+ All documents
Home > Documents > Solvent-Induced Polymorphic Nanoscale Transitions for 12-Hydroxyoctadecanoic Acid Molecular Gels

Solvent-Induced Polymorphic Nanoscale Transitions for 12-Hydroxyoctadecanoic Acid Molecular Gels

Date post: 02-Dec-2023
Category:
Upload: independent
View: 0 times
Download: 0 times
Share this document with a friend
7
Solvent-Induced Polymorphic Nanoscale Transitions for 12-Hydroxyoctadecanoic Acid Molecular Gels Songwei Wu, ,Jie Gao, ,Thomas J. Emge, ,# and Michael A. Rogers* ,,School of Environmental and Biological Sciences, Department of Food Science, Rutgers University, The State University of New Jersey, New Brunswick, New Jersey 08901, United States School of Arts and Science, Department of Chemistry and Chemical Biology, Rutgers University, The State University of New Jersey, Piscataway, New Jersey 08854, United States * S Supporting Information ABSTRACT: 12-Hydroxyoctadecanoic acid (12HSA) molec- ular gels have been reported to form self-assembled brillar network (SAFiNs) in organic solvents. For the rst time, dierent polymorphic forms for 12HSA molecular gels have been reported. 12HSA, in alkanes and thiols, have a hexagonal subcell spacing (4.1 Å) and are arranged in a multilamellar fashion with a distance greater than the bimolecular length of 12HSA (54 Å). This polymorphic form corresponded to SAFiN with CGC less than 1 wt %. 12HSA, in nitriles, aldehydes, and ketones, have a triclinic parallel subcell (4.6, 3.9, and 3.8 Å) and interdigitation of the lamellar structure (3844 Å). This polymorphic form corresponds to a less eective sphereultic supramolecular crystalline network, which immobilizes solvents at CGC greater than 1.5 wt %. INTRODUCTION Molecular organogels are thermally reversible, quasi-solid materials comprised of an organic liquid (usually 95%) and a gelator molecule that self-assemble via physical interactions, including hydrogen-bonding, 14 ππ stacking, 5 dipoledipole, 6,7 and London dispersion forces, 8 into a three-dimensional network. 911 Although the physical interactions between gelator molecules are central in understanding gelation, the solventgelator speci c (i.e., H-bonding) and nonspecic (dipoledipole, dipole-induced, and instantaneous dipole induced forces) intermolecular interactions are equally important. 12,13 The process of self-assembly, in molecular gels, is an intricate process that must balance the solubility and those intermolecular forces that control epitaxial growth into axially symmetric elongated aggregates. 10,1316 During assembly, individual molecules are driven to assemble by molecular self-recognition and intermolecular noncovalent interactions into oligomers, and subsequently these oligomers assemble into brillar aggregates immobilizing the solvent via capillary forces. 17,18 Herein, we present an investigation of the rst polymorphic transformation, for a molecular gel, induced by modifying the solvent with 12HSA as the gelator. In molecular gels, polymorphic transitions have only been noted in (R)-18-(n- alkylamino)octadecan-7-ols in CCL4 which undergoes a gelgel polymorphic transition during heating. 9 Several other transitions have been reported in molecular gels; however dierences lie at the supramolecular level of structure induced by crystallographic mismatches and not dierent polymorphic forms. 2,3,12,15,1922 12HSA, a structurally simple, highly eective low molecular weight gelator (LMOG), has been studied extensively for gelation kinetics 2,2325 and supramolecular structure forma- tion, 3,19,20,2628 as well as to monitor surface properties, 29 Received: January 22, 2013 Published: February 5, 2013 Table 1. Critical Gelator Concentrations from ref 13 and Peak Melting Temperatures Determined in Triplicate Using Dierential Scanning Calorimetry solvent CCG (wt %) melting temperature (°C) hexane 0.4 61.9 ± 0.01 heptane 0.3 62.5 ± 0.1 octane 0.3 61.6 ± 0.5 nonane 0.25 60.5 ± 0.9 decane 0.2 63.9 ± 0.1 tetradecane 0.2 64.8 ± 0.35 1-pentanethiol 0.5 NA 1-hexanethiol 0.45 45.7 ± 0.01 1-heptanethiol 0.45 49.1 ± 0.03 1-octanethiol 0.4 50.8 ± 0.04 1-decanethiol 0.3 51.9 ± 0.06 butylnitrile 2.1 64.2 ± 0.3 hexanenitrile 1.9 65.5 ± 1.2 heptylnitrile 1.5 58.4 ± 1.0 nonanenitrile 0.9 NA butylaldehyde 2.8 NA dodecylaldehyde 1.4 41.2 ± 0.57 4-heptone 2 NA 5-nonanone 2.1 32.8 ± 0.8 6-undecanone 1.6 45.6 ± 1.5 Article pubs.acs.org/crystal © 2013 American Chemical Society 1360 dx.doi.org/10.1021/cg400124e | Cryst. Growth Des. 2013, 13, 13601366
Transcript

