laboratory notes\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoJOURNAL OF
APPLIED
CRYSTALLOGRAPHY
ISSN: 1600-5767

An examination of the thermal expansion of urea using high-resolution variable-temperature X-ray powder diffraction

CROSSMARK_Color_square_no_text.svg

aInstitute of Particle Science and Engineering, School of Process Environment and Materials Engineering, University of Leeds, Leeds LS2 9JT, UK, and bCentre of Molecular and Interface Engineering, School of Engineering and Physical Sciences, Heriot-Watt University, Edinburgh EH14 4AS, UK
*Correspondence e-mail: k.j.roberts@leeds.ac.uk

(Received 10 August 2005; accepted 19 September 2005)

Variable-temperature high-resolution capillary-mode powder X-ray diffraction is used to assess changes in unit-cell dimensions as a function of temperature over the range 188–328 K. No evidence was found for any polymorphic transformations over this temperature range and thermal expansion coefficients for urea were found to be αa = (5.27 ± 0.26) × 10−5 K−1 and αc = (1.14 ± 0.057) × 10−5 K−1.

1. Introduction

Despite being of fundamental importance in many areas, information about the solid-state properties of molecular materials still remains surprisingly limited (Desiraiju, 2003[Desiraiju, G. (2003). J. Mol. Struct. 656, 5-15.]). In particular, detailed knowledge of the thermal expansion tensor for many low-symmetry organic materials is not always available, this despite the fact that structural change as a function of temperature can have ramifications for particle properties.

In the case of crystals not undergoing any polymorphic change, the effect of temperature on crystal structure is well described in terms of a linear thermal expansion coefficient. More specifically, in the case of anisotropic crystals such as urea, thermal expansion can be described by a linear thermal expansion coefficient along its primary crystallographic axes:

[\alpha = {1 \over {L_0 }}{{{\rm d}l} \over {{\rm d}T}} , \eqno (1)]

where L0 is a given cell parameter of the material (reference length) in a given direction and at a given temperature T0 (reference temperature).

In this short note, we report the determination of the thermal expansion coefficients of urea using temperature-programmed high-resolution X-ray powder diffraction. Urea, O=C(NH2)2, has applications in agrochemistry and nonlinear optics and is one of the simpler organic molecules. Urea is an unusual compound, having only eight atoms yet existing in a solid form under ambient conditions. Because of its simplicity, the crystal structure of urea was one of the first among the organic crystals to be determined by X-ray crystallographic methods (Vaughan & Donohue, 1952[Vaughan, P. & Donohue, J. (1952). Acta Cryst. 5, 530-535.], and references therein), being found to crystallize in the highly symmetrical tetragonal space group [P\bar{4} 2 _{1} m ] with two molecules in the unit cell of dimensions a = 5.661 and c = 4.712 Å. Urea crystal chemistry is dominated by hydrogen-bonding chains associated with a network of six strong intermolecular hydrogen bonds, four intrachain along the c axis and two interchain along the a/b axis in its first coordination sphere (see Docherty et al., 1993[Docherty, R., Roberts, K., Saunders, V., Black, S. & Davey, R. (1993). Faraday Discus. 95, 11-25.]), and is noteworthy for having its carbonyl group as an acceptor for four hydrogen bonds.

2. Experimental details

The calculation of the linear thermal expansion was carried out by monitoring the changes in the a and c cell parameters of a polycrystalline sample of urea, as determined using a Siemens D5000 high-resolution powder X-ray diffractometer equipped with an Oxford Instruments Cryocool liquid-nitrogen cold-flow cryostatic temperature controller. The urea powder, finely ground in a pestle and mortar and loaded into a 1 mm capillary, was examined over a temperature range from 188 to 328 K.

