research papers\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoSTRUCTURAL SCIENCE
CRYSTAL ENGINEERING
MATERIALS
ISSN: 2052-5206

High-pressure neutron diffraction study of L-serine-I and L-serine-II, and the structure of L-serine-III at 8.1 GPa

CROSSMARK_Color_square_no_text.svg

aSchool of Chemistry and Centre for Science at Extreme Conditions, The University of Edinburgh, King's Buildings, West Mains Road, Edinburgh EH9 3JJ, Scotland, and bISIS Pulsed Neutron and Muon Facility, Rutherford Appleton Laboratory, Chilton, Didcot OX11 0QX, England
*Correspondence e-mail: s.parsons@ed.ac.uk

(Received 23 November 2005; accepted 15 May 2006)

The hydrostatic compression of L-serine-d7 has been studied to 8.1 GPa by neutron powder diffraction. Over the course of this pressure range the compound undergoes two phase transitions, the first between 4.6 and 5.2 GPa, yielding L-serine-II, and the second between 7.3 and 8.1 GPa, yielding L-serine-III. All three polymorphs are orthorhombic, P212121, and feature chains of serine molecules connected via head-to-tail ND⋯O hydrogen bonds formed between ammonium and carboxylate groups. The chains are linked into a ribbon by a second set of ND⋯O hydrogen bonds. The hydroxyl moieties are distributed along the outer edges of the ribbon and in phase I they connect the ribbons into a layer by chains of OD⋯OD hydrogen bonds. The layers are connected together by a third set of ND⋯O hydrogen bonds, forming [R^3_4(14)] rings with substantial voids at their centres. In the transition from phase I to II these voids begin to close up, but at the cost of breaking the OD⋯OD chains. The OD⋯OD hydrogen bonds are replaced by shorter OD⋯O hydrogen bonds to carboxylate groups. At 7.3 GPa the O⋯O distance in the OD⋯O hydrogen bonds measures only 2.516 (17) Å, which is short, and we propose that the phase transition to phase III that occurs between 7.3 and 8.1 GPa relieves the strain that has built up in this region of the structure. The hydroxyl D atom now bifurcates between the OD⋯O contact that had been present in phase II and a new OD⋯O contact formed to a carboxylate in another layer. Hirshfeld surface fingerprint plots show that D⋯D interactions become more numerous, while hydrogen bonds actually begin to lengthen in the transition from phase II to III.

1. Introduction

Pressure is an important thermodynamic variable in the study of molecular systems as it provides a useful tool for changing the hierarchy of interactions in the solid state. For example, hydrogen-bonding interactions are usually regarded as being `structure-directing', but as pressure increases the need to pack molecules efficiently is expected to begin to assume more importance. The application of pressure, unlike temperature, also allows structural studies to be performed that depend on changes in volume alone, and not the simultaneous effects of volume and thermal energy (Fourme et al., 2001[Fourme, R., Kahn, R., Mezouar, M., Girard, E., Hoerentrup, C., Prange, T. & Ascone, I. (2001). J. Synchrotron Rad. 8, 1149-1156.]).

The importance of pressure is well known in marine biology and deep-sea diving physiology. The application of pressure to biological systems has been known to have major effects, from the deactivation of enzymes (Somero, 1992[Somero, G. N. (1992). Ann. Rev. Physiol. 54, 557-577.]) to killing bacteria (Johnson et al., 1974[Johnson, F. H., Eyring, H. & Stover, B. J. (1974). The Theory of Rate Processes in Biology and Medicine. New York: Wiley.]). In fact, the structural effects of pressure were observed in proteins by Bridgeman as early as 1914, when the white of an egg subjected to hydrostatic pressure at room temperature became coagulated in appearance, becoming much like a hard-boiled egg. Recent work by Fourme et al. (2001[Fourme, R., Kahn, R., Mezouar, M., Girard, E., Hoerentrup, C., Prange, T. & Ascone, I. (2001). J. Synchrotron Rad. 8, 1149-1156.]) has opened up the possibility of collecting high-resolution, high-pressure data on large macromolecules using short wavelength X-rays (ca 0.3 Å). More recently, the first crystal structure of a macromolecular assembly under high-pressure was published on the structure of cubic cowpea mosaic virus crystals (Girard et al., 2005[Girard, E., Kahn, R., Mezouar, M., Dhaussy, A.-C., Lin, T., Johnson, J. E. & Fourme, R. (2005). Biophys. J. 88, 3562-3571.]).

A recent goal in the field of small-molecule crystallography has been to gain an understanding of the changes which occur in molecular systems as a function of pressure at atomic resolution (Boldyreva, 2004[Boldyreva, E. V. (2004). NATO Science Series, II: Mathematics, Physics and Chemistry, edited by A. Katrusiak & P. F. McMillan, Vol. 140, pp. 495-512. Dordrecht: Kluwer Academic Publishers.]; Katrusiak, 2004[Katrusiak, A. (2004). NATO Science Series, II: Mathematics, Physics and Chemistry, edited by A. Katrusiak & P. F. McMillan, Vol. 140, pp. 513-520. Dordrecht: Kluwer Academic Publishers.]; Hemley & Dera, 2000[Hemley, R. J. & Dera, P. (2000). Rev. Mineral. Geochem. 41, 335-419.]). Amino acids have several advantages in this area, as they are polyfunctional and form well defined intermolecular interactions such as hydrogen bonds and dipolar contacts; the results also have some relevance to the macromolecular work described above. In recent years a number of high-pressure crystallographic studies have been carried out on amino acids and related compounds, including glycine (Dawson et al., 2005[Dawson, A., Allan, D. R., Belmonte, S. A., Clark, S. J., David, W. I. F., McGregor, P. A., Parsons, S., Pulham, C. R. & Sawyer, L. (2005). Cryst. Growth Des. 5, 1415-1427.]; Boldyreva et al., 2003[Boldyreva, E. V., Ahsbahs, H. & Weber, H.-P. (2003). Z. Kristallogr. 218, 231-236.]), cystine (Moggach, Allan, Parsons, Sawyer & Warren, 2005[Moggach, S. A., Allan, D. R., Parsons, S., Sawyer, L. & Warren, J. E. (2005). J. Synchrotron Rad. 12, 598-607.]) and cysteine (Moggach et al., 2006[Moggach, S. A., Allan, D. R., Clark, S. J., Gutmann, M. J., Parsons, S., Pulham, C. R. & Sawyer, L. (2006). Acta Cryst. B62, 296-309.]).

We recently described the effect of pressure on the anhydrous orthorhombic phase of L-serine, L-serine-I, which undergoes a single crystal to single crystal phase transition at 5.4 GPa to a new high-pressure phase, L-serine-II (Moggach, Allan, Morrison, Parsons & Sawyer, 2005[Moggach, S. A., Allan, D. R., Morrison, C. A., Parsons, S. & Sawyer, L. (2005). Acta Cryst. B61, 58-68.]). This transition occurs with no change in symmetry. The reduction in volume is achieved by the compression of interstitial voids. In addition, the strain present in some hydrogen-bonded interactions, which become very short just below the transition pressure, is also relieved during the phase transformation.

[Scheme 1]

Our previous study on serine was carried out using X-ray diffraction and certain inferences regarding H-atom positions were made on the basis of N and O positions. We now describe a study of the compression of L-serine by neutron powder diffraction, which enables statements on H-atom positions to be made with a greater degree of certainty. Recently, a second phase transition was observed in L-serine via a high-pressure Raman study, the transition taking place at 7.8 GPa (Kolesnik et al., 2005[Kolesnik, E. N., Goryainov, S. V. & Boldyreva, E. V. (2005). Dokl. Phys. Chem. 404, 61-64.]). Here we confirm this reported transition to a third high-pressure phase of serine, L-serine-III, at 8.1 GPa, and describe its structure for the first time.

2. Experimental

L-Serine-d7 (98.2 atom% D) was purchased from CDN isotopes (catalogue number D-6466). One small crystallite was removed directly from the sample bottle and shown to be L-serine-I by single-crystal diffraction (Kistenmacher et al., 1974[Kistenmacher, T. J., Rand, G. A. & Marsh, R. E. (1974). Acta Cryst. B30, 2573-2578.]). The absolute configuration at the α-carbon atom was taken as S.

2.1. Neutron powder diffraction at high pressure

Ambient-temperature, high-pressure neutron powder diffraction data were collected by the time-of-flight technique at the PEARL beamline high-pressure facility (HiPr) at ISIS. Approximately 50 mm3 of L-serine-d7 was gently ground into a fine powder and loaded into a Paris–Edinburgh (P–E) high-pressure cell. The P–E cell ram pressure was monitored and controlled by means of a computer-controlled hydraulic system. The sample was loaded into a null-scattering Ti–Zr alloy capsule gasket (Marshall & Francis, 2002[Marshall, W. G. & Francis, D. J. (2002). J. Appl. Cryst. 35, 122-125.]). A pellet of lead was included as a pressure marker.

Time-of-flight neutron powder diffraction data suitable for structure refinement were obtained by electronically focusing the 1080 individual detector element spectra of the main PEARL/HiPr 2θ = 90° detector bank. The summed pattern was then normalized with respect to the incident beam monitor and the scattering from a standard vanadium calibration sample. Lastly, the diffraction pattern intensities were corrected for the wavelength and scattering-angle dependence of the neutron attenuation by the P–E cell anvil (WC) and gasket (TiZr) materials.

In order to ensure that the pressure applied to a crystal held in a pressure cell is uniform, it is necessary to immerse the sample in a medium which displays hydrostatic behaviour throughout the pressure range of interest. A 4:1 mixture of methanol and ethanol is commonly used in high-pressure work, although L-serine has some solubility in this mixture. A 1:1 mixture of deuterated pentane and 2-propanol was therefore used for an initial survey. Data were collected in approximately 0.7 GPa steps from 0.1 GPa to a final pressure of 4.3 GPa. Peak broadening occurred above 4.3 GPa, suggesting that conditions became non-hydrostatic at high pressures.

Based on our previous work on crystal growth from solutions at high pressure, the solubility of L-serine was expected to decrease at elevated pressures (Fabbiani et al., 2005[Fabbiani, F. P. A., Allan, D. R., Marshall, W. G., Parsons, S., Pulham, C. R. & Smith, R. I. (2005). J. Cryst. Growth, 275, 185-192.]). In order to extend the pressure range for this study a second sample of L-serine was loaded with a 4:1 mixture of MeOD:EtOD as the hydrostatic medium. Diffraction data were collected in approximately 1 GPa steps from 4.6 GPa to a final pressure of 8.1 GPa. On increasing pressure to 5.2 GPa, the phase transition from L-serine-I to L-serine-II was observed, a transition which we have observed previously in a single-crystal high-pressure X-ray diffraction experiment on a non-deuterated sample of L-serine-I (Moggach, Allan, Morrison, Parsons & Sawyer, 2005[Moggach, S. A., Allan, D. R., Morrison, C. A., Parsons, S. & Sawyer, L. (2005). Acta Cryst. B61, 58-68.]). On increasing pressure further to 8.1 GPa, another phase transition was observed to a hitherto uncharacterized phase of L-serine which we have designated L-serine-III.

2.2. Structure solution and refinement

Powder-pattern indexing, structure solution and Rietveld refinements were carried out using TOPAS-A (Coelho, 2005[Coelho, A. (2005). Topas-A: General Profile and Structure Analysis Software for Powder Diffraction Data; https://members.optusnet.com.au/~alancoelho/ .]). Rietveld refinement of phases I and II was carried out starting from the 4.8 and 5.4 GPa coordinates of Moggach, Allan, Morrison, Parsons & Sawyer (2005[Moggach, S. A., Allan, D. R., Morrison, C. A., Parsons, S. & Sawyer, L. (2005). Acta Cryst. B61, 58-68.]) for L-serine-I and L-serine-II, respectively.

The structure of the new phase (L-serine-III) was indexed on an orthorhombic unit cell, a = 6.5574, b = 9.5397, c = 5.6101 Å, in the space group P212121 using the method of Coelho (2003[Coelho, A. (2003). J. Appl. Cryst. 36, 86-95.]). The structure was then solved by simulated annealing. The crystal structure of phase III could be re-solved to yield the same structure in repeated simulated annealing runs. The same solution was also obtained by the refinement of phase II coordinates: the difference between the phases largely lies in the orientation of the hydroxyl group and the two structures are similar enough that one can be refined into the other. During refinement the serine molecules were treated as variable metric rigid groups using the Z-matrix formalism available in TOPAS.

Lead (included in the sample as a pressure marker), tungsten carbide and nickel were also included in the refinement models. The weak tungsten carbide and nickel peaks both arise from the P–E cell anvil cores which consist of cemented WC with a Ni binder. The refined cell dimensions of lead were used to calculate the applied pressure using a Birch–Murnaghan equation-of-state (Birch, 1947[Birch, F. (1947). Phys. Rev. 71, 809-824.]) using Vo = 30.3128 Å3, Bo = 41.92 GPa, B′ = 5.72. These parameters were derived by Fortes (2004[Fortes, A. D. (2004). PhD Thesis: Computational and Experimental Studies of Solids in the Ammonia-Water System. Department of Earth Sciences, University College, University of London.]) as averages of the values in Kuznetsov et al. (2002[Kuznetsov, A. Z., Dmitriev, V., Dubrovinsky, L., Prakapenka, V. & Weber, H.-P. (2002). Solid State Commun. 122, 125-127.]), Miller & Schuele (1969[Miller, R. A. & Schuele, D. E. (1969). J. Phys. Chem. Solids, 30, 589-600.]) and Waldorf & Alers (1962[Waldorf, D. L. & Alers, G. A. (1962). J. Appl. Phys. 33, 3266-3269.]).

All intramolecular bond distances and angles were constrained to their ambient pressure values for all but the hydroxyl group, whose O—D distance and C—O—D angle were refined. All torsion angles were also refined. All N—D and C—D distances were constrained to 1.009 and 1.083 Å, respectively. All non-H atoms were refined with a common isotropic displacement parameter; the displacement parameters of the D atoms were equal to 1.5 times this value. The final Rietveld fits are shown in Fig. 1[link]. Unit-cell dimensions and refinement parameters for L-serine-I at 4.6 GPa, L-serine-II at 5.2 and 7.3 GPa, and L-serine-III at 8.1 GPa are included in Table 1[link].1 Fourier difference maps calculated after the removal of some atoms returned the positions of the same atoms again; the success of this test attests to the quality of the data measured.

Table 1
Experimental details

  4.5 GPa 5.2 GPa 7.3 GPa 8.1 GPa
Phase I I II III
Crystal data
Chemical formula C3D7NO3 C3D7NO3 C3D7NO3 C3D7NO3
Mr 112.11 112.11 112.11 112.11
Cell setting, space group Orthorhombic, P212121 Orthorhombic, P212121 Orthorhombic, P212121 Orthorhombic, P212121
Temperature (K) 298 298 298 298
a, b, c (Å) 8.3096 (3), 8.5950 (3), 5.4647 (3) 6.8700 (3), 9.6396 (5), 5.6058 (3) 6.6874 (3), 9.5815 (6), 5.5619 (3) 6.5477 (4), 9.5405 (6), 5.6092 (4)
V3) 390.29 (3) 371.24 (3) 356.38 (4) 350.40 (4)
Z 4 4 4 4
Dx (Mg m−3) 1.908 2.006 2.089 2.125
Radiation type Neutron Neutron Neutron Neutron
Specimen form, colour Powder (particle morphology: block), colourless Powder (particle morphology: block), colourless Powder (particle morphology: block), colourless Powder (particle morphology: block), colourless
Specimen size (mm3) 55 55 55 55
         
Data collection
Diffractometer HiPr/PEARL, ISIS HiPr/PEARL, ISIS HiPr/PEARL, ISIS HiPr/PEARL, ISIS
Data collection method Specimen mounting: TiZr capsule; mode: transmission; scan method: time-of-flight Specimen mounting: TiZr capsule; mode: transmission; scan method: time-of-flight Specimen mounting: TiZr capsule; mode: transmission; scan method: time-of-flight Specimen mounting: TiZr capsule; mode: transmission; scan method: time-of-flight
d-spacing range (Å) 0.6–4.2 0.6–4.3 0.6–4.1 0.6–4.2
         
Refinement
Refinement on I I I I
R factors and goodness-of-fit Rp = 0.0299, Rwp = 0.0414, Rexp = 0.1074, S = 0.39 Rp = 0.0291, Rwp = 0.0415, Rexp = 0.1099, S = 0.38 Rp = 0.0330, Rwp = 0.0474, Rexp = 0.1226, S = 0.39 Rp = 0.0346, Rwp = 0.0492, Rexp = 0.1291, S = 0.38
Profile function Voigt function convoluted with a back-to-back double exponential Voigt function convoluted with a back-to-back double exponential Voigt function convoluted with a back-to-back double exponential Voigt function convoluted with a back-to-back double exponential
No. of parameters 39 39 39 39
Weighting scheme Based on measured s.u.'s Based on measured s.u.'s Based on measured s.u.'s Based on measured s.u.'s
(Δ/σ)max <0.0001 <0.0001 <0.0001 <0.0001
Preferred orientation correction Preferred orientation was treated with a spherical harmonic expansion to 4th order Preferred orientation was treated with a spherical harmonic expansion to 4th order Preferred orientation was treated with a spherical harmonic expansion to 4th order Preferred orientation was treated with a spherical harmonic expansion to 4th order
[Figure 1]
Figure 1
Rietveld refinement plots of L-serine: (a) phase I at 4.6 GPa; (b), (c) phase II at 5.2 and 7.3 GPa; (d) phase III at 8.1 GPa. The blue and red lines correspond to the observed and calculated powder patterns, respectively; the difference plot is shown in grey.

Crystal structures were visualized using the programs MERCURY (Bruno et al., 2002[Bruno, I. J., Cole, J. C., Edgington, P. R., Kessler, M., Macrae, C. F., McCabe, P., Pearson, J. & Taylor, R. (2002). Acta Cryst. B58, 389-397.]) and DIAMOND (Crystal Impact, 2004[Crystal Impact (2004). DIAMOND, Version 3.0. Crystal Impact GbR, Postfach 1251, 53002 Bonn, Germany. https://www.crystalimpact.com/diamond .]). Analyses were carried out using PLATON (Spek, 2003[Spek, A. L. (2003). J. Appl. Cryst. 36, 7-13.]), as incorporated in the WIN-GX suite (Farrugia, 1999[Farrugia, L. J. (1999). J. Appl. Cryst. 32, 837-838.]). Hirshfeld surface analysis was performed using the program CrystalExplorer (Wolff et al., 2005[Wolff, S. K., Grimwood, D. J., McKinnon, J. J., Jayatilaka, D. & Spackman, M. A. (2005). CrystalExplorer1.5. University of Western Australia; https://www.Theochem.uwa.edu.au/crystal_explorer/ .]).

3. Results and discussion

3.1. The response of L-serine-I-d7 to 4.5 GPa

We have previously investigated the response of L-serine-I to pressure using single-crystal X-ray diffraction (Moggach, Allan, Morrison, Parsons & Sawyer, 2005[Moggach, S. A., Allan, D. R., Morrison, C. A., Parsons, S. & Sawyer, L. (2005). Acta Cryst. B61, 58-68.]). Boldyreva et al. (2005[Boldyreva, E. V., Kolesnik, E. N., Drebushchak, T. N., Ahsbahs, H., Beukes, J. A. & Weber, H.-P. (2005). Z. Kristallogr. 220, 58-65.]) also studied L-serine-I to 4.4 GPa, obtaining similar results. While deuterated materials can sometimes display different phase behaviour to their perhydrogenated analogues, the response of deuterated L-serine-d7-I is essentially the same as that of L-serine-h7-I up to 4.6 GPa. On increasing pressure to 5.2 GPa, the sample underwent a phase transition to L-serine-II. The phase transition from L-serine-I to L-serine-II is marked by a sudden decrease in volume (Fig. 2[link]).

[Figure 2]
Figure 2
Variation of the lattice parameters (a)–(c) and volume (d) of L-serine as a function of pressure (GPa).

3.2. Description of phases I and II

We have discussed the structures of L-serine-I and II in detail in Moggach, Allan, Morrison, Parsons & Sawyer (2005[Moggach, S. A., Allan, D. R., Morrison, C. A., Parsons, S. & Sawyer, L. (2005). Acta Cryst. B61, 58-68.]), and our motive for including a summary here is to highlight hydrogen-bonded motifs relevant to the development of L-serine-III, described below.

C(5) chains are formed in both L-serine-I and II along c by N1D5⋯O2 hydrogen bonds (Figs. 3[link]a and b). A second C(5) chain, generated from the first via a 21 operation, is linked to the first through N1D6⋯O1 hydrogen bonds, and together the N1D5⋯O2 and N1D6⋯O1 interactions form ribbons containing successive R33(11) ring motifs. The CD2OD side chains are distributed along the outside edges of these ribbons. In L-serine-I these link the ribbons into a layer via OD⋯OD⋯OD interactions (Fig. 3[link]a). In L-serine-II the ribbons are connected into a layer by O3D7⋯O2 hydrogen bonds formed between the alcohol and carboxyl moieties (Fig. 3[link]b). The graph-sets thus-formed are R33(13) and R23(13), and the layers are stacked along the a axis in both phases.

[Figure 3]
Figure 3
(a) L-Serine-I at 4.6 GPa, (b) L-serine-II at 5.2 GPa and (c) L-serine-III at 8.1 GPa as viewed along a. R33(13), R23(13) and R33(11) ring motifs are shown. ND⋯O and OD⋯O hydrogen bonds are drawn as black dotted lines, while CD⋯O interactions are drawn as orange dotted lines. For clarity, the bifurcated interaction N1D5⋯O1/O2 and CD⋯O interactions are shown in the top-left corner only. These layers are referred to as the A layers in the text. Colour scheme: C grey, H white, N blue and O red.

The layers are linked together in both phases I and II by N1D4⋯O2 interactions, which build another primary-level C(5) chain. The combination of the N1D5⋯O2 and N1D4⋯O2 C(5) chains along a and c builds a third set of secondary-level ring motifs, those having the descriptor [R^3_4(14)] (Figs. 4[link]a and b). In fact, these chains can be viewed as building up another layer, which is parallel to the ac plane. Overall, the structure consists of two sets of layers: one stacks along a and contains R33(11) and R33(13) ring motifs, the other stacks along b and contains R34(14) rings. In Moggach, Allan, Morrison, Parsons & Sawyer (2005[Moggach, S. A., Allan, D. R., Morrison, C. A., Parsons, S. & Sawyer, L. (2005). Acta Cryst. B61, 58-68.]) we referred to these as the A and B layers, respectively.

[Figure 4]
Figure 4
(a) L-Serine-I at 4.6 GPa, (b) L-serine-II at 5.2 GPa and (c) L-serine-III at 8.1 GPa as viewed along b. [R^3_4(14)] ring motifs are shown. ND⋯O and OD⋯O hydrogen bonds are drawn as black dotted lines, while CD⋯O interactions are drawn as orange dotted lines. For clarity, the bifurcated interaction N1D5⋯O1/O2 and CD⋯O interactions are shown in the top-left corner only. These layers are referred to as the B layers in the text. The colour scheme is the same as in Fig. 3[link].

3.3. Comparison of L-serine-I at 4.5 GPa and L-serine-II at 5.2 GPa

One difference between L-serine-I and -II is the position of the hydroxyl H atom. In our original X-ray study the position of this H atom was rather imprecisely determined, but our conclusions are confirmed by the present neutron study. The O⋯O and D⋯O distances in the O3D7⋯O2 hydroxyl–carboxyl hydrogen bonds in phase II are substantially shorter than the O3D7⋯O3 hydroxyl–hydroxyl hydrogen bonds in phase I (Table 2[link]).

Table 2
Hydrogen-bonding parameters (Å, °) in L-serine

Roman numerals refer to the phase of L-serine.

