Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/230884607

Single-crystal investigation of L-tryptophan with Z '=16

Article in Acta crystallographica. Section B, Structural science · October 2012


DOI: 10.1107/S0108768112033484 · Source: PubMed

CITATIONS READS

68 441

3 authors, including:

Carl Henrik Görbitz Graeme Day


University of Oslo University of Southampton
272 PUBLICATIONS 4,663 CITATIONS 255 PUBLICATIONS 12,506 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Carl Henrik Görbitz on 17 March 2014.

The user has requested enhancement of the downloaded file.


electronic reprint
Acta Crystallographica Section B
Structural
Science
ISSN 0108-7681
Editor: Sander van Smaalen

Single-crystal investigation of L-tryptophan with Z  = 16

Carl Henrik Görbitz, Karl Wilhelm Törnroos and Graeme M. Day

Acta Cryst. (2012). B68, 549–557

Copyright 
c International Union of Crystallography
Author(s) of this paper may load this reprint on their own web site or institutional repository provided that
this cover page is retained. Republication of this article or its storage in electronic databases other than as
specified above is not permitted without prior permission in writing from the IUCr.
For further information see http://journals.iucr.org/services/authorrights.html

Acta Crystallographica Section B: Structural Science publishes papers in structural chem-


istry and solid-state physics in which structure is the primary focus of the work reported.
The central themes are the acquisition of structural knowledge from novel experimental
observations or from existing data, the correlation of structural knowledge with physico-
chemical and other properties, and the application of this knowledge to solve problems
in the structural domain. The journal covers metals and alloys, inorganics and minerals,
metal-organics and purely organic compounds.

Crystallography Journals Online is available from journals.iucr.org


Acta Cryst. (2012). B68, 549–557 Carl Henrik Görbitz et al. · Single-crystal investigation
research papers
Acta Crystallographica Section B
Structural
Single-crystal investigation of L-tryptophan with Z0 =
Science 16
ISSN 0108-7681

Carl Henrik Görbitz,a* A complex, disorder-free structure in the space group P1 has Received 13 June 2012
Accepted 24 July 2012
Karl Wilhelm Törnroosb and been established for l-tryptophan, for which no crystal
Graeme M. Dayc structure has previously been available. The 16 molecules in
the asymmetric unit can be divided into two groups of eight;
a
one where the side chains have gauche orientations and one
Department of Chemistry, University of Oslo,
with trans orientations. Molecules within each group have
1033 Blindern, Oslo N-0315, Norway,
b
Department of Chemistry, University of Bergen,
almost identical molecular geometries. The unit-cell para-
Allégt. 41, Bergen N-5007, Norway, and meters mimic a hexagonal cell, but deviations from 90 for the
c
Department of Chemistry, University of cell angles  = 84.421 (4) and  = 87.694 (4) give a small tilt
Cambridge, Lensfield Road, Cambridge CB2 that rules out hexagonal symmetry. The hydrogen-bonding
1EW, England
pattern resembles that found in the crystal structure of the
racemic structure of dl-tryptophan, but a lower density
Correspondence e-mail: combined with longer hydrogen bonds and inter-aromatic
c.h.gorbitz@kjemi.uio.no interactions show that the enantiomeric structure is less
efficiently packed.

1. Introduction
Due to their obvious interest as the building units of proteins
and their role in many metabolic processes, amino acids were
among the first chiral compounds to be investigated with X-
ray diffraction methods (the structure of achiral glycine was
determined by Albrecht & Corey from an incomplete data set
in 1939). In 1950, when Shoemaker et al. published their work
on l-threonine, just 14 other detailed structures of organic
molecules crystallizing in Sohnke space groups were known
[according to the Cambridge Structural Database (CSD);
version 5.33 of November 2011; Allen, 2002], of which only
half dealt with chiral substances (i.e. not achiral or meso
forms). More amino acid structures appeared in the following
years (as hydrates for l-asparagine and l-arginine), the
sequence to some extent reflecting the ease with which high
quality crystals could be obtained. The initial rush thus ended,
in 1976, with the 17th amino acid, l-leucine, notorious for its
thin, flaky crystals. After a 14 year break (or 10 if the deter-
mination of the  form of l-glutamic acid is considered), a
crude structure (R factor 0.147) was revealed for phenylala-
nine (as the d enantiomer), for which crystals are plagued by
various, as yet undetermined, types of stacking disorders.
Recently, anhydrous structures were finally obtained for both
l-asparagine (Yamada et al., 2007) and l-arginine (Courvoisier
et al., 2012), meaning that prior to the present investigation
pure crystal structures were known for 18 out of the 20 stan-
dard amino acids, with l-tryptophan (l-Trp) and l-lysine still
missing (Table 1).
l-Trp has previously been investigated in complexes with
acetic acid (as a hydrate, GOMDAO: Li et al., 2009), formic
acid (MUGKAA: Hübschle et al., 2002; MUGKAA01: Scheins
# 2012 International Union of Crystallography et al., 2004), pyridine-2,4-dicarboxylic acid (as an ethanol
Printed in Singapore – all rights reserved solvate, NUQHIR: Di, 2010) and d-(R)-mandelate (as a 1.5

Acta Cryst. (2012). B68, 549–557 doi:10.1107/S0108768112033484 549


electronic reprint
research papers
hydrate, UGITAG: Fujii, 2009). Unravelling the solid-state Table 1
structure for pure l-Trp has, however, proved very difficult. First complete single-crystal structure determinations (including atom
coordinates) of the 20 standard amino acids according to the Cambridge
Khawas & Murti (1969) reported a unit cell with the para- Structural Database (CSD; Version 5.33 of November 2011; Allen, 2002).
meters a = 16.81, b = 17.90 and c = 6.90 Å in the orthorhombic
Amino acid† CSD refcode Reference
space group Pmmm (which in retrospect seems a bit strange
for a chiral compound). We have now obtained a single-crystal l-Alanine LALNIN Simpson & Marsh (1966)
structure of l-Trp (I), which is described here and compared l-Arginine‡ – Courvoisier et al. (2012)
l-Aparagine§ VIKKEG Yamada et al. (2007)
with the racemate dl-Trp (II) (space group P21/c with a = l-Aspartic acid} LASPRT Derissen et al. (1968)
18.90, b = 5.74, c = 9.31 Å,  = 101.8 , CSD refcode l-Cysteine (mon) LCYSTN Harding & Long (1968)
QQQBTP01: Bakke & Mostad, 1980; QQQBTP02: Hübschle l-Cysteine (ort) LCYSTN21 Kerr & Ashmore (1973)
l-Glutamine GLUTAM Cochran & Penfold (1952)
et al., 2004). l-Glutamic acid () LGLUAC03 Lehmann & Nunes (1980)
l-Glutamic acid () LGLUAC Hirokawa (1955)
Glycine ()†† GLYCIN02 Marsh (1958)
Glycine () GLYCIN Iitaka (1960)
Glycine () GLYCIN01 Iitaka (1961)
l-Histidine (ort) LHISTD10 Madden, McGandy &
Seeman (1972)
l-Histidine (mon) LHISTD01 Madden, McGandy,
Seeman, Harding &
Hoy (1972)
l-Isoleucine LISLEU Torii & Iitaka (1971)
l-Leucine LEUCIN Harding & Howieson
(1976)
l-Lysine – –
l-Methionine LMETON10 Torii & Iitaka (1973)
l-Trp is an essential amino acid for humans, meaning that it d-Phenylalanine SIMPEJ Weissbuch et al. (1990)
must be part of our diet. It is the least common of the standard l-Proline‡‡ PROLIN Kayushina & Vainshtein
(1965)
amino acids in proteins with an average occurrence of about l-Serine§§ LSERIN10 Benedetti et al. (1973)
1.3%, but Trp residues play important roles in protein stability l-Threonine LTHREO Shoemaker et al. (1950)
and recognition. Additionally, Trp is a biochemical precursor l-Tryptophan - This work
l-Tyrosine LTYROS10 Mostad et al. (1972)
for serotonin (a neurotransmitter), melatonin (a neuro- l-Valine LVALIN Torii & Iitaka (1970)
hormone) and niacin (vitamin B3 or nicotinic acid).
† Polymorphism is indicated by either (mon) and (ort) for monoclinic and orthorhombic
forms, respectively, or by greek letters. ‡ From powder data, not yet in the CSD,
dihydrate (ARGIND): Karle & Karle (1964). § Monohydrate (ASPARM): Kartha &
de Vries (1961). } Monohydrate (IJEQET): Umadevi et al. (2003). †† Solved and
2. Experimental refined from an incomplete data set by Albrecht & Corey (1939). ‡‡ Monohydrate
(RUWGEV): Janczak & Luger (1997). §§ Monohydrate (LSERMH): Frey et al.
2.1. Crystal preparation (1973).
From a saturated solution of (I) in water (approximately
10 mg ml1), 30 ml was deposited into a series of 30  6 mm
test tubes that were subsequently sealed with Parafilm1. A
needle was then used to prick a single small hole in the N—H = 0.88 (aromatic) or 0.91 Å (amino) and C—H = 0.95
Parafilm1 of each tube, after which it was placed inside a (aromatic), 0.99 (methylene) or 1.00 Å (methine), while
larger test tube filled with 1 ml of acetonitrile. The systems permitting free rotation for the amino groups. Uiso values were
were ultimately capped and left for 3 d at 293 K. l-Trp regu- set to 1.2Ueq of the carrier atom, or 1.5Ueq for amino groups.
larly precipitated as thin flakes, but one tube produced some In the absence of significant anomalous scattering effects,
crystals of a different habit, as rhombohedral prisms, and the 23 163 Friedel pairs were merged. Experimental details are
largest specimen was used for data collection. given in Table 2.
Alternative refinements utilized a significantly reduced
2.2. Data collection and structure refinement number of refinement parameters through the employment of
Data collection was carried out on a Bruker AXS APEX II SHELX EADP commands, forcing the use of the same
ULTRA rotating anode Pt135 CCD diffractometer, using anisotropic displacement parameters for atoms related by
graphite-monochromated Mo K radiation ( = 0.71073 Å) pseudotranslational symmetry. One such model (CIF available
for 0.3 scans over 182 in ! in four orthogonal ’ positions. as supplementary material1) yielded an R value of 0.088 for
Structure solution was carried out by direct methods 1338 parameters. The estimated standard uncertainties for
combined with iterative difference-Fourier synthesis (Shel- calculated geometrical parameters were unchanged compared
drick, 2008). with the unconstrained refinement, or even increased
The structure, with 16 molecules labelled A–P in the slightly.
asymmetric unit, was refined without constraints or restraints 1
Supplementary data for this paper are available from the IUCr electronic
on C, N or O positions. No positional disorder was indicated. archives (Reference: EB5018). Services for accessing these data are described
H atoms were introduced in theoretical positions with fixed at the back of the journal.

550 Carl Henrik Görbitz et al.  Single-crystal investigation Acta Cryst. (2012). B68, 549–557
electronic reprint
research papers
Table 2 Table 3
Experimental details. Torsion angles ( ) for l-Trp in crystal structures.
Crystal data Refcode† 1 ‡ 2‡ Ref
Chemical formula C11H12N2O2
Mr 204.23 (I)§ 173.4, 172.9 (4), 113.8, 112.2 (6), –
Crystal system, space group Triclinic, P1 174.1 (4) 115.2 (6)
Temperature (K) 123 (I) 79.0, 77.8 (5), 80.0 (5) 111.6, 109.6 (6), 113.6 (6) –
a, b, c (Å) 11.430 (3), 11.464 (4), 35.606 (9) (II) 166.8 107.9 (a)
, ,  ( ) 84.421 (4), 87.694 (4), 60.102 (2) GOMDAO 62.1 103.3 (b)
V (Å3) 4025.6 (19) MUGKAA01 53.5 104.1 (c)
Z, Z0 16 NUQHIR 58.5 114.4 (d)
Radiation type Mo K UGITAG 164.4 102.3 (e)
 (mm1) 0.095
Crystal size 0.62  0.28  0.14 References: (a) Hübschle et al. (2004); (b) Li et al. (2009); (c) Scheins et al. (2004); (d) Di
(2010); (e) Fujii (2009). † CSD (Allen, 2002). ‡ For (I) 1 and 2 are N1—C2—C3—
Data collection C5 and C2—C3—C5—C4, respectively (see Fig. 2). § First line is for the T family,
second line is for the G family. For each torsion angle the average for eight molecules is
Diffractometer Bruker APEXII ULTRA CCD
given, followed by the extreme values (in italic type face with calculated s.u.s).
Absorption correction Multi-scan
Tmin, Tmax 0.943, 0.987
No. of measured, independent and 66 471, 24 736, 19 659
observed [I > 2(I)] reflections
Rint 0.043 deviation for the best fit between O, N and C atoms in pairs of
(sin /)max (Å1) 0.721 molecules with the same conformation is typically around
0.03 Å.
Refinement
R[F2 > 2(F2)], wR(F2), S 0.085, 0.255, 1.13 Fig. 1 shows that the structure of (I) has a layered
No. of reflections 24 736 construction. Hydrogen bonding occurs in two separate
No. of parameters 2178 hydrophilic layers, including the polar heads of molecules A–
No. of restraints 3
H-atom treatment H-atom parameters constrained G and I–P, respectively. Between the hydrophilic layers are
 max,  min (e Å3) 0.53, 0.47 hydrophobic bilayers of the side chains, one composed of the
side chains from molecules E–L and one from the side chains
Computer programs used: APEX2 (Bruker, 2007a), SAINT-Plus (Bruker, 2007b),
SHELXTL (Sheldrick, 2008), SADABS (Sheldrick, 1996). of molecules A–D and M–P.

3. Results and discussion


3.1. Molecular structure, asymmetric unit and unit cell
The asymmetric unit and the unit cell are shown in Fig. 1.
The molecular conformation of l-Trp is essentially described
by the two torsion angles N1—C2—C3—C5 (1), which may
be gauche+, trans or gauche, and C2—C3—C5—C4 (2),
which may be positive or negative, giving rise to a total of six
basic conformations (combinations of 1 and 2). The torsion
angles listed in Table 3 show that the 16 l-Trp molecules in the
structure of (I) populate only two of these conformations;
molecules A, B, E, F, I, J, M and N have 1 = trans and 2 ’
114 (called the T family), while molecules C, D, G, H, K, L,
O and P have 1 = gauche and 2 ’ 112 (G family). The
neighbouring molecules B and C, shown in Fig. 2, are repre-
sentative examples of both conformations.
Each conformation found in the structure of (I) has been
observed for l-Trp in two previous structures, while the formic
acid solvate MUGKAA01 (Scheins et al., 2004) with 1 =
Figure 1
gauche+ and negative 2 provides the only example of a third The asymmetric unit, chosen so that all 16 molecules are within the unit
l-Trp conformation. cell, viewed approximately along the ab diagonal. H atoms bonded to C
The very narrow ranges for torsion angles given in Table 3 have been omitted for clarity. Molecules of the T family (trans side-chain
and a molecular overlay diagram (supplementary material) orientation, see text) are shown with dark grey C atoms; those of the G
family have light grey C atoms. Hydrogen bonds appear as dotted lines.
illustrate that molecular geometries within each of the The side chain of molecule G has been depicted in wireframe
conformational families are remarkably similar. Side-chain representation to relieve overlap with molecule F; an extra copy of
geometries and orientations are hardly distinguishable, while a molecule G with all atoms in white, called G0, at (x; 1 þ y; z) is included.
limited rotation is seen for the amino group, which is deter- The blue, dashed line represents the interface at the centre of the
hydrophobic bilayer, while the open arrow shows a pseudo twofold screw
mined by a single refinement parameter (no s.u.’s available axis. Various terms used in describing the structure are given at the
owing to constrained refinement). The calculated r.m.s. bottom.

Acta Cryst. (2012). B68, 549–557 Carl Henrik Görbitz et al.  Single-crystal investigation 551
electronic reprint
research papers
Table 4 (Görbitz et al., 2009). Three types of hydrogen
Hydrogen-bond distances (Å) and angles ( ) in the structures of (I) and (II) (Hübschle bonds are present in Fig. 3(a) (as well as in the
et al., 2004).
qualitatively identical pattern in Fig. 3b), forming
Type† H  O N  O N—H  O what can be described as a series of tapes with a
l-Trp (I) repeating unit of eight molecules, e.g.
N1—H1  O1 1 1.94, 1.90, 1.98‡ 2.804, 2.784 (6), 2.833 (5) 159, 152, 165 E:A:H:D:F:B:G:C, utilizing hydrogen-bond type 1
N1—H3  O1 2 2.00, 1.97, 2.04 2.902, 2.870 (6), 2.928 (6) 171, 161, 179 with C1—C2—N1—H1 = gauche and type 2
N1—H2  O2 3 1.82, 1.79, 1.86 2.711, 2.692 (6), 2.735 (6) 167, 161, 177
with C2—C1—N1—H3 = gauche+. Notably,
dl-Trp (II)§ molecules with similar conformations occur in
N15—H15A  O13 1 1.91} 2.814 172} pairs along the tapes, which are subsequently
N15—H15B  O13 2 1.94 2.826 165
N15—H15C  O14 3 1.80 2.707 175 interconnected by type 3 interactions with C1—
C2—N1—H2 = trans into perpendicular head-to-
† As identified in Figs. 3 and 4. ‡ For each parameter the average for 16 interactions is given, followed tail chains after an off-set of four amino acids in
by the extreme values (in italic type face with calculated s.u.s). § Atomic numbers retained from the
original contribution. } Values listed are after normalization of the N—H bond lengths to 0.91 Å as used the tape direction. This generates E:F:E:F and
in the refinement of (I). other equivalent sequences in Fig. 3(a), with

3.2. Hydrogen bonding


Fig. 3 shows the hydrogen-bonding pattern in the crystal
structure of (I). Amino acids with bulky side chains are often
unable to form good hydrogen-bonding patterns on their own,
presumably as this would introduce unacceptable steric
conflict, and rather depend on incorporating water molecules
as integral parts of the network. Examples include non-
aromatic compounds like tert-butylglycine monohydrate and
neopentylglycine monohydrate (Weissbuch et al., 1990)
[neopentyl is —CH2—C—(CH3)3] as well as amino acids with
aromatic side chains like 3-fluoro- and 2,4,5-trifluor-
ophenylalanine monohydrates (In et al., 2003), 4-nitropheny-
lalanine monohydrate (Dai & Fu, 2008) and -
methyltryptophan monohydrate (Cyr et al., 1999). l-Trp can
evidently do without water, but forms a pattern that is distinct
from those found for other layered amino acid structures

Figure 3
Hydrogen bonding within the two distinct hydrophilic layers in the crystal
structure of (I) with reference to Fig. 1. Side chains are not shown beyond
C (C3 in Fig. 2). The colour coding for C atoms is similar to Fig. 1, while
Figure 2 two different colours have been used in an arbitrary fashion for N and O
Molecule B with a trans orientation for 1 (N1—C2—C3—C5) and atoms to give each molecule in each layer a unique appearance. Views are
molecule C with a gauche orientation. Displacement ellipsoids have been perpendicular to the ab plane, i.e. along c*. The three numbers in (a)
drawn at the 50% probability level; H atoms are shown as spheres of identify three types of hydrogen bonds, while the rectangles highlight a
arbitrary size. hydrogen-bonded tape (red) and a head-to-tail chain (blue).

552 Carl Henrik Görbitz et al.  Single-crystal investigation Acta Cryst. (2012). B68, 549–557
electronic reprint
research papers
adjacent head-to-tail chains having opposite directions. The structure of the achiral compound 8-amino-1,4-diox-
Interestingly the hydrogen-bonding pattern of the racemate aspiro[4.5]decane-8-carboxylic acid (JAPJUF: Vela et al.,
(II), depicted in Fig. 4(a), resembles the pattern of (I) despite 1989), in space group Pna21, is also related to (I) with the
the fact that the characteristic tapes are missing. The dimer in hydrogen-bonding pattern shown in Fig. 4(b). The only
Fig. 4(a) is centrosymmetric and thus could not occur for l- essential modification compared with (I) is that the carboxy-
Trp, but head-to-tail chains are nevertheless very similar in late O atom involved in the head-to-tail chains also accepts a
both structures, the main difference being that the orientation second H atom, while for (I) this O atom is not involved in the
of chains occurs in pairs for (II), i.e. ""##"" rather than "#" tape motif.
for (I). The hydrogen bond lengths of (I) and (II) are
compared in Table 4 [a complete description of hydrogen-
bond geometry for (I) is available as supplementary material].
The hydrogen bond of type 2, with H—C—N—H = trans, is
clearly shorter for (II) than for (I). For the other two inter-
actions there are no significant differences. The type 1 inter-
action is more linear for (II) than for (I), but not shorter.

