Biquinoline Derivative - American Chemical Society

Sep 19, 2011 - (5) (a) Kusanagi, H. Polym. J. 1996, 28, 708. (b) Engelkamp, H.;. Middelbeek, S.; Nolte, R. J. M. Science 1999, 284, 785. (6) (a) Koert...
0 downloads 0 Views 4MB Size
ARTICLE pubs.acs.org/crystal

Acid Inclusion Properties of Helical Self-Assembly of a 5,50-Biquinoline Derivative Dipjyoti Kalita and Jubaraj B. Baruah* Department of Chemistry, Indian Institute of Technology Guwahati, Guwahati 781 039, Assam, India

bS Supporting Information ABSTRACT: With a view to understand the stacking of anions in the folds of the helical self-assembly, a new organic helical 5,50 -biquinoline derivative, bis-5,50 -[2-(quinolin-8-ylamino)ethanol], having a hydroxyl pendent group attached through a flexible ethylene tether is synthesized through a manganese(II) catalyzed CC bond formation reaction. The compound adopts a helical structure with a dihedral angle between two planes of quinoline rings as 60.08 and self-assembles with the aid of hydoxy groups and nitrogen atoms in the quinoline rings along with CH 3 3 3 O interactions. The helical self-assembly of bis-5,50 -[2-(quinolin-8-ylamino)ethanol] can encapsulate guests of two different shapes, namely trigonal planar (nitrate) and tetrahedral (perchlorate), without significantly losing its original helicity. Interactions of this helical molecule with perchloric acid and nitric acid are found to be similar in solution; however, in solid state perchloric acid forms its salt while nitric acid forms a co-crystal with it.

’ INTRODUCTION Many biological macromolecules containing nitrogenous bases adopt helical structures.1 The structure of deoxyribose nucleic acid proposed by Watson and Crick in 1953 is one of the important discoveries of helical structure.2 Inspired by commonly available helical conformations in biological macromolecules, there is a challenge to synthesize artificial systems having helical chainlike structures.3 Helical structures are generally guided by hydrogen bonds,4 van der Waals interactions,5 and dative bonds to metal ions.6 Synthesis of helicene-like molecules is one of the prime interests of modern research. Several methodologies have been developed for their synthesis;7ac further to this, anion induced changes in helicity of metal complexes7d,e are reported in the literature. Lehn and co-workers have reported temperature dependent interconversion between double and single helical structures in oligopyridine carboxamide.8 Theoretical studies have suggested formation of helical structures in peptide molecules.9 Coordination polymers having homochiral helices have been constructed, and their optical devices are also reported.10 Many chiral metal organic assemblies having helical structure are synthesized, and the origin of their chirality is being studied.11 The quinoline moiety is known to adopt different orientations in the crystal lattice that contribute to the helical structures of many organic12 and inorganic13 molecules. Further to this, molecules with the biquinoline moiety are used in the field of molecular recognition14 and as optical materials.15 Despite the interesting properties and wide synthetic protocols of 2,20 biquinoline derivatives, there is limited synthetic methodology and study on 5,50 -biquinoline derivatives.16 The synthesized biquinoline derivative 2 (Scheme 1) is chosen such that it has hydroxy functional groups with flexible ethylene tethers which would help in self-assembling to form an extended structure. It would also get easily protonated or hydrogen bonded to an acid r 2011 American Chemical Society

and expected to hold anions. Many receptors have preorganized anion binding sites, but preorganized spiral systems for anion binding are not studied in detail.17aj Anion binding is also studied to distinguish salt and co-crystal formation.18 We felt that helical molecules will have ready empty spaces in the folds of the extended helix to accommodate anions. With this background, we synthesized a new 5,50 -biquinoline derivative with an alcohol pendent group, namely bis-5,50 -[2-(quinolin-8-ylamino)ethanol] (2), and studied its anion binding properties.

’ RESULTS AND DISCUSSION The biquinoline derivative bis-5,50 -[2-(quinolin-8-ylamino)ethanol] (2) was synthesized in multiple steps. In the first step, methyl 2-(quinolin-8-ylamino)acetate, which is an ester derivative of 8-aminoquinoline, was prepared by reacting 8-aminoquinoline with methyl bromoacetate. The ester group of methyl 2-(quinolin-8-ylamino)acetate on reduction gave an alcohol, namely, 2-(quinolin-8-ylamino)ethanol (1). The 2-(quinolin-8ylamino)ethanol (1) thus obtained was treated with a catalytic amount of manganese(II) acetate to give dimerized derivative 2 (Scheme 1). It may be noted that 8-aminoquinoline does not form a CC bond by reaction with manganese(II) acetate, and the reaction is very specific for coupling of 1. All these compounds formed in different steps are characterized by spectroscopic techniques such as IR, NMR, and mass spectroscopy. Compound 2 has the desired proton signals for aromatic protons besides signals corresponding to the two methylene groups flanked by oxygen and nitrogen at 3.7 ppm and 2.5 ppm, respectively. The X-ray single crystal structure of compound 2 is also Received: August 12, 2011 Revised: September 10, 2011 Published: September 19, 2011 5131

dx.doi.org/10.1021/cg201052k | Cryst. Growth Des. 2011, 11, 5131–5138

Crystal Growth & Design

ARTICLE

Scheme 1. Synthetic Route of Compound 2

Figure 1. (a) Crystal structure of 2 (inset ORTEP diagram of the asymmetric unit of compound 2) (50% elipsoid probability); (b) view of the hydrogen bonded assembly of 2 along the a-crystallographic axis; (c) space-filling model of 2.

Figure 2. Short range interactions of compound 2.

determined. Compound 2 crystallizes in the monoclinic C2/c space group, with only half of the molecule appearing in the crystallographic asymmetric unit (inset of Figure 1a). The structure of the compound is shown in Figure 1a. The compound has a twisted conformation with a dihedral angle of 60.08 between the planes of the two quinoline rings. Generally, binol

derivatives have dihedral angles between the naphthalene rings19 in the range of two naphthyl rings 7879. The presence of a hydroxyl group and the basic nitrogen atoms in the quinoline ring allows the molecule to have a selfassembled structure through different weak interactions. In the crystal structure of the self-assembled structure of 2 when viewed 5132

dx.doi.org/10.1021/cg201052k |Cryst. Growth Des. 2011, 11, 5131–5138

Crystal Growth & Design

ARTICLE

Figure 3. Changes in absorption spectra of 2 (6.7  105 M in methanol) on addition of perchloric acid (102 M in methanol). Inset: Expansion in the visible region showing the small increase in absorption at 469 nm.

Figure 4. Changes in absorption spectra of 2 (6.7  105 M in methanol) on addition of nitric acid (102 M in methanol). Inset: Expansion in the visible region showing the small increase in absorption at 471 nm.

along the a-crystallographic axis, we have observed that the molecules adopt a single helical structure (Figure 1b). The short-range interactions that stabilize the helical arrangement of 2 include two intermolecular hydrogen bonds of the hydroxy group with the quinoline N atom and the amine N atom (Figure 2). A space-filling model of 2 shows it to have a tight packed structure devoid of vacant space other than the folds (Figure 1c). The crystal structure of compound 2 also reveals that two such single helixes are held together by a T-type CH 3 3 3 π interaction as shown in Figure 2. The distance between C3H3 and the centroid of the neighboring ring (C1C5C6C7C8C9) is found to be 2.713 Å. These helical structures self-assemble to make a two-dimensional sheet. The UVvisible spectra of compound 2 show two absorption maxima at 253 nm (ε = 1.79  104 mol1 cm1) and 368 nm (ε = 4.51  103 mol1 cm1). On gradual addition of perchloric acid (102 M in methanol) to a solution of 2 (6.7  105 M in methanol), both the absorption maxima decrease, which results in the formation of three isosbestic points at 268, 324, and 417 nm (Figure 3). The formation of these isosbestic points clearly suggests the existence of two chemical species at equilibrium in the solution state. There is also a slight increase in the absorbance at 469 nm on gradual increase in the acid concentration (Figure 3, inset). Similarly, in the case of titration of 2 with nitric acid, we have observed that the absorption spectra of 2 on gradual addition of

Figure 5. IR-spectra of 2, 2a, and 2b.

nitric acid pass through three isosbestic points at 268, 325, and 423 nm as shown in Figure 4. Thus it may be seen that in both cases the changes observed in solution for protonation are similar, or in other words there is no recognition of the two acids used. 5133

dx.doi.org/10.1021/cg201052k |Cryst. Growth Des. 2011, 11, 5131–5138

Crystal Growth & Design

ARTICLE

Figure 6. (a) ORTEP diagram of the asymmetric unit of salt 2a (50% ellipsoid probability); (b) helical self-assembled structure of salt 2a; (c) spacefilling model of 2a showing the encapsulation of the perchlorate anion.

Figure 7. Short-range interactions in 2a.

The IR-spectra (Figure 5) of compound 2 show a strong peak at 3394 cm1 for the OH stretching frequency and another peak at 1603 cm1 and 1578 cm1 characterestic of CdC and CdN. In the case of the perchlorate salt (2a), we have observed three strong peaks at 1144, 1111, and 1088 cm1 that are characteristic of the perchlorate anion in addition to the peaks of the parent compound 2. Generally, an ionic perchlorate gives a broad single perchlorate absorption peak around 1100 cm1. The splitting of the perchlorate peak may be attributed to its possible distortion in an supramolecular environment. In the case of the co-crystal with nitric acid (2b) a strong peak at 1384 cm1 characterestic of the nitrate group is observed. However, the crystallographic structure shows that on treatment with perchloric acid compound 2 forms a salt (2a), where the two quinoline N atoms are protonated, whereas in the case of nitric acid a co-crystal (2b) is formed. The salt 2a crystallizes in

an orthorhombic space group Fddd. Its asymmetric unit contains half of the protonated compound 2 and a perchlorate anion (Figure 6a). The helical structure of 2 is retained after the inclusion of the perchlorate anion in the fold of the helix (Figure 6b). The perchlorate anion does not disturb the assembly of intermolecular hydrogen bonds of 2; instead the anion is encapsulated inside the cavity formed by the helical arrangement of salt 2a as shown in Figure 6c. In this case also the compound adopts a twisted geometry with a dihedral angle of 67.04 between the planes of the two quinoline rings. One of the oxygen atoms of the perchlorate anion is found to be disordered, and a disorder model is fixed by sharing the occupancy between two atoms. Perchlorate anions are known to exist in disordered form in a supramolecular environment.20 Such disorder may be attributed to the flexible interacting environment of a host for multiple orientations, thereby adopting an average occupancy in terms of oxygen atoms at a particular site. 5134

dx.doi.org/10.1021/cg201052k |Cryst. Growth Des. 2011, 11, 5131–5138

Crystal Growth & Design In case of salt 2a, the hydroxy group of the molecule acts as a hydrogen atom acceptor in a strong intermolecular hydrogen bond with the protonated quinoline N atom (N1H 3 3 3 O1). The perchlorate anion is held together by a number of CH 3 3 3 O interactions (C4H4 3 3 3 O2 and C11H11B 3 3 3 O4) and hydrogen bonding interaction (O1H1 3 3 3 O5 and O1H1 3 3 3 O3) as shown in Figure 7. Co-crystal 2b crystallizes in the orthorhombic space group C2221 with half of compound 2, a nitric acid molecule, and a water molecule in the crystallographic asymmetric unit (Figure 8a). Co-crystal 2b also adopts a helical shape, although we did not observe the same type of intermolecular hydrogen bonds as we observed in the case of 2 (Figure 8b). The nitric acid and the water molecules are held in between the parent molecules as shown in the space-filling model in Figure 8c. The formation of a co-crystal with nitric acid by 2 can be justified from the crystallographic evidence. The bond angle — C1N1C2 (Figure 8a) of the quinoline ring of 2b is found to be 119.21; however, this bond angle is 124.43 in the case of the salt 2a, where the quinoline N atom is protonated, which is in accordance with the bond angles of protonated and unprotonated quinoline derivatives reported earlier.21 This suggests that the quinoline N atom is not protonated in the case of 2b. Further, while refining the X-ray crystal structure, the hydrogen atom was located in the difference Fourier maps and the residual electron density was observed near the O atom and assigned to a proton accordingly, suggesting it to be a nitric acid instead of nitrate anion. The NO bond distances of the nitrate molecule are also not the same

ARTICLE

(dN 3 3 3 O (Å), N3 3 3 3 O2, 1.26; N3 3 3 3 O3, 1.15; N3 3 3 3 O4, 1.16), which suggests it to be nitric acid; however, these bond distances do not match exactly the structure of free nitric acid derived from microwave spectra, where the bond distances are reported22 as 1.40, 1.19, and 1.21 Å. Two nitric acid and two water molecules are encapsulated per parent molecule through intermolecular hydrogen bonds (O2 H2A 3 3 3 N2, N2H2N 3 3 3 O2) and CH 3 3 3 O interactions (C2H2 3 3 3 O4) as shown in Figure 9. The helical chains are further held together by weak CH 3 3 3 N interactions (C8 H8 3 3 3 N3). The selected hydrogen bond parameters of 2, 2a, and 2b are shown in Table 1. Structural studies along with fluorescence emission of several structurally simple nonhelical quinoline based amide derivatives have been reported earlier.23 The solid state fluorescence spectra of the salt (2a) and the cocrystal (2b) are studied, and it is found that the salt (2a) shows emission maxima at 517 nm whereas the co-crystal (2b) shows Table 1. Hydrogen Bond Parameters of 2, Salt 2a, and Co-crystal 2ba DH 3 3 3 A

dDH

dH 3 3 3 3 A

dD 3 3 3 A

(Å)

(Å)

(Å)

— DH 3 3 3 A (deg)

Compound 1 O(1)H(1) 3 3 3 N(1) [i] N(2)H(2N) 3 3 3 O(1)

0.82

2.07

2.875(3)

167.0

0.82(2)

2.46(2)

2.769(3)

103.8(17)

O(1)H(1) 3 3 3 O(5) O(2)H(2A) 3 3 3 N(1)

0.82

1.78

2.584(18)

166.0

0.82

1.99

2.808(12)

173.0

N(2)H(2N) 3 3 3 O(2) N(2)H(2N) 3 3 3 N(1) C(2)H(2) 3 3 3 O(4)

0.86

2.17

3.030(13)

178.0

0.86

2.41

2.740(12)

103.0

0.93

2.49

3.140(2)

127.0

C(4)H(4) 3 3 3 O(3) [ii]

0.93

2.56

3.446(13)

160.0

O(1)H(1) 3 3 O(1)H(1) 3 3

3 N(2)

0.82

2.58

2.906(10)

105.0

3 O(5) [iii]

0.82

2.57

3.225(18)

138.0

0.93(3)

2.45(7)

2.771(11)

101.0(5)

0.93(3)

1.85(3)

2.750(10)

164.0(7)

Salt 2a

Co-crystal 2b

N(1)H(1N) 3 3 3 N(2) N(1)H(1N) 3 3 3 O(1) [iv] N(2)H(2N) 3 3 3 N(1) N(2)H(2N) 3 3 3 O(1) [iv] C(2)H(2) 3 3 3 O(4) [iii]

Figure 8. (a) ORTEP diagram of the asymmetric unit of 2b (50% ellipsoid probability); (b) helical structure of 2b; (c) space-filling model showing the encapsulation of nitric acid and water molecules.

0.86

2.46

2.771(11)

102.0

0.86

2.26

3.084(9)

160.0

0.93

2.43

3.157(14)

135.0

C(4)H(4) 3 3 3 O(2) [v] 0.93 C(11)H(11B) 3 3 3 O(4) [iv] 0.97

2.55

3.424(11)

157.0

2.43

3.309(19)

151.0

Symmetry parameters: (i) 1/2  x, 1/2  y, 1  z; (ii) 1/2  x, 1/2 + y, 1 /2  z; (iii) x, 1/4  y, 1/4  z; (iv) x, 1/2  y, 1/2  z; (v) 1/4  x, y, 1 /4  z. a

Figure 9. Short range interactions in co-crystal 2b. 5135

dx.doi.org/10.1021/cg201052k |Cryst. Growth Des. 2011, 11, 5131–5138

Crystal Growth & Design

ARTICLE

Table 2. Crystallographic Parameters of 2, 2a, and 2b 2

2a

2b

formula

C22H22N4O2

C22H24Cl2N4O10

C22H24N6O10

CCDC no.

838324

838325

838326

mol. wt.

374.44

575.35

532.47

crystal system

monoclinic

orthorhombic

orthorhombic

space group

C2/c

Fddd

C2221

temp/K

296

296

296

wavelength/Å

0.71073

0.71073

0.71073

a/Å b/Å

29.310(14) 7.026(3)

51.844(11) 8.1994(19)

14.640(5) 20.521(5)

c/Å

9.173(4)

24.616(6)

8.322(2)

α/deg

90.00

90.00

90.00

β/deg

106.564(18)

90.00

90.00

γ/deg

90.00

90.00

90.00

V/Å3

1810.5(15)

10464(4)

2500.2(12)

Z

4

16

4

density/(mg m3) abs coeff/mm1

1.374 0.091

1.461 0.310

1.415 0.114

abs correction

none

none

none

F(000)

792

4768

1112

total no. of reflections

1652

2320

2008

reflections, I > 2σ(I)

1011

975

855

max. 2θ/deg

51.0

50.0

49.46

35 e h g 35

61e h g 61

16 e h g 16

ranges (h, k, l)

7 e k g 8 11 e l g 10

9 e k g 9 29 e l g 29

22 e k g 22 9 e l g 9

complete to 2θ (%)

97.6

100.0

94.5

refinement method

full-matrix least-squares on F2

full-matrix least-squares on F2

full-matrix least-squares on F2

data/restraints/parameters

1652/0/132

2320/9/186

2008/0/173

GOF (F2)

0.902

1.972

1.274

R indices [I > 2σ(I)]

0.0445

0.1216

0.1038

R indices (all data)

0.0763

0.2256

0.2274

the emission maxima at 509 nm upon excitation at 380 nm. The fluorescence spectra of protonated quinoline derivatives show emission maxima at around 510 nm.23 However, compound 2 showing fluorescence emission at 507 nm makes it difficult to distinguish the protonated and unprotonated forms in this particular case. From the above observations it is clear that the salt or co-crystal formation from two strong acids could not be distinguished in solution, whereas in the solid state the structures of 2a and 2b differ. The formations of co-crystal and salt are related to pKa,24 but from the present study we observe that two strong acids result in either of them. Thus, it is the packing preference in the solid state that favors the state as protonated or a molecule of crystallization, but finally they are in the lattice due to electrostatic interactions further supported by other directional weak interactions. In conclusion, we have synthesized a new organic helical 5,50 biquinoline derivative to establish its self-assembling properties in the solid state. We have also shown that the helical molecules can encapsulate guests of two different shapes without losing the original helicity. The interactions of these helical molecules with perchloric acid and nitric acid are found to be similar in solution; however, in solid state perchloric acid forms its salt while nitric acid forms a co-crystal with it.

’ EXPERIMENTAL SECTION All reagents and solvents were purchased commercially and were used without further purification, unless otherwise stated. 1H NMR data were recorded with a Varian 400 MHz FTNMR spectrometer. The FT-IR spectra were recorded using a PerkinElmer spectrum one spectrometer in the KBr pellets in the range 4000400 cm1. The UV/vis spectra were recorded using a PerkinElmer Lambda 750 spectrometer. For the UV titration, a solution of compound 2 in methanol was prepared and a constant aliquot (10 μL) of different guest acids (102 M in methanol) was added to the host solution. UV spectra of the solution were recorded after every addition of the guest acids. The X-ray single crystal diffraction data were collected at 296 K with Mo Kα radiation (λ = 0.71073 Å) using a Bruker Nonius SMART CCD diffractometer equipped with a graphite monochromator. The SMART software was used for data collection and also for indexing the reflections and determining the unit cell parameters; the collected data were integrated using SAINT software. The structures were solved by direct methods and refined by full-matrix least-squares calculations using SHELXTL software. All the non-H atoms were refined in the anisotropic approximation against F2 of all reflections. The H-atoms, except those attached to oxygen atoms, were placed at their calculated positions and refined in the isotropic approximation; those attached to oxygen were located in the difference Fourier maps and refined with isotropic displacement coefficients. The crystallographic parameters of the compounds are 5136

dx.doi.org/10.1021/cg201052k |Cryst. Growth Des. 2011, 11, 5131–5138

Crystal Growth & Design tabulated in Table 2. One of the oxygen atoms of the perchlorate anion of 2a is disordered, and the disorder model is fixed by sharing the occupancy between two oxygen atoms in a 1:1 ratio.

Synthesis of 2-(Quinolin-8-ylamino)ethanol (1) and Bis-5,50 [2-(quinolin-8-ylamino)ethanol] (2). 8-Aminoquinoline (0.725 g,

5 mmol) was dissolved in dry acetone (20 mL), and potassium carbonate (1.0 g, 7.5 mmol) was added to it. The solution was stirred at room temperature for 15 min, then methyl bromoacetate (0.76 g, 5 mmol) was added, and the reaction mixture was then refluxed at 70 C for 12 h (The progress of the reaction was monitored at regular intervals using TLC). After completion of the reaction, the reaction mixture was then filtered and the solvent was then removed under reduced pressure to obtain the ester, methyl 2-(quinolin-8-ylamino)acetate, as a yellow semisolid. In the next step, the ester was reduced to the corresponding alcohol 2-(quinolin-8-ylamino)ethanol (1) by reduction with sodium borohydride at room temperature in methanol and tetrahydrofuran solvent (1:1 v/v) as per conventional procedure. The alcohol was purified by column chromatography using a silica column with ethyl acetate in hexane (15%) as eluent. Yield: 67%. 1H NMR (CDCl3, 400 MHz): 8.7 (d, J = 4.0 Hz, 1H); 8.0 (d, J = 8.4 Hz, 1H); 7.3 (m, 2H); 7.0 (d, J = 8.0 Hz, 1H); 6.7 (d, J = 8.0 Hz, 1H); 3.9 (t, J = 5.2 Hz, 2H); 3.5 (t, J = 6.0 Hz, 2H). LC-MS [M + 1]: 189.01. The crude alcohol 1 (0.564 g, 3 mmol) was then dissolved in 20 mL of methanol, and a catalytic amount of Mn(O2CCH3)2 3 4H2O (0.036 g, 0.03 mmol) was added. The reaction mixture was stirred for about 8 h at room temperature to obtain bis-5,50 -[2-(quinolin-8-ylamino)ethanol] (2) as a yellow solid. The product was further purified by column chromatography using a silica column with ethyl acetate in hexane (15%) as eluent. Yield: 61%. IR (KBr, cm1): 3394 (s), 2929 (w), 2854 (w), 1603 (m), 1578 (s), 1518 (s), 1472 (m), 1440 (m), 1379 (m), 1358 (s), 1071 (m), 817 (m). 1 HNMR (DMSO-d6, 400 MHz): 8.7 (s, 1H), 7.6 (d, J = 7.6 Hz, 1H), 7.3 (d, J = 7.6 Hz, 1H), 6.8 (d, J = 7.6 Hz, 1H), 6.6 (s, 1H), 4.9 (s, 1H), 3.7 (m, 2H), 2.5 (m, 2H). LC-MS [M + 1]: 375.34.

Synthesis of the Salt and the Co-crystal 2a and 2b. Compound 2 was dissolved in a dilute solution of perchloric acid (0.3 M) and nitric acid (1 M) and allowed to stand for 3 days to obtain the crystals of their respective salt (2a) and co-crystal (2b). 2a: Yield: 72%. IR (KBr, cm1): 3398 (bs), 2932 (w), 2854 (w), 1627 (s), 1573 (s), 1459 (m), 1407 (m), 1307 (m), 1144 (s), 1111 (s), 1088 (s), 806 (m), 772 (m). 2b: Yield: 68%. IR (KBr, cm1): 3440 (bs), 1628 (s), 1574 (w), 1523 (w), 1458 (w), 1384 (s), 1123 (w), 1059 (m), 815 (m).

’ ASSOCIATED CONTENT Supporting Information. X-ray crystallographic files (CIF) of 2, 2a, and 2b. This information is available free of charge via the Internet at http://pubs.acs.org/.

bS

’ AUTHOR INFORMATION Corresponding Author

*E-mail: [email protected].

’ ACKNOWLEDGMENT The authors thank the Department of Science and Technology, New Delhi, India, for use of the X-ray facility, and one of the authors (D.K.) is thankful to the Council of Scientific and Industrial Research, New Delhi, India, for a senior research fellowship. ’ REFERENCES (1) (a) Anfinsen, C. B. Science 1973, 181, 223. (b) Venkatchalam, C. M.; Ramachandran, G. N. Annu. Rev. Biochem. 1969, 38, 45.

ARTICLE

(2) Watson, J. D.; Crick, F. C. H. Nature 1953, 171, 737. (3) (a) Nelson, J. C.; Saven, J. G.; Moore, J. S.; Wolynes, P. G. Science 1997, 277, 1793. (b) Piguet, C.; Bernardinelli, G.; Hopfgartner, G. Chem. Rev. 1997, 97, 2005. (c) Albrecht, M. Chem. Rev. 2001, 101, 3457. (4) (a) Blasio, B. D.; Benedetti, E.; Pavone, V.; Pedone, C. Biopolymers 1989, 28, 203. (b) Langs, D. A. Science 1988, 241, 188. (5) (a) Kusanagi, H. Polym. J. 1996, 28, 708. (b) Engelkamp, H.; Middelbeek, S.; Nolte, R. J. M. Science 1999, 284, 785. (6) (a) Koert, M.; Harding, M.; Lehn, J.-M. Nature 1990, 346, 339. (b) Hasenknopf, B.; Lehn, J.-M.; Kneisel, B. O.; Baum, G.; Fenske, D. Angew. Chem., Int. Ed. 1996, 35, 1838. (c) Bell, T. W.; Jousselin, H. Nature 1994, 367, 441. (7) (a) Urbano, A. Angew. Chem., Int. Ed. 2003, 42, 3986. (b) Nuckolls, C.; Katz, T. J.; Katz, G.; Collings, P. J.; Castellanos, L. J. Am. Chem. Soc. 1999, 121, 79. (c) Johnson, C. A., II; Berryman, O. B.; Sather, A. C.; Zakharov, L. N.; Haley, M. M.; Johnson, D. W. Cryst. Growth Des. 2009, 9, 4247. (d) Wenzel, M.; Jameson, G. B.; Ferguson, L. A.; Knapp, Q. W.; Forgan, R. S.; White, F. J.; Parsons, S.; Tasker, P. A.; Plieger, P. G. Chem. Commun. 2009, 3606. (e) Miyake, H.; Yoshida, K.; Sugimoto, H.; Tsukube, H. J. Am. Chem. Soc. 2004, 126, 6524. (8) Berl, V.; Huc, I.; Khoury, R. G.; Krische, M. J.; Lehn, J.-M. Nature 2000, 407, 720. (9) Schramm, P.; Hofmann, H.-J. J. Pept. Sci. 2010, 16, 276. (10) Han, L.; Hong, M. C.; Luo, R. H.; Lin, Z. Z.; Yuan, D. Q. Chem. Commum. 2003, 2580. (11) (a) Yan, B.; Capracotta, M. D.; Maggard, P. A. Inorg. Chem. 2005, 44, 6509. (b) Maggard, P. A.; Stern, C. L.; Poeppelmeier, K. R. J. Am. Chem. Soc. 2001, 123, 7742. (c) Hong, M. C. Cryst. Growth Des. 2007, 7, 10. (12) (a) Jiang, H.; Leger, J.-M.; Dolain, C.; Guionneau, P.; Huc, I. Tetrahedron 2003, 59, 8365. (b) Gan, Q.; Bao, C.; Kauffmann, B.; Lard, A. G.; Xiang, J.; Liu, S.; Huc, I.; Jiang, H. Angew. Chem., Int. Ed. 2008, 47, 1715. (13) (a) Yuan, G.; Shao, K.-Z.; Wang, X.-L.; Lan, Y.-Q.; Zhao, Y.-H.; Su, Z.-M. Inorg. Chem. Commun. 2008, 11, 1246. (b) Prema, D.; Wiznycia, A. V.; Scott, B. M. T.; Hilborn, J.; Desper, J.; Levy, C. J. Dalton Trans. 2007, 4788. (14) Liu, Y.; Song, Y.; Wang, H.; Zhang, H.-Y.; Wada, T.; Inoue, Y. J. Org. Chem. 2003, 68, 3687. (15) (a) Song, Y.; Chen, Y.; Liu, Y. J. Photochem. Photobiol., A: Chem. 2005, 173, 328. (b) Pucci, D.; Crispini, A.; Ghedini, M.; Szerb, E. I.; Deda, M. L. Dalton Trans. 2011, 40, 4614. (16) Laurence, D.; Iii, T.; Papadimitrakopoulos, F. Macromol. Symp. 1998, 125, 143. (17) (a) Sessler, J. L.; Barkey, N. M.; Pantos, G. D.; Lynch, V. M. New J. Chem. 2007, 646. (b) Begum, R.; Kang, S. O.; Bowman-James, K. Angew. Chem., Int Ed. 2006, 45, 7882. (c) Bianchi, A., Bowman-James, K., Gracia Esana, E., Eds. Supramolecular Chemistry of Anions; Wiley-VCH: Weinheim, Germany, 1997; (d) Atwood, J. L.; Holman, K. T.; Steed, J. W. J. Chem. Soc., Chem. Commun. 1996, 1401. (e) Steed, J. W. Chem. Commun. 2006, 2637. (f) Martinez-Manez, R.; Sancenon, F. Chem. Rev. 2003, 103, 4419. (g) Sukasai, C.; Tuntulani, T. Chem. Soc. Rev. 2003, 32, 192. (h) Beer, P. D.; Gale, P. A. Coord. Chem. Rev. 2001, 213, 79. (i) Atwood, J. L.; Holman, K. T.; Steed, J. W. Chem. Commun. 1996, 1401. (j) Schmidtchen, F. P.; Berger, M. Chem. Rev. 1997, 97, 1609. (18) (a) Stanton, M. K.; Bak, A. Cryst. Growth Des. 2008, 8, 3856. (b) Bhogala, B. R.; Basavoju, S.; Nangia, A. CrystEngComm 2005, 7, 551. (c) Mohmed, S.; Tocher, D. A.; Vickers, M.; Karamertzanis, P. G.; Price, S. L. Cryst. Growth Des. 2009, 9, 2881. (d) Singh, D.; Bhattacharyya, P.; Baruah, J. B. Cryst. Growth Des. 2010, 10, 348. (19) Liu, G.-H.; Xue, Y.-N.; Yao, M.; Fang, H.-B.; Yu, H.; Yang, S.-P. J. Mol. Struct. 2008, 875, 50. (20) (a) Yue, X.-T.; Nie, J.-J.; Li, Y.-T.; Wu, Z.-Y.; Yan, C.-W. Acta Crystallogr. 2011, C67, m115. (b) Ghazzali, M.; Ohrstrom, L.; Lincoln, P.; Langer, V. Acta Crystallogr. 2008, C64, m243. (c) Karabyk, H.; Gokc, A. G.; Kun, N. C.; Aygun, M.; Kzlkus, Y. T. J. Chem. Cryst. 2009, 39, 279. (d) Katrusiak, A.; Dolska, M.; Urjasz, H.; Grech, E.; Brezinski, B. J. Mol. Struct. 2002, 610, 73. 5137

dx.doi.org/10.1021/cg201052k |Cryst. Growth Des. 2011, 11, 5131–5138

Crystal Growth & Design

ARTICLE

(21) (a) Kalita, D.; Baruah, J. B. CrystEngComm 2010, 12, 1562. (b) Kalita, D.; Sarma, R.; Baruah, J. B. CrystEngComm 2009, 11, 803. (22) Cox, A. P.; Riveros, J. M. J. Chem. Phys. 1965, 42, 3106. (23) (a) Karmakar, A.; Sarma, R. J.; Baruah, J. B. CrystEngComm 2007, 9, 379. (b) Karmakar, A.; Baruah, J. B. Supramol. Chem. 2008, 20, 667. (24) (a) Childs, S. L.; Stahly, G. P.; Park, A. Mol. Pharmaceutics 2007, 4, 323. (b) Bucar, D.-K.; Henry, R. F.; Lou, X.; Borchardt, T. B.; Zhang, G. G. Z. Chem. Commun. 2007, 525. (c) Karki, S.; Friscic, T.; Jones, W.; Motherwell, W. D. S. Mol. Pharmaceutics 2007, 4, 347. (d) Perakyla, M. J. Am. Chem. Soc. 1998, 120, 12895.

5138

dx.doi.org/10.1021/cg201052k |Cryst. Growth Des. 2011, 11, 5131–5138