The resultant powder data were analysed with the lattice parameters calculated using the program DICVOL (Boultif & Louer, 1991[Boultif, F. & Louer, D. (1991). J. Appl. Cryst. 24, 987-993.]). The quality of the lattice parameters derived was crossed-checked through full-pattern fitting using the program GSAS (Larson & von Dreele, 1994[Larson, A. & von Dreele, B. (1994). GSAS, LosAlamos National Laboratory, New Mexico, USA.]). Using the values obtained from DICVOL during pattern indexing resulted in a comparatively low value for Rwp (9%), and hence the change of lattice parameters after the refinement was not found to be significant, reflecting a difference between the refined and the experimental values of about 0.01%. Hence, the DICVOL data was found to be of sufficient quality for calculation of the thermal expansion coefficient.

The thermal expansion was calculated along the a and c axes via equation (1)[link] with reference temperature and cell parameters from previous studies (Swaminathan et al., 1984[Swaminathan, S., Craven, B. M. & McMullan, R. K. (1984). Acta Cryst. B40, 300-3006.]), from which the parameters at 150 K are a = 5.59 and c = 4.69 Å.

3. Results and discussion

The lattice expansion as a function of temperature is given in Fig. 1[link]. No significant variation in cell parameters, reflecting polymorphic variation, was found to occur, confirming previous observations on the polymorphic behaviour of urea. The thermal expansion coefficient over the temperature range 188–328 K was measured to be (5.27 ± 0.26) × 10−5 K−1 in the a/b direction, and (1.14 ± 0.057) × 10−5 K−1 in the c direction, which correlates well with the values determined from previously published data over a temperature range of 12–123 K (Swaminathan et al., 1984[Swaminathan, S., Craven, B. M. & McMullan, R. K. (1984). Acta Cryst. B40, 300-3006.]). In this range, the coefficients of thermal expansion, calculated in the same way as used for the data presented here, were found to be (3.74 ± 0.187) × 10−5 K−1 in the a/b direction and (1.21 ± 0.0605) × 10−5 K−1 in the c direction.

[Figure 1]
Figure 1
The lattice expansion [(ll0)/l0, dimensionless] is represented as a function of the temperature. The coefficient of thermal expansion was determined from the slope of the straight line used to fit the experimental data.

Examining the above results, it is noteworthy that the coefficient of thermal expansion in the a/b direction is larger, for both temperature regions examined, compared with that in the c direction. This is expected, reflecting weaker intermolecular interactions associated with the hydrogen bonds along a/b axis (interchain) compared with those along the c axis (intrachain) (Docherty et al., 1993[Docherty, R., Roberts, K., Saunders, V., Black, S. & Davey, R. (1993). Faraday Discus. 95, 11-25.]), and hence the a/b axis allows for the greater thermal expansion. The comparison of the expansion coefficient for the two different temperature ranges, i.e. this study and that of Swaminathan et al., indicates a larger change of the thermal expansion coefficient in the a/b direction for the higher temperature range, again reflecting the weaker nature of the intermolecular interactions associated with the hydrogen bonding compared with the stronger ones along the c direction. In contrast, the change of the coefficient of thermal expansion in the c direction for the two temperature ranges is quite similar, being within the range of the error for these studies.

A literature survey (Table 1[link]) of thermal expansion data for organic crystals reveals that available data are limited for this class of materials. The thermal expansion of urea is comparable with other urea-based compounds (Meng & Lu, 1998[Meng, F. & Lu, K. (1998). J. Mater. Sci. Lett. 33, 265-268.]) but is larger than other aromatic hydrocarbons such as benzophenone (Girdwood, 1998[Girdwood, S. (1998). PhD thesis, University of Stathclyde, Glasgow, UK.]). This is not unexpected given the dominance of the van der Waals and hydrogen-bonding contribution to the crystal lattice energy.

Table 1
Results of a literature survey on previously measured coefficients of thermal expansion for some related organic crystals

Compound Coefficient of thermal expansion (K−1) Reference
α-Chloro­acetic acid (nonlinear) 7.7 × 10−5 Hoseneder & Hartel (2001[Hoseneder, R. & Hartel, G. (2001). Mater. Chem. Phys. 68, 278-279.])
MODPA (4,5-methyl-1,3,4-oxadiazin-2-yl-N,N-di­methylphenyl­amine) 1.8 × 10−4 Franco & Reck (2002[Franco, O. & Reck, G. (2002). J. Phys. Chem. Solids, 63, 1805-1813.])
DPO (2,5-di­phenyl-1,3,4-oxa­diazine) 1.9 × 10−4 Franco & Reck (2002[Franco, O. & Reck, G. (2002). J. Phys. Chem. Solids, 63, 1805-1813.])
L-Alanine 1.64 × 10−7, 12.07 × 10−7 Misoguti (1996[Misoguti, L. (1996). Opt. Mater. 6, 147-152.])
Hexamine 6 × 10−7 Fomo (1974[Fomo, C. (1974). J. Cryst. Growth, 21, 61-64.])
UDT (urea-d-tartaric acid) 5.23 × 10−5, 3.86 × 10−5, 3.57 × 10−5 Meng & Lu (1998[Meng, F. & Lu, K. (1998). J. Mater. Sci. Lett. 33, 265-268.])
Benzophenone 1.77 × 10−7, 3.00 × 10−7, 1.48 × 10−7 Girdwood (1998[Girdwood, S. (1998). PhD thesis, University of Stathclyde, Glasgow, UK.])
Urea 5.27 × 10−5, 1.14 × 10−5 This work

Acknowledgements

This work, which formed part of the PhD thesis of one of us (PM), has been carried out as part of the Chemical Behaving Badly initiative, a collaborative project funded by the UK's EPSRC funding body, via Grant GR/L/68797, together with industrial support from Astra Charnwood, BASF, GlaxoWellcome, ICI, Malvern Instruments Ltd, Pfizer, SmithKlineBeecham, and Zeneca. We gratefully acknowledge these sponsors and all the members of this academic/industrial team, notably the industrial coordinator L. J. Ford.

References

First citationBoultif, F. & Louer, D. (1991). J. Appl. Cryst. 24, 987–993.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationDesiraiju, G. (2003). J. Mol. Struct. 656, 5–15.  Web of Science CrossRef Google Scholar
First citationDocherty, R., Roberts, K., Saunders, V., Black, S. & Davey, R. (1993). Faraday Discus. 95, 11–25.  CrossRef CAS Web of Science Google Scholar
First citationFomo, C. (1974). J. Cryst. Growth, 21, 61–64.  CrossRef Web of Science Google Scholar
First citationFranco, O. & Reck, G. (2002). J. Phys. Chem. Solids, 63, 1805–1813.  Web of Science CSD CrossRef CAS Google Scholar
First citationGirdwood, S. (1998). PhD thesis, University of Stathclyde, Glasgow, UK.  Google Scholar
First citationHoseneder, R. & Hartel, G. (2001). Mater. Chem. Phys. 68, 278–279.  Web of Science CrossRef Google Scholar
First citationLarson, A. & von Dreele, B. (1994). GSAS, LosAlamos National Laboratory, New Mexico, USA.  Google Scholar
First citationMeng, F. & Lu, K. (1998). J. Mater. Sci. Lett. 33, 265–268.  CrossRef CAS Web of Science Google Scholar
First citationMisoguti, L. (1996). Opt. Mater. 6, 147–152.  CrossRef Web of Science Google Scholar
First citationSwaminathan, S., Craven, B. M. & McMullan, R. K. (1984). Acta Cryst. B40, 300–3006.  CSD CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationVaughan, P. & Donohue, J. (1952). Acta Cryst. 5, 530–535.  CrossRef CAS IUCr Journals Google Scholar

© International Union of Crystallography. Prior permission is not required to reproduce short quotations, tables and figures from this article, provided the original authors and source are cited. For more information, click here.

Journal logoJOURNAL OF
APPLIED
CRYSTALLOGRAPHY
ISSN: 1600-5767
Follow J. Appl. Cryst.
Sign up for e-alerts
Follow J. Appl. Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds