Next Article in Journal
Effect of Time Delay on Laser-Triggered Discharge Plasma for a Beyond EUV Source
Next Article in Special Issue
Homochirality: A Perspective from Fundamental Physics
Previous Article in Journal
Intelligent Supply Chain Information System Based on Internet of Things Technology under Asymmetric Information
Previous Article in Special Issue
Сhiral and Racemic Fields Concept for Understanding of the Homochirality Origin, Asymmetric Catalysis, Chiral Superstructure Formation from Achiral Molecules, and B-Z DNA Conformational Transition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Overview of Low-Temperature Heat Capacity Data for Zn2(C8H4O4)2.C6H12N2 and the Salam Hypothesis

1
Nikolaev Institute of Inorganic Chemistry, Siberian Branch, Russian Academy of Sciences, Lavrentyev Av., 3, RU-630090 Novosibirsk, Russia
2
Novosibirsk State University, Pirogova Street, 2, RU-630090 Novosibirsk, Russia
*
Author to whom correspondence should be addressed.
Symmetry 2019, 11(5), 657; https://doi.org/10.3390/sym11050657
Submission received: 29 March 2019 / Revised: 8 May 2019 / Accepted: 9 May 2019 / Published: 11 May 2019
(This article belongs to the Special Issue Possible Scenarios for Homochirality on Earth)

Abstract

:
The review presents the progress in the analysis of low-temperature heat capacity of the metal-organic framework Zn2(C8H4O4)2.C6H12N2 (Zn-DMOF). In Zn-DMOF, left-twisted D3(S) and right-twisted D3(R) DABCO molecules (C6H12N2) can transform into each other by tunneling to form a racemate. Termination of tunneling leads to a phase transition in the subsystem of twisted molecules. It is suggested that Zn-DMOF may be considered a model system to study the mechanisms of phase transitions belonging to the same type as hypothetical Salam phase transitions.

1. Introduction

According to the Salam hypothesis, a small parity-violating energy difference (PVED) between amino acid molecules along with the Bose-Einstein (BE) condensation, makes the less stable right enantiomers tunnel into the more stable left enantiomers, by changing their structural forms. This process was described as a second-order phase transition, which is an analog of the Bardeen-Cooper-Schrieffer (BCS) phase transition; therefore, physical properties such as heat capacity and magnetic susceptibility should change during this phase transition according to the BCS laws [1,2]. Even though no such phase transitions have been found in the crystals of known amino acids, the building material of living organisms, the systems demonstrating the BE condensation of chiral molecules are still of interest. A model of BE condensation was developed for a gas of non-interacting chiral molecules to determine the PVED contribution from low-temperature heat capacity data [3]. The present review summarizes low-temperature heat capacity data, which indicate that the BE condensation may work in a subsystem of triethylenediamine (DABCO) molecules (C6H12N2) in the metal-organic framework Zn2(C8H4O4)2.C6H12N2 (Zn-DMOF) and that the mechanism of Salam phase transitions remains possible. In Zn-DMOF, the enantiomers are represented by left- and right-twisted DABCO molecules, which transform into each other as a result of tunneling.

2. Structure of DABCO Molecule in Zn2(C8H4O4)2·C6H12N2

Triethylenediamine (DABCO) appears in the form of two conformational isomers with D3h and D3 point group symmetries, depending on intermolecular interactions. Also, a quasi-D3h form of DABCO is possible due to strong vibrations of the molecule around the C3 axis. The molecules with the D3 symmetry, which can be left-twisted D3(S) or right-twisted D3(R), are considered to be chiral isomers (enantiomers) [4].
Above 223 K, the crystal structure of the metal-organic framework Zn-DMOF is tetragonal, with space group P4/mmm [5]. The horizontal planes are formed by terephthalate anions [C8H4O4]2− (BDC2−) which are linked to {Zn2} pairs by carboxylate anions. The vertical edges are formed by DABCO molecules (linkers), the point symmetry of which does not contain a 4-fold rotational symmetry axis (Figure 1). This is the reason why DABCO molecules are orientationally disordered; moreover, D3(S) and D3(R) forms can transform into each other (by activation or tunneling) [4,6]. Calorimetry, nuclear magnetic resonance, and X-ray structural analysis data provide evidence of the presence of phase transitions in Zn-DMOF at ~14, ~60, and ~130 K [7,8,9,10,11,12].

3. Mobility of DABCO Molecules in Zn2(C8H4O4)2·C6H12N2

In Zn-DMOF, BDC2 anions and DABCO molecules are involved in activation mobility. According to the nuclear magnetic resonance (NMR) studies of the activation mobility of BDC2 anions, the [C6H4] groups of BDC2 anions rotate about the C2 axis through an angle of 180° (flipping) [13,14,15]. No effect of BDC2 flipping on the mobility of DABCO in Zn-DMOF was discovered [13,16,17].
According to the detailed analysis of the temperature behavior of the spin-lattice relaxation times of hydrogen nuclei (1H NMR T1(T)), D3(S) and D3(R) forms of DABCO can make up a racemic mixture, and their mirror symmetry may be broken during the phase transition at ~60 K [6,11,12]. The time decay of nuclear magnetic moments (M) of hydrogen atoms in DABCO was analyzed to find the distribution of DABCO molecules over different states. Above ~165 K, the time decay of M is a single exponential function characterized by a single value T1. In this case, DABCO molecules with D3 and D3h symmetries reorient similarly, their proton spins constitute a single spin system, the activation barrier is equal to ~4 kJ/mol. Between 165 and 60 K, the time decay of M is a biexponential function containing two values T1, each corresponding to a certain fraction of nuclear spins in M. The ratio of these fractions is estimated to be ⅓:⅔. The ⅓·M fraction corresponds to 1H spins of DABCO molecules of the D3h symmetry, the mobility of which is characterized by a short value T1SH. The ⅔·M fraction corresponds to 1H spins of the sum of S- and R-forms of DABCO. In this case, these forms are indistinguishable due to tunneling transitions, so the above fraction (⅔·M) represents the racemic state of DABCO molecules and is characterized by a single value T1L of a larger magnitude. During the phase transition at 60 K and down to 25 K, the behavior of T1L is interpreted as the termination of tunneling between energy degenerate quantum states of R- and S- forms of DABCO, and their fractions in M remain equal to each other (⅓:⅓). Below 25 K, the decay M is nonexponential and can be conventionally characterized by three values T1. So, the phase transition at ~14 K is associated with the redistribution of DABCO molecules over different energy states characterized by contributions ¼·M, ¼·M, and ½·M and the appearance of a chiral polarized state [11].
Note that the racemate state was also reported for 1,4-bis(carboxyethynyl)bicyclo[2.2.2]octane (BABCO) molecules, which are analogs of DABCO. It was shown that the ratio of left- and right- twisted forms of BABCO can be controlled by light [18,19]. The disorder of DABCO and its analogs, which causes phase transitions, is also observed in other systems [20,21,22].
Thus, the 1H NMR T1(T) data testify that phase transitions are associated with the mobility of DABCO molecules. The analysis of the function M provides quantitative data on the distribution of DABCO molecules over different states at various temperatures. However, it is still unclear how these states are structurally realized in Zn-DMOF. Low-temperature heat capacity data for Zn-DMOF may be used to clarify this problem.

4. Low-Temperature Heat Capacity in Zn2(C8H4O4)2·C6H12N2

All phase transitions in Zn-DMOF (at ~14, ~60, and ~130 K) are second-order phase transitions [10]. Table 1 shows the maximum values of the anomalous parts of heat capacity ΔCp = Cp − CpL, where Cp is the heat capacity of the substance and CpL is the regular part of heat capacity "in the absence of phase transitions". The entropy of the phase transitions is shown in Table 2.
The obtained data indicate that the absorbed atoms of 4He affect the states of D3(S) and D3(R) forms of DABCO (phase transitions ~14 and ~60 K) and do not affect the ordering and disordering of BDC2− anions during the phase transition at 130 K. This result can be explained by the fact that the structure of DABCO is flexible [23,24] as compared to that of BDC2− anions and can therefore be deformed in the presence of adjoining 4He atoms, whereas the structure of BDC2− anions remains unchanged.
The temperature dependence of heat capacity of Zn-DMOF is almost linear between the phase transitions at ~14, ~60, and ~130 K and above 130 K [8,10,11] to indicate the presence of a one-dimensional elastic continuum [8,12]. Figure 2 shows the comparison of experimental [8] and tabulated (Tarasov model) [25,26] heat capacity values using the fitting parameters obtained in [11].
According to the XRD data, the crystal lattice of Zn-DMOF expands along the a and b axes (dL/LdT(a,b) = −9.59·106 K1) and shrinks along the c axis (dL/LdT(c) = 12.2·106 K1) as the temperature decreases to ~130 K and below ~130 K |dL/LdT(c)| > |dL/LdT(a,b)|(dL/LdT is the coefficient of thermal expansion) [9]. Hence, the interactions in the -Zn-DABCO-Zn- chain directed along the c axis are assumingly stronger than BDC2-[Zn2]4+-BDC2- interactions in the ab plane, which determines a one-dimensional elastic continuum for the behavior of the heat capacity. The phase transition at ~130 K was interpreted as an order-disorder phase transition associated with a change in the relative spatial arrangement of BDC2 anions, while the DABCO molecules preserve their activation mobility and remain disordered [9].
Linear regions of heat capacity in Zn-DMOF were analyzed using the Stockmayer-Hecht model for the heat capacity of chain crystalline polymers [26,27]. The model assumes that molecular groups in a chain vibrate as single units connected by strong intrachain bonds, while interchain interactions are neglected. The temperature dependence of the volumetric heat capacity Cv is expressed in terms of two relationships, Cv/(3Nk) and T/Tm, where N is the number of repeated vibrating units, Tm = hνm/k, h is the Planck constant, k is the Boltzmann constant, and νm is the maximum frequency of stretching vibrations in the chain. The repeated vibrating unit along the c axis in Zn-DMOF consists of two Zn atoms and one DABCO molecule ({Zn2DABCO}) [5].
Experimental smoothed values Cp obtained as functions of temperature in [8] were represented on a log-log plot and fitted by best tabulated values Cv/(3Nk) for each T/Tm assuming that Cp−Cv is small [27] (Figure 3). As a result, it was found that the vibrating chain is formed by ~38–39 {Zn2DABCO} units above 130 K, by ~30 units at 60–130 K, and by ~12 units at 14–60 K (Table 3). Below 14 K, the heat capacity obeys the ~T3 law (Figure 3) to indicate that interchain interactions become stronger and the lattice vibrational modes become three-dimensional [8,25].
The values νm ~ 1250 cm−1 and νm ~ 765 cm−1 fall into the region of stretching vibrations of DABCO, and νm ~ 285 cm−1 fall into the region of Zn-N and Zn-Zn stretchings (Table 3) [28]. Thus, the obtained values νm correspond to the stretchings in the chains, in accordance with the model [27].
As can be seen, the values obtained from the analysis of Cp for N {Zn2DABCO} units correlate with fractions (M) in different phases of Zn-DMOF, if N and M values above 130 K are taken as a unit (Table 3). The obtained quantitative agreement between NMR data and the analysis of heat capacity suggests the following conclusions. Above 130 K, the chains consisting of ~39 {Zn2DABCO} units contain DABCO molecules with D3(S), D3(R), and D3h symmetries. At 60-130 K, the longest chains (~29 {Zn2DABCO} units) contain only D3 forms in the racemic state. The vibrations of these chains make the largest contribution to the heat capacity, while the vibrations of the chains consisting of D3h forms make no contribution practically, due to their shorter size. Finally, below ~60 K there are three types of chains (~12 {Zn2DABCO} units) of the same length but containing three different DABCO forms (D3(S), D3(R), and D3h). The size of the chains below 14 K cannot be estimated, since the heat capacity is no more linear at these temperatures.

5. Heat Capacity Behavior during the Phase Transition at 60 K and the Salam Hypothesis

The PVED values for the DABCO molecule and the [Zn2DABCO]4+ cation were obtained in [29]. The difference between the energies of mirror isomers is as small as ~5·10−16 kJ/mol (~5.2·10−18 eV) for DABCO and an order of magnitude higher (~5·10−15 kJ/mol or ~5.2·10−17 eV) for the [Zn2DABCO]4+ cation. Therefore, the contribution of PVED increases in the presence of Zn2+ cations and is determined mainly by the contribution of zinc cations. This contribution increases if Zn2+ cations are replaced by heavier cations Cd2+ and Hg2+ [30]. Hence, it can be assumed that the PVED breaking of mirror symmetry between D3(S) and D3(R) forms of DABCO may be caused by their external environment in the Zn-DMOF structure.
If the symmetry breaking during the phase transition at ~60 K takes place in the chains containing only D3(S) and D3(R) forms of DABCO, then, according to the Salam hypothesis, the behavior of heat capacity must correspond to the behavior of heat capacity during the superconducting phase transition [1,2].
In fact, it was discovered that the temperature behavior of heat capacity of Zn-DMOF is an exponential function below ~60 K (Figure 4) [31]. The behavior of heat capacity in the region of second-order phase transitions was studied using the values of the anomalous part of heat capacity ΔCp = Cp − CpL, and the behavior of CpL was described using the Tarasov model [25]. Figure 3 shows the obtained ΔCp values.
The region below ~60 K is of particular interest, since it is associated with the termination of tunneling between D3(S) and D3(R) forms of DABCO as the temperature decreases [11,12]. Based on the hypothesis suggested in [1,2], a study was carried out to verify the compliance of heat capacity ΔCp to the exponential dependence ~ exp(−Δ/T), гдe Δ = 1.76·Tc (Δ is the energy gap at 0 K). Figure 4 shows ΔCp as a function of 1/T in the temperature region 15 K < T < 60 K. As can be seen, a good agreement with the exponential law is achieved for the parameter Δ equal to ~56 K (or ~5 ·10−3 eV). [31]. The obtained value Δ turned out to be almost twice as small as expected (~106 K for Tc = 60 K). There is probably some inaccuracy with the parameters determining function CpL, which may cause the error of determining the Δ value. However, the detected exponential behavior of ΔCp below 60 K signifies the presence of a BE condensation. The amplitude of ΔCp during the phase transition at is 60 K ≈ 10 J/mol/K (Table 1), which corresponds to the thermal energy jump (ΔCp·Tc) ≈ 600 J/mol (or 6·10−3 eV), which agrees well with Δ. The value of (ΔCp·Tc) is 1015 times bigger than the PVED (~5.2·10−18 eV) of one DABCO molecule, but it can be explained by the phenomenon of BE condensation [1,2].
However, neither Cp data nor 1H NMR T1(T) data show any energy difference between D3(S) and D3(R) forms of DABCO below 60 K (according to the Salam hypothesis, the ratio between D3(S) and D3(R) forms of DABCO should change). Apparently, the energy difference between D3(S) and D3(R) forms remains negligible and can be observed only at lower temperatures, when the thermal energy of the crystal approaches zero [6]. Indeed, according to 1H NMR T1(T) data, the decay M as a function of time shows anomalous behavior below 25 K [12], but it is not manifested in the Cp behavior until the phase transition at ~14 K.

6. Conclusions

The Salam hypothesis is considered impossible during the phase transitions in amino acid crystals, since the barriers between L- and D-forms of alanine involve intramolecular bond breaking and are as high as ~200 kJ/mol [32]. In Zn-DMOF, the activation barrier between D3(S) and D3(R) forms of DABCO is estimated to be ~4 kJ/mol [7] and 5 kJ/mol [27] according to NMR data and quantum chemical calculations, respectively. Thus, this barrier is ~40 times smaller than the barrier between L- and D-forms of alanine [30]. The NMR data indicate the presence of tunneling between D3(S) and D3(R) forms of DABCO. The tunneling splitting for the DABCO molecule in the free state is estimated to be ~6 cm1− (~8.6 K) [24], which is comparable to the temperature range of observed phase transitions in Zn-DMOF. The behavior of heat capacity below 60 K corresponds to the heat capacity during the BE condensation. According to the NMR data, still lower temperatures are associated with a redistribution of DABCO with different symmetries over energy states to form a chiral polarized state. In the model system [Zn2DABCO]4+, the R-form is most favorable due to the PVED [29,30], but it is currently unclear which symmetry of the chains built of {Zn2DABCO} units corresponds to the most energetically favorable state. The method of resonant X-ray diffraction with circularly polarized X-rays [33] or optical methods seem to be most preferable for use at low and extra-low temperatures.
We believe that metal-organic frameworks or related compounds containing enantiomers in the racemic state (not necessarily amino acid molecules) may be considered as model systems to study Salam phase transitions. Our studies were aimed at revealing the effects of chirality stabilization in isomeric molecules in solids at low temperatures with the goal of exploring the idea of the cold scenario of life origin on the Earth.

Author Contributions

S.K., analysis of heat capacity data, writing and editing of the manuscript, M.R., editing of the manuscript and discussion, D.P., discussion, I.M.: discussion.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Salam, A. The role of chirality in the origin of life. J. Mol. Evol. 1992, 33, 105–113. [Google Scholar] [CrossRef]
  2. Salam, A. Chirality, phase transitions and their induction in amino acids. Phys. Lett. B 1992, 288, 153–160. [Google Scholar] [CrossRef]
  3. Bargueňo, P.; Perez de Tudela, R.; Miret-Artes, S.; Gonzalo, I. An alternative route to detect parity violating energy differences through Bose–Einstein condensation of chiral molecules. Phys. Chem. Chem. Phys. 2011, 13, 806–810. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Kozlova, S.G.; Mirzaeva, I.V.; Ryzhikov, M.R. DABCO molecule in the M2(C8H4O4)2·C6H12N2 (M = Co, Ni, Cu, Zn) metal-organic frameworks. Coord. Chem. Rev. 2018, 376, 62–74. [Google Scholar] [CrossRef]
  5. Dybtsev, D.N.; Chun, H.; Kim, K. Rigid and flexible: A highly porous metal–organic framework with unusual guest-dependent dynamic behavior. Angew. Chem. Int. Ed. 2004, 43, 5033–5036. [Google Scholar] [CrossRef]
  6. Kozlova, S.G.; Gabuda, S.P. Thermal properties of Zn2(C8H4O4)2•C6H12N2 metal-organic framework compound and mirror symmetry violation of dabco molecules. Sci. Rep. 2017, 7, 11505. [Google Scholar] [CrossRef]
  7. Gabuda, S.P.; Kozlova, S.G.; Samsonenko, D.G.; Dybtsev, D.N.; Fedin, V.P. Quantum Rotations and Chiral Polarization of Qubit Prototype Molecules in a Highly Porous Metal–Organic Framework: 1H NMR T1 Study. J. Phys. Chem. C 2011, 115, 20460–20465. [Google Scholar] [CrossRef]
  8. Paukov, I.E.; Samsonenko, D.G.; Pishchur, D.P.; Kozlova, S.G.; Gabuda, S.P. Phase transitions and unusual behavior of heat capacity in metal organic framework compound Zn2(C8H4O4)2 N2(CH2)6. J. Solid State Chem. 2014, 220, 254–258. [Google Scholar] [CrossRef]
  9. Kim, Y.; Haldar, R.; Kim, H.; Koo, J.; Kim, K. The guest-dependent thermal response of the flexible MOF Zn2(BDC)2(DABCO). Dalton Trans. 2016, 45, 4187–4192. [Google Scholar] [CrossRef] [PubMed]
  10. Pishchur, D.P.; Kompankov, N.B.; Lysova, A.A.; Kozlova, S.G. Order-disorder phase transitions in Zn2(C8H4O4)2·C6H12N2 in atmospheres of noble gases. J. Chem. Thermodyn. 2019, 130, 147–153. [Google Scholar] [CrossRef]
  11. Gabuda, S.P.; Kozlova, S.G. Chirality-related interactions and a mirror symmetry violation in handed nano structures. J. Chem. Phys. 2014, 141, 044701. [Google Scholar] [CrossRef]
  12. Gabuda, S.P.; Kozlova, S.G. Abnormal difference between the mobilities of left- and right-twisted conformations of C6H12N2 roto-symmetrical molecules at very low temperatures. J. Chem. Phys. 2015, 142, 234302. [Google Scholar] [CrossRef]
  13. Sabylinskii, A.V.; Gabuda, S.P.; Kozlova, S.G.; Dybtsev, D.N.; Fedin, V.P. 1H NMR refinement of the structure of the guest sublattice and molecular dynamics in the ultrathin channels of [Zn2(C8H4O4)2(C6H12N2)]·n(H3C)2NCHO. J. Struct. Chem. 2009, 50, 421–428. [Google Scholar] [CrossRef]
