the amount of such phonons present, leading to an increase in both τ and l (l = vPhτ), which both increase exponentially with the function exp(cθD/T). Due to this exponential behavior, a small change in θD results in drastic changes in τ and l. Looking at the experimental results (Figure Figure22a), while 1 with a lower θD (67.2 K) displays a gradual exponential curve, poly-1 with a higher θD (83.3 K) shows pronounced exponential behavior. The Debye temperatures of substances comprehensively reflect their inherent structural parameters. In the present system, before and after topochemical polymerization, the types and number of elements are the same, the crystal structures are isomorphous, and the change in molecular structure is rather small. Therefore, the change in θD between 1 and poly-1 is exclusively due to the formation of covalent bonds. This provides an increase in θD, which in turn results in a significant improvement in the heat transport properties below θD.
Conclusions
Thanks to a molecule discovered by Wudl et al. that exhibits a single-crystal-to-single-crystal topochemical polymerization with a yield of >99% and the μTWA method, which allows for accurate measurements of the thermal diffusivity of small single crystals, we have successfully shown for the first time the difference in thermal conduction between molecules assembled by intermolecular forces and those linked by covalent bonds. Covalent bonds and intermolecular forces are essential in organic materials, while heat is a ubiquitous energy. Thus, understanding of their relationship is important for the development of functional organic materials, especially those for thermal management. It is intuitively understandable that polymers have better heat transport properties than low-molecular-weight assemblies, but in reality, even with the ingenuity of polymer synthesis, attempts to improve their heat transport properties have often been unsuccessful. This work, which clearly demonstrates the importance of the Debye temperature, provides useful guidelines for the design of highly thermally conductive organic and polymeric materials.
Acknowledgments
This work was supported by the Japan Science and Technology Agency (JST) CREST (JPMJCR18I4 for T.F. and JPMJCR19I3 for J.M.). This work was also supported in part by the Cooperative Research Program of “Network Joint Research Center for Materials and Devices” from MEXT, Japan.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.4c11849.
Experimental details and methods including single-crystal X-ray analysis, specific heat measurements, and thermal diffusivity measurements using a microtemperature wave analysis (μTWA) (PDF)
References
- Wang X.; Wang W.; Yang C.; Han D.; Fan H.; Zhang J. Thermal transport in organic semiconductors. J. Appl. Phys. 2021, 130, 170902.10.1063/5.0062074. [CrossRef] [Google Scholar]
- Takehara R.; Kubo N.; Ryu M.; Kitani S.; Imajo S.; Shoji Y.; Kawaji H.; Morikawa J.; Fukushima T. Insights into Thermal Transport through Molecular π-Stacking. J. Am. Chem. Soc. 2023, 145, 22115–22121. 10.1021/jacs.3c07921. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
- Okada Y.; Uno M.; Nakazawa Y.; Sasai K.; Matsukawa K.; Yoshimura M.; Kitaoka Y.; Mori Y.; Takeya J. Low-temperature thermal conductivity of bulk and film-like rubrene single crystals. Phys. Rev. B 2011, 83, 11330510.1103/PhysRevB.83.113305. [CrossRef] [Google Scholar]
- Moore A. L.; Shi L. Emerging challenges and materials for thermal management of electronics. Mater. Today 2014, 17, 163–174. 10.1016/j.mattod.2014.04.003. [CrossRef] [Google Scholar]
- Zhang H.; Yao Y.; Payne M. M.; Anthony J. E.; Brill J. W. Thermal diffusivities of functionalized pentacene semiconductors. Appl. Phys. Lett. 2014, 105, 07330210.1063/1.4894094. [CrossRef] [Google Scholar]
- Wang H.; Yu C. Organic thermoelectrics: materials preparation, performance optimization, and device integration. Joule 2019, 3, 53–80. 10.1016/j.joule.2018.10.012. [CrossRef] [Google Scholar]
- Wu S.; Yan T.; Kuai Z.; Pan W. Thermal conductivity enhancement on phase change materials for thermal energy storage: A review. Energy Storage Mater. 2020, 25, 251–295. 10.1016/j.ensm.2019.10.010. [CrossRef] [Google Scholar]