Solvent-Induced Polymorphic Nanoscale Transitionsfor 12-Hydroxyoctadecanoic Acid Molecular GelsSongwei Wu,†,⊥ Jie Gao,†,⊥ Thomas J. Emge,‡,# and Michael A. Rogers*,†,⊥

†School of Environmental and Biological Sciences, Department of Food Science, Rutgers University, The State University ofNew Jersey, New Brunswick, New Jersey 08901, United States‡School of Arts and Science, Department of Chemistry and Chemical Biology, Rutgers University, The State University ofNew Jersey, Piscataway, New Jersey 08854, United States

*S Supporting Information

ABSTRACT: 12-Hydroxyoctadecanoic acid (12HSA) molec-ular gels have been reported to form self-assembled fibrillarnetwork (SAFiNs) in organic solvents. For the first time,different polymorphic forms for 12HSA molecular gels havebeen reported. 12HSA, in alkanes and thiols, have a hexagonalsubcell spacing (∼4.1 Å) and are arranged in a multilamellarfashion with a distance greater than the bimolecular length of12HSA (∼54 Å). This polymorphic form corresponded toSAFiN with CGC less than 1 wt %. 12HSA, in nitriles,aldehydes, and ketones, have a triclinic parallel subcell (∼4.6,3.9, and 3.8 Å) and interdigitation of the lamellar structure (38−44 Å). This polymorphic form corresponds to a less effectivesphereultic supramolecular crystalline network, which immobilizes solvents at CGC greater than 1.5 wt %.

■ INTRODUCTION

Molecular organogels are thermally reversible, quasi-solid materialscomprised of an organic liquid (usually ≥95%) and a gelatormolecule that self-assemble via physical interactions, includinghydrogen-bonding,1−4 π−π stacking,5 dipole−dipole,6,7 andLondon dispersion forces,8 into a three-dimensional network.9−11

Although the physical interactions between gelator molecules arecentral in understanding gelation, the solvent−gelator specific (i.e.,H-bonding) and nonspecific (dipole−dipole, dipole-induced, andinstantaneous dipole induced forces) intermolecular interactionsare equally important.12,13 The process of self-assembly, inmolecular gels, is an intricate process that must balance thesolubility and those intermolecular forces that control epitaxialgrowth into axially symmetric elongated aggregates.10,13−16 Duringassembly, individual molecules are driven to assemble by molecularself-recognition and intermolecular noncovalent interactions intooligomers, and subsequently these oligomers assemble into fibrillaraggregates immobilizing the solvent via capillary forces.17,18

Herein, we present an investigation of the first polymorphictransformation, for a molecular gel, induced by modifying thesolvent with 12HSA as the gelator. In molecular gels,polymorphic transitions have only been noted in (R)-18-(n-alkylamino)octadecan-7-ols in CCL4 which undergoes a gel−gel polymorphic transition during heating.9 Several othertransitions have been reported in molecular gels; howeverdifferences lie at the supramolecular level of structure inducedby crystallographic mismatches and not different polymorphicforms.2,3,12,15,19−22

12HSA, a structurally simple, highly effective low molecularweight gelator (LMOG), has been studied extensively for

gelation kinetics2,23−25 and supramolecular structure forma-tion,3,19,20,26−28 as well as to monitor surface properties,29

Received: January 22, 2013Published: February 5, 2013

Table 1. Critical Gelator Concentrations from ref 13 andPeak Melting Temperatures Determined in Triplicate UsingDifferential Scanning Calorimetry

solvent CCG (wt %) melting temperature (°C)

hexane 0.4 61.9 ± 0.01heptane 0.3 62.5 ± 0.1octane 0.3 61.6 ± 0.5nonane 0.25 60.5 ± 0.9decane 0.2 63.9 ± 0.1tetradecane 0.2 64.8 ± 0.351-pentanethiol 0.5 NA1-hexanethiol 0.45 45.7 ± 0.011-heptanethiol 0.45 49.1 ± 0.031-octanethiol 0.4 50.8 ± 0.041-decanethiol 0.3 51.9 ± 0.06butylnitrile 2.1 64.2 ± 0.3hexanenitrile 1.9 65.5 ± 1.2heptylnitrile 1.5 58.4 ± 1.0nonanenitrile 0.9 NAbutylaldehyde 2.8 NAdodecylaldehyde 1.4 41.2 ± 0.574-heptone 2 NA5-nonanone 2.1 32.8 ± 0.86-undecanone 1.6 45.6 ± 1.5