  14. Gallyamov, M.R.; Moroz, N.K.; Kozlova, S.G. NMR line shape for a rectangular configuration of nuclei. Appl. Magn. Reson. 2011, 41, 477–482. [Google Scholar] [CrossRef]
  15. Khudozhitkov, A.E.; Kolokolov, D.I.; Stepanov, A.G.; Bolotov, V.A.; Dybtsev, D.N. Metal-cation-independent dynamics of phenylene ring in microporous MOFs: A 2H solid-state NMR study. J. Phys. Chem. C 2015, 119, 28038–28045. [Google Scholar] [CrossRef]
  16. Kozlova, S.G.; Pishchur, D.P.; Dybtsev, D.N. Phase transitions in a metal–organic coordination polymer: [Zn2(C8H4O4)2(C6H12N2)]·with guest molecules. Thermal effects and molecular mobility. Phase Trans. 2017, 90, 628–636. [Google Scholar] [CrossRef]
  17. Burtch, N.C.; Torres-Knoop, A.; Foo, G.S.; Leisen, J.; Sievers, C.; Ensing, B.; Dubbeldam, D.; Walton, K.S. Understanding DABCO nanorotor dynamics in isostructural metal–organic frameworks. J. Phys. Chem. Lett. 2015, 6, 812–816. [Google Scholar] [CrossRef] [PubMed]
  18. Lemouchi, C.; Mézière, C.; Zorina, L.; Simonov, S.; Rodríguez-Fortea, A.; Canadell, E.; Wzietek, P.; Auban-Senzier, P.; Pasquier, C.; Giamarchi, T.; et al. Design and evaluation of a crystalline hybrid of molecular conductors and molecular rotors. J. Am. Chem. Soc. 2012, 134, 7880–7891. [Google Scholar] [CrossRef]
  19. Lemouchi, C.; Iliopoulos, K.; Zorina, L.; Simonov, S.; Wzietek, P.; Cauchy, T.; Rodríguez-Fortea, A.; Canadell, E.; Kaleta, J.; Michl, J.; et al. Crystalline arrays of pairs of molecular rotors: Correlated motion, rotational barriers, and space-inversion symmetry breaking due to conformational mutations. J. Am. Chem. Soc. 2013, 135, 9366–9376. [Google Scholar] [CrossRef]
  20. Shi, X.; Luo, J.; Sun, Z.; Li, S.; Ji, C.; Li, L.; Han, L.; Zhang, S.; Yuan, D.; Hong, M. Switchable dielectric phase transition induced by ordering of twisting motion in 1,4-diazabicyclo[2.2.2]octane chlorodifluoroacetate. Cryst. Growth Des. 2013, 13, 2081–2086. [Google Scholar] [CrossRef]
  21. Yao, Z.-S.; Yamamoto, K.; Cai, H.-L.; Takahashi, K.; Sato, O. Above room temperature organic ferroelectrics: Diprotonated 1,4-diazabicyclo[2.2.2] octane shifts between two 2-chlorobenzoates. J. Am. Chem. Soc. 2016, 138, 12005–12008. [Google Scholar] [CrossRef]
  22. Chen, L.-Z.; Huang, D.-D.; Ge, J.-Z.; Pan, Q.-J. Reversible ferroelastic phase transition of N-chloromethyl-1,4-diazabicyclo[2.2.2]octonium trichlorobromoaquo copper(II). Inorg. Chem. Commun. 2014, 45, 5–9. [Google Scholar] [CrossRef]
  23. Nizovtsev, A.S.; Ryzhikov, M.R.; Kozlova, S.G. Structural flexibility of DABCO. Ab initio and DFT benchmark study. Chem. Phys. Lett. 2017, 667, 87–90. [Google Scholar] [CrossRef]
  24. Mathivon, K.; Linguerri, R.; Hochlaf, M. Systematic theoretical studies of the interaction of 1,4-diazabicyclo [2.2.2]octane (DABCO) with rare gases. J. Chem. Phys. 2013, 139, 164306. [Google Scholar]
  25. Tarasov, V.V. Heat Capacity of Anisotropic Solids. Zhurnal Fiz. Khimii 1950, 24, 111–128. (In Russian) [Google Scholar]
  26. Wunderlich, B.; Baur, H. Heat Capacities of Liner High Polymers; Springer: Berlin, Germany, 1970. [Google Scholar]
  27. Stockmayer, W.H.; Hecht, C.E. Heat capacity of chain polymeric crystals. J. Chem. Phys. 1953, 21, 1954–1958. [Google Scholar] [CrossRef]
  28. Tan, K.; Nijem, N.; Canepa, P.; Gong, Q.; Li, J.; Thonhauser, T.; Chabal, Y.J. Stability and hydrolyzation of metal organic frameworks with paddle-wheel SBUs upon hydration. Chem. Mater. 2012, 24, 3153–3167. [Google Scholar] [CrossRef]
  29. Mirzaeva, I.V.; Kozlova, S.G. Computational estimation of parity violation effects in a metal-organic framework containing DABCO. Chem. Phys. Lett. 2017, 687, 110–115. [Google Scholar] [CrossRef]
  30. Mirzaeva, I.V.; Kozlova, S.G. Parity violating energy difference for mirror conformers of DABCO linker between two M2+ cations (M = Zn, Cd, Hg). J. Chem. Phys. 2018, 149, 214302. [Google Scholar] [CrossRef] [PubMed]
  31. Kozlova, S.G. Behavior of the heat capacity at second-order phase transitions in the [Zn2(C8H4O4)2·C6H12N2] metal-organic framework compound. JETP Lett. 2016, 104, 253–256. [Google Scholar] [CrossRef]
  32. Sullivan, R.; Pyda, M.; Pak, J.; Wunderlich, B.; Thompson, J.R.; Pagni, R.; Pan, H.J.; Barnes, C.; Schwerdtfeger, P.; Compton, R. Search for electroweak interactions in amino acid crystals. II. The Salam hypothesis. J. Phys. Chem. A 2003, 107, 6674–6680. [Google Scholar] [CrossRef]
  33. Tanaka, Y.; Kojima, T.; Takata, Y.; Chainani, A.; Lovesey, S.W.; Knight, K.S.; Takeuchi, T.; Oura, M.; Senba, Y.; Ohashi, H.; et al. Determination of structural chirality of berlinite and quartz using resonant x-ray diffraction with circularly polarized x-rays. Phys. Rev. B 2010, 81, 144104. [Google Scholar] [CrossRef]
Figure 1. The structure of Zn2(C8H4O4)2·C6H12N2 (Zn-DMOF), space group P4/mmm. Positions of carbon atoms in triethylenediamine (DABCO) molecules are disordered [5]. Hydrogen atoms are omitted for clarity. DABCO and BDC2− structures are shown in the insets. (Compiled from Figure 1 in [10] and Figure 1 in [1]).
Figure 1. The structure of Zn2(C8H4O4)2·C6H12N2 (Zn-DMOF), space group P4/mmm. Positions of carbon atoms in triethylenediamine (DABCO) molecules are disordered [5]. Hydrogen atoms are omitted for clarity. DABCO and BDC2− structures are shown in the insets. (Compiled from Figure 1 in [10] and Figure 1 in [1]).
Symmetry 11 00657 g001
Figure 2. Tabulated (T/TD) experimental (+) and calculated (solid lines) dependences of heat capacity (Cp) for Zn-DMOF and one-dimensional elastic continuums (Tarasov model [25]). Vertical dashed lines show the temperatures of phase transitions. Solid blue, green, and red lines corresponds to the Debye temperatures of 1490 K, 2230 K, and 2950 K, respectively (according to the data reported in [11]).
Figure 2. Tabulated (T/TD) experimental (+) and calculated (solid lines) dependences of heat capacity (Cp) for Zn-DMOF and one-dimensional elastic continuums (Tarasov model [25]). Vertical dashed lines show the temperatures of phase transitions. Solid blue, green, and red lines corresponds to the Debye temperatures of 1490 K, 2230 K, and 2950 K, respectively (according to the data reported in [11]).
Symmetry 11 00657 g002
Figure 3. Log-log plot of the Zn-DMOF heat capacity versus temperature. Experimental (crosses) and calculated values of heat capacity at 14.7–57.4 K (blue lines), 130.1–72.6 K (green lines), and 299.6–141.6 K (red lines).
Figure 3. Log-log plot of the Zn-DMOF heat capacity versus temperature. Experimental (crosses) and calculated values of heat capacity at 14.7–57.4 K (blue lines), 130.1–72.6 K (green lines), and 299.6–141.6 K (red lines).
Symmetry 11 00657 g003
Figure 4. Temperature dependence of ΔCp (in gram-atom units) for Zn-DMOF (left) and ΔCp plotted as a function of 1/T below ~60 K (right). ΔCp is shown on the logarithmic scale (according to the data from Figures 2 and 4 in [31]).
Figure 4. Temperature dependence of ΔCp (in gram-atom units) for Zn-DMOF (left) and ΔCp plotted as a function of 1/T below ~60 K (right). ΔCp is shown on the logarithmic scale (according to the data from Figures 2 and 4 in [31]).
Symmetry 11 00657 g004
Table 1. ΔCp (J/mol/K) values at the phase transitions in Zn-DMOF under various pressures of the heat-exchange gas 4He (P·105,Pa).
Table 1. ΔCp (J/mol/K) values at the phase transitions in Zn-DMOF under various pressures of the heat-exchange gas 4He (P·105,Pa).
P~14 K~60 K~130 K
0.516.0 ± 0.48.0 ± 0.223.0 ± 0.3
1.525.0 ± 0.411.0 ± 0.223.0 ± 0.3
Table 2. Entropies ΔS/R of the phase transitions in the region of critical temperatures (Tc, K) under various pressures of the heat-exchange gas 4He (P·105,Pa) in Zn-DMOF. R is the universal gas constant.
Table 2. Entropies ΔS/R of the phase transitions in the region of critical temperatures (Tc, K) under various pressures of the heat-exchange gas 4He (P·105,Pa) in Zn-DMOF. R is the universal gas constant.
Tc~14~60~130
PΔS/RΔS/RΔS/R
0.510.42 ± 0.050.14 ± 0.020.30 ± 0.04
1.520.28 ± 0.040.23 ± 0.020.30 ± 0.04
Table 3. Calculated parameters for Zn-DMOF. M · is the nuclear magnetic moment, N · is the number of {Zn2DABCO} units normalized with respect to corresponding values above 130 K.
Table 3. Calculated parameters for Zn-DMOF. M · is the nuclear magnetic moment, N · is the number of {Zn2DABCO} units normalized with respect to corresponding values above 130 K.
Region of Fit, K299.6–141.6130.1–72.657.4–14.7
νm, cm−11250765285
N~38.5~28.9~12.0
N · 1~0.75~0.31
M · 1~0.67~0.33

Share and Cite

MDPI and ACS Style

Kozlova, S.; Ryzhikov, M.; Pishchur, D.; Mirzaeva, I. Overview of Low-Temperature Heat Capacity Data for Zn2(C8H4O4)2.C6H12N2 and the Salam Hypothesis. Symmetry 2019, 11, 657. https://doi.org/10.3390/sym11050657

AMA Style

Kozlova S, Ryzhikov M, Pishchur D, Mirzaeva I. Overview of Low-Temperature Heat Capacity Data for Zn2(C8H4O4)2.C6H12N2 and the Salam Hypothesis. Symmetry. 2019; 11(5):657. https://doi.org/10.3390/sym11050657

Chicago/Turabian Style

Kozlova, Svetlana, Maxim Ryzhikov, Denis Pishchur, and Irina Mirzaeva. 2019. "Overview of Low-Temperature Heat Capacity Data for Zn2(C8H4O4)2.C6H12N2 and the Salam Hypothesis" Symmetry 11, no. 5: 657. https://doi.org/10.3390/sym11050657

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop