Next Article in Journal
Cobalt and Iron Cyano Benzene Bis(Dithiolene) Complexes
Previous Article in Journal
First-Principles Calculations of P-B Co-Doped Cluster N-Type Diamond
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Low Temperature Raman Spectroscopy of Tetrahydrofuran: Phonon Spectra Compared to Matrix Isolation Spectra in Air

by
Vlasta Mohaček-Grošev
Center of Excellence for Advanced Materials and Sensing Devices, Research Unit New Functional Materials, Ruđer Bošković Institute, Bijenička cesta 54, 10000 Zagreb, Croatia
Crystals 2024, 14(5), 468; https://doi.org/10.3390/cryst14050468
Submission received: 22 April 2024 / Revised: 9 May 2024 / Accepted: 14 May 2024 / Published: 16 May 2024
(This article belongs to the Section Organic Crystalline Materials)

Abstract

:
The conformation of tetrahydrofuran (THF) molecules in vapor has been the subject of considerable computational and experimental studies, the most recent by Park and Kwon stated that the difference between the most stable, twisted C2 conformer and the bent Cs conformer is 17 ± 15 cm−1. Because of low symmetry, all modes from both conformers are allowed in the Raman and infrared spectra. In 1982, Aleksanyan and Antipov observed the emergence of two Raman bands at 249 and 303 cm−1 at 20 K, while only one band at 293 cm−1 was present in solid THF at 142. They assigned the 249 cm−1 band to the restricted pseudorotational motion of THF in the solid state, because on heating, the band diminishes and is too weak to be observed near melting point (at 142 K). Cadioli et al. reported a study of the vibrational spectrum of tetrahydrofuran, giving a complete assignment of all bands including those present in the low-temperature Raman spectrum at 85 K and infrared bands observed at 90 K. They assigned the band at 242 cm−1 in the Raman spectrum at 85 K as an overtone of the lowest normal mode (pseudorotational mode), while the 299 cm−1 band in the same spectrum was assigned as a radial mode. In the following, low-temperature Raman spectra of solid THF together with the Raman matrix isolated spectrum of THF in air will be presented and compared to published data. Our results indicate that the band observed at 245 cm−1 at 10 K is too strong to be assigned as an overtone, since its intensity is of the same magnitude as the 299 cm−1 band.

1. Introduction

The non-rigidity of small molecules has captured the interest of scientists because it is essential for explaining their chemical preferences, such as conformations of amino acids in peptides [1], for the determination of the composition of sugars in solutions [2], and for explaining the origin of modern twist–bend nematic liquid crystals [3], among others. As pointed out by Park and Kwon, structural variations of ribose having C2′-endo conformation in B-deoxyribonucleic acid and C3′-endo conformation in ribonucleic acid have an impact on their biological function [4]. Tetrahydrofuran is produced and used in large quantities reaching nearly half a million tons per year [5], because it can solvate both polar and non-polar compounds [6] and is a candidate for hydrogen storage due to its ability to form clathrate hydrates [7].
The conformation of tetrahydrofuran (THF) molecules in vapor has been the subject of considerable computational and experimental studies [8,9,10,11], the most recent of which stated that the difference between the most stable twisted C2 conformer and the bent Cs conformer is 17 ± 15 cm−1 [4]. Because of low symmetry, all modes are allowed in both the Raman and the infrared spectrum. However, the transitions between pseudorotational levels for THF in the gas phase require special attention, since the two lowest normal modes, pseudorotational and radial transition, display a multitude of sub-transitions [12,13,14,15,16].
In the solid state, it has been repeatedly established by both X-ray [5,17] and neutron scattering [18] that THF exists in the twisted C2 conformation. It crystallizes in the C2/c space group with four molecules per unit cell, each molecule being on the rotation axis of symmetry of the second order [17]. Aleksanyan and Antipov observed the emergence of two Raman bands at 249 and 303 cm−1 at 20 K, while only one band at 293 cm−1 was present in solid THF at 142 K [19]. They assigned the two bands to the out-of-plane skeletal vibrations of THF in restricted non-planar conformation. Cadioli et al. reported a study on the vibrational spectrum of tetrahydrofuran, giving a complete assignment of all bands including those present in low-temperature Raman spectrum at 85 K and infrared bands observed at 90 K [20]. They assigned the band at 242 cm−1 in the Raman spectrum at 85 K as an overtone of ν33, which is the lowest normal mode (pseudorotational mode), while the 299 cm−1 band in the same spectrum was assigned as a radial mode (ν17). In a subsequent publication, Gallinella et al. presented vibrational spectra THF-d8, assigning the two Raman bands observed in the solid at 125 and 199 cm−1 to 0 →1 and 0 → 2 jumps of the pseudorotational motion [21].
Recently, an FTIR matrix isolation experiment was conducted where THF molecules were immobilized in N2, and the conclusion based on repeated annealing was that both Cs and C2 were present in the matrix independently of the annealing process [22]. Also, Park and Kwon used mass spectroscopy combined with vacuum UV photoionization to produce a THF cation in the D0 state, which was predominantly in the C2 conformation, and analyzed data for both C2 and Cs conformations in the ground S0 state of the neutral molecule [4]. They determined the conformational stabilities of the twisted C2 and bent Cs conformers to be 17 ± 15 cm−1, and stated that they can equilibrate under the molecular beam conditions applied in the experiment [4].
In the following, low-temperature Raman spectra of solid THF together with the Raman matrix isolated spectrum of THF in air will be presented and compared to published data. To the best of the authors’ knowledge, no previous Raman matrix isolation spectra of tetrahydrofuran have been published so far. Our results indicate that the band observed at 245 cm−1 at 10 K is too strong to be assigned as an overtone, as was assigned in [20].

