SURMOF Induced Morphological Crystal Engineering of Substituted

Sep 11, 2018 - (MOF) as an example. The prepared SURMOF was used to study morphological crystal engineering of substituted benzamides. Pre-...
0 downloads 0 Views 4MB Size
Article pubs.acs.org/crystal

Cite This: Cryst. Growth Des. 2018, 18, 7048−7058

SURMOF Induced Morphological Crystal Engineering of Substituted Benzamides Geetha Bolla and Allan S. Myerson* Department of Chemical Engineering, Massachusetts Institute of Technology, 77 Massachusetts Avenue, E19-502b, Cambridge, Massachusetts 02139, United States

Crystal Growth & Design 2018.18:7048-7058. Downloaded from pubs.acs.org by UNIV OF SOUTH DAKOTA on 11/07/18. For personal use only.

S Supporting Information *

ABSTRACT: Well-established, self-assembled monolayers are further explored and extended to SURMOFs as basic template heterosurfaces by using a basic stranded HKUST metal−organic framework (MOF) as an example. The prepared SURMOF was used to study morphological crystal engineering of substituted benzamides. Prepared SURMOF substrates surfaces were analyzed by scanning electron microscopy and confirmed through thin film X-ray diffraction. Here we proved that the MOF surface exhibited adsorption of the substituted benzamides nucleus during both nucleation and growth directed due to the weak supramolecular interactions. MOF induced nucleation was investigated and showed control over the substituted benzamides for morphological engineering. Hence the concept of MOF induced morphological crystal engineering is adapted to the field of morphology crystal engineering and showed the additional advantage of heterogeneous nucleation.



growth, which finally leads to changes in crystal morphology. Hence, an efficient method for kinetic resolution of racemic conglomerates by crystallization in the presence of “tailor-made” additives was described and explained as stereoselective adsorption of the additive at the surface of the growing crystals of the enantiomer resulting in a drastic decrease in their rate of growth, which thus leads to preferential crystallization of the opposite enantiomer (“rule of reversal”).2 This was proven by the crystallization of the conglomerates (R,S)-glutamic acid hydrochloride, (R,S)-threonine, (R,S)-p-hydroxyphenyl) glycine p-toluene sulfonate, and (R,S)-asparagine hydrate in the presence of other amino acids as additives. As another example, a dramatic change in morphology in the presence of carboxylic acid additives on amide crystals, RCONH2/RCO2H, e.g., asparagine/aspartic acid, was proven because the carboxyl O atom is a much weaker proton acceptor than the corresponding amide O atom. Hence, there is a large energy loss in replacing an amide−amide hydrogen bond by an amide−acid hydrogen bond at the site of the additive. The system aspartic acid as an additive was added on asparagine, and the crystals that are prismatic in shape induce growth of {010} as plates due to retardation of growth along the b axis. Often this method is limited in practical experiments due to the additive as an impurity and the excessive amount of additive needed.10 In another direction, recently heterogeneous surface medication methods were developed extensively through self-assembled

INTRODUCTION A crystal cannot be formed instantaneously; i.e., nucleation occurs at first, followed by growth, during growth, and each crystal has its own growth history. Hence, understanding the crystal structure and morphology crystal engineering of small organic molecules has become the grand challenge for solid-state chemists for the last couple of decades through various methods.1−3 However, the fundamental understanding with proper examples on molecular faces and the ability to design crystals with desired propertiesthe whole concept falling under the umbrella of crystal engineeringare still at a rudimentary level in organic or inorganic crystals; hence, there is an emergent need due to the potential applications.4−6 In addition, what has been developed to understand or control biomineral crystal formation mechanisms with optical properties called biomineralization7−9 has become fascinating with a dual advantage in improved fabrication of synthetic materials and further help to solve many serious pathological problems involving mineralization. Consequently, numerous methods have been developed to understand crystal growth; among them, a “tailor-made” additive (a molecule very similar in structure) induced crystal mechanism due to surface adsorption of a particular crystal face via weak supramolecular interaction in solution was initiated recently based on crystal engineering principles.3 Reported results show that such additives had a dramatic effect on crystal growth and habit due to the preferential adsorption of the additive on specific crystal faces called active sites for nucleation and growth. In fact, generally, a bound additive perturbs the regular growth of upcoming layers based on attachment energies and allows them different directional © 2018 American Chemical Society

Received: August 13, 2018 Revised: September 9, 2018 Published: September 11, 2018 7048

DOI: 10.1021/acs.cgd.8b01214 Cryst. Growth Des. 2018, 18, 7048−7058

Crystal Growth & Design

Article

Figure 1. (a) SURMOFs basic design with multiple layers in the present study. (b) Preparation method for SURMOFs used in the present study, and this was called dipping layer by layer (LBL-D).