3.3. Aromatic interactions and pseudosymmetry structure,


asymmetric unit and unit cell
If all l-Trp molecules had the same side-chain orientation,
the screw axis along the ab diagonal involving molecules I–P
readily visible in Fig. 1, would have been proper rather than
pseudo (a similar pseudo-twofold screw along the b axis
operates on the polar heads of the molecules A–H). Evidently,
such a simple arrangement is not compatible with an efficient
stacking of the aromatic side chains, as the approximate top
view in Fig. 5 shows that side chains instead are related by
pseudo-glide planes. The presence of such pseudo-glide planes
is rendered possible by the occurrence of two different side
chain conformations in the structure of (I), and results in a

Figure 4
Hydrogen-bonding patterns in the crystal structures of (a) dl-Trp (II)
(Hübschle et al., 2004) and (b) an aminocyclohexanecarboxylic acid (Vela
et al., 1989), both curtailed beyond C. Labels and coloured rectangles
appear as in Fig. 3: (a) with an additional hydrogen-bonded dimer
highlighted in (a).

Figure 6
(a) Representative section of a hydrophobic region in the structure of (I)
showing interactions between the indole groups of the Trp side chains of
molecules A, B, C and D. C atoms are coloured according to side-chain
Figure 5 conformation as in Fig. 1, while two different colours have been used for
A pseudo twofold screw axis for molecules I–P operating on the polar N atoms to distinguish individual molecules. (b) Interactions between
heads of the amino acid molecules (inside circle). The hydrophobic side side chains in the structure of (II) (Hübschle et al., 2004) with l or d giving
chains are instead related by horizontal pseudo-glide planes (dashed amino acid chirality. The H  C( ) distances (in Å) were measured after
lines). In this illustration molecules belonging to different conformational normalization of the N—H bond to 0.88 Å and are directly comparable
families are not only distinguished by the colour of the C atoms, but with the corresponding distances in (a). The dashed horizontal lines show
members of the T family also have O and N atoms in darker colour. a true crystallographic glide plane in (b) and a pseudo glide plane in (a).

Acta Cryst. (2012). B68, 549–557 Carl Henrik Görbitz et al.  Single-crystal investigation 553
electronic reprint
research papers
stacking arrangement of aromatic groups that is amazingly understanding of the crystal build-up of (I) to verify that the
similar to that found for the racemate (II), Fig. 6. P1 space group with Z0 = 16 is not an artefact.
The A and B molecules in Fig. 6(a) belong to the T family The stack of molecules I–P in Figs. 1 and 6 is essentially
and have the same function as the conformationally identical obtained from two similar repeating units M:I:O:K and
l-enantiomers in Fig. 6(b), while the C and D molecules N:J:P:L with conformations T:T:G:G. Thus, the operational
belong to the G family and play the role of the D enantiomers structural unit of (I) consists of four independent molecules.
in (II). The representative lengths of the two shortest N— Potentially, such units could be used to build triclinic, mono-
H  C( ) interactions indicate, however, that hydrophobic clinic or even orthorhombic crystal structures with Z0 = 4, all
groups are slightly less efficiently packed in the crystal struc- retaining the hydrogen-bonding pattern and aromatic inter-
ture of the pure enantiomer. Together with the longer type 2 actions (within hydrophobic monolayers) of (I). These hypo-
hydrogen bond in (I), this contributes to a slightly lower thetical packing arrangements furthermore have in common
density for (I) than for (II), calculated values being 1.348 (T = that they would include just a single type of hydrophobic
123 K) and 1.371 g cm3 (T = 173 K), respectively. bilayer or interface, which is the default for enantiomeric and
racemic layered amino acid structures. This includes the
structure of (II), compared in Fig. 7 with the structure of
(I).
3.4. Unit cell revisited: crystal construction principles In Fig. 8 the focus is on the dashed hydrophobic interfaces
With only eight entries in the CSD (Allen, 2002), crystal in Fig. 7. The surfaces of (II) shown in Figs. 8(a) and (b) are
structures of organic molecules with Z0  16 are exceedingly related by inversion symmetry, which reverses the indicated
rare. Accordingly, whenever a crystal structure with a Z0 value directions of the C—C bond vectors (for enantiomeric
of this magnitude is presented, suspicion will immediately be structures a reversal of monolayer directions within a double
raised concerning potential failure to include missing layer is usually accomplished by a twofold screw operation).
symmetry. It is therefore imperative to obtain a fundamental Putting the two surfaces together in (c) makes the ‘knobs-in-
holes’ fit easily visible. In contrast, the surface generated by
molecules E, F, G and H in Fig. 8(e) is qualitatively obtained
from the IJKL surface in Fig. 8(d) by first a 180 rotation
around the horizontal x axis (C—C bond vector) and then
an in-plane anti-clockwise 60 rotation around the z axis. If the
resulting interface shown in detail in Fig. 8(f) and labelled 1 in
Fig. 7(b) had been the only way to fit the two surfaces, the
structure of (I) would have belonged to the hexagonal space
group P65 with Z0 = 8 (two units of four molecules) and
approximate cell parameters a0 = b0 = 11.44 Å [= (a + b)/2] and
c0 = 106.22 Å (= csin sin 3). There appears to be no
structures in the CSD with hexagonal symmetry and such a
60 twist between distinct layers, but the dimeric amino acid
l-cystine (Dahaoui et al., 1999), in space group P6122 with cell
parameters a = b = 5.412 (1) and c = 55.956 (1) Å, gives an
impression of such an arrangement, a conceptual difference to
(I) being that hydrophobic bilayers are covalently linked by
S—S bonds.
What makes the structure of (I) particularly intriguing,
however, is that there is not just a single way to fit two
hydrophobic surfaces, but also a second alternative that is
operational at interface 2 in Fig. 7(b). An understanding of the
subtle difference between 1 and 2 can be gained by comparing
Figs. 8(d)–(f) with Figs. 8(g)–(i). The IJKL surface in Fig. 8(d)
Figure 7
(a) Hydrophobic bilayers in the crystal structure of (II), viewed along the and the ABCD surface in Fig. 8(g) are indistinguishable. The
b axis. For each bilayer C atoms in the top and bottom monolayer are same is also true for the EFGH and MNOP surfaces in Figs.
coloured in light grey and dark grey, respectively. The pink, dashed lines, 9(e) and (h), but while the final in-plane rotation in Fig. 8(e) to
separating structurally identical surfaces, show the central interface in give the fit shown in Fig. 8(f) is anti-clockwise, it is clockwise in
each unit cell. Hydrophilic layers appear as shaded rectangles (amino and
carboxylate groups are not shown), while C atoms are small spheres and Fig. 8(h), resulting in a new fit shown in Fig. 8(i). In this way all
C—C bonds are yellow. The arrow shows the viewing direction in Fig. 8. 16 independent molecules (four units of four molecules)
(b) Similar for (I) with a view along the a axis. (Note: in Figs. 7 and 8 side- obtain a unique set of neighbours with respect to inter-
chain colour does not reflect conformation.) For each monolayer the
molecular interactions. Note for instance in Fig. 8 how mole-
molecules contributing a side chain are identified, e.g. A, B, C, D for the
layer at the top. Two distinct interfaces, labelled 1 and 2, separate cules A, B, I and J, which all belong to the T family, have
structurally different surfaces. dissimilar contacts with overlapping molecules O (G), N (T),

554 Carl Henrik Görbitz et al.  Single-crystal investigation Acta Cryst. (2012). B68, 549–557
electronic reprint
research papers
E (T) and H (G), respectively. The  60 rotation of layers manner (neither twinning nor disorder was indicated in the
leaves the triclinic unit cell of (I) some hexagonal-like traits, refinement).
like a and b axes of about the same length and a  angle close
to 60 (Table 2), but as the alternating +60 and 60 rotations 3.5. Systematic absences in the diffraction pattern
along the c axis operate with centres of rotation that are
slightly translocated in the other two directions,  and  are The presence of pseudo-screw axes and glide planes
shifted significantly away from 90 (Table 2). described above leads to certain groups of reflections being
Two other structures have crystallized in the space group P1 systematically weak or absent, but much more important for
with Z0 = 16 (2,2-aziridine-dicarboxamide, BIPCOS01: the observed diffraction pattern is translational pseudosym-
Brückner, 1982; cholesterol at 310 K, CHOEST21, Hsu et al., metry. It is useful in this connection to regard first each of the
2002), but none are divided into hydrophobic and hydrophilic two blocks in Fig. 1 independently, and then turn to the
layers with this kind of relationship between groups of complete structure.
molecules. From Fig. 3(a) it is clear that the A–H block has close to
The mystery of (I) is of course why and how hardly perfect translational symmetry along the vertical b axis,
distinguishable surfaces should interact in two different ways effectively cutting the repeat unit into b/2. Reflections with k
in a crystal structure, and indeed doing so in a fully ordered odd accordingly do not have a contribution from this block,
Fig. 9(a). In a similar manner, the
pattern shown in Fig. 4(b) is face-
centred, meaning that it gives no
contributions to reflections with h +
k odd, Fig. 9(b). When combining
information from Figs. 9(a) and (b)
in Fig. 9(c), we find that reflections
that do not fulfil any of the condi-
tions hkl: k = 2n or hkl: h + k = 2n,
e.g. (0,1,0) or (2,1,1), do not have
significant contributions from any of
the two blocks and thus are missing.
This explains the unusual
‘systematic absences’ for the
triclinic space group P1 observed in
the diffraction pattern of (I), Fig.
9(d).

3.6. Search for polymorphs


To establish whether the single
crystal used for data collection was
the result of a freak occurrence of
another polymorph than usually
obtained for l-Trp, additional crys-
tallizations were performed from
various solvents, uniformly giving
the familiar, very thin flakes. Cell
parameters determined form two
such specimens matched those
listed in Table 2.
Furthermore, X-ray powder
diffraction data were collected for a
typical sample, and the observed
pattern fitted the simulated pattern
of the single-crystal structure [illus-
tration available as supplementary
Figure 8 material, room temperature cell
(a) and (b) Monolayers in the structure of (II). The arrows indicate the directions of the C—C bond
parameters are a = 11.5223 (13), b =
vectors; l and d give amino acid chirality. (c) Fit between the two monolayers to give a full bilayer, atoms
in the bottom monolayer (a) are displayed here as a space-fill representation. (d)–(f) Similar drawings 11.5156 (12), c = 35.877 (6) Å,  =
for the two independent bilayers in the structure of (I). See text for details. 84.393 (9),  = 87.681 (7),  =

Acta Cryst. (2012). B68, 549–557 Carl Henrik Görbitz et al.  Single-crystal investigation 555
electronic reprint
research papers
59.987 (8) , V = 4102.1 (10) Å3]. Although other polymorphs We would then be looking for some unified pseudo-translation
certainly may exist, we thus see no indication of such in our between A, B, E, F, I, J, M, N (T family) and similarly for C, D,
experiments. G, H, K, L, O, P (G family). Both sets have similar relation-
ships to the leading molecule. In the molecular refinement the
3.7. A modulated structure? A molecule was used as a motive with which the remaining
family members were related by rotations and translations
Structures of organic compounds with high Z0 values may
A 0; 0; 0; xA; yA; zA
sometimes be described (and potentially better understood) as
B 0; 0; 0; xA; yA þ 1=2; zA
commensurately (or incommensurately) modulated structures
E 60; 0; 180; xE; yE; zE
while applying the higher-dimensional superspace approach
F 60; 0; 180; xE; yE þ 1=2; zE
(Schönleber, 2011). Examples include 4,40 -dimethyl-2-
I 60; 0; 0; xI; yI; zI
hydroxy-6-oxocyclohexene-1-carboxylic acid with Z0 = 2.5
J 60; 0; 0; xI þ 1=2; yI þ 1=2; zI
(five half-molecules; Duncan et al., 2002) as well as various
M 0; 0; 180; xM; yM; zM
crown complexes such as [Cu(H2O)2(15-crown-5)](NO3)2 with
N 0; 0; 180; xM þ 1=2; yM þ 1=2; zM
Z0 = 10 (Schönleber & Chapuis, 2004; Hao et al., 2005). It was
thus appropriate to check whether this could also be the case So for A + B and E + F we have the pseudo-translation
for the structure of (I). (0,1/2,0), which really would lead to supercell effects and the
Several tests were made using JANA2006 (Petrı́cek et al., possibility of a description as being modulated, but for I + J
2006), with the conclusion that there is in fact no reasonable and M + N the pseudo translation is different (1/2,1/2,0).
way to apply the superspace theory and handle the structure Accordingly, there is no chance to use the formalism for
as modulated. There are 16 independent molecules in the modulated structures. A similar scheme is valid for molecules
space group P1, in two different conformations (see above). of the G family.

4. Conclusion
The elusive solid-state structure of
l-tryptophan (l-Trp) has been
determined from a single-crystal X-
ray analysis based on diffraction
data collected from a specimen of
unusually high quality. The packing
arrangement, incorporating a
hydrogen-bonding pattern not
previously observed for amino
acids, mimics the structure of dl-
Trp, where fully ordered l-Trp
molecules belonging to two
different conformations called T
and G play the roles of the l- and
d-enantiomers in the crystal struc-
ture of the racemate. This gives rise
to extended pseudo-symmetry and
together with the occurrence of two
independent types of hydrophobic
interfaces in the crystal bring the
number of molecules in the asym-
metric unit up to 16, a very rare
phenomenon indeed. There is no
indication that this is in fact a
modulated structure. From a theo-
retical point of view, the packing is
unique in that it, for the triclinic
space group P1, gives apparent
Figure 9 systematic absences based on the
(a)–(c) Simulated diffraction patterns of (I) for a hk layer. Atoms in the A–H block give contributions
condition hkl: h + k = 2n or k = 2n.
only to the reflections coloured in blue in (a), while the I–P block (b) give rise to the red flections in (b).
In the combined pattern in (c) reflections with contributions from both blocks are coloured in black. (d) To investigate the energy gained
A projection of the experimentally observed hk0 layer. by l-Trp in increasing Z0 from 1 or

556 Carl Henrik Görbitz et al.  Single-crystal investigation Acta Cryst. (2012). B68, 549–557
electronic reprint
research papers
2, as seen for the other enantiomerically pure amino acids, to 8 Hao, X., Siegler, M. A., Parkin, S. & Brock, C. P. (2005). Cryst.
or 16, and to study the alternative low Z0 structures that are Growth Des. 5, 2225–2232.
Harding, M. M. & Howieson, R. M. (1976). Acta Cryst. B32, 633–634.
available to the molecule, we are in the process of performing
Harding, M. M. & Long, H. A. (1968). Acta Cryst. B24, 1096–1102.
extended structure prediction calculations and energy calcu- Hirokawa, S. (1955). Acta Cryst. 8, 637–641.
lations on the observed structure described here. Crystal Hsu, L.-Y., Kampf, J. W. & Nordman, C. E. (2002). Acta Cryst. B58,
structures with multiple, flexible molecules in the asymmetric 260–264.
unit pose a particular challenge for theoretical structure Hübschle, C. B., Dittrich, B. & Luger, P. (2002). Acta Cryst. C58,
o540–o542.
prediction methods, usually based on global lattice energy
Hübschle, C. B., Messerschmidt, M. & Luger, P. (2004). Cryst. Res.
minimization (Day, 2011), due to the high dimensionality of Technol. 39, 274–278.
the energy surface that must be explored (van Eijck & Kroon, Iitaka, Y. (1960). Acta Cryst. 13, 35–45.
2000). These studies will be discussed in a future paper, now in Iitaka, Y. (1961). Acta Cryst. 14, 1–10.
preparation. In, Y., Kishima, S., Minoura, K., Nose, T., Shimohigashi, Y. & Ishida,
T. (2003). Chem. Pharm. Bull. 51, 1258–1263.
Janczak, J. & Luger, P. (1997). Acta Cryst. C53, 1954–1956.
The authors thank Dr David Wragg, Department of Karle, I. L. & Karle, J. (1964). Acta Cryst. 17, 835–841.
Chemistry, University of Oslo, Norway, for collecting and Kartha, G. & de Vries, A. (1961). Nature, 192, 862–863.
Kayushina, R. L. & Vainshtein, B. K. (1965). Kristallografiya, 10, 833–
refining the powder XRD data, and Dr Václav Petřı́ček, 844.
Institute of Physics, Academy of Sciences of the Czech Kerr, K. A. & Ashmore, J. P. (1973). Acta Cryst. B29, 2124–2127.
Republic, Czech Republic, for testing the experimental data Khawas, B. & Krishna Murti, G. S. R. (1969). Acta Cryst. B25, 1006–
set with respect to potential modulation. 1009.
Lehmann, M. S. & Nunes, A. C. (1980). Acta Cryst. B36, 1621–
1625.
References Li, J., Liang, Z.-P. & Tai, X. S. (2009). Personal communication.
Albrecht, G. & Corey, R. B. (1939). J. Am. Chem. Soc. 61, 1087–1103. Madden, J. J., McGandy, E. L. & Seeman, N. C. (1972). Acta Cryst.
Allen, F. H. (2002). Acta Cryst. B58, 380–388. B28, 2377–2382.
Bakke, Ø. & Mostad, A. (1980). Acta Chem. Scand. B, 34, 559–570. Madden, J. J., McGandy, E. L., Seeman, N. C., Harding, M. M. & Hoy,
Benedetti, E., Pedone, C. & Sirigu, A. (1973). Gazz. Chim. Ital. 103, A. (1972). Acta Cryst. B28, 2382–2389.
555–561. Marsh, R. E. (1958). Acta Cryst. 11, 654–663.
Brückner, S. (1982). Acta Cryst. B38, 2405–2408. Mostad, A., Nissen, H. M. & Romming, C. (1972). Acta Chem. Scand.
Bruker (2007a). APEX2. Bruker AXS, Inc., Madison, Wisconsin, 26, 3819–3833.
USA. Petrı́cek, V., Dusek, M. & Palatinus, L. (2006). JANA2006. Institute
Bruker (2007b). SAINT Bruker AXS, Inc., Madison, Wisconsin, of Physics, Praha, Czech Republic.
USA. Scheins, S., Dittrich, B., Messerschmidt, M., Paulmann, C. & Luger, P.
Cochran, W. & Penfold, B. R. (1952). Acta Cryst. 5, 644–653. (2004). Acta Cryst. B60, 184–190.
Courvoisier, E., Williams, P. A., Lim, G. K., Hughes, C. E. & Harris, Schönleber, A. (2011). Z. Kristallogr. 226, 499–517.
K. D. M. (2012). Chem. Commun. 48, 2761–2763. Schönleber, A. & Chapuis, G. (2004). Ferroelectrics, 305, 99–102.
Cyr, L. V., Newton, M. G. & Phillips, R. S. (1999). Bioorg. Med. Chem. Sheldrick, G. M. (1996). SADABS. University of Göttingen,
7, 1497–1503. Germany.
Dahaoui, S., Pichon-Pesme, V., Howard, J. A. K. & Lecomte, C. Sheldrick, G. M. (2008). Acta Cryst. A64, 112–122.
(1999). J. Phys. Chem. A, 103, 6240–6250. Shoemaker, D. P., Donohue, J., Schomaker, V. & Corey, R. B. (1950).
Dai, W. & Fu, D.-W. (2008). Acta Cryst. E64, o1446. J. Am. Chem. Soc. 72, 2328–2349.
Day, G. M. (2011). Crystallogr. Rev. 17, 3–52. Simpson, H. J. & Marsh, R. E. (1966). Acta Cryst. 20, 550–555.
Derissen, J. L., Endeman, H. J. & Peerdeman, A. F. (1968). Acta Cryst. Torii, K. & Iitaka, Y. (1970). Acta Cryst. B26, 1317–1326.
B24, 1349–1354. Torii, K. & Iitaka, Y. (1971). Acta Cryst. B27, 2237–2246.
Di, K. (2010). Acta Cryst. E66, o1125–o1126. Torii, K. & Iitaka, Y. (1973). Acta Cryst. B29, 2799–2807.
Duncan, L. L., Patrick, B. O. & Brock, C. P. (2002). Acta Cryst. B58, Umadevi, K., Anitha, K., Sridhar, B., Srinivasan, N. & Rajaram, R. K.
502–511. (2003). Acta Cryst. E59, o1073–o1075.
Eijck, B. P. van & Kroon, J. (2000). Acta Cryst. B56, 535–542. Vela, M. A., McLaughlin, M. L. & Fronczek, F. R. (1989). Acta Cryst.
Frey, M. N., Lehmann, M. S., Koetzle, T. F. & Hamilton, W. C. (1973). C45, 1091–1093.
Acta Cryst. B29, 876–884. Weissbuch, I., Frolow, F., Addadi, L., Lahav, M. & Leiserowitz, L.
Fujii, I. (2009). Anal. Sci. 25, 35–36. (1990). J. Am. Chem. Soc. 112, 7718–7724.
Görbitz, C. H., Vestli, K. & Orlando, R. (2009). Acta Cryst. B65, 393– Yamada, K., Hashizume, D., Shimizu, T. & Yokoyama, S. (2007). Acta
400. Cryst. E63, o3802–o3803.

Acta Cryst. (2012). B68, 549–557 Carl Henrik Görbitz et al.  Single-crystal investigation 557
View publication stats
electronic reprint

You might also like