- Hoshino N.; Akutagawa T. Contrasting temperature dependences of isostructural one-dimensional ferroelectric crystals NH4HSO4 and RbHSO4 in terms of thermal conductivities. J. Chem. Phys. 2020, 2020 (153), 194503.10.1063/5.0028153. [Abstract] [CrossRef] [Google Scholar]
- Gueye M. N.; Vercouter A.; Jouclas R.; Guérin D.; Lemaur V.; Schweicher G.; Lenfant S.; Antidormi A.; Geerts Y.; Melis C.; Cornil J.; Vuillaume D. Thermal conductivity of benzothieno-benzothiophene derivatives at the nanoscale. Nanoscale 2021, 13, 3800–3807. 10.1039/D0NR08619C. [Abstract] [CrossRef] [Google Scholar]
- Dou L.; Zheng Y.; Shen X.; Wu G.; Fields K.; Hsu W. C.; Zhou H.; Yang Y.; Wudl F. Single-crystal linear polymers through visible light–triggered topochemical quantitative polymerization. Science 2014, 343, 272–277. 10.1126/science.1245875. [Abstract] [CrossRef] [Google Scholar]
- Morikawa J.; Hashimoto T. Thermal diffusivity of aromatic polyimide thin films by temperature wave analysis. J. Appl. Phys. 2009, 105, 113506.10.1063/1.3116509. [CrossRef] [Google Scholar]
- Orie A.; Morikawa J.; Hashimoto T. Micro-scale thermal diffusivity measurements of banded spherulites of poly-(l-lactic acid) using a thermo-electric micro sensor. Thermochimica acta 2012, 532, 148–151. 10.1016/j.tca.2011.08.017. [CrossRef] [Google Scholar]
- Ryu M.; Takamizawa S.; Morikawa J. Thermal diffusivity of organosuperelastic soft crystals during stress-induced phase transition. Appl. Phys. Lett. 2021, 119, 251902.10.1063/5.0055707. [CrossRef] [Google Scholar]
- Callaway J. Model for lattice thermal conductivity at low temperatures. Phys. Rev. 1959, 113, 1046–1051. 10.1103/PhysRev.113.1046. [CrossRef] [Google Scholar]
- Holland M. G. Analysis of lattice thermal conductivity. Phys. Rev. 1963, 132, 2461–2471. 10.1103/PhysRev.132.2461. [CrossRef] [Google Scholar]
- Slack G. A. Nonmetallic crystals with high thermal conductivity. J. Phys. Chem. Solids 1973, 34, 321–335. 10.1016/0022-3697(73)90092-9. [CrossRef] [Google Scholar]
- Fulem M.; Laštovka V.; Straka M.; Růžička K.; Shaw J. M. Heat capacities of tetracene and pentacene. J. Chem. Eng. Data 2008, 53, 2175–2181. 10.1021/je800382b. [CrossRef] [Google Scholar]
- Venkataraman G.; Sahni V. C. External vibrations in complex crystals. Rev. Mod. Phys. 1970, 42, 409–470. 10.1103/RevModPhys.42.409. [CrossRef] [Google Scholar]
- Debye P. On the theory of specific heats. Ann. Phys. 1912, 39, 789–839. 10.1002/andp.19123441404. [CrossRef] [Google Scholar]
- Einstein A. Planck’s theory of radiation and the theory of heat capacity of solids. Ann. Phys. 1907, 22, 180–190. [Google Scholar]
- Since the Cp curve fitting was successful in the temperature range of 300 to 2 K, it is reasonable to assume that the vph values derived from the obtained θD values do not change significantly in the measurement temperature range.
- Ashcroft N.; Mermin N.. Solid State Physics; Saunders College Publishing: Fort Worth, 1976. [Google Scholar]
- Klemens P. G. The thermal conductivity of dielectric solids at low temperatures (theoretical). Proc. R. Soc. London 1951, A208, 108–133. 10.1098/rspa.1951.0147. [CrossRef] [Google Scholar]
Citations & impact
This article has not been cited yet.
Impact metrics
Alternative metrics
Discover the attention surrounding your research
https://www.altmetric.com/details/169633500
Similar Articles
To arrive at the top five similar articles we use a word-weighted algorithm to compare words from the Title and Abstract of each citation.
Planning Implications Related to Sterilization-Sensitive Science Investigations Associated with Mars Sample Return (MSR).
Astrobiology, 22(s1):S112-S164, 19 May 2022
Cited by: 1 article | PMID: 34904892
Thermal Transport in Quasi-1D van der Waals Crystal Ta2Pd3Se8 Nanowires: Size and Length Dependence.
ACS Nano, 12(3):2634-2642, 27 Feb 2018
Cited by: 7 articles | PMID: 29474086
Crystal Growth and Thermal Properties of Quasi-One-Dimensional van der Waals Material ZrSe3.
Micromachines (Basel), 13(11):1994, 17 Nov 2022
Cited by: 0 articles | PMID: 36422424 | PMCID: PMC9693893
Straightforward measurement of anisotropic thermal properties of a Bi2Se3 single crystal.
J Phys Condens Matter, 30(11):115701, 21 Mar 2018
Cited by: 1 article | PMID: 29469058
Funding
Funders who supported this work.
Core Research for Evolutional Science and Technology (2)
Grant ID: JPMJCR18I4
Grant ID: JPMJCR19I3