2. Experimental

Liquid tetrahydofuran was purchased from VWR company; it had a purity > 99% and was used as such. The experimental setup for Raman matrix isolation experiments has been described previously [23,24]; therefore, we briefly stress the main points. A closed-cycle helium cryostat is used for the cooling of the cold finger in the sample chamber onto which a gold-plated surface is fixed. Once the lowest attainable temperature is achieved (10 K), the valve that lets the mixture of air and tetrahydrofuran vapor is opened. Within several minutes, a solid film of frozen air and THF molecules is formed on the gold-plated surface, and Raman spectral acquisition undertaken. Since the vapor pressure of tetrahydrofuran is 18.8 kbar at 295 K [25], at atmospheric pressure of 101,325 Pa, there are 5.4 molecules of nitrogen or oxygen per one molecule of tetrahydrofuran present. Their vibrations, however, are not overlapping vibrational bands of tetrahydrofuran, since they are observed at 2328 cm−1 (N2) and at 1553 cm−1 (O2). For cooling, we used a CCS 350 Janis Research closed cycle He cryostat with Lake Shore 331 temperature controller. Spectra were recorded with a T64000 Horiba-JobinYvon Raman spectrometer, operating in triple subtractive mode. Raman spectra were recorded in multiwindow option, from 10 to 3200 cm−1 for the polycrystalline sample and from 50 to 3200 cm−1 for the liquid and matrix isolated THF. The laser power of the 532 nm DPSS laser (ChangChun Industries Ltd., Changchun, China) applied to the sample was 7 mW. Pure THF spectra of polycrystalline THF were obtained after a small amount of sample was sealed in a glass tube and mounted in the sample chamber of the cryostat. In Figure 1, an overview of Raman spectra of THF in liquid (295 K, 150 K) and crystal phases (120 K, 10 K) is given. In Figure 2, the high wavenumber regions (2400–3100 cm−1) of the spectra are compared for matrix isolated, liquid and crystalline THF, while the middle parts of the same spectra are shown in Figure 3 and Figure 4.

3. Computational Details

Normal modes of tetahydrofuran in C2 and Cs conformation (Figure 5) were calculated using Gaussian09 [26] at the b3lyp/6-31++G(d,p) level of theory. All frequencies turned out positive. The difference in the ground state energies of the two conformers was only 5.7·10−5 Ha, or 0.46 meV. The C2 conformer was found to have lower energy.