Article

pubs.acs.org/crystal

© 2013 American Chemical Society 1360 dx.doi.org/10.1021/cg400124e | Cryst. Growth Des. 2013, 13, 1360−1366

solvent polarity,15,30 the influence of minor components,31 andeffects of chemical structure.30,32−36 Zhu and Dordick haveeloquently articulated that the heterogeneous nature of organogelsconsists of critical interactions between the solvent and gelator,and it is those interactions that govern the physical behavior andgelation process.12 Molecular assembly begins, in LMOGs, once acritical concentration is reached, below which the gelator is solubleand exists as monomers in solution.12 Once the criticalconcentration is reached, the gelation process begins leading toa decrease of gelator in the solution phase.12 Further addition ofgelator molecules causes the solvent to be immobilized but doesnot affect the soluble concentration of the gelator.12 Since gelatorsolubility is strongly dependent on the nature of the solvent, moststudies stop at correlating critical gelator concentration (CGC) tosolubility parameters.12,13 Typically, studies that examine the effectof solvent often focus on solubility parameters because excellentcorrelations between solubility and CGC exist and nuisances inmolecular assembly affecting the CCG remain relatively unknown.In this current work, we examine 12HSA in several solvents tostudy the nanostructure and microstructure of the molecular gels.

■ METHODSSolvent selection criteria were maintained as simply as possible: thealiphatic chain being linear, saturated, the functional group located in

the primary position, and the solvent must be in a liquid state between10 and 30 °C. The only exceptions to these selection criteria were theketones where the functional group was located in the exact middle ofthe molecule. Apolar solvents included the aliphatics, and polarsolvents were subdivided into three categories: aldehydes, ketones, andnitriles; as well, the solvents capable of hydrogen bondingand forminga molecular gel were only the thiols. A complete list of the employedsolvents and their solubility parameters may be found in Gao et al.13

All solvents and R-12-hydroxystearic acid (12HSA) were obtainedfrom Sigma-Aldrich (Cherry Hill, NJ, USA) with a purity greaterthan 95%.

The supramolecular structure was imaged using a Linkham imagingstation (Linkham, Surrey, England) equipped with a Q imagining 2560 ×1920 pixel CCD camera (Micropublisher, Surrey, Canada) and a 10×Olympus lens (0.25 N.A.) (Olympus, Tokyo, Japan). Samples wereplaced on a glass slide with a coverslip on top of the sample. The slide wastransferred into a peltier temperature control stage (LTS120, Linkham,Surrey, England) and heated to 80 °C and was slowly cooled (2 °C/minto 20 °C) to observe supramolecular network formation.

The X-ray diffraction (XRD) or wide-angle X-ray scattering(WAXS) patterns of several samples containing 12HSA gel in differentsolvents were obtained by use of a Bruker HiStar area detector and anEnraf-Nonius FR571 rotating anode X-ray generator equipped withRigaku Osmic mirror optic system (∼0.06 deg 2θ nominal dispersionfor Cu Kα; l = 1.5418 Å) operating at 40 kV and 40 mA. All of the datawere collected at room temperature over a period of about 300 s. Thesample-to-detector distance was 10.0 cm, and the standard spatialcalibration was performed at that distance. Scans were 4 deg wide inomega (ω) with fixed detector, or Bragg, angle (2θ) of 0 deg, and fixedplatform (f and c) angles of 0 and 45 deg, respectively. In all cases, thecount rate for the area detector did not exceed 100 000 cps.

The carbonyl (∼1700 cm−1) and hydroxyl (∼3200 cm−1) signalswere measured using a Thermo Nicolet FT-IR and an attenuated totalrefraction (ATR) prism (Thermo Fisher Scientific, MA, USA). 256scans were collected at a resolution of 4 cm−1

Figure 1. Brightfield microscope images of molecular gels of 2 wt %12-hydroxyoctadecanoic acid in hexane (A) and dodecane (B). Widthof micrograph is 120 μm.

Figure 2. Brightfield microscope images of molecular gels of 2 wt %12-hydroxyoctadecanoic acid in acetonitrile (A) and octanenitrile (B),butylaldehyde (C), decanal (D), heptone (E), and undecanone (F).Width of micrograph is 120 μm.

Crystal Growth & Design Article

dx.doi.org/10.1021/cg400124e | Cryst. Growth Des. 2013, 13, 1360−13661361

■ DISCUSSIONRecent work found that the CGC of 12HSA is stronglyinfluenced by solvent parameters varying between 0.2 wt % tonearly 3.0 wt % (Table 1).13 Solvent parameters influence thesupramolecular structure of gelators, where the interactionbetween the solvent and gelator induce changes in thethickness, the number of junction zones, and the sense ofhelical twist.12,18 For symmetrical trehalose diesters, a class of

super gelators, the gelation number varies between 12000 and100 depending on the solvent.12 In this work, Zhu and Dordickillustrate that the fiber thickness and morphology influences thegelation number, which correlates to the CCG.12 In the solventsselected for this study with 12-HSA as the molecular gelator, notonly did the CGC vary but also the melting temperature of themolecular gels from 65 to 32 °C (Table 1). The melting tem-peratures were determined from the peak of the differential scanningcalorimety thermograms (Supplemental Figures 1−3, SupportingInformation). Melting temperatures could not be obtained for allsamples because several solvents were too volatile. Previously, it hasbeen established that the melting temperature of the molecular gel isdependent on the supramolecular structure of the fibers, wherethinner fibers melt at lower temperatures than thicker fibers; in otherwords, as the crystalline perfection increases so does the meltingpoint.28,30 However, these changes in melting points are relativelyminor compared to the changes we observed in this study.On a supramolecular level, the fibrillar structure of 12HSA

molecular gels varies depending on the solvent (Figures 1−3).For 12HSA molecular gels in alkanes (Figure 1A,B) the aspectratio of the nanofibers is very large. The large aspect ratio andthin fiber morphology result in a high surface area and which inturn allows for very low CGC and high gelation numbers. Asthe alkane chain length increases, the CGC decreases and thefibers length increases (Figure 1). In select thiols, capable offorming weak hydrogen bonds, the cohesive energy densityincreases (varying between 17 to 18 MPa1/2) compared toalkanes (varying between 15 and 16 MPa1/2) due to the weakhydrogen bonding and polar component.13,37 The supra-molecular structure is affected by the change in solubilityparameters to produce shorter fibers, which are thinner than inthe alkanes (Figure 3A,B). Although there is a drastic change inthe supramolecular structure, the overall fiber morphology isstill very effective at entraining the solvent forming a molecular gelat low CGCs (Table 1). In polar solvents (nitriles (Figure 2A,B);aldehydes (Figure 2C,D); ketones (Figure 2E,F)), the supra-molecular structure varies greatly compared to the alkanesand thiols. The cohesive energy density is higher in nitriles(17−25 MPa1/2), aldehydes (17−18 MPa1/2), and ketones(18−20 MPa1/2) compared to the alkanes and thiols. Thesupramolecular structure varies significantly in the differentclasses of polar solvents (Figure 3). Although all of the solventsstill form molecular gels, the CGC is considerably higher,ranging between 0.9 and 2.8 wt % and the opacity of the gels

Figure 3. Brightfield microscope images of molecular gels of 2 wt %12-hydroxyoctadecanoic acid in pentanethiol (A) and decanethiol (B).Width of micrograph is 120 μm.

Figure 4. Vertically offset wide-angle (A) and short-angle subcell spacings (B) for 2 wt % 12HSA in aliphatic solvents. The diffractrograms frombottom to the top are hexane, heptane, octane, nonane, decane, dodecane, tetradecane.

Crystal Growth & Design Article

dx.doi.org/10.1021/cg400124e | Cryst. Growth Des. 2013, 13, 1360−13661362

increases. The supramolecular structure also changes from highaspect ratio fibers to spherultic-like aggregates. In acetonitrile(Figure 3A), radial growth from central nuclei occurs untiladjacent crystals impede their growth. As the chain length ofthe nitriles increases, the supramolecular crystal structurebecomes thinner lowering the CGC (Figure 3B). 12HSA moleculargels in aldehydes (Figure 3C,D) and ketones (Figure 3E,F) havethe highest CGC and opacity.13 The supramolecular crystallinestructure does not effectively entrain the solvent due to the

colloidal crystal structure. Although there are some fiber-likestructures, the supramolecular network does not resemble what isobserved in either alkanes (Figure 1) or thiols (Figure 2). At first, itwas assumed that the solvent parameters varied the supramolecularstructures; however, this did not explain the differences in themelting temperatures. This compelled us to examine the possibilitythat there may be differences at the nano level of structure,meaning different polymorphic forms may exist within moleculargels of 12HSA.

Figure 5. Vertically offset FT-IR spectra focusing on the carboxylic acid region (A) and hydroxyl region (B), using the air as the background. Thediffractrograms from bottom to the top are hexane, heptane, octane, nonane, decane, dodecane, tetradecane.

Figure 6. Vertically offset wide-angle (A) and short-angle subcell spacings (B) for 2 wt % 12HSA in thiol-based solvents. The diffractrograms frombottom to the top are hexanethiol, hexanethiol, and decanethiol.

Figure 7. Vertically offset FT-IR spectra focusing on the carboxylic acid region (A) and hydroxyl region (B), using the air as the background. Thediffractrograms from bottom to the top are hexanethiol, hexanethiol, and decanethiol.

Crystal Growth & Design Article

dx.doi.org/10.1021/cg400124e | Cryst. Growth Des. 2013, 13, 1360−13661363

With drastic differences in the microstructure, the nano-structure was probed to study if there was a correlation with thecrystal morphology. XRD was used to measure the wide- andshort-angle spacings of the molecular gels. The X-raydiffractograms for all alkanes had very similar diffractionpatterns with a short-angle spacing of 53.5 Å and an hlk higherorder reflection at 15.2 Å (Figure 4A). The presence of thehigher order reflection indicates a multi-lamellar crystalconfiguration. The lamellar thickness of 12HSA shouldcorrelate to approximately twice the extended molecular lengthof 12HSA (∼46 Å).30 In alkanes it appears that a swollenlamellar network is present, which has been previouslyreported.26 The wide-angle spacing at 4.3 Å corresponds tohexagonal (∼4.1 Å) subcell spacings (Figure 4B).38,39

Observing the FT-IR spectrogram, the peak corresponding tothe carboxyl acid dimerization at 1690 cm−1 was observed,which is typical for hydroxyoctadecanoic acid molecular gels(Figure 5A).2,40,41 The peak corresponding to the hydroxylhydrogen bonding, appearing as a broad peak at 3200 cm−1,indicates that the hydroxyl groups at position 12 are involved inhydrogen bonds (Figure 5B).Although the network structure was finer in molecular gels

containing thiols than was observed in alkanes, the entrainment ofthe solvent by the fibrillar network was similar (Figures 1 and 2).In selected thiols, similar short-angle spacings (Figure 6A) andwide-angle spacings (Figure 6B) are observed indicating ahexagonal spacing with a multilamellar crystal lattice equal to anextended bimolecular length. The FT-IR spectrograms indicate acyclic dimer between the carboxylic acid head groups (Figure 7A),and the broad weak peak at 3200 cm−1 indicates that there is anabsence of free hydroxyl groups (Figure 7B).In polar solvents, including nitriles, aldehydes, and ketones

(Figure 3), the CGC (CGC > 1.5 wt %) drastically increasedcompared to the thiols and alkanes (CCG < 1.0 wt %) (Table 1).This corresponds to changes in the supramolecular structuresobserved in the aforementioned solvents (Figures 1−3). In polarsolvents (which have higher cohesive energy densities), the fibrillarcrystal morphology is replaced with a spherultic-like structure that

is less effective at entraining solvents. On a nanoscale level,considerable changes in the molecular arrangement occurred. Theshort-angle spacings indicate a decrease of the lamellar spacingfrom ∼53 Å to 44 Å for nitriles (Figure 8A) and ketones (Figure 8E)and ∼39 Å for aldehydes (Figure 8C). These changes are alsoreflected in the hlk higher order reflections. This reduction in thelamellar spacing is near the minimum of the bimolecular lengthfor 12HSA (∼46 Å) in nitriles and ketones and is significantlyshorter for the aldehydes suggesting that there is an inter-digitation of 12HSA in the multilamellar structures. With thechange in the short-angle spacing, a polymorphic transition isalso observed in the subcell spacing. The hexagonal subcellspacing observed in the fibrillar aggregates (∼4.1 Å) gives wayto a triclinic parallel subcell (strong peak at 4.6 Å, and two weakpeaks at 3.9 Å and 3.8 Å) in nitriles, aldehydes, and ketones(Figure 8A,C,D). In the aldehydes, only the 4.6 and 3.8 Å peakis visible, but the presence of the peak at 3.8 Å indicates that itis unlikely to be in the hexagonal polymorphic form.

Figure 8. Vertically offset wide-angle (A, C, E) and short-angle subcellspacings (B, D, F) for 2 wt % 12HSA in polar solvents. Thediffractrograms from bottom to the top are (A, B) butylnitrile,hexanenitrile, heptylnitrile, and nonanenitrile; (C, D) butylaldehydeand dodecylaldehyde; (E, F) heptone, nonanone, and undecone.

Figure 9. Vertically offset FT-IR spectra focusing on the carboxylicacid region (A) and hydroxyl region (B), using the air as thebackground. The diffractrograms from bottom to the top are (A, B)butylnitrile, hexanenitrile, heptylnitrile, and nonanenitrile; (C, D)butylaldehyde and dodecylaldehyde; (E, F) heptone, nonanone, andundecone.

Figure 10. Schematic diagram of two polymorphic forms (hexagonaland triclinic parallel) of 12HSA in various solvents capable of formingmolecular gels. The box represents the side view of the subcell that ispictured next to the molecular structures.

Crystal Growth & Design Article

dx.doi.org/10.1021/cg400124e | Cryst. Growth Des. 2013, 13, 1360−13661364

Although the wide and short-angle spacings are similar fornitriles, aldehydes, and ketones, there are differences in relationto the noncovalent interactions between the 12HSA molecules(Figure 9). FT-IR shows that 12HSA in nitriles formscarboxylic acid dimers (1690 cm−1) (Figure 9A). A broadpeak is observed at 3600 cm−1 (Figure 9B), indicating thathydroxyl groups, at position 12, form hydrogen bonds. Unlikethe nitriles, the carboxylic acid groups in aldehydes (Figure 9C)and ketones (Figure 9E) do not form cyclic dimers (∼1720 cm−1).This in part would explain why 12HSA molecule gels in aldehydesand ketones melt at temperatures much less than the other threesets of selected solvents. As well, the broad peak between 3200and 3600 cm−1 has the diffuse pattern, suggesting that at leastsome of the hydroxyl groups are involved in hydrogen bonds.However, a new small peak is superimposed on the broad peakindicating there is substantial free hydroxyl groups in the moleculargels with aldehydes and ketones used as solvents. For the firsttime, two polymorphic forms (i.e., a hexagonal and triclinicparallel) are present in molecular gels (Figure 10), which result inchanges at the microscopic and macroscopic levels of structure.

■ CONCLUSIONS12HSA is one of the most studied molecular gelators, in partdue to its molecular simplicity, inexpensive cost, and versatilityin gelling numerous solvents. With so many studies on 12HSAin molecular gels, it is amazing that no studies on poly-morphism have been reported. In alkanes and thiols, ahexagonal subcell spacing (∼4.1 Å) and a multilamellar crystalmorphology with a distance between lamella greater than thebimolecular length of 12HSA was observed. This polymorphicform corresponded to molecular gels with CGC less than 1 wt% and a supramolecular self-assembled fibrillar network. 12HSAmolecular gels in nitriles, aldehydes, and ketones have a triclinicparallel subcell (∼ 4.6, 3.9, and 3.8 Å) and interdigitation inthe lamellar. This polymorphic form is far less effective atimmobilizing solvents with CGC greater than 1.5 wt % and asphereultic supramolecular crystalline network.

■ ASSOCIATED CONTENT*S Supporting InformationDifferential scanning thermograms of the melting profiles aresupplied. This material is available free of charge via theInternet at http://pubs.acs.org.

■ AUTHOR INFORMATIONCorresponding Author*E-mail: [email protected].

Author Contributions⊥These authors contributed equally.

Author Contributions#Provided X-ray analysis.

NotesThe authors declare no competing financial interest.

■ ACKNOWLEDGMENTSThis project was supported by the HATCH Program throughthe New Jersey Agriculture Research Station (NJAES).

■ ABBREVIATIONS12HSA, 12-hydroxyoctadecanoic acid

■ REFERENCES(1) Kuwahara, T.; et al. Crystal structure of DL-12-hydroxystearicacid. Chem. Lett. 1996, 25, 435−436.(2) Lam, R.; et al. A molecular insight into the nature ofcrystallographic mismatches in self-assembled fibrillar networksunder non-isothermal crystallization conditions. Soft Matter 2010, 6(2), 404−408.(3) Li, J. L.; et al. Architecture of a Biocompatible SupramolecularMaterial by Supersaturation-Driven Fabrication of its Network. J. Phys.Chem. B 2005, 109, 24231−24235.(4) Rogers, M. A.; et al. Multicomponent Hollow Tubules FormedUsing Phytosterol and γ-Oryzanol-Based Compounds: An Under-standing of Their Molecular Embrace. J. Phys. Chem. 2010, 114, 8278−8295.(5) Moffat, J. R.; Smith, D. K. Controlled Self-Sorting in theAssembly of ‘Multi-Gelator’ Gels. Chem. Commun. 2009, 3, 316−318.(6) Brotin, T.; Devergne, J. P.; Fages, F. PhotostationaryFluorescence Emission and Time Resoved Spectroscopy of Symmetri-cally Disubstituted Anthracenes on the Meso and Side Rings: TheUnusual Behavior of the 1,4 Derivative. Photochem. Photobiol. 1992,55, 349−358.(7) Terech, P.; Furman, I.; Weiss, R. G. Structures of OrganogelsBased Upon Cholesteryl 4-(2-Anthryloxy)Butanoate, A HighlyEfficient Luminescing Gelator - Neutron and X-Ray Small-AngleScattering Investigations. J. Phys. Chem. 1995, 99 (23), 9558−9566.(8) Toro-Vazquez, J. F.; et al. Relationship Between MolecularStructure and Thermo-mechanical Properties of Candelilla Wax andAmides Derived from (R)-12-Hydroxystearic Acid as Gelators ofSafflower Oil. Food Biophysics 2010, 5 (3), 193−202.(9) Mallia, V. A.; et al. Reversible Phase Transitions within Self-Assembled Fibrillar Networks of (R)-18-(n-Alkylamino)octadecan-7-ols in Their Carbon Tetrachloride Gels. J. Am. Chem. Soc. 2011, 133,15045−15054.(10) Weiss, R. G., Terech, P. Introduction. In Molecular Gels:Materials with Self-Assembled Fibrillar Networks; Weiss, R. G., Terech,P., Eds.; Springer: Dordrecht, The Netherlands, 2006; p 1−13.(11) George, M.; Weiss, R. G. Molecular Organogels. Soft MatterComprised of Low-Molecular-Mass Organic Gelators and OrganicLiquids. Acc. Chem. Res. 2006, 39, 489−497.(12) Zhu, G.; Dordick, J. S. Solvent Effect on Organogel Formationby Low Molecular Weight Molecules. Chem. Mater. 2006, 18, 5988−5995.(13) Gao, J.; Wu, S.; Rogers, M. A. Harnessing Hansen SolubilityParameters to Predict Organogel Formation. J. Mater. Chem. 2012, 22,12651−12658.(14) Suzuki, M.; et al. Effects of hydrogen bonding and van der Waalsinteractions on organogelation using designed low-molecular-weightgelators and gel formation at room temperature. Langmuir 2003, 19(21), 8622−8624.(15) Rogers, M. A.; Marangoni, A. G. Solvent-Modulated Nucleationand Crystallization Kinetics of 12-Hydroxystearic Acid: A Non-isothermal Approach. Langmuir 2009, 25 (15), 8556−8566.(16) Raynal, M.; Bouteiller, L. Organogel Formation Rationalized byHansen Solubility Parameters. Chem. Commun. 2011, 47, 8271−8273.(17) Hardy, J. G.; Hirst, A. R.; Smith, D. K. Exploring molecularrecognition pathways in one- and two-component gels formed bydendritic lysine-based gelators. Soft Matter 2012, 8, 3399−3406.(18) Edwards, W.; Lagadec, C. A.; Smith, D. K. Solvent-gelatorinteractions–using empirical solvent parameters to better understandthe self-assembly of gel-phase materials. Soft Matter 2011, 7, 110−117.(19) Wang, R. Y.; et al. Architecture of Fiber Network: FromUnderstanding to Engineering of Molecular Gels. J. Phys. Chem. B2006, 10, 25797−25802.(20) Li, J. L.; et al. Nanoengineering of a Biocompatible Organogelby Thermal Processing. J. Phys. Chem. B 2009, 113 (15), 5011−5015.(21) Liu, X. Y.; Sawant, P. D. Determination of the FractalCharacteristic of Nanofiber-Network Formation in SupramolecularMaterials. ChemPhysChem 2002, 4, 374−377.

Crystal Growth & Design Article

dx.doi.org/10.1021/cg400124e | Cryst. Growth Des. 2013, 13, 1360−13661365

(22) Huang, X.; et al. Distinct Kinetic Pathways Generate OrganogelNetworks with Contrasting Fractality and Thixotopic Properties. J.Am. Chem. Soc. 2006, 128, 15341−15352.(23) Terech, P. Kinetics of Aggregation in a Steroid Derivative/Cyclohexane Gelifying System. J. Colloid Interface Sci. 1985, 107 (1),244−255.(24) Rogers, M. A.; Marangoni, A. G. Non-Isothermal Nucleationand Crystallization of 12-Hydroxystearic Acid in Vegetable Oils. Cryst.Growth Des. 2008, 8 (12), 4596−4601.(25) Lam, R. S. H.; Rogers, M. A. Experimental Validation of theModified Avrami Model for Non-Isothermal Crystallization Con-ditions. Cryst. Eng. Commun. 2010, 13, 866−875.(26) Rogers, M. A.; Wright, A. J.; Marangoni, A. G. Engineering theoil binding capacity and crystallinity of self-assembled fibrillarnetworks of 12-hydroxystearic acid in edible oils. Soft Matter 2008, 4(7), 1483−1490.(27) Rogers, M. A.; Wright, A. J.; Marangoni, A. G. Nanostructuringfiber morphology and solvent inclusions in 12-hydroxystearic acid/canola oil organogels. Curr. Opin. Colloid Interface Sci. 2009, 14 (3),223−223.(28) Rogers, M. A.; Wright, A. J.; Marangoni, A. G. Nanostructuringfiber morphology and solvent inclusions in 12-hydroxystearic acid/canola oil organogels. Curr. Opin. Colloid Interface Sci. 2009, 14 (1),33−42.(29) Menger, F. M.; et al. Chain-Substituted Lipids in Mono-molecular Films. Effect of Polar Substituents on Molecular Packing.Langmuir 1989, 5, 833−838.(30) Mallia, V. A.; et al. Robust Organogels from Nitrogen-Containing Derivatives of (R)-12-Hydroxystearic Acid as Gelators:Comparisons with Gels from Stearic Acid Derivatives. Langmuir 2009,25 (15), 8615−8625.(31) Fameau, A. L.; et al. 12-Hydroxystearic acid lipid tubes undervarious experimental conditions. J. Colloid Interface Sci. 2010, 341 (1),38−47.(32) Grahame, D. A. S.; et al. Influence of Chirality on the Modes ofSelf-Assembly of 12-Hydroxystearic Acid in Molecular Gels of MineralOil. Soft Matter 2011, 7, 7359−7365.(33) Huang, X.; Weiss, R. G. Molecular organogels of the sodium saltof (R)-12-hydroxystearic acid and their templated syntheses ofinorganic oxides. Tetrahedron 2007, 63 (31), 7375−7385.(34) Sakurai, T.; et al. A Comparative Study on Chiral and Racemic12-Hydroxyoctadecanoic Acids in the Slutions and Aggregation States:Does the Racemic Form Really Form a Gel? Bull. Chem. Soc. Jpn. 2010,83 (2), 145−149.(35) Terech, P.; et al. Organogels and Areogels of Racemic andChiral 12-Hydroxyoctadecanoic Acid. Langmuir 1994, 10 (10), 3406−3418.(36) Wright, A. J.; Marangoni, A. G. Time, temperature, andconcentration dependence of ricinelaidic acid-canola oil organo-gelation. J. Am. Oil Chem. Soc. 2007, 84 (1), 3−9.(37) Hansen, C. M. Hansen Solubility Parameters, 2nd ed.; CRCPress: Boca Raton, FL, 2007.(38) Marangoni, A. G. Crystallography, in Fat Crystal Networks;Marangoni, A. G., Ed.; Marcel Dekker: New York, 2005; p 1−20.(39) Sato, H.; Ueno, S. Polymorphism in Fats and Oils, in Bailey’sIndustrial Oil and Fat Products; Shahidi, F., Ed.; John Wiley & Sons,Inc: New York, 2005; p 77−120.(40) Lin-Vien, D., et al., The Handbook of Infrared and RamanCharacteristic Frequencies of Organic Molecules; London: AcademicPress, 1991.(41) Rogers, M. A.; Pedersen, T.; Quaroni, L. Hydrogen-BondingDensity of Supramolecular Self-Assembled Fibrillar Networks ProbedUsing Synchrotron Infrared Spectromicroscopy. Cryst. Growth Des.2009, 9 (8), 3621−3625.

Crystal Growth & Design Article

dx.doi.org/10.1021/cg400124e | Cryst. Growth Des. 2013, 13, 1360−13661366


Recommended