Figure 2. (a−f) SEM images of the prepared SURMOF thin film surface in different experiments, which showed that the surface is at the microlevel with uniform MOF distribution.

they are limited in their success rate.11−15 Here we have extended SAMs to SURMOFs (SAM + MOFs, Figure 1) which are principally the deposition of the MOFs on SAM heterosurfaces. Recently a prominent incremental result was observed, and it proved that MOFs are highly porous and it

monolayers (SAM) for polymorph screening and morphological crystal engineering because the presence of a foreign surface in a crystallizing system can influence the pathway of crystallization for the simple reason that a nucleation barrier can be lowered as a reduction in the interfacial free energy occurs, but 7049

DOI: 10.1021/acs.cgd.8b01214 Cryst. Growth Des. 2018, 18, 7048−7058

Crystal Growth & Design

Article

Scheme 1. Model Systems Such As Benzamides Used in This Study

showed applications in various directions of the science.16−20 SURMOFs have been studied recently with different applications such as photovoltaics, CO2 reduction, memory devices, supercapacitors, and batteries.17 The advantage of controlled orientation of the MOFs on the basis of the ground SAM functional group would allow preferential growth of the target functional group of small organic molecules, but this young branch of chemistry not yet been explored in the direction of morphological engineering. As MOFs are highly porous crystalline materials, their impact and contribution on the surface as a heterogeneous layer are quite effective when compared with the usual SAM surface. In addition, they can lead to different nucleation and growth directions than the usual crystallization path. With this prime idea, here we started our studies and discussed how SURMOF can influence crystal morphological crystal engineering of benzamides (BZAs). The supreme selection of BZAs was justified due to their lack of a complexation nature during crystallization time with selected HKUST MOFs.



Figure 3. (a) Morphology calculated based on attachment energy. (b) Calculated morphology with hydrogen bond preferences in faces based on BFDH predictions. Single crystal face indexing experiment results of BZA. (c, d) Face indexing of the BZA single crystals with their faces grown on solution crystallization without the MOF surface suggests plate morphology with (001), (011), and (101) morphologically important faces as major faces. (e, f) Face indexing of the BZA crystals grown on the surface of the MOF crystallization suggests needle morphology with (001) and (011) morphologically important faces as major faces and (101) was decreased.

RESULTS AND DISCUSSION The template SURMOF crystallization method involved three major steps. As the first step, the SAMs were prepared using gold substrates and thiol solutions, and second, a well-known, reported three-dimensional (3D) HKUST MOF,21 which consists of trimestic acid (H3BTC) and copper acetate, was formed from solutions using layer by layer dipping16 (LBL-D). These SURMOF substrates were designed to investigate how the template functionalization of these inorganic, ordered, highly porous self-assembled layers of MOF on the surface (SURMOF) can influence the nucleation of functional organic molecules; all the preparation details are discussed in Supporting Information, section 1. The prepared SURMOF was confirmed by thin film X-ray diffraction (Figure S1a,b), topochemical analysis was carried out by scanning electron microscopy (SEM), and atomic force microscopy was reported recently in our earlier paper.22 However, here also we have confirmed the surface again with SEM analysis for different batches (Figure 2). The images of Figure 2a,b are one batch for 10 layers and they showed the porous MOF surface distribution; Figure 2c−f shows different batches in lower to higher magnification, suggesting uniform distribution. In addition, various places on the surface SEM image analysis suggests that the whole substrate film was distributed in an uniform way. However, they are micrometers in length. As part of our efforts, the third step is crystallization of small organics, and to double check the viability of our SURMOF approach here we have chosen benzamide (BZA), 4-amino benzamide (ABZA), and 4-hydroxy benzamide (HBZA). Indeed, no complexation was observed between BZAs and pristine HKUST MOF during slurry experiments at room temperature in 4−6 h (Figure S1). However, in our previous result,