4. Results and Discussion

Comparing the Raman spectra of matrix isolated THF with those of liquid (Figure 2, Figure 3 and Figure 4 and Figure 6), one observes the better resolving of vibrational transitions, especially in the 1100–1500 cm−1 interval (Figure 3). Four normal modes have their corresponding transitions at 1492, 1487, 1466, and 1452 cm−1, while the 1506 cm−1 band was assigned as the combination ν15 + ν16 (Table 1). Whereas in the liquid, a few broad bands can be identified between 1100 and 1300 cm−1, in the matrix one observes bands at 1371, 1343, 1317, and 1296 cm−1, which are assigned as ν7, ν24, ν8, and ν25 of the twisted conformer, C2. The band at 1287 cm−1 is assigned as ν25 of the THF molecule at the second matrix site. Bands at 1179, 1227 and 1253 cm−1 observed in Raman spectrum of liquid have their counterparts in 1176, 1239 and 1246 cm−1 bands observed for matrix isolated THF (Figure 3). The strongest band in liquid is observed at 914 cm−1, while in the matrix isolated THF there are bands at 872, 894, 906, 925, and 961 cm−1 (Figure 4, Table 1). Stocka et al. assigned the bands at 1013.9 and 1079.1 cm−1 to Cs conformers, and these bands had their intensity diminished after annealing [22]. We observed a weak band at 1066 cm−1 and assigned it to ν28 of the C2 conformer.
The bands in spectrum of polycrystalline THF are much narrower, showing splitting due to molecular packing in the unit cell. The crystal structure of tetrahydrofuran was solved by Luger and Buschmann [17]. THF crystallizes in the C2/c space group, having Z = 4, and unit cell dimensions equal to a = 607.6(9) pm, b = 891.4(9) pm, c = 774.3(11) pm and β = 106.11(12)0 (see Figure 7). Each THF molecule in the crystal lies on a C2 axes and has a C2 symmetry, with a twisted conformation. Later, David and Ibberson performed reinvestigations of the THF crystal structure using neutron scattering, and showed that the THF crystal remains stable down to 5 K [18]. Our assignment of Raman modes follows the procedure of Berenblut et al. on the C2/c structure of gypsum [27], where two molecules make a motif and Z = 2. The total number of degrees of freedom is 156, and if one removes three acoustical degrees of freedom (their symmetry is A u     2 B u ), there are 153 vibrations left.
The resulting decomposition of the reducible representation of THF phonons without acoustic modes is
Γ v i b = 38   A g   40   B g   37   A u     38 B u
This result was derived from characters of a reducible representation, which is stated in the following. The only atoms that are placed on a symmetry element are oxygen atoms (four of them are on a proper C2 axes in a unit cell), giving the characters of the following reducible representation: χ(E) =156, χ(C2) = − 4, χ(i) = 0, χ(σ)= 0.
Of these modes, there are twelve vibrational and nine translational modes expected below 250 cm−1. Usually called external phonons, their symmetry is found via correlation tables to be
Γ e x t = 4   A g   8   B g   3   A u     6   B u
The two lowest internal modes, restricted puckering motion and radial mode, are observed at 245 and 299 cm−1 at 10 K (Table 1, Figure 6).
Of particular interest to us is the low-frequency part of the THF Raman spectra; for the polycrystalline solid, lattice modes usually appear below 200 cm−1, while in the spectra of liquid and matrix isolated THF, a Rayleigh wing starts gaining in intensity (see Figure 6). Both in the spectrum of liquid and in the spectrum of matrix isolated THF, there is a single weak and broad line at 283 cm−1, but in the spectrum of polycrystalline THF at 10 K, there are two medium weak bands, appearing at 245 and 299 cm−1 (Figure 8), in accordance with the work of Aleksanyan and Antipov [19] who observed them at 249 and 303 cm−1 at 20 K. The assignments provided by Cadioli et al. [20], which we verified by repeating the calculation of normal modes of twisted THF using Gaussian09 [26] at the b3lyp/6-31++G(d,p) level of theory, are shown in Table 1. Previous studies on the structure of the free THF molecule claim that either the Cs (envelope) [5,10] or the C2 (twisted) form is the most stable conformation [4,11,20]. The matrix isolation technique enables one to observe THF vibrational transitions in the environment more closely resembling that of a gas than that of a liquid, providing better resolution of the observed bands. In particular, while calculation predicts two bands at 581 and 672 cm−1 for C2 conformer, for the Cs conformer, the corresponding modes are predicted at 645 and 651 cm−1. The observed bands in the Raman spectrum of matrix isolated THF are at 577 and 662 cm−1 (Table 1, Figure 6). Stocka et al. presented FTIR matrix isolated spectra of THF in the 600–3025 cm−1 interval, and did not report any bands below 600 cm−1 [22]. Instead, they followed the behavior of the 1073.1 and 898.9 cm−1 bands on annealing, and concluded that both conformers are present in their matrix. Comparing the Raman spectra of matrix isolated THF with those of crystals (Figure 2, Figure 3 and Figure 4), one finds a close resemblance of band positions and intensities, confirming our conclusion that only C2 is present in the matrix.
In the majority of organic compounds, vibrational bands observed in liquid usually split into several new bands, reflecting the strength of the intermolecular interactions experienced in solid. The difference in wavenumbers observed between new split bands is called crystal splitting, and can range from a few cm−1 to several tens of wavenumbers, depending on the type of molecular motion from which the band originates. In tetrahydrofuran crystals, Raman bands at 586 and 667 cm−1 are narrow compared to 244 and 299 cm−1 bands (Figure 8). While the 586 cm−1 band corresponds to C-O-C bending coupled to CH2 rocking, the 667 cm−1 is predominantly C-O-C bending vibration. No crystal splitting Ag/Bg is observed for these two bands in crystal; therefore, we do not think it likely that 245 and 299 cm−1 bands are the Ag and Bg components of the radial mode observed at 283 cm−1 in liquid and in the matrix. The normal mode of C2 THF having the lowest wavenumber is calculated to lie at 42 cm−1, while the following higher mode (referred to as the “radial“mode [12]) is at 251 cm−1. While the value obtained for the radial mode is closer to the 299 cm−1 band observed in crystal (Figure 6), the band at 244 cm−1, which we assign to the quasi pseudorotational motion of THF in solid, is at much higher wavenumbers than calculated in the harmonic approximation for the twisted conformer in the free state.
Temperature dynamics of large amplitude motions can provide insights into molecular dynamics. For example, the internal rotation of methyl group in toluene and nitromethane was studied in detail by Cavagnat et al. [29]. The 0A → 1A and 0E → 1E transitions of internal rotation in CH3NO2 that are observed in low-temperature Raman spectra show a weakening of intensity on heating, and merge into a single band at 50 K [29]. Another small carbon ring compound, cyclobutane, has two solid phases: a plastic crystalline phase I just below the melting point, which occurs at 182 K, and crystal phase II, into which phase I transforms below 145 K [30]. Durig et al. recorded low-temperature Raman spectra of cyclobutane, and discovered that the puckering mode, which is observed as a strong band at 257 cm−1 at 10 K, gradually loses its intensity and practically vanishes in the vicinity of transition into orientationally disordered phase I [31]. The same behavior is observed for the 244 cm−1 tetrahydrofuran band (Figure 6), confirming the work of Aleksanyan and Antipov [19].
Small globular molecules tend to form plastic crystals, structures in which molecules have their centres of mass ordered on lattice points, but having orientational disorder to a certain degree [32]. Often, the fast-cooling of a plastically crystalline state produces a glassy crystal, which has its positional order preserved but orientational order frozen. Angell et al. proposed a guideline by which one can estimate whether a compound is prone to form a glassy crystal: if the ratio of the temperature of the boiling point Tb to temperature Tm of melting Tb/Tm is greater than 2, a compound may form a glassy crystal [33]. For tetrahydrofuran, Tb is 339 K [34], and the melting point as determined by Lebedev et al. is Tm = (164.9 ± 0.2) K, which gives us Tb/Tm of 2.055 [35]. One may thus question the possibility of the existence of a glassy state in THF—such a type was predicted by the molecular dynamics simulation method employed by Rong-Ri et al. [36], but ruled out after several attempts to achieve a vitrification in THF proved unsuccessful [35]. Cosidering that such a phase was found by Raman spectroscopy to exist in nitromethane in a very narrow temperature interval [37] and was not detected by the classical measurement of heat capacity [38], there is a possibility that tetrahydrofuran possesses such a phase as well.

5. Conclusions

Low-temperature Raman spectroscopy was applied to study the dynamics of tetrahydrofuran in the polycrystalline phase, and as a matrix isolated compound in air. The Raman spectrum of tetrahydrofuran isolated in air matrix was recorded for the first time, showing a better resolution of bands than is found in liquid. The width of bands observed in the spectrum of THF in air matrix, together with the observed low-frequency 283 cm−1 band, suggests that THF molecules undergo free pseudorotational motion, which is hindered in crystal. The temperature dependence of the 244 cm−1 band observed at 10 K in the spectrum of crystalline THF shows a progressive diminishing of its intensity towards melting point. This band was assigned as the hindered pseudorotational mode of the C2 conformer present in the crystal.

Funding

This work was partially supported by Centre of Excellence for Advanced Materials and Sensors, a project co-financed by the Croatian government and the European Union through the European regional development fund e The Competitivness and Cohesion Operational Programme (KK.01.1.1.01). Calculations using Gaussian09 were performed at University of Zagreb Computing Centre SRCE.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The author declares no conflict of interests.

References

  1. Grdadolnik, J.; Mohacek-Grosev, V.; Baldwin, R.L.; Avbelj, F. Populations of the three major backbone conformations in 19 amino acid dipeptides. Proc. Natl. Acad. Sci. USA 2011, 108, 1794–1798. [Google Scholar] [CrossRef] [PubMed]
  2. Angyal, S.J. The Composition of Reducing Sugars in Solution. Adv. Carboh. Chem. Biochem. 1984, 42, 15–68. [Google Scholar] [CrossRef]
  3. Arukawa, Y.; Komatsu, K.; Feng, J.; Zhu, C.; Tsuji, H. Distinct twist-bend nematic phase behaviors associated with the esther-linked liquid crystal dimers. Mater. Adv. 2021, 2, 261–272. [Google Scholar] [CrossRef]
  4. Man Park, S.; Ho Kwon, S. Deciphering the conformational preference and ionization dynamics of tetrahydrofuran: Insights from advanced spectroscopic techniques. J. Chem. Phys. 2024, 160, 114308. [Google Scholar] [CrossRef] [PubMed]
  5. Boese, A.D.; Boese, R. Tetrahydrothiophene and Tetrahydrofuran, Computational and X-ray Studies in the Crystalline Ph/ase. Cryst. Growth Des. 2015, 15, 1073–1081. [Google Scholar] [CrossRef]
  6. Bowron, D.T.; Finney, J.L.; Soper, A.K. The structure of liquid tetrahydrofuran. J. Am. Chem. Soc. 2006, 128, 5119–5126. [Google Scholar] [CrossRef] [PubMed]
  7. Lee, H.; Lee, J.-W.; Kim, D.Y.; Park, J.; Seo, Y.-T.; Zeng, H.; Moudrakovski, I.L.; Ratcliffe, C.I.; Ripmeester, J.A. Tuning clathrate hydrates for hydrogen storage. Nature 2005, 434, 743–746. [Google Scholar] [CrossRef] [PubMed]
  8. Ford, T.A. The evolution of the structural, vibrational and electronic properties of the cyclic ethers—On ring size. An ab initio study. J. Mol. Struct. 2014, 1073, 125–133. [Google Scholar] [CrossRef]
  9. Yang, T.; Su, G.; Ning, C.; Deng, J.; Wang, F.; Zhang, S.; Ren, X.; Huang, Y. New Diagnostic of the Most Populated Conformer of Tetrahydrofuran in the Gas Phase. J. Phys. Chem A 2007, 111, 4927–4933. [Google Scholar] [CrossRef]
  10. Rayon, V.M.; Sordo, J.A. Pseudorotation motion in tetrahydrofuran: An ab initio study. J. Chem. Phys. 2005, 122, 204303. [Google Scholar] [CrossRef]
  11. Cremer, D.; Pople, J. Molecular orbital theory of the electronic structure of organic compounds. XXIII. Pseudorotation in saturated five-membered ring compounds. J. Amer. Chem. Soc. 1975, 97, 1358–1367. [Google Scholar] [CrossRef]
  12. Sont, W.N.; Wieser, H. The radial transtion in tetrahydrofuran. J. Raman Spectrosc. 1981, 11, 334–338. [Google Scholar] [CrossRef]
  13. Greenhouse, J.A.; Strauss, H.L. Spectroscopic Evidence for Pseudorotation. II. The Far-Infrared Spectra of Tetrahydrofuran and 1,3-Dioxolane. J. Chem. Phys. 1969, 50, 124–134. [Google Scholar] [CrossRef]
  14. Engerholm, G.G.; Luntz, A.C.; Gwinn, W.D.; Harris, D.O. Ring Puckering in Five-Membered Rings. II. The Microwave Spectrum, Dipole Moment, and Barrier to Pseudorotation in Tetrahydrofuran. J. Chem. Phys. 1969, 50, 2446–2457. [Google Scholar] [CrossRef]
  15. Mamleev, A.H.; Gunderova, L.N.; Galeev, R.V. Microwave Spectrum and Hindered Pseudorotation of Tetrahydrofuran. J. Struct. Chem. 2001, 42, 365–370. [Google Scholar] [CrossRef]
  16. Melnik, D.G.; Gopalakrishnan, S.; Miller, T.A.; de Lucia, F.C. The absorption spectroscopy of the lowest pseudorotational states of tetrahydrofuran. J. Chem. Phys. 2003, 118, 3589–3599. [Google Scholar] [CrossRef]
  17. Luger, P.; Buschmann, J. Twist conformation of Tetrahydrofuran in the Crystal. Angew. Chem. Int. Ed. Engl. 1983, 22, 410. [Google Scholar] [CrossRef]
  18. David, W.I.F.; Ibberson, R.M. A reinvestigation of the structure of tetrahydrofuran by high-resolution neutron powder diffraction. Acta Cryst. C 1992, 48, 301–303. [Google Scholar] [CrossRef]
  19. Aleksanyan, V.T.; Antipov, B.G. On the intramolecular dynamics of five-membered ring compounds in solid state. In Raman Spectroscopy Linear and Nonlinear; Lascombe, J., Huong, P., Eds.; John Wiley & Sons: New York, NY, USA, 1982; pp. 605–606. [Google Scholar]
  20. Cadioli, B.; Gallinella, E.; Coulombeau, C.; Jobic, H.; Berthier, G. Geometric Structure and Vibrational Spectrum of Tetrahydrofuran. J. Phys. Chem. 1993, 97, 7844–7856. [Google Scholar] [CrossRef]
  21. Gallinella, E.; Cadioli, B.; Flament, J.P.; Berthier, G. A scaled force field for tetrahydrofuran and its isotopomers from quantum mechanical calculations and infrared and Raman spectra. J. Mol. Struct. 1994, 315, 137–148. [Google Scholar] [CrossRef]
  22. Stocka, J.; Čeponkus, J.; Šablinskas, V.; Rodziewicz, P. Conformational diversity of the THF molecule in N2 matrix by means of FTIR matrix isolation experiment and Car-Parrinello molecular dynamics simulations. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2020, 238, 118425. [Google Scholar] [CrossRef] [PubMed]
  23. Mohaček-Grošev, V.; Furić, K.; Vujnović, V. Raman study of water deposited in solid argon matrix. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2022, 269, 120770. [Google Scholar] [CrossRef]
  24. Furić, K.; Volovšek, V. Water ice at low temperatures and pressures: New Raman results. J. Mol. Struct. 2010, 976, 174–180. [Google Scholar] [CrossRef]
  25. Available online: https://webbook.nist.gov/cgi/cbook.cgi?ID=C109999&Mask=4&Type=ANTOINE&Plot=on (accessed on 19 April 2024).
  26. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. (Eds.) Gaussian 09, Revision A.02; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  27. Berenblut, B.J.; Dawson, P.; Wilkinson, G.R. The Raman spectrum of gypsum. Spectrochim. Acta 1971, 27A, 1849–1863. [Google Scholar] [CrossRef]
  28. Macrae, C.F.; Bruno, I.J.; Chisholm, J.A.; Edgington, P.R.; McCabe, P.; Pidcock, E.; Rodriguez-Monge, L.; Taylor, R.; van de Streek, J.; Wood, P.A. Mercury CSD 2.0—New Features for the Visualization and Investigation of Crystal Structures. J. Appl. Cryst. 2008, 41, 466–470. [Google Scholar] [CrossRef]
  29. Cavagnat, D.; Lascombe, J.; Lassegues, J.C.; Horsewill, A.J.; Heidemann, A.; Suck, J.B. Neutron and Raman scattering studies of the methyl dynamics in solid toluene and nitromethane. J. De Phys. 1984, 45, 97–105. [Google Scholar] [CrossRef]
  30. Stein, A.; Lehmann, C.W.; Luger, P. Crystal Structure of Cyclobutane at 117 K. J. Am. Chem. Soc. 1992, 114, 7684–7687. [Google Scholar] [CrossRef]
  31. Durig, J.R.; Sullivan, J.F.; Durig, D.T. Low-frequency vibrational spectra of molecular crystals. XXII: Cyclobutane-d0 and cyclobutane-d8. Mol. Cryst. Liq. Cryst. 1979, 54, 51–58. [Google Scholar] [CrossRef]
  32. Sherwood, J.N. (Ed.) The Plastically Crystalline State, Orientationally-Disordered Crystals; John Wiley & Sons: Hoboken, NJ, USA, 1979. [Google Scholar]
  33. Angell, C.A.; Busse, L.E.; Cooper, E.I.; Kadiyale, R.K.; Dworkin, A.; Ghelfenstein, M.; Szwarc, H.; Vassal, A. Glasses and glassy crystals from molecular and molecular ionic systems. J. Chim. Phys. Paris 1985, 82, 267–274. [Google Scholar] [CrossRef]
  34. NIH National Library of Medicine. Available online: https://pubchem.ncbi.nlm.nih.gov/compound/Tetrahydrofuran (accessed on 1 May 2024).
  35. Lebedev, B.V.; Rabinovich, I.B.; Milov, I.V.; Lityagor, V.Y. Thermodynamic properties of tetrahydrofuran from 8 K to 322 K. J. Chem. Thermodyn. 1978, 10, 321–329. [Google Scholar] [CrossRef]
  36. Rong-Ri, T.; Xin, S.; Lin, H.; Feng-Shon, Z. Liquid glass transition of tetrahydrofuran and 2-methyltetrahydrofuran. Chin. Phys. B 2012, 21, 086402. [Google Scholar] [CrossRef]
  37. Mohaček Grošev, V.; Stelzer, F.; Jocham, D. Internal rotation dynamics of nitromethane at low temperatures. J. Mol. Struct. 1999, 476, 181–189. [Google Scholar] [CrossRef]
  38. Cavagnat, D.; Brom, H.; Nugteren, P.R. Methyl internal rotation in partially deuterated molecular solids: The NO2CHD2 and NO2CH2D systems. J. Chem. Phys. 1987, 87, 801–808. [Google Scholar] [CrossRef]
Figure 1. Raman spectra of THF in liquid (295 K, 150 K) and crystal phases (120 K, 10 K) from 10 to 1600 cm−1 (solid) and 50 to 1600 cm−1 (liquid). The spectra of liquids do not show the top of the very strong band at 914 cm−1 in order to provide a better view of weaker bands (see Figure 4).
Figure 1. Raman spectra of THF in liquid (295 K, 150 K) and crystal phases (120 K, 10 K) from 10 to 1600 cm−1 (solid) and 50 to 1600 cm−1 (liquid). The spectra of liquids do not show the top of the very strong band at 914 cm−1 in order to provide a better view of weaker bands (see Figure 4).
Crystals 14 00468 g001
Figure 2. Raman spectra of THF in matrix (10 K), liquid (295 K) and crystal phases (10 K) from 2400 to 3100 cm−1.
Figure 2. Raman spectra of THF in matrix (10 K), liquid (295 K) and crystal phases (10 K) from 2400 to 3100 cm−1.
Crystals 14 00468 g002
Figure 3. Raman spectra of THF in matrix (10 K), liquid (295 K) and crystal phases (10 K) from 1000 to 1700 cm−1.
Figure 3. Raman spectra of THF in matrix (10 K), liquid (295 K) and crystal phases (10 K) from 1000 to 1700 cm−1.
Crystals 14 00468 g003
Figure 4. Raman spectra of THF in matrix (10 K), liquid (295 K) and crystal phases (10 K) from 580 to 1300 cm−1.
Figure 4. Raman spectra of THF in matrix (10 K), liquid (295 K) and crystal phases (10 K) from 580 to 1300 cm−1.
Crystals 14 00468 g004
Figure 5. Optimized geometrical structures of the most stable C2 conformer and the Cs conformer. Oxygen atom is shown in red.
Figure 5. Optimized geometrical structures of the most stable C2 conformer and the Cs conformer. Oxygen atom is shown in red.
Crystals 14 00468 g005
Figure 6. Raman spectra of THF in matrix (10 K) and liquid (295 K) from 50 to 700 cm−1 compared to Raman spectrum of polycrystalline THF at 10 K from 10 to 700 cm−1.
