Seed-Assisted Effects on Solution-Mediated Phase Transformation: A

Dec 18, 2017 - However, when the nucleation step plays an important role in the subsequent transformation process, we question if the transformation r...
0 downloads 10 Views 4MB Size
Article Cite This: Ind. Eng. Chem. Res. 2018, 57, 784−793

pubs.acs.org/IECR

Seed-Assisted Effects on Solution-Mediated Phase Transformation: A Case Study of L‑Histidine in Antisolvent Crystallization Ya Liu,†,§ Shijie Xu,†,§ Yumin Liu,†,§ Mingyang Chen,†,§ Yifu Chen,†,§ Shichao Du,†,§ Yaping Wang,†,§ Panpan Sun,†,§ Mengmeng Sun,†,§ Peng Shi,†,§ and Junbo Gong*,†,‡,§ †

School of Chemical Engineering and Technology, State Key Laboratory of Chemical Engineering and ‡Tianjin Key Laboratory for Modern Drug Delivery and High-Efficiency, School of Pharmaceutical Science and Technology, Tianjin University, Tianjin 300072, China § The Co-Innovation Center of Chemistry and Chemical Engineering of Tianjin, Tianjin 300072, China S Supporting Information *

ABSTRACT: In this study, the effects of crystal nucleation− growth on subsequent solvent-mediated phase transformation experiments of L-histidine (L-his) in antisolvent crystallization were performed at 298.15 K. The unexpected acceleration of the overall transformation rate in antisolvent crystallization with solution-mediated phase transformation was found for a methanol volume fraction of 0.35 < x2 < 0.60. Interestingly, concomitant polymorphs were obtained for a methanol volume fraction of 0.30 ⩽ x2 ⩽ 0.65 in antisolvent crystallization, whereas only form B was observed for other volume fractions of methanol, which indicated that the concomitant polymorphic phenomenon was the main reason for the accelerated transformation rate in antisolvent crystallization. Furthermore, seed-assisted experiments and suspended solution-mediated phase transformation experiments were designed to uncover the role of the nucleation of form A accompanying with form B in the subsequent transformation process.

1. INTRODUCTION The molecular arrangements in crystals during nucleation and growth frequently display various solid-state forms with different chemical and physical properties (e.g., bioavailability, morphology, purity, and so on).1,2 At certain temperatures (T) and pressures (P), there is only one thermodynamically stable form, whereas other forms are metastable forms that could transform into the most stable form via solid-state polymorphic transformation or solution-mediated polymorphic transformation.3−5 In industry, it is extremely important to precisely control crystallization of polymorphs to obtain highly functional crystals.6,7 Solvent-mediated polymorphic transformation is an efficient technique to obtain the most stable polymorph. It is also useful to examine the relative stability of polymorphs and eliminate the less stable polymorph to ensure the phase purity.8 For these purposes, many researchers studied the effect of numerous parameters, such as solvent, temperature, additive, solid loading and scale of operation on the kinetic and mechanism of solvent-mediated polymorphic transformation.4,6,9−13 In addition, the overall phase transformation of polymorphic forms can be governed by either the dissolution of the metastable phase or the nucleation and growth of the stable phase.14,15 It is well-known that crystal nucleation is the start of phase transformation in crystallization and has the utmost impact on product quality aspects such as crystal size and polymorphic form.16 In addition, a fundamental understanding © 2017 American Chemical Society

of nucleation is as well needed for the rigorous control and prediction of the crystalline product quality of any crystallization processes on an industrial scale.17 However, there are few reports about investigation of the phase transformation in the antisolvent crystallization process, which tends to be even more complicated because of the prior nucleation process. The model compound used in this study is L-histidine (Abbreviated as L-his, Figure 1), which is one of the most essential and naturally occurring amino acids.18 It often serves as a precursor for many chemicals and is extensively used in many fields, such as food, pharmaceutical, and feed industries. Moreover, it has an important implication upon several biological activities, e.g., transmission of metal elements in

Figure 1. Chemical structure of L-his. Received: Revised: Accepted: Published: 784

November 17, 2017 December 17, 2017 December 18, 2017 December 18, 2017 DOI: 10.1021/acs.iecr.7b04762 Ind. Eng. Chem. Res. 2018, 57, 784−793

Article

Industrial & Engineering Chemistry Research

2.3. Establishment of PXRD Calibration Curve of Form A and Form B of L-his for Quantitative Analysis. PXRD can be used to determine the relative weight percentages of the crystalline phases in the quantitative analysis of a binary polymorphic mixture. It is known that PXRD diffractions can be affected by various factors such as type of sample holder, rotation of the sample, powder packing, and preferred orientation effects.28−30 Different amounts of form A and form B were weighed out to a total mass of 0.0347 g. To avoid mixing inhomogeneity, the standard samples were mixed gently for 2 min using an agate mortar and pestle before PXRD acquisition. The standard samples were then placed on a zero background holder and scanned from 5° to 35° (2θ) at a step size of 0.02° and a scanning rate of 8°/min. The peak intensity of form A at 17.75° (2θ) and that of form B at 17.25° (2θ) were used, respectively. According to the change of the corresponding characteristic peak intensity, the relative mass fraction of each form in their binary mixtures can be calculated by using eq 1.

biological bases and neurotransmitting or neuromodulating in the mammalian central nervous system.19,20 L-his is known to crystallize in two forms: A and B, with form A being stable.20−27 Madden et al. first reported the crystal structures of form A and form B in 1972.24,25 Kitamura et al. studied the formation and transformation of polymorphs of Lhis by rapidly cooling ethanol and water mixtures and an aqueous solution. It was concluded that both form A and form B arose simultaneously because nucleation and growth rates of the two forms were approximately equal.21,22 Roelands et al. studied the formation of polymorphs of L-his as a function of supersaturation ratio and interfacial energy in ethanol antisolvent crystallization. They suggested that the polymorphic fraction of form B increased with increasing supersaturation ratio.23 Wantha et al. determined the growth rate of form A and the dissolution rate of form B of L-his in aqueous solution.26 Moreover, the nucleation and growth mechanisms of form B of 27 L-his in the water−ethanol system were also determined. Besides, Roelands et al. and Kitamura et al. measured the transformation rate of form B into form A as a function of ethanol fraction, x2 = 0.2, 0.3, and 0.4. It can be concluded that the transformation was more rapid with decreasing ethanol volume fractions.22,23 However, when the nucleation step plays an important role in the subsequent transformation process, we question if the transformation rate always changes monotonously with the ratio of the antisolvent? In this work, we report the results of our studies on the formation of polymorphs and the solventmediated transformation rate as a function of the methanol fraction in the L-his−water system. Furthermore, two kinds of suspended polymorphic transformations, including adding form B to a binary water + methanol saturated solution and introducing form A into a binary water + methanol saturated solution, were designed to investigate the effects of the formation of polymorphs on subsequent the solvent-mediated transformation process. A polarizing microscope and the PXRD characterization were used to observe the whole polymorphic transformation process.

xA =

IA IA + IB

(1)

where xA is the mass fraction of form A, IA and IB represent the intensity of characteristic diffraction peaks of form A and form B, respectively. 2.4. Solubility Measurements. A gravimetric method was employed to determine the solubility of both form A and form B of L-his in methanol + water binary solvent mixtures at 298.15 K. Approximately 25−30 mL of the binary solvent mixtures with volume fraction being known was transferred into a 50 mL conical flask, whose temperature was controlled by an air bath shaker with accuracy u(T) = 0.1 K (type HNY200R, Tianjin Honor Instrument Co., Ltd.). When the actual temperature of binary solvent mixtures was the same as the setting temperature, an excessive amount of both forms was added into the thermostatic solvent mixture, respectively. The solution was stirred at a speed of 260 r/min for at least 11 h to reach the solid−liquid equilibrium. Then the agitation was stopped, and the solution system was kept still for more than 1 h under the original temperature to allow the undissolved particles to settle down completely before sampling. A preheated or precooled organic membrane (0.45 μm) was used to filter the upper clear solution. And the filtered liquor was transferred into a drying Petri dish, which was weighed in advance. After that, the Petri dish with filter liquor was weighed immediately and put into a vacuum drying oven at 323.15 K for about 50 h. Then the Petri dish with dried solute was reweighed several times until the mass of glass dish did not change. Meanwhile, the residual undissolved crystals in the conical flasks were separated and the forms of the solids were confirmed by PXRD. During the experiments, all masses were determined by using an analytic balance (ML204/02, MettlerToledo) with standard uncertainty u(B) = 0.0001 g. To minimize experimental errors, each solubility measurement was repeated three times, and the mean value was used as the final result. According to the measured data, the mass fraction solubility (x1) of L-his in mixed solvents can be calculated by using eq 2, and the volume fraction of methanol (x2) in the binary solvent mixtures is calculated by eq 3: m x1 = 1 ms (2)

2. EXPERIMENTAL SECTION 2.1. Materials. Form A of L-his with stated mass fraction purity higher than 0.99, was purchased from Tokyo Chemical Industry Co., Ltd. Organic solvents (methanol and ethanol) were purchased from Tianjin Jiangtian Chemical Reagent Co., Ltd. The solvents were analytical grade with mass fraction purity higher than 0.995. All materials above were used without any further treatment. Form B crystals of L-his were obtained by rapid antisolvent crystallization at 298.15 K as described in the previous work.21−23 A saturated aqueous solution of L-his was prepared and then the pure ethanol was added rapidly to reach a volume fraction x2 = 0.5 for producing the form B. Deionized water (conductivity 0.65. The crystals were further observed with the help of an optical microscope, and all of them were plate-like, as shown in Figure 3a. When the methanol volume fraction ranges from 0.35 to 0.65, the nucleation of form A accompanying form B was observed. Both plate-like form B and needle-like form A were observed by optical microscope, as depicted in Figure 3b. The crystals harvested from the above concomitant nucleation33 experiments were quantitatively analyzed on the basis of the calibration curve to determine the mass fraction of form A. It is evident from Figure 6 that, for 0.35 ≤ x2 ≤ 0.45,

Figure 7. Induction time versus the methanol volume fraction at 298.15 K.

demonstrated a trend that the induction time decreased with increasing the antisolvent volume fraction. It varied from 19 min at x2 = 0.15 to approximately 30 s at x2 = 0.8. For x2 = 0.15−0.45, the induction time decreased sharply with increasing methanol volume fraction, whereas for x2 = 0.5− 0.8, no significant reduction could be measured. Care was taken that solution and antisolvent were mixed rapidly, tmix < 3 s. Thus, we can observe that the induction time was much longer than the mixing time in the batch experiments. In other words, the nucleation and crystal growth started at a uniform supersaturation in the crystallizer.34 Besides, the polymorphic transformation process from form B to form A would not happen within several days if there was no seed of form A (as shown in section 3.5, Figure 10). The transformation time is one to several orders of magnitude longer than the induction time. Therefore, it can be confirmed that the formation of form A during the antisolvent crystallization experiments discussed above is not due to the transformation from form B to form A but because of directly nucleation from solution. Thus, concomitant nucleation of both forms of A and B should be the real reason for concomitant polymorphism at methanol volume fraction of 0.35 ≤ x2 ≤ 0.65 in antisolvent crystallization. 3.5. Transformation Experiments in Binary Methanol + Water Mixed Saturated Solutions. The overall transformation from form B to form A in antisolvent crystallization with solution-mediated phase transformation was carried out at 298.15 K. The overall transformation time is shown in Figure 8 as a function of methanol volume fraction, x2 = 0.1−0.6. The antisolvent crystallization experiments were monitored by a polarizing microscope. The optical micrographs (x2 = 0.4) are shown in Figure 9. During the process of transformation, form B, which possessesd a thin plate-like shape, gradually became needle-like form A. Then the plate-like crystals diminished until form B eventually disappeared. Interestingly, an unexpected phenomenon that the overall transformation rate in antisolvent crystallization with solution-mediated phase transformation dramatically increases was observed for methanol volume fraction 0.35 < x2 < 0.60, as illustrated in Figure 8. Then, as described in section 3.3, we believe that this interesting phenomenon may be ascribed to the nucleation of form A when the methanol volume fraction ranges from 0.35 to 0.60.

Figure 6. Mass fraction of form A in the concomitant crystals as a function of methanol volume fraction at 298.15 K.

the polymorphic fraction of form A increased with the increasing methanol volume fraction whereas, for x2 > 0.45, the polymorphic fraction of form A decreased. The maximum fraction of form A is about 0.117 at about x2 = 0.4. 3.4. Induction Time versus Mixing Time versus Transformation Time. The induction time versus the methanol volume fraction was presented in Figure 7, which 788

DOI: 10.1021/acs.iecr.7b04762 Ind. Eng. Chem. Res. 2018, 57, 784−793

Article

Industrial & Engineering Chemistry Research

Figure 8. Overall transformation time in antisolvent crystallization and solution-mediated phase transformation as a function of the methanol volume fraction in binary solvents at 298.15 K.

Figure 10. Phase transformation time from form B to completely transform to form A as a function of methanol volume fraction in the methanol + water binary saturated solution at 298.15 K.

Therefore, the suspended polymorphic transformation from form B to form A at 298.15 K by adding form B into binary water + methanol saturated solution was designed to validate if the transformation rate will be accelerated without the nucleation step. The results for the complete transformation are shown in Figure 10. It is clear that the complete transformation time in suspended phase transformation increases with the increasing volume fraction of methanol in Figure 10. Generally, the transformation rate is higher in the solvent giving a higher solubility.4,8,35 Form Figure 5 and Table 1, it can be found that the solubility and the solubility differences of form A and form B decrease with the increasing proportion of water in water + methanol binary solvents. From Table 1, it can be calculated that the solubility difference at x2 = 0.5 is only 0.88 g/1000 g solution. Therefore, there will be a smaller driving force for the polymorphic transformation with the increasing proportion of methanol in the mixtures, which is consistent with the result in Figure 10. Namely, the sudden increase of the transformation rate did not appear in this kind of suspended solution-mediated phase transformation which further confirms that the nucleation step may play an indispensable role in our antisolvent crystallization with the transformation process from form B to form A.

When Figure 8 is compared with Figure 10, the overall transformation rates in antisolvent crystallization with solutionmediated phase transformation are significantly faster than that of suspended phase transformation for methanol volume fractions x2 = 0.4, 0.45, and 0.5. The whole antisolvent crystallization experimental process consists of two stages: one is the nucleation−growth stage and the other is the solutionmediated phase transformation. Undoubtedly, the nucleationand-growth process in antisolvent crystallization would have the utmost impact on subsequent solution-mediated phase transformation. Simultaneously, at an intermediate methanol volume fraction (0.35−0.65), the nucleation of form A accompanying form B was observed, as discussed in section 3.4. Therefore, it is reasonable to assume that the concomitant crystallization in the preceding nucleation−growth stage has accelerated the transformation process of the later phase transformation. The samples collected at different times during the nucleation-and-growth and phase transformation process were determined by PXRD. In this paper, the effects of different nucleation results on the subsequent phase transformation are described, for example, at x2 = 0.3, 0.4. The results are shown in Figures 11 and 12, respectively. From Figure 11, it is obvious

Figure 9. Images (magnified 20×) showing the nucleation-and-growth and phase transformation process at 298.15 K, x2 = 0.4: 4 min, 13 min, 25.5 min, 52 min, 14.5 h, 15.25 h. 789

DOI: 10.1021/acs.iecr.7b04762 Ind. Eng. Chem. Res. 2018, 57, 784−793

Article

Industrial & Engineering Chemistry Research

4. DISSCUSSION It is well-known that solution-mediated transformations can be described by three essential processes: dissolution of the metastable solid, nucleation of the stable solid phase, and growth of the stable phase. Polymorphic transformations are not certain because of the complication of these three steps, each of which or a combination of which can be the ratedetermining step. O’Mahony et al.36 defined four principal scenarios in which dissolution, growth, dissolution−nucleation, or nucleation−growth can determine the overall transformation rate. The polymorphic transformation at methanol volume fractions of 0.2, 0.25, and 0.3 are shown in Figure 13, Figure S2, and Figure S3, respectively.

Figure 11. PXRD patterns of the samples collected at 298.15 K for x2 = 0.3: 6 min, 40 min, 5 h, 17 h, 21 h, 24.25 h.

Figure 13. Antisolvent crystallization and solution-mediated phase transformation of form A to form B at a methanol volume fraction of 0.2. (■)The solution concentration profile, as measured via the ultraviolet spectroscopic method, and (●) form A composition as measured by PXRD;. Figure 12. PXRD patterns of the samples collected at 298.15 K for x2 = 0.4: 4 min, 18 min, 11.5 h, 13.5 h, 14.5 h, 15.25 h.

Figure 13 shows the solution concentration remains at the solubility of the metastable polymorph B in the process of solution-mediated polymorphic transformation (SMPT) from form B to 100% form A at a methanol volume fraction of 0.2. There was a considerable lag time of approximately 180 min prior to the nucleation and growth of form A in the plateau region and there was the second lag time of approximately 583 min before the fast transformation. The solid phase had completely transformed to the stable polymorph A after about 1000 min. Nucleation and then growth of the stable polymorph occurred at the expense of the dissolving metastable polymorph. Where the total rate of growth is compensated for by the much faster total rate of dissolution, a plateau in solution concentration/supersaturation is achieved. Figure S2 and S3 show similar transformations. The evidence of a plateau region in both of the SMPT experiments presented here showed that the transformations were controlled by growth of the stable polymorph. The results show that the transformation process is controlled by the nucleation and growth of the stable form A and belongs to“scenario d” as outlined by O’Mahony et al. From Table S1, it can be found that the supersaturation ratio SA (calculated on the basis of the solubility of form A) increases with increasing the antisolvent volume fraction in antisolvent crystallization, and the lag time prior to the nucleation and

that the PXRD patterns at the initial time fit perfectly with the PXRD patterns of form B, which has a characteristic peak at 17.25°. At about 40 min, PXRD patterns of form A came to appear with a weak characteristic peak at 17.75°. The characteristic peak at 17.25° became slightly obvious until 5 h. The evolution of the relative intensity of the characteristic peak represented a transformation process from form B to form A. Along with the decreased of the characteristic peak intensity of form B, the intensity of the characteristic peak of form A increased gradually, and the whole phase transformation process lasted about 24 h. However, the PXRD patterns have the characteristic peaks 17.25° and 17.75° at the initial time in Figure 12, which confirmed the appearance of concomitant polymorphism (x2 = 0.4), and the whole phase transformation process lasted about 15 h. From Figures 11 and 12, it is clear that the nucleation induction period of the stable form A took a long time. A small amount of form A was produced for x2 = 0.4 at the initial time in antisolvent crystallization whereas the crystallization of the crystals was basically form B for x2 = 0.3. Moreover, the nucleation induction period of the stable form A was significantly shorter, and ultimately the entire transformation time was shorter for x2 = 0.4. 790

DOI: 10.1021/acs.iecr.7b04762 Ind. Eng. Chem. Res. 2018, 57, 784−793

Article

Industrial & Engineering Chemistry Research growth of form A in the plateau region increases as well. As a result, the time point of rapid conversion of the SMPT is delayed, and finally, the complete transformation time in phase transformation increases with the increasing volume fraction of methanol. In addition, as can be seen from Figure 14 and Table

Figure 15. (Black squares) results of the solution-mediated phase transformation which introduced the form A crystals as a function of the methanol volume fraction and the mass fraction of form A at 298.15 K. (Blue squares) amount of introduced crystal A in the remaining suspension (both form B and form A) to the saturated solution.

verify the rationality of the hypothesis that form A included in the concomitant polymorph has seed-assisted effects on the subsequent transformation. On the basis of the analysis above, the key problem is understanding and controlling the nucleation behavior of L-his in methanol−water solution. In general, according to Ostwald’s rule of stages, the stable form is readily obtained at low supersaturation, whereas the metastable form is the kinetic favorable form at high supersaturation, and concomitant polymorphism may be produced at moderate supersaturation. However, in this case, the nucleation of form A accompanying with form B was observed at moderate supersaturation, and the metastable form produced at both low supersaturation and high supersaturation, which is a very interesting phenomenon. The mechanism of self-assembly of L-histidine in solution appears to contradict with the classical nucleation theory, which should be ascribed to the structure and density fluctuations of the precritical nuclei in the supersaturated solution. Therefore, in the future, it needs more work to investigate the nucleation mechanism and nucleation kinetics of L-histidine in methanol− water mixed solvent system to provide fundamental theory support for understanding the solution-mediated phase transformation and controlling the crystallization outcome.

Figure 14. Counts versus the time for different methanol volume fractions.

S1, the total counts of equilibrium particles of the crystal form B in suspension increases with the increasing volume fraction of antisolvent. However, the counts of larger particle size decreases. The SMPT experiments were monitored by a polarizing microscope. From Figure S4, one can clearly observe the needle-like stable form A nucleated and grown on or near the surface of the plate-like metastable form B. It seems plausible to assume that the number density of form B in the solution influences the nucleation point of form A and then affects the nucleation of form A and the entire solutionmediated phase transformation. As a consequence, the complete transformation time in phase transformation increases with the increasing volume fraction of methanol in Figure 8. When the methanol volume fraction x2 rises to 0.4−0.5, the complete transformation time decreases sharply, and in this paper we attribute it to the concomitant polymorphism. Due to the presence of form A in concomitant nucleation, the nucleation metastable zone width of form A is significantly shortened, thus leading to the higher nucleation rate of crystal form A and the decreased concentration of the mother liquor, which further promotes the dissolution of B. Therefore, the polymorphic transformation of L-his from form B to form A increases dramatically. Furthermore, the suspended polymorphic transformation from form B to form A at 298.15 K by introducing form A into binary water + methanol saturated solution was designed to validate if the transformation rate will be accelerated in the concomitant nucleation region. The amount of crystal A introduced was based on the quantitative results of Figure 6 and the results of the suspended solution-mediated phase transformation are shown in Figure 15. Again, it is evident that the phase transformation time was significantly shortened, as shown in Figure 15. The transformation time is shortened with the increase of the amount of form A and basically corresponds to the overall transformation time at the same methanol volume fraction in the antisolvent crystallization with solution-mediated phase transformation in Figure 8. The experimental results

5. CONCLUSIONS In summary, the quantitative polymorphic transformation experiments of L-his in antisolvent crystallization and solution-mediated phase transformation were performed at 298.15 K. Concomitant polymorphism of form A and form B was obtained for a methanol volume fraction of 0.30 ≤ x2 ≤ 0.65 in antisolvent crystallization whereas only form B was observed for other volume fractions of methanol. Accordingly, the overall transformation time in antisolvent crystallization with solution-mediated phase transformation is significantly reduced to one-tenth of suspended solution-mediated phase transformation for methanol volume fraction 0.35 < x2 < 0.60. It was confirmed by seeded transformation experiments and the PXRD results in the whole phase transformation process. The results obtained in this work have some guidance function on the effect of the nucleation−growth of the polymorphic 791

DOI: 10.1021/acs.iecr.7b04762 Ind. Eng. Chem. Res. 2018, 57, 784−793

Article

Industrial & Engineering Chemistry Research

(12) Lu, J.; Rohani, S. Polymorphic Crystallization and Transformation of the Anti-viral/HIV Drug Stavudine. Org. Process Res. Dev. 2009, 13, 1262−1268. (13) Schöll, J.; Bonalumi, D.; Lars Vicum, A.; Mazzotti, M.; Müller, M. In Situ Monitoring and Modeling of the Solvent-Mediated Polymorphic Transformation of l-Glutamic Acid. Cryst. Growth Des. 2006, 6, 881−891. (14) O’Mahony, M. A.; Seaton, C. C.; Croker, D. M.; Veesler, S.; Rasmuson, Å. C.; Hodnett, B. K. Investigation into the Mechanism of Solution-Mediated Transformation from FI to FIII Carbamazepine: The Role of Dissolution and the Interaction between Polymorph Surfaces. Cryst. Growth Des. 2013, 13, 1861−1871. (15) Davey, R. J.; Cardew, P. T. Rate Controlling Processes in Solvent-Mediated Phase Transformations. J. Cryst. Growth 1986, 79, 648−653. (16) Mullin, J. W. Crystallization, fourth ed.; ButterworthHeinemann, 2001; pp 536−575. (17) Kulkarni, S. A.; Meekes, H.; ter Horst, J. H. Polymorphism Control through a Single Nucleation Event. Cryst. Growth Des. 2014, 14, 1493−1499. (18) Kopple, J. D.; Swendseid, M. E. Evidence that Histidine is an Essential Amino Acid in Normal and Chronically Uremic Man. J. Clin. Invest. 1975, 55, 881−91. (19) Sasmal, M.; Maiti, T. K.; Bhattacharyya, T. K. Ultra-Low Level Detection of L-Histidine Using Solution-Processed ZnO Nanorod on Flexible Substrate. IEEE T. Nanobiosci. 2015, 14, 634−640. (20) Liu, Y.; Wang, Y.; Liu, Y.; Xu, S.; Chen, M.; Du, S.; Gong, J. Solubility of L-histidine in Different Aqueous Binary Solvent Mixtures from 283.15 to 318.15 K with Experimental Measurement and Thermodynamic Modelling. J. Chem. Thermodyn. 2017, 105, 1−14. (21) Kitamura, M. Crystallization Behavior and Transformation Kinetics of L-histidine Polymorphs. J. Chem. Eng. Jpn. 1993, 26, 303− 307. (22) Kitamura, M.; Furukawa, H.; Asaeda, M. Solvent Effect of Ethanol on Crystallization and Growth Process of L-histidine Polymorphs. J. Cryst. Growth 1994, 141, 193−199. (23) Roelands, C. P. M.; Jiang, S.; Kitamura, M.; ter Horst, J. H.; Kramer, H. J. M.; Jansens, P. J. Antisolvent Crystallization of the Polymorphs of L -Histidine as a Function of Supersaturation Ratio and of Solvent Composition. Cryst. Growth Des. 2006, 6, 955−963. (24) Madden, J. J.; Mcgandy, E. L.; Seeman, N. C.; Harding, M. M.; Hoy, A. The Crystal Structure of the Monoclinic Form of L -Histidine. Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 1972, 28, 2382−2389. (25) Madden, J. J.; Mcgandy, E. L.; Seeman, N. C. The Crystal Structure of the Orthorhombic Form of L -(+)-Histidine. Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 1972, 28, 2377− 2382. (26) Wantha, L.; Flood, A. E. Growth and Dissolution Kinetics of A and B Polymorphs of L -Histidine. Chem. Eng. Technol. 2015, 38, 1022−1028. (27) Wantha, L. Determining the Nucleation and Growth Mechanisms of B Polymorph of L-Histidine by Induction Time Measurement. Chem. Eng. Technol. 2016, 39, 1289−1294. (28) Li, Y.; Chow, P. S.; Tan, R. B. Quantification of Polymorphic Impurity in an Enantiotropic Polymorph System Using Differential Scanning Calorimetry, X-ray Powder Diffraction and Raman Spectroscopy. Int. J. Pharm. 2011, 415, 110−118. (29) Qiu, J. B.; Li, G.; Sheng, Y.; Zhu, M. R. Quantification of Febuxostat Polymorphs using Powder X-ray Diffraction Technique. J. Pharm. Biomed. Anal. 2015, 107, 298. (30) Croker, D. M.; Hennigan, M. C.; Maher, A.; Hu, Y.; Ryder, A. G.; Hodnett, B. K. A Comparative Study of the Use of Powder X-ray Diffraction, Raman and Near Infrared Spectroscopy for Quantification of Binary Polymorphic Mixtures of Piracetam. J. Pharm. Biomed. Anal. 2012, 63, 80−86. (31) Kuldipkumar, A.; Kwon, G. S.; Zhang, G. G. Determining the Growth Mechanism of Tolazamide by Induction Time Measurement. Cryst. Growth Des. 2007, 7, 234−242.

crystallization behavior on subsequent transformation experiments in antisolvent crystallization.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.iecr.7b04762. Calibration curve, antisolvent crystallization and solution-mediated phase transformation of form A to form B, morphology during the polymorphic transformation, supersaturation, particle number and transformation time, and standard errors and confidence intervals (PDF)



AUTHOR INFORMATION

Corresponding Author

*J. Gong. Tel: 86-22-27405754. Fax: + 86-22-27314971. Email: [email protected]. ORCID

Shichao Du: 0000-0002-8369-2983 Junbo Gong: 0000-0002-3376-3296 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The authors are grateful for the financial support of National Natural Science Foundation of China (81361140344, 21676179 and 21376164), National 863 Program (2015AA021002). Major Project of Tianjin (15JCZDJC33200) and Major National Scientific Instrument Development Project 21527812.



REFERENCES

(1) Hollingsworth, M. D. Crystal Engineering: from Structure to Function. Science 2002, 295, 2410−2413. (2) Hilfiker, R., Ed. Polymorphism: in the Pharmaceutical Industry; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, 2006. (3) Jiang, S.; Ter Horst, J. H.; Jansens, P. J. Concomitant Polymorphism of o-Aminobenzoic Acid in Antisolvent Crystallization. Cryst. Growth Des. 2008, 8, 37−43. (4) Yang, L.; Hao, H.; Zhou, L.; Wei, C.; Hou, B.; Xie, C.; Yin, Q. Crystal Structures and Solvent-Mediated Transformation of the Enantiotropic Polymorphs of 2,3,5-Trimethyl-1,4-diacetoxybenzene. Ind. Eng. Chem. Res. 2013, 52, 17667−17675. (5) Brittain, H. G. Polymorphism in Pharmaceutical Solids; CRC Press: Boca Raton, FL, 2009. (6) Kitamura, M. Strategy for Control of Crystallization of Polymorphs. CrystEngComm 2009, 11, 949−964. (7) Jiang, S.; Jansens, P. J.; ter Horst, J. H. Mechanism and Kinetics of the Polymorphic Transformation of o-Aminobenzoic Acid. Cryst. Growth Des. 2010, 10, 2123−2128. (8) Gu, C. H.; Young, V. G.; Grant, D. J. W. Polymorph Screening: Influence of Solvents on the Rate of Solvent-Mediated Polymorphic Transformation. J. Pharm. Sci. 2001, 90, 1878−1890. (9) Du, W.; Yin, Q.; Hao, H.; Bao, Y.; Zhang, X.; Huang, J.; Li, X.; Xie, C.; Gong, J. Solution-Mediated Polymorphic Transformation of Prasugrel Hydrochloride from Form II to Form I. Ind. Eng. Chem. Res. 2014, 53, 5652−5659. (10) Ferrari, E. S.; Davey, R. J. Solution-Mediated Transformation of α to β l-Glutamic Acid: Rate Enhancement Due to Secondary Nucleation. Cryst. Growth Des. 2004, 4, 1061−1068. (11) Croker, D.; Hodnett, B. K. Mechanistic Features of Polymorphic Transformations: The Role of Surfaces. Cryst. Growth Des. 2010, 10, 2806−2816. 792

DOI: 10.1021/acs.iecr.7b04762 Ind. Eng. Chem. Res. 2018, 57, 784−793

Article

Industrial & Engineering Chemistry Research (32) Mangin, D.; Puel, F.; Veesler, S. Polymorphism in Processes of Crystallization in Solution: A Practical Review. Org. Process Res. Dev. 2009, 13, 1241−1253. (33) Bernstein, J.; Davey, R. J.; Henck, J. O. Concomitant Polymorphs. Angew. Chem., Int. Ed. 1999, 38, 3440. (34) Jiang, S.; ter Horst, J. H.; Jansens, P. J. Concomitant Polymorphism of o-Aminobenzoic Acid in Antisolvent Crystallization. Cryst. Growth Des. 2008, 8, 37−43. (35) Davey, R. J.; Blagden, N.; Righini, S.; Alison, H.; Ferrari, E. S. Nucleation Control in Solution Mediated Polymorphic Phase Transformations: The Case of 2,6-Dihydroxybenzoic Acid. J. Phys. Chem. B 2002, 106, 1954−1959. (36) O’Mahony, M. A.; Maher, A.; Croker, D. M.; Rasmuson, Å. C.; Hodnett, B. K. Examining Solution and Solid State Composition for the Solution-Mediated Polymorphic Transformation of Carbamazepine and Piracetam. Cryst. Growth Des. 2012, 12, 1925−1932.

793

DOI: 10.1021/acs.iecr.7b04762 Ind. Eng. Chem. Res. 2018, 57, 784−793