acetaminophen (APAP) was chosen as an example organic to explore crystalline forms as a proof-of-concept, and we reported the crystallization of APAP Form II with block morphology for the first time.22 In an extended direction here, we studied the morphological changes and crystal growth directions of BZAs (Figure 2) on the SURMOF surface as discussed in detail. Example 1. Benzamide. The morphology of a matured crystal is defined by its relative rates of growth in different directions, which means the faster the growth in a given direction, the smaller the face at the perpendicular. Hence, if growth is inhibited in a direction perpendicular to a particular crystal face, the area of that face is expected to increase relative to the areas of other faces of the same crystal. Addadi et al. reported that BZA crystallizes from ethanol as plate-like crystals, but in the presence of an additive such as benzoic acid they observed retardation of crystal face growth along the crystallographic b axis, and it ended as needle crystals.1 In detail, BZA (compared with reported Refcode BZAMID11) crystallizes in the monoclinic crystal system (P21/c) with cyclic homodimer motifs to form stable (001) layers with major morphological importance. Further phenyl C−H and π functional interacted via weak van der Waals interactions between phenyl groups along the other 7050

DOI: 10.1021/acs.cgd.8b01214 Cryst. Growth Des. 2018, 18, 7048−7058

Crystal Growth & Design

Article

Figure 4. (a, b) View of the crystal face (001), (c, d) view of the crystal face (011), (e, f) view of the crystal face (101), calculated from Materials Studio. This suggests that decreasing morphological importance of the (101) face is due to the lack of hydrogen bond polar groups.

Table 1. Attachment Energy Calculations of the BZA Crystal Form I by Using the Dreiding Force Field Method crystal face

Eatt (total) kcal/mol

Eatt (vdW) kcal/mol

Eatt (electrostatic) kcal/mol

total facet area

% total facet area

(0 0 2) (0 1 1) (1 0 1̅) (1 0 1) (0 1 2)

−14.91 −44.50 −49.70 −47.89 −43.55

−13.37 −30.93 −30.32 −29.36 −36.24

−1.54 −13.57 −19.38 −18.52 −7.31

1.537996 × 104 4.453315 × 103 2.069583 × 103 3.309268 × 103 1.604570 × 103

57.35217626 16.60650455 7.71751644 12.34032763 5.98347513

directions and it ended as plate-like crystals. In the present case, with an additive, for example, benzoic acid the formation of acidamide heterosynthon is possible through energy loss at an active site which can lead to an overall loss in energy BZA along b, and hence morphology transformed two-dimensional (2D) crystals into one dimensional. Apart from benzoic acid, there are a few other additives that were studied such as o-/mtoluamide, and it resulted in inhibition of growth along the crystallographic a direction due to the adsorption in the b direction.1 Here we have calculated the morphology based on the attachment energy, and next BFDH predictions (Figure 3a,b) suggest plate morphology of BZA. BZA crystals were grown with solution crystallization in EtOH (1−10 mg/mL), and further the surface of the designed SURMOF substrate was considered for face indexing experiments. Solution crystallization without the MOF surface resulted in single crystals showing plate morphology with (001), (011), (101) morphologically important faces as major faces, whereas crystals grown on the surface of the MOF crystallization suggest needle morphology with (001), (011) morphologically important faces as major faces and (101) was retarded. Packing of BZA clearly showed the (001) face along the b direction, BZA dimers N−H···O and along the crystallographic a direction, and N−H···O catemers of BZA

extended via supramolecular interaction (hydrogen bond). Hence the SURMOF induced BZA crystal was morphologically engineered by stopping the (001) crystal face along the b direction. This might be attributed to the weak interactions between the HKUST MOF porous surface and amide CONH2. However, 2D slice crystal faces calculated in Materials Studio showed (Figure 4a,b) that the (001) face of the BZA faces exposed to amide as our next (011) face (Figure 4c,d) has no hydrogen bond groups, whereas the (101) crystal face showed alternative amide groups (Figure 4e,f). Calculated attachment energies (Table 1) are for (001), (011), (101) −14.9, −44.50, −47.89 kcal/mol, which showed that the growth rate of the second and third crystal faces difference is not high (∼3 kcal/mol). Hence controlling one face with modified SURMOF surface crystallization is an alternative to the additive induced crystallization as the result matches to attain plate to needle crystals of BZA. Example 2. 4-Amino Benzamide. ABZA crystals grew initially without substrate in EtOH (1−10 mg/mL) solvent through solution crystallization, and the observed crystals showed prismatic, block morphology with (100) and (110) as major morphologically important crystal faces. The results are further supported by attachment energy and BFDH morphology predictions and showed block habit (100) as a major crystal 7051

DOI: 10.1021/acs.cgd.8b01214 Cryst. Growth Des. 2018, 18, 7048−7058

Crystal Growth & Design

Article

Figure 5. (a−r) ABZA crystal morphology experiments through predictions and face indexing experiments with and without the SURMOF surface.

face (Figure 5a,b). Crystals 1, 2, and 3 were grown from low, medium, and high concentrated solutions without HKUST MOF surfaces and were picked for face monitoring experiments. ABZA crystallizes in monoclinic, P21 through primary amide chains N−H···O along the b axis. Further, these opposite chains interact to form 2D sheets, and these sheets form next N···H−O interactions with the amine to give a 3D structure (Figure 5b). Here we have re-collected the single crystal data of ABZA as the reported data unit cell orientation

unique axis is different from the present studies; all the crystallographic data are displayed in Table S1. Hence it is a trivial situation that a hetero substrate can interact in three different motifs like at inertial chains, second at the opposite chains site, and finally at the amide active site at the time of the birth of the crystal (nucleation time). Interestingly these crystals showed initially the (100) face as the major face and (110), (111), (010), (101), (001), and (111) as minor faces at low dilution. Further (100) crystal faces started disappearing, and 7052

DOI: 10.1021/acs.cgd.8b01214 Cryst. Growth Des. 2018, 18, 7048−7058

Crystal Growth & Design

Article

Table 2. Attachment Calculations of the ABZA crystal face

Eatt (total) kcal/mol

Eatt (vdW) kcal/mol

Eatt (electrostatic) kcal/mol

total facet area

% total facet area

(0 0 1) (1 0 0) (1 0 1̅) (1 0 1) (0 1 1) (0 1̅ 1) (1 1 0) (1 1̅ 0) (1 1 1̅) (1 1̅ 1̅)

−16.53 −13.55 −17.56 −23.57 −25.36 −25.36 −19.55 −19.55 −23.15 −23.15

−3.89 −9.06 −9.40 −10.93 −14.90 −14.90 −14.80 −14.80 −15.13 −15.13

−12.64 −4.48 −8.15 −12.64 −10.45 −10.45 −4.74 −4.74 −8.01 −8.01

1.357478 × 103 1.407212 × 103 725.36305004

24.27382879 25.16313667 12.97062225

38.15940387 38.15940387 819.04188985 819.04188985 193.94900169 193.94900169

0.68234964 0.68234964 14.64574597 14.64574597 3.46811054 3.46811054

Figure 6. View of the ABZA crystal faces calculated from Materials Studio.

also further showed the (111) faces as a minor morphological importance. We have calculated the hkl slices of the ABZA faces to understand the growth mechanism. The crystal faces which were attributed through solution crystallization (100), (001), (110) crystal faces (from low to high concentrations) exhibited amide and amide group motifs toward the surface (Figure 6). Hence, MOF pores can block their surface, which allows other faces to grow and further leads to different growth compared to normal experiments. Example 3. 4-Hydroxy Benzamide. HBZA crystals (compared with reported refcode VIDMAX) were grown initially without substrate in EtOH (1−10 mg/mL) through solution crystallization as such as BZA and ABZA, and the observed, well-grown single crystals showed rod morphology with (001), (011), and (010) with equal morphologically important crystal faces, and the crystals grew along the crystallographic a axis. The results are further supported by attachment energy and BFDH morphology predictions and showed block habit (100) as a major crystal face (Figure 7a,b). At first, crystals 1 and

the (110) face started growing as the major face at a higher concentration; (111) finally ended as block morphology (Figure 5c−h). As our main focus is to understand the crystal morphology influence due to the MOF surface, here we grew ABZA crystals on the surface of the SURMOF heterogeneous substrate through a solution phase epitaxy crystallizations method. The grown crystals labeled as 4, 5, 6, 7, and 8 (Figure 5i−r) showed needles on the surface and block morphology at the edges, (111) as the major face and (001), (110), (100), (011), and (010), which clearly confirmed that the nucleation of ABZA on MOF surface is different from normal solvent crystallization experiments and it did not follow the attachment energy during the crystal growth (Table 2). This result suggests that the N−H···O interactions of the amide chains and amide to amine acted as active sites with the MOF surface, which finally directed the crystal faces which have high attachment energies ending as needle morphology (retarded along the b axis and allowed along the crystallographic a axis). Indeed, at the edge grown crystals showed needle morphology with the (100) face as the primary face and 7053

DOI: 10.1021/acs.cgd.8b01214 Cryst. Growth Des. 2018, 18, 7048−7058

Crystal Growth & Design

Article

Figure 7. (a, b) HBZA crystal faces calculated from Materials Studio by attachment energy and BFDH morphology along with their interactions. (c−n) Single crystal face indexing experiments of the HBZA crystals grown with and without MOF substrate. This suggests that solution crystallization gave rod morphology with a near growth rate of (011), (010), (001), whereas with MOF small to big plates (001) as the major face. In addition, at the edge all the faces have an equal growth rate ending with block morphology.

here are in detail displayed in Supporting Information, sections 2 and 3. Powder X-ray Diffraction Analysis. In order to know the 2D orientation growth of BZA, ABZA, and HBZA bulk crystals, thin film X-ray diffraction (XRD) experiments were performed. Crystals grown from solution crystallization showed plate morphology and resulted in PXRD with (001), (011), (101); (2θ ∼ 7, 17, 18) crystal faces as major morphological importance. A gradual decrease and selective diffraction of the (002), 2θ of BZA at 7.851 was observed for the crystals grown on the surface MOF; this confirmed the 1D growth (Figure 9a). ABZA crystals grown at low dilution showed unique (00l) 2θ values, and further medium to high concentrated solution crystallization experiments resulted in (hk0) 2θ peaks (Figure 9b), which supported the single crystal face indexing experiments (Figure 5). However, the thin films and the crystals grown on the designed surface resulted in an additional 2θ, which correspond to other crystals and confirmed the retardation of the previously existing faces (Figure 9c). Further the same was done for HBZA, and this confirmed that the growth of the crystals in solution ended with (0kl), (00l) faces, whereas in the presence of SURMOF crystals grew (00l) as the major face exclusively (Figure 9d,e).

2 were grown from solutions without the HKUST MOF surface and showed the (001), (010), and (011) faces as major faces with equal importance, and growth happened along the a axis via N−H···O (Figure 7b-ii) supramolecular interactions. Second, crystals on the surface of the SURMOF heterogeneous substrate through the solution phase epitaxy crystallization method were studied carefully with multiple crystals at various places of the designed surface. These crystals showed plate morphology as (001) with major and (011), (010), (100), and (101) with minor morphological importance. This clearly shows the MOF induced grown crystals labeled as 3 and 4 (Figure 7) are extended along the crystallographic c axis through N−H···O on the surface and block morphology at the edges. Edge crystals were balanced of MOF and solution crystallization faces ended with similar morphological importance as the existing crystal faces, which clearly confirmed that the nucleation of HBZA on the MOF surface happened along the c axis, whereas in solution the crystals grew along the crystallographic a axis. Controlling the growth was successful here by using our designed SURMOFs. All the faces’ exposure toward the surface and their attachment energies are shown in Figure 8 and Table 3. Indeed, here the MOF surface grown crystals did not show the attachment energy calculated stable faces (001) and (011) of −56.49 and −48.40 kcal/mol. Further, all the crystals studied 7054

DOI: 10.1021/acs.cgd.8b01214 Cryst. Growth Des. 2018, 18, 7048−7058

Crystal Growth & Design

Article

Figure 8. View of the HBZA crystal faces calculated from Materials Studio.

Table 3. Attachment Energies of the ABZA Crystal Faces Calculated From Materials Studio crystal face

Eatt (total) kcal/mol

Eatt (vdW) kcal/mol

Eatt (electrostatic) kcal/mol

total facet area

% total facet area

(0 0 2) (0 1 1) (1 0 0) (0 2 0) (1 1 0) (1 1 1̅) (1 0 2) (1 1 1)

−56.49 −48.40 −62.41 −49.29 −64.09 −76.85 −64.72 −68.14

−19.31 −19.67 −44.25 −35.98 −40.71 −43.19 −38.66 −41.43

−37.17 −28.73 −18.16 −13.30 −23.38 −33.65 −26.06 −26.71

1.005028 × 104 2.126161 × 104 5.937345 × 103 4.085920 × 103 8.319058 × 103

18.37105511 38.86442752 10.85296318 7.46871626 15.20653285

4.290678 × 103 762.23955157

7.84299574 1.39330935

Morphological Crystal Engineering and Possible Growth on the Heterosurface. It is a well-established fact that many of the organic molecules are crystalline and they have well-defined external and internal structures. The internal structure deals with the arrangement of molecules in the crystal lattice, and this is called polymorphism.23 The second one is the external structure which reveals the morphology of the crystal, stating the shape of crystal without changing the

internal structure, and this is called morphological crystal engineering.1,2 It is also a well-known fact that the alteration of the external and internal structure could further affect the physicochemical stability of active pharmaceutical ingredients (APIs).24−28 This phenomenon of altering the internal and external structure with crystal engineering principles has gained huge attention in the scientific community in recent times due to their applications ranging from biochemistry 7055

DOI: 10.1021/acs.cgd.8b01214 Cryst. Growth Des. 2018, 18, 7048−7058

Crystal Growth & Design

Article

Figure 9. Powder, single crystal, film XRD comparison studies. (a) A change in relative peak intensities was observed from 2D plate crystals of BZA to needles. (b) ABZA commercial and crystals grown at low to high concentration comparison suggest the growth of the (110) face. LC = low concentration, HC = high concentration. (c) ABZA crystals and film on the surface confirmed the induced faces of (00l). (d, e) HBZA single crystals grown from solution and on surface and also film, which again confirmed the (00l) face and crystal morphology rod (with equal relative growth in 001, 011) to 2D plates with the (001) face was observed.

various morphologies. Here Model I resulted in BZA-like additive inducing crystallization, and Model II was observed successfully for ABZA and HBZA.

to optoelectronics. Consider a molecule as A which can undergo crystallization with n number of molecules. At first it forms a nucleus at birth, and then it will grow to form a single crystal. In addition, it can form polymorphs and is able to produce changes in morphology, and hence the displayed image Figure 10a depicts how the MOF surface can influence crystal formation. Normally if the crystal is a platelike shape, Model I, with two growth rates considered as K1 and K2 (K1 = K2) to result in faces f with surface S, it can adsorb a tailor-made additive or hetero surface at the active crystal face, and finally it will give new morphology, generally needles (K1 ≠ K2) Figure 10b. In addition, consider other models as Model II: if the crystal grows in a prismatic or acicular shape, the possible growth rates K1 = K2, K1 ≫ K2, K1 ≪ K2 end with three different shapes. The influenced growth can happen and would lead to block crystals, Figure 10 c.1,2,29 Models are modified from refs 1, 2, and 29. As our initial goal was to crystallize the organics on the designed hetero surface, here it is SURMOF to help us understand the influence of the crystal growth mechanism, and here we have proposed a model, Figure 10d, based on BZA examples. As shown in Figure 10b−d, the yellow line represents the MOF surface which can block the crystal faces at the surfaces and allow growth in other directions to lead to



CONCLUSIONS HKUST SURMOFs were used as heterosurfaces to study the morphological crystal engineering of the substituted BZAs, and the concept of MOF induced morphological crystal engineering was adapted successfully to the field of crystal engineering. In addition, it has shown the advantage of heterogeneous nucleation here with substituted BZAs and proved that the MOF surface exhibited adsorption of the substituted BZAs nucleus during both nucleation and growth directed due to the weak supramolecular interactions. Here in present study, the (001) face of BZA was inhibited in the b direction and allowed growth in the a direction to be observed. The second example ABZA resulted in needle crystals rather than prismatic or block morphology. Next to that, the third example HBZA showed interesting plate 2D morphology with the (002) crystal face as the major morphological importance, whereas in solution crystallization it was (001), (011), and (010) in near equal growth rate. Here we have observed for all the three examples successful morphological crystal engineering with the SURMOF heterogeneous 7056

DOI: 10.1021/acs.cgd.8b01214 Cryst. Growth Des. 2018, 18, 7048−7058

Crystal Growth & Design

Article

Figure 10. (a) An additive or heterosurface (present study SURMOF) can retard the growth of the α-phase to the affiliated α-type clusters (An)α and allow the (An)β nuclei to develop into macroscopic crystals {An}β of the metastable phase. (b) Growth mechanism type I, where crystal face modification from plates to needles is possible. (c) Growth mechanism type II, possible morphology alteration in various morphologies like needles, plates, prisms, pyramid type, and blocks. (d) Simple scheme for surface induced morphology with plate and rod morphology crystals. Surface of the porous MOF interacts with upcoming organics, and it can influence the nucleation and growth, which finally alter the morphology.



surface. However, still there is a future scope for a library of predesigned surfaces with different organic crystals.



(1) Addadi, L.; Berkovitch-Yellin, Z.; Weissbuch, I.; van Mil, J.; Shimon, L. J. W.; Lahav, M.; Leiserowitz, L. Growth and Dissolution of Organic Crystals with “Tailor-Made” Inhibitors- Implications in Stereochemistry and Materials Science. Angew. Chem., Int. Ed. Engl. 1985, 24, 466−485. (2) Addadi, L.; Weinstein, S.; Gati, E.; Weissbuch, I.; Lahav, M. Resolution of Conglomerates with the Assistance of Tailor-made Impurities. Generality and Mechanistic Aspects of the “Rule of Reversal”. A New Method for Assignment of Absolute Configuration. J. Am. Chem. Soc. 1982, 104, 4610−4617. (3) Weissbuch, I.; Popovitz-biro, R.; Lahav, M.; Leiserowitz, L.; Rehovot. Understanding and Control of Nucleation, Growth, Habit, Dissolution and Structure of Two and Three-Dimensional Crystals Using ’Tailor-Made’ Auxiliaries. Acta Crystallogr., Sect. B: Struct. Sci. 1995, 51, 115−148. (4) Desiraju, G. R. Crystal Engineering: From Molecule to Crystal. J. Am. Chem. Soc. 2013, 135, 9952−9967. (5) Desiraju, G. R. Crystal engineering. From molecules to materials. J. Mol. Struct. 2003, 656, 5−15. (6) Desiraju, G. R.; Vittal, J. J.; Ramanan, A. Crystal Engineering: A Text Book; World Scientific: Singapore, 2011. (7) Gur, D.; Palmer, B. A.; Weiner, S.; Addadi, L. Light Manipulation by Guanine Crystals in Organisms: Biogenic Scatterers, Mirrors, Multilayer Reflectors and Photonic Crystals. Adv. Funct. Mater. 2017, 27, 1603514. (8) Hirsch, A.; Palmer, B. A.; Elad, N.; Gur, D.; Weiner, S.; Addadi, L.; Kronik, L.; Leiserowitz, L. Biologically Controlled Morphology and Twinning in Guanine Crystals. Angew. Chem., Int. Ed. 2017, 56, 9420− 9424. (9) Palmer, B. A.; Hirsch, A.; Brumfeld, V.; Aflalo, E. D.; Pinkas, I.; Sagi, A.; Rosenne, S.; Oron, D.; Leiserowitz, L.; Kronik, L.; Weiner, S.; Addadi, L. Optically functional isoxanthopterin crystals in the mirrored eyes of decapod crustaceans. Proc. Natl. Acad. Sci. U. S. A. 2018, 115, 2299−2304. (10) Wang, J. L.; Berkovitch-yellin, Z.; Leiserowitz, L. Location of ‘Tailor-Made’ Additives in the Crystal and their Effect on Crystal Habit.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.cgd.8b01214. Material and methods; single crystal face indexing experiments; two-dimensional crystal faces exposed based on Materials Studio calculations; Table S1. Single crystal data of the ABZA; PXRD experiments (PDF) Accession Codes

CCDC 1854513 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_ [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



REFERENCES

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Geetha Bolla: 0000-0002-2726-8352 Allan S. Myerson: 0000-0002-7468-8093 Funding

G.B. and A.M. thank INDO-US for a fellowship and Novartis for their financial support. Notes

The authors declare no competing financial interest. 7057

DOI: 10.1021/acs.cgd.8b01214 Cryst. Growth Des. 2018, 18, 7048−7058

Crystal Growth & Design

Article

A Study on the Host-Additive System L-Asparagine-L-aspartic Acid Monohydrate. Acta Crystallogr., Sect. B: Struct. Sci. 1985, 41, 341−348. (11) Love, J. C.; Estroff, L. A.; Kriebel, J. K.; Nuzzo, R. G.; Whitesides, G. M. Self-Assembled Monolayers of Thiolates on Metals as a Form of Nanotechnology. Chem. Rev. 2005, 105, 1103−1169. (12) Yang, X.; Sarma, B.; Myerson, A. S. Polymorph Control of Micro/Nano-Sized Mefenamic Acid Crystals on Patterned SelfAssembled Monolayer Islands. Cryst. Growth Des. 2012, 12, 5521− 5528. (13) Kang, J. F.; Zaccaro, J.; Ulman, A.; Myerson, A. Nucleation and Growth of Glycine Crystals on Self-Assembled Monolayers on Gold. Langmuir 2000, 16, 3791−3796. (14) Lee, A. Y.; Lee, I. S.; Dette, S. S.; Boerner, J.; Myerson, A. S. Crystallization on Confined Engineered Surfaces: A Method to Control Crystal Size and Generate Different Polymorphs. J. Am. Chem. Soc. 2005, 127, 14982−14983. (15) Hiremath, R.; Basile, J. A.; Varney, S. W.; Swift, J. A. Controlling Molecular Crystal Polymorphism with Self-Assembled Monolayer Templates. J. Am. Chem. Soc. 2005, 127, 18321−18327. (16) Falcaro, P.; Ricco, R.; Doherty, C. M.; Liang, K.; Hill, A. J.; Styles, M. J. MOF positioning technology and device Fabrication. Chem. Soc. Rev. 2014, 43, 5513−5560. (17) Liu, J.; Wöll, C. Surface-supported metal−organic framework thin films: fabrication methods, applications, and challenges. Chem. Soc. Rev. 2017, 46, 5730−5770. (18) Shekhah, O.; Wang, H.; Kowarik, S.; Schreiber, F.; Paulus, M.; Tolan, M.; Sternemann, C.; Evers, F.; Zacher, D.; Fischer, R. A.; Wöll, C. Step-by-Step Route for the Synthesis of Metal−Organic Frameworks. J. Am. Chem. Soc. 2007, 129, 15118−15119. (19) Hermes, S.; Schroder, F.; Chelmowski, R.; Wöll, C.; Fischer, R. A. Selective Nucleation and Growth of Metal−Organic Open Framework Thin Films on Patterned COOH/CF3-Terminated Self-Assembled Monolayers on Au(111). J. Am. Chem. Soc. 2005, 127, 13744−13745. (20) Furukawa, H.; Cordova, K. E.; O’Keeffe, M.; Yaghi, O. M. The Chemistry and Applications of Metal-Organic Frameworks. Science 2013, 341, 1230444. (21) Arslan, H. K.; Shekhah, O.; Wohlgemuth, J.; Franzreb, M.; Fischer, R. A.; Wöll, C. High-Throughput Fabrication of Uniform and Homogenous MOF Coatings. Adv. Funct. Mater. 2011, 21, 4228−4231. (22) Bolla, G.; Myerson, A. S. SURMOF induced polymorphism and crystal morphological engineering of acetaminophen polymorphs: advantage of heterogeneous nucleation. CrystEngComm 2018, 20, 2084−2088. (23) McCrone, W. C. In Physics and Chemistry of the Organic Solid State; Fox, D., Labes, M. M., Weissberger, A., Eds.; Wiley Interscience: New York, 1965; Vol. 2, pp 725−767. (24) Bolla, G.; Nangia, A. Pharmaceutical cocrystals: walking the talk. Chem. Commun. 2016, 52, 8342−8360. (25) Modi, S. R.; Dantuluri, A. K. R.; Perumalla, S. R.; Sun, C. C.; Bansal, A. K. Effect of Crystal Habit on Intrinsic Dissolution Behavior of Celecoxib Due to Differential Wettability. Cryst. Growth Des. 2014, 14, 5283−5292. (26) Modi, S. R.; Dantuluri, A. K. R.; Puri, V.; Pawar, Y. B.; Nandekar, P.; Sangamwar, A. T.; Perumalla, S. R.; Sun, C. C.; Bansal, A. K. Impact of Crystal Habit on Biopharmaceutical Performance of Celecoxib. Cryst. Growth Des. 2013, 13, 2824−2832. (27) Duggirala, N. K.; Perry, M. L.; Almarsson, Ö .; Zaworotko, M. J. Pharmaceutical cocrystals: along the path to improved medicines. Chem. Commun. 2016, 52, 640−655. (28) Karanam, M.; Dev, S.; Choudhury, A. R. New Polymorphs of Fluconazole: Results from Cocrystallization Experiments. Cryst. Growth Des. 2012, 12, 240−252. (29) Addadi, L.; Berkovitch-Yellin, Z.; Domb, N.; Gati, E.; Lahav, M.; Leiserowitz, L. Resolution of conglomerates by stereoselective habit modifications. Nature 1982, 296, 21−26.

7058

DOI: 10.1021/acs.cgd.8b01214 Cryst. Growth Des. 2018, 18, 7048−7058