Figure 6. Raman spectra of THF in matrix (10 K) and liquid (295 K) from 50 to 700 cm−1 compared to Raman spectrum of polycrystalline THF at 10 K from 10 to 700 cm−1.
Crystals 14 00468 g006
Figure 7. Crystal packing of tetrahydrofuran in the C2/c space group. Image prepared with the Mercury program [28]. Violet lines correspond to glide planes, green lines to screw axis and proper C2 axes and yellow dots to centres of inversion. Oxygen atoms are on proper C2 axes.
Figure 7. Crystal packing of tetrahydrofuran in the C2/c space group. Image prepared with the Mercury program [28]. Violet lines correspond to glide planes, green lines to screw axis and proper C2 axes and yellow dots to centres of inversion. Oxygen atoms are on proper C2 axes.
Crystals 14 00468 g007
Figure 8. Temperature dependence of low-frequency Raman spectra of polycrystalline THF obtained at 120 K, 110 K, 90 K, 80 K and 10 K from 10–600 cm−1.
Figure 8. Temperature dependence of low-frequency Raman spectra of polycrystalline THF obtained at 120 K, 110 K, 90 K, 80 K and 10 K from 10–600 cm−1.
Crystals 14 00468 g008
Table 1. Observed bands in Raman spectra of tetrahydrofuran in matrix (10 K), crystal (10 K) and liquid (295 K). The assignment of fundamentals is taken from Cadioli et al. [20].
Table 1. Observed bands in Raman spectra of tetrahydrofuran in matrix (10 K), crystal (10 K) and liquid (295 K). The assignment of fundamentals is taken from Cadioli et al. [20].
Observed Bands in Raman Spectra of
Tetrahydrofuran
Matrix Isolated
10 K
Assignment
C2 Conformer
Crystal 10 K
C2 Conformer
Liquid 295 KAssignment
from [20]
3019 w
3003 mw, sh
2992 s2988 s, sh, br2 ν5
2984 s, br ν12983 m
2972 mw, sh
2964 s, brν1
2947 s, brν22951 ms
2943 m2943 s, brν2
2935 ms, brν32935 m, sh
2924 ms, sh
2915 s, brν3
2881 s, brν42880 m2878 sν4
2860 vvs2864 s, sh
2729 wν22 + ν262732 m2721 mwν22 + ν26
2676 wν5 + ν102676 w2661 wν5 + ν10
2328 vs. N2
1553 s, O2
1506 mwν15 + ν16 1511 m
1492 mν5 1496 m
1487 m, shν221487 w1489 m, brν5
1474 ms1475 m, br, shν22
1466 mν6 1466 ms
1452 mν231453 vw1451 m, brν23
1371 wν7 1375 mw1367 wν7
1343 wν24 1344 m1337 wν24
1317 vwν8 1315 w
1296 wν251304 w
1287 w, shν25 second site 1288 mw, vbr, shν25
1246 mν261254 ms1253 m, vbr, shν26
1239 m, shν91244 ms
1214 vwν9 second site 1227 mν9
1190 m
1181 w1179 w, brν10
1176 wν101173 w
1147 wν11 1145 m1140 w, shν11
1066 wν28 1058 m1072 mwν28
1030 mν121040 s1030 m, brν12
961 wν29 961 mw949 w, shν29
925 sν13 929 vs
920 w
906 mν30910 w914 vsν13
894 mν14883 vs902 s, shν30
872 wν31878 m, sh
868 m
848 m844 w, shν15
838 wν15 841 m
662 wν16 667 m654 w, brν16
577 wν32 586 m598 w, brν32
299 mw, br ν17 radial mode
283 vw, brν17 284 wν17 radial mode
245 mw, asym, br ν33
quasi pseudorotational
mode
136 mw lattice mode
125 m lattice mode
108 m lattice mode
94 w lattice mode
80 mw lattice mode
69 mw lattice mode
24 w lattice mode
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mohaček-Grošev, V. Low Temperature Raman Spectroscopy of Tetrahydrofuran: Phonon Spectra Compared to Matrix Isolation Spectra in Air. Crystals 2024, 14, 468. https://doi.org/10.3390/cryst14050468

AMA Style

Mohaček-Grošev V. Low Temperature Raman Spectroscopy of Tetrahydrofuran: Phonon Spectra Compared to Matrix Isolation Spectra in Air. Crystals. 2024; 14(5):468. https://doi.org/10.3390/cryst14050468

Chicago/Turabian Style

Mohaček-Grošev, Vlasta. 2024. "Low Temperature Raman Spectroscopy of Tetrahydrofuran: Phonon Spectra Compared to Matrix Isolation Spectra in Air" Crystals 14, no. 5: 468. https://doi.org/10.3390/cryst14050468

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