Pressure (GPa) 4.6I 5.2II 7.3II 8.1III
N1D4⋯O2i        
N1⋯O2 2.743 (11) 2.821 (17) 2.764 (18) 2.836 (15)
D4⋯O2 1.839 (11) 1.877 (19) 1.85 (2) 1.890 (16)
∠N1D4O2 147.3 (5) 154.3 (12) 149.3 (12) 154.8 (7)
N1D5⋯O1ii        
N1⋯O1 3.012 (5) 3.071 (4) 3.008 (4) 3.085 (3)
D5⋯O1 2.215 (8) 2.243 (8) 2.168 (8) 2.304 (7)
∠N1D5⋯O1 134.8 (5) 138.4 (5) 139.6 (5) 133.3 (4)
N1D5⋯O2ii        
N1⋯O2 2.714 (10) 2.823 (10) 2.829 (11) 2.776 (9)
D5⋯O2 1.762 (13) 1.875 (13) 1.902 (15) 1.796 (11)
∠N1D5O2 155.7 (5) 155.2 (9) 151.3 (10) 162.9 (4)
N1D6⋯O1iii        
N1⋯O1 2.631 (14) 2.642 (15) 2.640 (17) 2.694 (18)
D6⋯O1 1.683 (14) 1.733 (14) 1.735 (16) 1.710 (18)
∠N1D6O1 154.6 (5) 147.8 (6) 147.1 (7) 164.2 (8)
O3D7⋯O3iv        
O3⋯O3 2.780 (5)
D7⋯O3 1.954 (18)
∠O3D7O3 156 (2)
O3D7⋯O2iv        
O3⋯O2 2.576 (16) 2.516 (17) 2.638 (15)
D7⋯O2 1.67 (3) 1.56 (3) 2.05 (3)
<O3D7O2 168 (3) 174 (3) 122 (2)
O3D7⋯O1v        
O3⋯O1 2.775 (13)
D7⋯O1       1.95 (3)
∠O3D7O1       152 (2)
C2D1⋯O1vi        
D1⋯O1 2.499 (14) 2.438 (16) 2.350 (17) 2.341 (16)
∠C2D1O1 119.2 (4) 111.1 (4) 109.7 (4) 105.9 (4)
C2D1⋯O2vi        
D1⋯O2 2.385 (15) 2.338 (16) 2.322 (14)
∠C2D1O2   159.6 (4) 157.9 (4) 154.7 (4)
C3D2⋯O2ii        
D2⋯O2 2.389 (12) 2.297 (12) 2.378 (11)
∠C3D2O2 128.2 (7) 129.7 (8) 129.2 (7)
C3D2⋯O3iv        
D2⋯O3 2.41 (3) 2.44 (3) 2.39 (3) 2.32 (2)
∠C3D2O3 114.5 (5) 131.3 (8) 132.5 (9) 152.7 (7)
C3D3⋯O3vii        
D3⋯O3 2.37 (3) 2.32 (2) 2.26 (2)
∠C3D3O3 132.9 (10) 134.1 (11) 157.2 (12)
C3D3⋯O1vi        
D3⋯O1 2.420 (15) 2.36 (2) 2.32 (2) 2.229 (17)
∠C3D3O1 105.7 (6) 109.5 (7) 109.5 (7) 113.4 (6)
Symmetry codes: (i) [-{1\over 2}+x,{3\over 2}-y,1-z]; (ii) x,y,1+z; (iii) [{3\over 2}-x,2-y,{1\over 2}+z]; (iv) [{3\over 2}-x,1-y,{1\over 2}+z]; (v) [1-x,-{1\over 2}+y,{3\over 2}-z]; (vi) [{1\over 2}+x,{3\over 2}-y,1-z]; (vii) [{3\over 2}-x,1-y,-{1\over 2}+z].

The conformations of the [R^3_4(14)] rings in L-serine-I and -II represent a second difference between the crystal structures of the two phases. Voids observed in L-serine-I-d7 at 4.6 GPa at the centres of the [R^3_4(14)] ring motifs are closed-up on undergoing the transition (Figs. 5[link]a and b). This is accomplished by:

  • (i) a displacement along the +c and −c directions by successive C(5) chains formed by N1D5⋯O2 hydrogen bonds, and

  • (ii) a shortening of the distance between the chains along the a direction (Figs. 6[link]a and b).

The opposite directions of displacement of the C(5) chains break the O3D7⋯O3 hydroxyl–hydroxyl hydrogen bonds which are replaced by O3D7⋯O2 hydroxyl–carboxylate hydrogen bonds in phase II.
[Figure 5]
Figure 5
Space-filling plots for [R^3_4(14)] ring motifs in (a) L-serine-I at 4.6 GPa, (b) L-serine-II at 5.2 GPa, (c) L-serine-II at 7.3 GPa and (d) L-serine-III at 8.1 GPa as viewed along b. The CD2OD side groups have been removed for clarity. The colour scheme is the same as in Fig. 3[link].
[Figure 6]
Figure 6
Graphs showing the distances between the centroids of the molecules in neighbouring C(5) chains along (a) c and (b) a.

CD⋯O contacts are also more numerous in L-serine-II than in L-serine-I. All C—H bonds act as donors in two hydrogen-bonding interactions which span the R23(13) and [R^3_4(14)] rings (Table 2[link], Figs. 3[link]b and 4[link]b).

3.4. L-Serine-II between 5.2 and 7.3 GPa

As the pressure on L-serine-II is increased from 5.2 to 7.3 GPa, the largest compression occurs along the a direction, with the a axis decreasing in length by 2.7%; the b and c axes reduce by only 0.6 and 0.8%, respectively. The relative displacement of the N1D5⋯O2 C(5) chains and the reduction in the distance between the chains both continue up to 7.3 GPa, leading to further compression of the voids within [R^3_4(14)] ring motifs (Figs. 6[link]a and b, and 5[link]c).

In the most compressible ND⋯O hydrogen bond, N1D5⋯O1, the D⋯O distance decreases by 0.075 (11) Å between 5.2 and 7.3 GPa. This hydrogen bond is the longest of its type at 5.2 GPa, and its shortening increases the bifurcation in the interaction N1D5⋯O1/O2. N1D6⋯O1, which is short anyway at 5.2 GPa, remains relatively unchanged. The largest change occurs for O3D7⋯O2, which was formed in the transition from phase I to phase II: by 7.3 GPa the O⋯O distance in this contact measures only 2.516 (17) Å, which is very short. Although it is short, O3D7⋯O2 is unsymmetrical, with O3—D7 and D7⋯O2 measuring 0.97 (2) and 1.56 (3) Å, respectively.

The shortest OH⋯O hydrogen bond found under ambient pressure conditions [2.474 (2) Å] occurs in the CSD refcode AMHACY (Leban, 1974[Leban, I. (1974). Cryst. Struct. Commun. 3, 237-239.]). This, however, corresponds to a symmetrical hydrogen bond where the H atom is shared equally between adjacent carboxyl groups from neighbouring molecules. Strong hydrogen bonds of this type are common when formed between acids and complementary bases (Steiner, 2002[Steiner, T. (2002). Angew. Chem. Int. Ed. 41, 48-76.]). If this search is limited to hydrogen-bonding interactions between a free hydroxyl group donor (bonded to an alkyl CH2) and carboxyl group acceptor, the shortest interaction observed under ambient conditions measures 2.584 (2) Å for GOLWIN (Tusvik et al., 1999[Tusvik, P. H., Mostad, A., Dalhus, B. & Rosenqvist, E. (1999). Acta Cryst. C55, 1113-1115.]). The O3D7⋯O2 hydrogen bond therefore appears to be pushed to below this lower limit observed under ambient pressure conditions, and on increasing pressure further the structure undergoes another phase transition to a second high-pressure phase, L-serine-III.

3.5. Structure of L-serine-III at 8.1 GPa

The phase transition from L-serine-II to L-serine-III, unlike that from phase I to II, occurs with only a modest decrease in volume (Fig. 2[link]d); the volume per non-H atom at 8.1 GPa is only 12.5 Å3. Overall the structure is closely related to the structures of phases I and II (Figs. 3[link]c and 4[link]c), and the structure can again be considered in terms of layers which stack along the a and b directions.

The [R^3_4(14)] ring motifs in the B layers appear longer and thinner than they do in L-serine-I and -II (Figs. 4[link]b and c), and over the course of the II-to-III phase transition the relative displacement of neighbouring N1D5⋯O2 C(5) chains in the +c and −c directions continues (Fig. 6[link]a). The associated compression along the a direction also continues (Fig. 6[link]b) and at 8.1 GPa there are no visible voids within the [R^3_4(14)] rings (Fig. 5[link]d). Small voids present at the centre of R23(13) ring motifs of the A layers remain essentially unchanged through the transition, while no observable voids can be observed at the centres of R33(11) ring motifs in phase II or III.

The phase transition from L-serine-I to L-serine-II was, in part, characterized by a change in the hydrogen bonds formed by the hydroxyl group, O3D7. A further change in hydrogen bonding in this region of the structure also occurs in the transition from phase II to phase III. In phase II the hydrogen bonding from O3D7 was confined to one A layer, but in phase III the hydrogen bonding from D7 becomes bifurcated between the same O3D7⋯O2 interaction that was present in phase II, and a new O3D7⋯O1 hydrogen bond which connects different A layers together. This change in the hydrogen-bonding environment of the O3D7 group may explain the findings of Kolesnik et al. (2005[Kolesnik, E. N., Goryainov, S. V. & Boldyreva, E. V. (2005). Dokl. Phys. Chem. 404, 61-64.]), who observed that the largest changes in the Raman spectra of L-serine on undergoing the transition from phase II to III occurred in bands at 1250 (HOC bend) and 1430 cm−1 (CO2 sym. str). When viewed in projection the A and B layers seem very similar in phases II and III: the difference lies in the way they are connected together and this is clearer in the c-axis projection illustrated in Fig. 7[link].

[Figure 7]
Figure 7
(a) L-Serine-II at 5.2 GPa and (b) L-serine-III at 8.1 GPa as viewed along c. A and B layers as referred to in the text are labelled. ND⋯O and OD⋯O hydrogen bonds are drawn as black dotted lines, while CD⋯O interactions are drawn as orange dotted lines. For clarity, CD⋯O interactions are shown in the top-left corner only. Note the formation of the O1D7⋯O1 interaction in L-serine-III at 8.1 GPa. The colour scheme is the same as in Fig. 3[link].

The O3⋯O2 distance within the O3D7⋯O2 hydrogen bonds in phase II is very short [2.516 (17) Å] at 7.3 GPa and the structure presumably experiences significant strain in this region. It seems reasonable to suggest that the II-to-III transition is driven by the need to alleviate close contacts about O3 in L-serine-II as pressure is increased beyond 7.3 GPa.

In general, at elevated pressures the influence of hydrogen bonds as structure-directing interactions is expected to diminish relative to the need to pack molecules more efficiently and several of the ND⋯O and OD⋯O hydrogen bonds in phase III at 8.1 GPa are longer than those in phase II at 7.3 GPa. In L-serine-II the hydrogen-bonding capacity of the C—D bonds is saturated and so the transition from phase II to III does not result in an increase in number of CD⋯O interactions, with the six interactions present in L-serine-II, still observed in L-serine-III.

3.6. Changes in CD⋯O and D⋯D interactions: Hirshfeld fingerprint plot analysis

McKinnon et al. (2004[McKinnon, J. J., Spackman, M. A. & Mitchell, A. S. (2004). Acta Cryst. B60, 627-668.]) have recently described how a Hirshfeld surface analysis can assist in the interpretation of differences between related crystal structures. Hirshfeld surfaces are created by partitioning space within a crystal structure into regions where the electron density from a sum of spherical atoms (the promolecule) dominates over the sum of the electron density of the crystal (the procrystal). The surfaces surround each molecule in a crystal structure, enabling various properties relating to packing to be calculated. Two such properties are de (the distance from the surface to the nearest atom in another molecule) and di (the distance from the surface to the nearest atom in the molecule itself). Plots of de against di are referred to as fingerprint plots and they encode information on the overall three-dimensional packing characteristics in a two-dimensional diagram. Fingerprint plots for the structures obtained in this study are shown in Fig. 8[link].

[Figure 8]
Figure 8
Two-dimensional fingerprint plots for (a) L-serine-I at 0.1 GPa, (b) L-serine-I at 4.6 GPa, (c) L-serine-II at 5.2 GPa, (d) L-serine-II at 7.3 GPa and (e) L-serine-III at 8.1 GPa.

On increasing the pressure in L-serine-I to between 0.1 to 4.6 GPa, the drive to compress the hydrogen-bonding interactions is represented clearly on the fingerprint plots with de values driven to shorter distances (Figs. 8[link]a and b). In particular, the two long spikes present in all fingerprint plots represent ND⋯O and OD⋯O hydrogen-bonding interactions, which can clearly be observed moving toward smaller values of de in L-serine-I (Figs. 8[link]a and b). The number of points at larger values of de and di also decrease on increasing pressure, giving a more compact plot at 4.6 GPa; the fingerprint plot is still more compact at 5.2 GPa (Fig. 8[link]c). This illustrates the compression of voids within R33(11) and [R^3_4(14)] ring motifs within A and B layers, respectively. A `hump' appears between the spikes at 4.5 GPa, becoming more prominent at 5.2 GPa. This corresponds to close D⋯D contacts. The shortest of these contacts at 4.6 GPa, D3⋯D6, forms between A layers between D atoms from CD2 and ammonium groups of adjacent molecules, respectively. D⋯D contacts become more numerous at 5.2 GPa, for example a D3⋯D6 distance formed along the a direction decreases from 2.19 to 2.08 Å.

On increasing the pressure further to 7.3 GPa, the fingerprint plots look much the same, although a slight movement of the NH⋯O/OD⋯O spikes toward the lower values of de can be observed (Figs. 8[link]c and d). The `hump' between the spikes at 7.3 GPa also appears to sharpen and move to larger values of de. This is accounted for by the movement of the hydroxyl OD group, which at 5.2 GPa forms the shortest D⋯D contact (D7⋯D3, 1.91 Å). On increasing the pressure to 7.3 GPa, this close contact is actually relieved by the hydroxyl group, which twists, making the O3D7⋯O2 hydrogen bond more linear, while increasing the D7⋯D3 distance (1.98 Å). This movement of the hydroxyl group can be quantified by comparing the D7—O3—C3—C2 torsion angle at 5.2 GPa (−178.5°) and 7.3 GPa (+172.3°). Nevertheless, on increasing pressure to 7.3 GPa, the number of close D⋯D contacts does increase, represented in Fig. 8[link](d) as an increase in the green colour within the `hump' region of the plot.

At 8.1 GPa the spikes corresponding to ND⋯O hydrogen bonds appear at larger values of de; this is caused by the increased lengths of ND⋯O and OD⋯O contacts in phase III referred to above (Figs. 8[link]d and e). We have suggested above that the II-to-III transition is driven by the need to relieve short intermolecular contacts, and the similarity of the volumes of phases II and III suggests that the transition does not result in markedly more efficient packing. However, there is a pronounced increase in the number of close D⋯D contacts between the spikes in the fingerprint plot shown in Figs. 8[link](d) and (e). Close contacts between H atoms are characteristic of the fingerprint plots obtained for alkanes (McKinnon et al., 2004[McKinnon, J. J., Spackman, M. A. & Mitchell, A. S. (2004). Acta Cryst. B60, 627-668.]) and in serine their gradually increasing prominence illustrates much more graphically than tables of data the tendency for efficient packing to compete with hydrogen bonding as structure-directing phenomena.

4. Summary

We have previously described a single-crystal X-ray study in which we observed a single crystal to single crystal phase transition from L-serine-I to an orthorhombic high-pressure phase, L-serine-II, between 4.8 and 5.4 GPa. The same transition has also been observed by neutron powder diffraction in perdeuterated L-serine between 4.6 and 5.2 GPa. The I-to-II transition is accompanied by a large reduction in volume, achieved by opposite displacements of successive C(5) chains formed by ND⋯O hydrogen bonds and a shortening of the distance between the chains. Overall, this closes up the voids present in hydrogen-bonded ring motifs, enabling the reduction in volume to occur. The opposite direction of the displacements of the C(5) chains breaks the OD⋯O hydroxyl–hydroxyl hydrogen bonds present in phase I. These are replaced by OD⋯O hydroxyl–carboxylate interactions in phase II. The change in position of the hydroxyl H atom was clearly demonstrated in this study.

Compression of L-serine-II from 5.2 to 7.3 GPa caused the O⋯O distance in the newly formed OD⋯O hydroxyl–carboxylate interaction to decrease to a very low 2.516 (17) Å. This is very short and it is possible that relief of the strain associated with this contact is responsible for a phase transition to a third form of serine as the pressure is increased from 7.3 to 8.1 GPa. In phase III the short OD⋯O hydroxyl–carboxylate interaction is replaced by a bifurcated interaction in which the OD interacts with two carboxylate moieties with O⋯O distances of 2.638 (15) and 2.775 (13) Å. The trend towards opposite displacements of successive C(5) chains formed by N1D5⋯O2 hydrogen bonds and a shortening of the distance between the chains continued through the II-to-III transition leading to very efficient packing of chains in phase III.

The Hirshfeld fingerprint plots show the changes in the balance of intermolecular interactions as a function of pressure. The increasing importance of D⋯D interactions at high pressure is particularly apparent. Hydrogen bonds seem to be longer in L-serine-III than L-serine-II, and the increase in D⋯D interactions revealed in the fingerprint plots point to the compromise between hydrogen bonding and efficient packing that must be reached at elevated pressures.

Supporting information


Computing details top

Figures top
[Figure 1]
[Figure 2]
[Figure 3]
[Figure 4]
[Figure 5]
[Figure 6]
[Figure 7]
[Figure 8]
(serine45) L-serine-d7 top
Crystal data top
C3D7NO3Z = ?
Mr = 112.11Cell parameters included in refinement
Orthorhombic, P212121Dx = 1.908 Mg m3
Hall symbol: P 2ac 2abNeutron radiation, λ = ? Å
a = 8.3096 (3) ÅT = 298 K
b = 8.5950 (3) ÅParticle morphology: Block
c = 5.4647 (3) ÅColourless
V = 390.29 (3) Å3?, ? × ? × ? mm
Data collection top
HiPr/PEARL, ISIS
diffractometer
Scan method: Time-of-flight
Radiation source: Pulsed neutron sourceAbsorption correction: empirical (using intensity measurements)
Calculated using custom ISIS software.
Specimen mounting: TiZr CapsuleTmin = ?, Tmax = ?
Data collection mode: transmission
Refinement top
Least-squares matrix: selected elements onlyProfile function: Voigt function convoluted with a back-to-back double exponential.
Rp = 2.98539 parameters
Rwp = 4.1400 restraints
Rexp = ?Weighting scheme based on measured s.u.'s
χ2 = 0.148(Δ/σ)max < 0.001
? data pointsBackground function: Chebychev polynomial
Excluded region(s): excluded short TOF region below d = 0.6 Ang.Preferred orientation correction: Preferred orientation was treated with a spherical harmonic expansion to 4th order:
sh_c20 -0.01856`_0.01347 sh_c22p -0.11996`_0.01681 sh_c40 -0.04522`_0.01736 sh_c42p 0.03326`_0.01975 sh_c44p -0.04617`_0.02528
Crystal data top
C3D7NO3V = 390.29 (3) Å3
Mr = 112.11Z = ?
Orthorhombic, P212121Neutron radiation, λ = ? Å
a = 8.3096 (3) ÅT = 298 K
b = 8.5950 (3) Å?, ? × ? × ? mm
c = 5.4647 (3) Å
Data collection top
HiPr/PEARL, ISIS
diffractometer
Scan method: Time-of-flight
Specimen mounting: TiZr CapsuleAbsorption correction: empirical (using intensity measurements)
Calculated using custom ISIS software.
Data collection mode: transmissionTmin = ?, Tmax = ?
Refinement top
Rp = 2.985? data points
Rwp = 4.14039 parameters
Rexp = ?0 restraints
χ2 = 0.148
Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzBiso*/Beq
C10.8924170.8044830.3497981.04 (7)
C20.9292780.7519200.6101781.04 (7)
C30.9305160.5769740.6310201.04 (7)
N10.8056860.8130860.7815561.04 (7)
O10.7771800.8923640.3166501.04 (7)
O20.9835210.7534660.1874451.04 (7)
O30.7772040.5139130.5619521.04 (7)
D11.0458220.7931810.6706661.56 (10)
D20.9563950.5475810.8196731.56 (10)
D31.0216270.5274460.5126411.56 (10)
D40.7009560.7567150.7520691.56 (10)
D50.8429520.7950690.9549851.56 (10)
D60.7911860.9280380.7512671.56 (10)
D70.7378520.4867510.7050431.56 (10)
Geometric parameters (Å, º) top
O1—C11.23N1—D61.01
O2—C11.25C1—C21.53
O3—C31.43C2—C31.50
O3—D70.879 (16)C2—D11.01
N1—C21.49C3—D21.01
N1—D41.01C3—D31.01
N1—D51.01
C3—O3—D7101.3 (15)N1—C2—C3108
C2—N1—D4109C1—C2—C3112
C2—N1—D5109O3—C3—C2111
C2—N1—D6109N1—C2—D1108
D4—N1—D5110C1—C2—D1111
D4—N1—D6110C3—C2—D1107
D5—N1—D6110O3—C3—D2110
O1—C1—O2126O3—C3—D3108
O1—C1—C2118C2—C3—D2108
O2—C1—C2116C2—C3—D3111
N1—C2—C1110D2—C3—D3109
(serine52) L-serine-d7 top
Crystal data top
C3D7NO3Z = ?
Mr = 112.11Cell parameters included in refinement
Orthorhombic, P212121Dx = 2.006 Mg m3
Hall symbol: P 2ac 2abNeutron radiation, λ = ? Å
a = 6.8700 (3) ÅT = 298 K
b = 9.6396 (5) ÅParticle morphology: Block
c = 5.6058 (3) ÅColourless
V = 371.24 (3) Å3?, ? × ? × ? mm
Data collection top
HiPr/PEARL, ISIS
diffractometer
Scan method: Time-of-flight
Radiation source: Pulsed neutron sourceAbsorption correction: empirical (using intensity measurements)
Calculated using custom ISIS software.
Specimen mounting: TiZr CapsuleTmin = ?, Tmax = ?
Data collection mode: transmission
Refinement top
Least-squares matrix: selected elements onlyProfile function: Voigt function convoluted with a back-to-back double exponential.
Rp = 2.90739 parameters
Rwp = 4.1460 restraints
Rexp = ?Weighting scheme based on measured s.u.'s
χ2 = 0.142(Δ/σ)max < 0.001
? data pointsBackground function: Chebychev polynomial
Excluded region(s): excluded TOF regions below d = 0.6 Ang and above 4.3 AngPreferred orientation correction: Preferred orientation was treated with a spherical harmonic expansion to 4th order:
sh_c20 -0.00637`_0.01473 sh_c22p -0.03216`_0.01970 sh_c40 -0.14753`_0.02502 sh_c42p 0.05236`_0.03201 sh_c44p 0.07515`_0.02941
Crystal data top
C3D7NO3V = 371.24 (3) Å3
Mr = 112.11Z = ?
Orthorhombic, P212121Neutron radiation, λ = ? Å
a = 6.8700 (3) ÅT = 298 K
b = 9.6396 (5) Å?, ? × ? × ? mm
c = 5.6058 (3) Å
Data collection top
HiPr/PEARL, ISIS
diffractometer
Scan method: Time-of-flight
Specimen mounting: TiZr CapsuleAbsorption correction: empirical (using intensity measurements)
Calculated using custom ISIS software.
Data collection mode: transmissionTmin = ?, Tmax = ?
Refinement top
Rp = 2.907? data points
Rwp = 4.14639 parameters
Rexp = ?0 restraints
χ2 = 0.142
Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzBiso*/Beq
C10.8528470.7863040.4538460.89 (6)
C20.9037130.7508190.7113310.89 (6)
C30.8753360.5985240.7620630.89 (6)
N10.7777870.8305430.8776390.89 (6)
O10.7736720.8985880.4128180.89 (6)
O20.8961600.6972080.3018140.89 (6)
O30.6719410.5648130.7679780.89 (6)
D11.0539230.7754790.7517991.34 (9)
D20.9406940.5763770.9336931.34 (9)
D30.9440800.5360440.6253801.34 (9)
D40.6381420.8012340.8537801.34 (9)
D50.8194770.8105631.0467721.34 (9)
D60.7922280.9327030.8426581.34 (9)
D70.6426140.4738080.8018621.34 (9)
Geometric parameters (Å, º) top
O1—C11.23N1—D61.01
O2—C11.25C1—C21.53
O3—C31.43C2—C31.50
O3—D70.920 (18)C2—D11.01
N1—C21.49C3—D21.01
N1—D41.01C3—D31.01
N1—D51.01
C3—O3—D7115.7 (11)N1—C2—C3108
C2—N1—D4109C1—C2—C3112
C2—N1—D5109O3—C3—C2111
C2—N1—D6109N1—C2—D1108
D4—N1—D5110C1—C2—D1111
D4—N1—D6110C3—C2—D1107
D5—N1—D6110O3—C3—D2110
O1—C1—O2126O3—C3—D3108
O1—C1—C2118C2—C3—D2108
O2—C1—C2116C2—C3—D3111
N1—C2—C1110D2—C3—D3109
(serine73) L-serine-d7 top
Crystal data top
C3D7NO3Z = ?
Mr = 112.11Cell parameters included in refinement
Orthorhombic, P212121Dx = 2.089 Mg m3
Hall symbol: P 2ac 2abNeutron radiation, λ = ? Å
a = 6.6874 (3) ÅT = 298 K
b = 9.5815 (6) ÅParticle morphology: Block
c = 5.5619 (3) ÅColourless
V = 356.38 (4) Å3?, ? × ? × ? mm
Data collection top
HiPr/PEARL, ISIS
diffractometer
Scan method: Time-of-flight
Radiation source: Pulsed neutron sourceAbsorption correction: empirical (using intensity measurements)
Calculated using custom ISIS software.
Specimen mounting: TiZr CapsuleTmin = ?, Tmax = ?
Data collection mode: transmission
Refinement top
Least-squares matrix: selected elements onlyProfile function: Voigt function convoluted with a back-to-back double exponential.
Rp = 3.29939 parameters
Rwp = 4.7350 restraints
Rexp = ?Weighting scheme based on measured s.u.'s
χ2 = 0.149(Δ/σ)max < 0.001
? data pointsBackground function: Chebychev polynomial
Excluded region(s): excluded short TOF region abovePreferred orientation correction: Preferred orientation was treated with a spherical harmonic expansion to 4th order:
sh_c20 -0.02855`_0.01650 sh_c22p 0.06921`_0.02025 sh_c40 -0.21078`_0.02999 sh_c42p 0.15177`_0.03335 sh_c44p 0.14470`_0.03575
Crystal data top
C3D7NO3V = 356.38 (4) Å3
Mr = 112.11Z = ?
Orthorhombic, P212121Neutron radiation, λ = ? Å
a = 6.6874 (3) ÅT = 298 K
b = 9.5815 (6) Å?, ? × ? × ? mm
c = 5.5619 (3) Å
Data collection top
HiPr/PEARL, ISIS
diffractometer
Scan method: Time-of-flight
Specimen mounting: TiZr CapsuleAbsorption correction: empirical (using intensity measurements)
Calculated using custom ISIS software.
Data collection mode: transmissionTmin = ?, Tmax = ?
Refinement top
Rp = 3.299? data points
Rwp = 4.73539 parameters
Rexp = ?0 restraints
χ2 = 0.149
Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzBiso*/Beq
C10.8538000.7828030.4622970.90 (7)
C20.9059310.7494300.7227710.90 (7)
C30.8736870.5971170.7789980.90 (7)
N10.7786740.8327860.8881510.90 (7)
O10.7692370.8942820.4184120.90 (7)
O20.9017200.6929430.3109540.90 (7)
O30.6640920.5652090.7850250.90 (7)
D11.0608460.7730960.7622491.35 (10)
D20.9398080.5762280.9530661.35 (10)
D30.9435730.5320270.6437901.35 (10)
D40.6348220.8033510.8675731.35 (10)
D50.8230440.8153981.0589371.35 (10)
D60.7937610.9348910.8476231.35 (10)
D70.6299170.4676980.7993251.35 (10)
Geometric parameters (Å, º) top
O1—C11.23N1—D61.01
O2—C11.25C1—C21.53
O3—C31.43C2—C31.50
O3—D70.97 (2)C2—D11.01
N1—C21.49C3—D21.01
N1—D41.01C3—D31.01
N1—D51.01
C3—O3—D7116.1 (13)N1—C2—C3108
C2—N1—D4109C1—C2—C3112
C2—N1—D5109O3—C3—C2111
C2—N1—D6109N1—C2—D1108
D4—N1—D5110C1—C2—D1111
D4—N1—D6110C3—C2—D1107
D5—N1—D6110O3—C3—D2110
O1—C1—O2126O3—C3—D3108
O1—C1—C2118C2—C3—D2108
O2—C1—C2116C2—C3—D3111
N1—C2—C1110D2—C3—D3109
(serine81) L-serine-d7 top
Crystal data top
C3D7NO3Z = ?
Mr = 112.11Cell parameters included in refinement
Orthorhombic, P212121Dx = 2.125 Mg m3
Hall symbol: P 2ac 2abNeutron radiation, λ = ? Å
a = 6.5477 (4) ÅT = 298 K
b = 9.5405 (6) ÅParticle morphology: Block
c = 5.6092 (4) ÅColourless
V = 350.40 (4) Å3?, ? × ? × ? mm
Data collection top
HiPr/PEARL, ISIS
diffractometer
Scan method: Time-of-flight
Radiation source: Pulsed neutron sourceAbsorption correction: empirical (using intensity measurements)
Calculated using custom ISIS software.
Specimen mounting: TiZr CapsuleTmin = ?, Tmax = ?
Data collection mode: transmission
Refinement top
Least-squares matrix: selected elements onlyProfile function: Voigt function convoluted with a back-to-back double exponential.
Rp = 3.46339 parameters
Rwp = 4.9210 restraints
Rexp = ?Weighting scheme based on measured s.u.'s
χ2 = 0.145(Δ/σ)max < 0.001
? data pointsBackground function: Chebychev polynomial
Excluded region(s): excluded short TOF region abovePreferred orientation correction: Preferred orientation was treated with a spherical harmonic expansion to 4th order:
sh_c20 0.04468`_0.02140 sh_c22p 0.13833`_0.01829 sh_c40 0.02676`_0.02850 sh_c42p 0.12883`_0.02934 sh_c44p 0.18664`_0.02870
Crystal data top
C3D7NO3V = 350.40 (4) Å3
Mr = 112.11Z = ?
Orthorhombic, P212121Neutron radiation, λ = ? Å
a = 6.5477 (4) ÅT = 298 K
b = 9.5405 (6) Å?, ? × ? × ? mm
c = 5.6092 (4) Å
Data collection top
HiPr/PEARL, ISIS
diffractometer
Scan method: Time-of-flight
Specimen mounting: TiZr CapsuleAbsorption correction: empirical (using intensity measurements)
Calculated using custom ISIS software.
Data collection mode: transmissionTmin = ?, Tmax = ?
Refinement top
Rp = 3.463? data points
Rwp = 4.92139 parameters
Rexp = ?0 restraints
χ2 = 0.145
Fractional atomic coordinates and isotropic or equivalent isotropic displacement parameters (Å2) top
xyzBiso*/Beq
C10.7786600.7864080.4860570.69 (7)
C20.8370690.7497560.7415620.69 (7)
C30.7996680.5969720.7948610.69 (7)
N10.7145520.8343590.9119240.69 (7)
O10.6785280.8938870.4508550.69 (7)
O20.8379060.7031540.3295040.69 (7)
O30.5847530.5675630.7980300.69 (7)
D10.9970690.7703520.7757461.03 (10)
D20.8652120.5743640.9679091.03 (10)
D30.8704220.5313470.6607071.03 (10)
D40.5649600.8219530.8742291.03 (10)
D50.7435980.8010801.0792681.03 (10)
D60.7538450.9361520.8952561.03 (10)
D70.5370070.4922760.8768001.03 (10)

Experimental details

(serine45)(serine52)(serine73)(serine81)
Crystal data
Chemical formulaC3D7NO3C3D7NO3C3D7NO3C3D7NO3
Mr112.11112.11112.11112.11
Crystal system, space groupOrthorhombic, P212121Orthorhombic, P212121Orthorhombic, P212121Orthorhombic, P212121
Temperature (K)298298298298
a, b, c (Å)8.3096 (3), 8.5950 (3), 5.4647 (3)6.8700 (3), 9.6396 (5), 5.6058 (3)6.6874 (3), 9.5815 (6), 5.5619 (3)6.5477 (4), 9.5405 (6), 5.6092 (4)
V3)390.29 (3)371.24 (3)356.38 (4)350.40 (4)
Z????
Radiation typeNeutron, λ = ? ÅNeutron, λ = ? ÅNeutron, λ = ? ÅNeutron, λ = ? Å
µ (mm1)????
Specimen shape, size (mm)?, ? × ? × ??, ? × ? × ??, ? × ? × ??, ? × ? × ?
Data collection
DiffractometerHiPr/PEARL, ISIS
diffractometer
HiPr/PEARL, ISIS
diffractometer
HiPr/PEARL, ISIS
diffractometer
HiPr/PEARL, ISIS
diffractometer
Specimen mountingTiZr CapsuleTiZr CapsuleTiZr CapsuleTiZr Capsule
Data collection modeTransmissionTransmissionTransmissionTransmission
Scan methodTime-of-flightTime-of-flightTime-of-flightTime-of-flight
Absorption correctionEmpirical (using intensity measurements)
Calculated using custom ISIS software.
2θ values (°)2θmin = ? 2θmax = ? 2θstep = ?2θmin = ? 2θmax = ? 2θstep = ?2θmin = ? 2θmax = ? 2θstep = ?2θmin = ? 2θmax = ? 2θstep = ?
Refinement
R factors and goodness of fitRp = 2.985, Rwp = 4.140, Rexp = ?, χ2 = 0.148Rp = 2.907, Rwp = 4.146, Rexp = ?, χ2 = 0.142Rp = 3.299, Rwp = 4.735, Rexp = ?, χ2 = 0.149Rp = 3.463, Rwp = 4.921, Rexp = ?, χ2 = 0.145
No. of data points????
No. of parameters39393939

Selected geometric parameters (Å, º) for (serine45) top
O1—C11.23N1—D51.01
O2—C11.25N1—D61.01
O3—C31.43C2—D11.01
O3—D70.879 (16)C3—D21.01
N1—C21.49C3—D31.01
N1—D41.01
C3—O3—D7101.3 (15)N1—C2—C3108
C2—N1—D4109O3—C3—C2111
C2—N1—D5109N1—C2—D1108
C2—N1—D6109C1—C2—D1111
D4—N1—D5110C3—C2—D1107
D4—N1—D6110O3—C3—D2110
D5—N1—D6110O3—C3—D3108
O1—C1—O2126C2—C3—D2108
O1—C1—C2118C2—C3—D3111
O2—C1—C2116D2—C3—D3109
N1—C2—C1110
Selected geometric parameters (Å, º) for (serine52) top
O1—C11.23N1—D51.01
O2—C11.25N1—D61.01
O3—C31.43C2—D11.01
O3—D70.920 (18)C3—D21.01
N1—C21.49C3—D31.01
N1—D41.01
C3—O3—D7115.7 (11)N1—C2—C3108
C2—N1—D4109O3—C3—C2111
C2—N1—D5109N1—C2—D1108
C2—N1—D6109C1—C2—D1111
D4—N1—D5110C3—C2—D1107
D4—N1—D6110O3—C3—D2110
D5—N1—D6110O3—C3—D3108
O1—C1—O2126C2—C3—D2108
O1—C1—C2118C2—C3—D3111
O2—C1—C2116D2—C3—D3109
N1—C2—C1110
Selected geometric parameters (Å, º) for (serine73) top
O1—C11.23N1—D51.01
O2—C11.25N1—D61.01
O3—C31.43C2—D11.01
O3—D70.97 (2)C3—D21.01
N1—C21.49C3—D31.01
N1—D41.01
C3—O3—D7116.1 (13)N1—C2—C3108
C2—N1—D4109O3—C3—C2111
C2—N1—D5109N1—C2—D1108
C2—N1—D6109C1—C2—D1111
D4—N1—D5110C3—C2—D1107
D4—N1—D6110O3—C3—D2110
D5—N1—D6110O3—C3—D3108
O1—C1—O2126C2—C3—D2108
O1—C1—C2118C2—C3—D3111
O2—C1—C2116D2—C3—D3109
N1—C2—C1110
 

Footnotes

1Supplementary data for this paper are available from the IUCr electronic archives (Reference: BS5027 ). Services for accessing these data are described at the back of the journal.

Acknowledgements

We thank Dr Alan Coehlo for his advice on the use of TOPAS and Dr Josh McKinnon for his help with the program CrystalExplorer. We also thank EPSRC for funding and CCLRC for provision of neutron beam-time at ISIS.

References

First citationBirch, F. (1947). Phys. Rev. 71, 809–824.  CrossRef CAS Web of Science Google Scholar
First citationBoldyreva, E. V. (2004). NATO Science Series, II: Mathematics, Physics and Chemistry, edited by A. Katrusiak & P. F. McMillan, Vol. 140, pp. 495–512. Dordrecht: Kluwer Academic Publishers.  Google Scholar
First citationBoldyreva, E. V., Ahsbahs, H. & Weber, H.-P. (2003). Z. Kristallogr. 218, 231–236.  Web of Science CrossRef CAS Google Scholar
First citationBoldyreva, E. V., Kolesnik, E. N., Drebushchak, T. N., Ahsbahs, H., Beukes, J. A. & Weber, H.-P. (2005). Z. Kristallogr. 220, 58–65.  Web of Science CSD CrossRef CAS Google Scholar
First citationBruno, I. J., Cole, J. C., Edgington, P. R., Kessler, M., Macrae, C. F., McCabe, P., Pearson, J. & Taylor, R. (2002). Acta Cryst. B58, 389–397.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationCoelho, A. (2003). J. Appl. Cryst. 36, 86–95.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationCoelho, A. (2005). Topas-A: General Profile and Structure Analysis Software for Powder Diffraction Data; https://members.optusnet.com.au/~alancoelho/Google Scholar
First citationCrystal Impact (2004). DIAMOND, Version 3.0. Crystal Impact GbR, Postfach 1251, 53002 Bonn, Germany. https://www.crystalimpact.com/diamondGoogle Scholar
First citationDawson, A., Allan, D. R., Belmonte, S. A., Clark, S. J., David, W. I. F., McGregor, P. A., Parsons, S., Pulham, C. R. & Sawyer, L. (2005). Cryst. Growth Des. 5, 1415–1427.  Web of Science CSD CrossRef CAS Google Scholar
First citationFabbiani, F. P. A., Allan, D. R., Marshall, W. G., Parsons, S., Pulham, C. R. & Smith, R. I. (2005). J. Cryst. Growth, 275, 185–192.  Web of Science CSD CrossRef CAS Google Scholar
First citationFarrugia, L. J. (1999). J. Appl. Cryst. 32, 837–838.  CrossRef CAS IUCr Journals Google Scholar
First citationFortes, A. D. (2004). PhD Thesis: Computational and Experimental Studies of Solids in the Ammonia–Water System. Department of Earth Sciences, University College, University of London.  Google Scholar
First citationFourme, R., Kahn, R., Mezouar, M., Girard, E., Hoerentrup, C., Prange, T. & Ascone, I. (2001). J. Synchrotron Rad. 8, 1149–1156.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationGirard, E., Kahn, R., Mezouar, M., Dhaussy, A.-C., Lin, T., Johnson, J. E. & Fourme, R. (2005). Biophys. J. 88, 3562–3571.  Web of Science CrossRef PubMed CAS Google Scholar
First citationHemley, R. J. & Dera, P. (2000). Rev. Mineral. Geochem. 41, 335–419.  CrossRef Google Scholar
First citationJohnson, F. H., Eyring, H. & Stover, B. J. (1974). The Theory of Rate Processes in Biology and Medicine. New York: Wiley.  Google Scholar
First citationKatrusiak, A. (2004). NATO Science Series, II: Mathematics, Physics and Chemistry, edited by A. Katrusiak & P. F. McMillan, Vol. 140, pp. 513–520. Dordrecht: Kluwer Academic Publishers.  Google Scholar
First citationKistenmacher, T. J., Rand, G. A. & Marsh, R. E. (1974). Acta Cryst. B30, 2573–2578.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationKolesnik, E. N., Goryainov, S. V. & Boldyreva, E. V. (2005). Dokl. Phys. Chem. 404, 61–64.  Web of Science CrossRef Google Scholar
First citationKuznetsov, A. Z., Dmitriev, V., Dubrovinsky, L., Prakapenka, V. & Weber, H.-P. (2002). Solid State Commun. 122, 125–127.  Web of Science CrossRef CAS Google Scholar
First citationLeban, I. (1974). Cryst. Struct. Commun. 3, 237–239.  CAS Google Scholar
First citationMarshall, W. G. & Francis, D. J. (2002). J. Appl. Cryst. 35, 122–125.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationMcKinnon, J. J., Spackman, M. A. & Mitchell, A. S. (2004). Acta Cryst. B60, 627–668.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationMiller, R. A. & Schuele, D. E. (1969). J. Phys. Chem. Solids, 30, 589–600.  CrossRef CAS Web of Science Google Scholar
First citationMoggach, S. A., Allan, D. R., Morrison, C. A., Parsons, S. & Sawyer, L. (2005). Acta Cryst. B61, 58–68.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationMoggach, S. A., Allan, D. R., Parsons, S., Sawyer, L. & Warren, J. E. (2005). J. Synchrotron Rad. 12, 598–607.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationMoggach, S. A., Allan, D. R., Clark, S. J., Gutmann, M. J., Parsons, S., Pulham, C. R. & Sawyer, L. (2006). Acta Cryst. B62, 296–309.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationSomero, G. N. (1992). Ann. Rev. Physiol. 54, 557–577.  CrossRef CAS Google Scholar
First citationSpek, A. L. (2003). J. Appl. Cryst. 36, 7–13.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationSteiner, T. (2002). Angew. Chem. Int. Ed. 41, 48–76.  Web of Science CrossRef CAS Google Scholar
First citationTusvik, P. H., Mostad, A., Dalhus, B. & Rosenqvist, E. (1999). Acta Cryst. C55, 1113–1115.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationWaldorf, D. L. & Alers, G. A. (1962). J. Appl. Phys. 33, 3266–3269.  CrossRef CAS Web of Science Google Scholar
First citationWolff, S. K., Grimwood, D. J., McKinnon, J. J., Jayatilaka, D. & Spackman, M. A. (2005). CrystalExplorer1.5. University of Western Australia; https://www.Theochem.uwa.edu.au/crystal_explorer/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 logoSTRUCTURAL SCIENCE
CRYSTAL ENGINEERING
MATERIALS
ISSN: 2052-5206
Follow Acta Cryst. B
Sign up for e-alerts
Follow Acta Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds