Hydrogen Bonding Paradigm in the Formation of Crystalline KH2PO4

Apr 24, 2017 - The hydrogen bonding paradigm in the formation of crystalline KH2PO4 from aqueous solution has been clarified by tracking the variation...
0 downloads 0 Views 2MB Size
Article pubs.acs.org/crystal

Hydrogen Bonding Paradigm in the Formation of Crystalline KH2PO4 from Aqueous Solution Congting Sun,† Xiaoyan Chen,† and Dongfeng Xue*,† †

State Key Laboratory of Rare Earth Resource Utilization, Changchun Institute of Applied Chemistry, Chinese Academy of Sciences, Changchun 130022, China

Downloaded via DURHAM UNIV on July 11, 2018 at 10:24:29 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: Revealing the hydrogen bonding paradigm is critical to clarify the formation mechanism of hydrogen bonded materials. The nucleation process of a typical nonlinear optical crystal, KH2PO4, is identified by in situ molecular vibration spectroscopy, which effectively demonstrates the oriented role of hydrogen bonding in local structure engineering. On the basis of the vibrational evolution of hydrogen bonds, the partition of different periods in the formation of crystalline KH2PO4 from aqueous solution becomes clear. In KH2PO4 aqueous solution, there are hydrated status, (H2PO4−)n aggregations, and prenucleation clusters. The prenucleation clusters exist in the solution with a metastable status over a period of time, and then they will transform into crystalline ones within a short time. Two distinct roles of P−O···H−O−P hydrogen bonding in the formation of crystalline KH2PO4 have been distinguished. At the initial stage of aggregation formation, P−O bond in H2PO4− group guides the P−O···H−O−P hydrogen bonding, leading to the H2PO4− that retains C2v symmetry, whereas P−O···H−O−P hydrogen bonding guides the twist and rotation of H2PO4− groups in prenucleation clusters, promoting the local structural evolution of (H2PO4−)n from C2v to D2d and the formation of crystalline KH2PO4 nuclei. The present work deepens the hydrogen bonding effect that can warrant much space to adjust the chemical bonding environment in constructing crystallographic frames.



INTRODUCTION Chemical bonds have been proposed to describe the interactions between adjacent atoms/ions, which dominate the relative position of atoms/ions in crystalline materials. Both crystal constituents and crystallographic structure determine the materials properties.1,2 Materials construction rules are established on the basis of understanding of intermolecular interactions in the framework of crystal packing and are further used to design novel crystalline solids with desired physicochemical properties by regulating the chemical bonding and composition constituents.3,4 The nature of optical nonlinearity challenges the design of nonlinear optical crystals.5−7 The Phillips−Van Vechten−Levine−Xue (PVLX) bond theory has been proposed, which uncovers the origin of nonlinearity of nonlinear optical crystals, that is, the total optical nonlinearities generally originate from the contributions of each type of constituent chemical bond.8−10 Among numerous constituent chemical bonds, hydrogen bond possesses directionality and stoichiometry,11,12 which has long been recognized in adjusting crystal packing and materials properties.13−15 The contributions of hydrogen bonding in some typical crystals with hydrogen bonds such as urea,5 ice,10 HIO3,16,17 KH2PO4 (KDP),17 NH4H2PO4 (ADP),16,17 K2La(NO3)5·2H2O,16,18 K2Ce(NO3)5·2H2O,18 K[B5O6(OH)4]· 2H2O,16 and Na2SeO4·H2SeO3·H2O2 to the total optical nonlinearity have been calculated. The calculated results demonstrate that hydrogen bonding play a very important © 2017 American Chemical Society

role in contributing to the total optical properties, indicating the possibility of employing hydrogen bonding effects in nonlinear optical crystal engineering. In the formation of crystalline materials, the competition and cooperation between multiple parameters dominate the growth behaviors, such as the crystallographic structure and anisotropic growth. In order to obtain high-quality crystalline materials, the integration among multiple parameters is required. To date, different crystal growth theories and models have been developed to extract the critical parameters.19,20 Chemical bonding theory of single crystal growth indicates that the chemical bonding process at the growing interface dominates the anisotropic growth rate of single crystals, which has been applied to the growth of both nanocrystals and bulk crystals.21,22 Among various types of chemical bonding at the growing interface, weak bonds play a critical role in determining the growth rate.23 By focusing on the variation of hydrogen bonding architectures between (H2PO4−)n clusters, the morphology evolution has been calculated for KH2PO4 single crystals in aqueous solution.24 Moreover, the morphology diagram of urea has been calculated in which the relative rate regions for different anisotropic geometries of crystalline urea are scaled on the basis of anisotropic hydrogen bonding Received: January 26, 2017 Revised: April 17, 2017 Published: April 24, 2017 3178

DOI: 10.1021/acs.cgd.7b00139 Cryst. Growth Des. 2017, 17, 3178−3184

Crystal Growth & Design

Article

between urea molecules.25,26 For the nonlinear optical crystals with hydrogen bonds, controlling the hydrogen bonding process at the growing interface facilitates the high quality crystal growth. Potassium dihydrogen phosphate (KH2PO4, KDP) is a wellknown hydrogen bonded crystal, which possesses special ferroelectrics, nonlinear optical, and electro-optic properties.27,28 KDP crystallizes in tetragonal structure, where each tetrahedral H2PO4− is hydrogen bonded to the other four H2PO4−.29 Previous studies show that phosphate has selectivity and high affinity as metal receptors or selectively binds organic molecules;30−33 therefore, the phosphates can affect biological and ecological systems significantly.34,35 For example, it has been reported that the phosphate ions and water molecules can be used to tune the metastability of amorphous calcium carbonates in organisms.36 Both experimental observations and theoretical calculations have been carried out to investigate the clusters such as hydrated ions and atomic groups and the interactions in H2PO4−−H2O and H2PO4−−H2PO4−.37−39 By combining the infrared multiple photon dissociation (IRMPD) spectra with ab initio molecular dynamics (AIMD), it has been described the anharmonic and dynamic effect on the structure and size of H2PO4−(H2O)n clusters and the formation of hydrogen bond network around a H2PO4− through solute− solvent.40 Rajbanshi et al. found H2PO4− self-association and the generation of extended polymeric structure with the maximum number of hydrogen bonds per anion during the crystallization process.26 Therefore, clarifying the hydrogen bonding is critical in disclosing the formation mechanism of crystalline KDP from aqueous solution system. In this work, both in situ Raman and attenuated total reflection-infrared (ATR-IR) spectroscopies are carried out to clarify the role of hydrogen bonding in the formation of KDP from aqueous solution. The vibration bands of H−O in water, H−O association in H2PO4−, and P−O···H−O−P between H2PO4− groups can be identified and real-time tracked. On the basis of the vibrational evolution of these distinct hydrogen bonds, different periods in the nucleation process of KDP from aqueous solution becomes clear. Moreover, the role of hydrogen bonding in the formation of crystalline KDP can also be distinguished.



completely. Time-dependent ATR-IR spectra can provide the hydrogen bonding characteristics in the formation of KDP crystallographic structure framework.



RESULTS AND DISCUSSION During the process in which crystalline KDP nuclei are formed, hydrogen bonding between H2PO4− groups is critical to the construction of structural framework. At the initial stage, both hydrated H2PO4− and K+ ions exist in KDP aqueous solution. With the successive volatilization of water, (H 2 PO 4 − ) n aggregations are formed by P−O···H−O−P hydrogen bonding. Owing to the substituent of H−O···H−O−P by P−O···H−O− P, the symmetry of H2PO4− group will change. Raman spectroscopy becomes a useful tool to identify the group symmetry since the group theory has been successfully applied to molecular vibration spectroscopy. H2PO4− group is Raman active, but the measurement time for Raman spectrum ranging from 200 to 1600 cm−1 is about 3.5 min. It is too longer to in situ observe the intermediate species or structures in KDP nucleation. In order to overcome this technique drawback, we carried out successive Raman scanning in a shorter wavenumber range of the aim group, i.e., H2PO4− group, which can be finished with 1 min. Figure 1 shows in situ Raman spectra of

Figure 1. In situ Raman spectra of KDP aqueous solution with different concentrations at 300−700 cm−1. (A) Time-dependent Raman spectra of KDP solution with the concentration of 1.46 M. (B) Time-dependent Raman spectra of KDP solution with the concentration of 1.84 M.

EXPERIMENTAL SECTION

Time-dependent Raman spectra of KDP aqueous solution were recorded using a Jobin-Yvon Horiba T64000 Raman triple grating spectrometer (Horiba Ltd., France), and green line Ar+ laser with 514.5 nm radiation was used as the excitation source. KDP aqueous solutions were prepared at the room temperature of 20 °C with the concentration of 1.46, 1.64, 1.84, and 2.02 mol/L, respectively. At the first stage of Raman measurements, the frequency was calibrated by the Raman band of silicon at 520 cm−1. Five microliters of KDP solution was added on the glass substrate, and the laser spot was ∼1 μm in diameter after being focused in the center of the aqueous droplet. Afterward, the successive Raman measurements were carried out using Mapping properties mode per 1 min in a particular wavenumber region. In in situ Raman experiments, the Raman spectra were recorded with a wavelength precision of 1 cm−1 in the frequency range 200−1600 cm−1 by using the NGSLabSpec software. Timedependent Raman spectra were collected to identify the chemical composition and group symmetry evolution during KDP nucleation. In situ ATR-IR spectral experiments were carried out by using an ATR cell with diamond wafer in Nicolet 6700 FT−IR spectrometer at 20 °C. The absorption spectral range is 4000−525 cm−1. Successive IR absorption spectra can be obtained with increasing the measurement time until the aqueous solution transforms into crystalline solid

KDP solution at 300−700 cm−1. Raman vibration bands at 380, 508, and 556 cm−1 can be assigned to the bending vibration of P(OH)2 (v2(PO4)), twisting vibration of PO2, and bending vibration of PO2 (v4(PO4)), respectively (Figure S1, Table S1). In aqueous state, only v2(PO4) and v4(PO4) bands can be observed in Raman spectra. With the proceeding of nucleation process, PO2 twisting vibration band is obvious, and v2(PO4) and v4(PO4) bending vibration bands respectively shift toward lower and higher wavenumbers. This can be attributed to the symmetry transformation of H2PO4− group from C2v to S4. Moreover, the appearance of PO2 twisting vibration band indicates the rotation of H2PO4− group before the formation of crystalline KDP nuclei. In order to study the role of P−O···H−O−P hydrogen bonding in guiding the molecular structure evolution of H2PO4− group, we further track the stretching vibration bands of P(OH)2 and PO2 in Raman spectra with the 3179

DOI: 10.1021/acs.cgd.7b00139 Cryst. Growth Des. 2017, 17, 3178−3184

Crystal Growth & Design

Article

wavenumber ranging from 800 to 1300 cm−1 (Table S1). As shown in Figure 2A, vs(P(OH)2) shifts toward higher

positions of vs(P(OH)2) at 0 s are 879, 881, 883, and 885 cm−1 for KDP solution with the concentration of 1.46, 1.64, 1.84, and 2.02 mol/L, respectively. The Raman spectra show that the (H2PO4−)n aggregations prefer to be formed in KDP solution with higher concentration, which can shorten the time for structural transformation in KDP nucleation process. On the basis of group theory, the stretching vibration of H2PO4− group, P−O···H−O−P hydrogen bonding, O−H stretching in H2PO4−, as well as the stretching, bending, and wagging vibration of H2O and O−H combination tone in H2O are IR-active (Figure S4). In situ ATR-IR spectroscopy was used to observe the formation of crystalline KDP nuclei in aqueous solution system (shown in Figure S5). The detection time continues 560 s, and per 40 s we can obtain an ATR-IR spectrum ranging from 525 to 4000 cm−1. At 560 s, there is no absorption bands of H2O in the ATR-IR spectrum, illustrating only crystalline KDP exists on the diamond wafer. Until 520 s, we can find the absorption bands of H2O. For H2O, its stretching vibration would be influenced by its surrounding hydrogen bonding. At nucleation stage, the stretching vibration band of H2O (a broad peak at ∼3300 cm−1) slightly shifts toward higher wavenumber (Figure 3). As shown in Figure 3A,

Figure 2. In situ Raman spectroscopy of KDP aqueous solution with different concentrations at 800−1300 cm−1. (A) Time-dependent Raman spectra of KDP solution with the concentration of 1.46 M. (B) Time-dependent Raman spectra of KDP solution with the concentration of 1.84 M.

wavenumber with prolonging the detection time. During the nucleation process, once (H2PO4−)n aggregations are formed, some H−O···H−O−P and H−O−H···O−P hydrogen bonds are broken, while P−O···H−O−P hydrogen bonds are formed. According to the Pauling electronegativity scale (EN), EN(H) (2.18) is slight smaller than EN(P) (2.19), and H−O···H−O− P hydrogen bonding is stronger than P−O···H−O−P hydrogen bonding. Moreover, we can also deduce that H−O−H···O−P hydrogen bonding is stronger than P−O···H−O−P hydrogen bonding. Consequently, vs(P(OH)2) will shift toward higher wavenumbers under the weaker P−O···H−O−P hydrogen bonding interaction. However, PO2 will shift toward lower wavenumbers owing to the stronger H−O in H2PO4− group when H−O···H−O−P is replaced by P−O···H−O−P hydrogen bonding, which favors the following rotation and twisting of PO4 frame to realize the symmetry evolution. The Raman observations can demonstrate the formation of P−O···H−O−P hydrogen bonding between H2PO4− groups in solution system by breaking the hydrogen bonding in hydrated H2PO4−. The appearance of v1(PO4) and v3(PO4) vibration (the stretching vibration of PO4 group with S4 symmetry) in Raman spectra means the transformation of H2PO4− from C2v symmetry to S4 symmetry owing to the intermediate position of H between two H2PO4− groups, indicating the formation of crystalline nuclei in KDP solution. Furthermore, we also compare the time-dependent Raman spectra of KDP aqueous solution respectively with the concentration of 1.46, 1.64, 1.84, and 2.02 mol/L (Figures 1, 2, S2, and S3). It can be found that the time when v1(PO4) and v3(PO4) bands appear decreases with increasing the solution concentration. In solution system, higher KDP concentration means lower weight of H2O molecules. Consequently, the number of H2O molecules around the H2PO4− group decreases with increasing KDP concentration, and the weight of H−O··· H−O−P and H−O−H···O−P hydrogen bonding becomes less. This will induce the aggregation between hydrated H2PO4− groups by P−O···H−O−P hydrogen bonding, which can be reflected by the positions of vs(P(OH)2) of H2PO4−. The

Figure 3. Variations of O−H stretching vibration of H2O in saturated KDP aqueous solution from 0 to 520 s. (A) Difference spectrum of IR absorbance spectra of H2O stretching vibration in KDP aqueous solution respectively at 520 and 0 s. (B) O−H stretching vibration bands of H2O in KDP solution within the wavenumber region of 3340−3400 cm−1. (C) O−H stretching vibration bands of H2O in KDP solution within the wavenumber region of 3240−3300 cm−1.

the difference spectrum of ATR-IR absorbance spectra of KDP aqueous solution respectively at 520 and 0 s shows a clear shift of v(O−H) toward higher wavenumbers. This can be attributed to the successive formation of P−O···H−O−P in hydrated (H2PO4−)n aggregations because both H−O···H−O−P and H−O−H···O−P hydrogen bonding are stronger than P−O··· H−O−P hydrogen bonding. With the formation of P−O···H− O−P hydrogen bonding, H−O vibration energy in H2O will become stronger, leading to the shift of H−O stretching vibration toward higher wavenumber. From the time-dependent O−H stretching vibration in ATR-IR spectra, we can deduce the increasing number of H2PO4− in the cluster. In previous studies, the stretching vibration band of O−H in H2PO4− group and combination tone band of H2O are always neglected. In the present work, we track these IR absorption bands in order to obtain more hydrogen bonding information 3180

DOI: 10.1021/acs.cgd.7b00139 Cryst. Growth Des. 2017, 17, 3178−3184

Crystal Growth & Design

Article

group, while IR absorption band at ∼2125 cm−1 can be assigned to the O−H combination tone band in H2O (Table S1). Increasing in situ observation time up to 80 s, no obvious O−H combination tone band in H2O can be found in the spectrum, indicating the decrease of H2O content in KDP solution. The v(O−H) band in H2PO4− group shifts toward lower wavenumber. This can be attributed to the increased P− O···H−O−P hydrogen bonding at the nucleation stage of KDP (Figure 5). In aqueous status, hydrated H2PO4− and K+ ions exist, and O−H combination tone band in H2O is obvious. Later, (H2PO4−)n aggregations are formed by P−O···H−O−P hydrogen bonding. Since the H2PO4− group retains C2v symmetry, the P−O···H−O−P hydrogen bonding around central H2PO4− group possesses C2v symmetry. In such a case, the H2PO4− group with C2v symmetry guides the initial P−O···H−O−P hydrogen bonding. Once the prenucleation clusters are formed, the symmetric evolution begins, P−O···H− O−P hydrogen bonding becomes stronger (Figure 5) and O− H in H2PO4− group becomes weaker (Figure 4). This allows the molecular structure transformation of KDP from C2v to D2d. In this stage, P−O···H−O−P hydrogen bonding guides the movement of group relative position. Figure 5A shows the time-dependent ATR-IR spectra of δ(P−O···H−O−P) in KDP solution. We plot the position of δ(P−O···H−O−P) against measurement time (Figure S6), and three zones can be divided obviously. As shown in Figure 5B, when t ≤ 80 s, δ(P−O···H−O−P) shifts toward higher wavenumbers. It is notable that there is a sharp shift of δ(P− O···H−O−P) within 40 s ≤ t ≤ 80 s. Before 40 s, δ(P−O···H− O−P) shifts toward a higher wavenumber due to the formation of (H2PO4−)n aggregations. Later, K+ incorporates into the (H2PO4−)n framework. Owing to the lower EN(K) (0.82) than EN(H) (2.18), P−O···H−O−P hydrogen bonding becomes

in KDP solution. As shown in Figure 4, the IR absorption band at ∼2380 cm−1 can be assigned to the v(O−H) in H2PO4−

Figure 4. Evolution of O−H in H2PO4− group and H2O at the nucleation stage for saturated KDP aqueous solution. (A) Timedependent O−H combination tone bands of H2O and v(O−H) in H2PO4− in saturated KDP aqueous solution. (B) Schematics of O−H bonding evolution under the interactions between H2PO4− groups. The blue dotted lines are drawn to guide the assignment of IR absorption bands of hydrogen bond in H2PO4− and H2O.

Figure 5. In situ observation of P−O···H−O−P hydrogen characteristics between H2PO4− groups in saturated KDP aqueous solution. (A) Timedependent δ(P−O···H−O−P) bands in saturated KDP aqueous solution. (B) Position evolution of δ(P−O···H−O−P) bands in nucleation of KDP from aqueous solution. Three stages can be clearly distinguished, and the red dotted lines are drawn to guide the partition of formation process of crystalline KDP. (C) Schematics of structural variations of (H2PO4−)n framework. Blue dashed lines in (A) are used to guide the shift of δ(P−O··· H−O−P) bands. 3181

DOI: 10.1021/acs.cgd.7b00139 Cryst. Growth Des. 2017, 17, 3178−3184

Crystal Growth & Design

Article

from 80 to 480 s, the structural transformation from 480 to 520 s, and the formation of crystalline KDP nuclei after 520 s. From both Figure 6A,B, similar evolution trends in KDP nucleation can be found, indicating the consistency of mesoscale structural evolution reflected by different vibration bands of the constituent groups. We further select the typical ATR-IR spectra that can reflect different status at the nucleation state, as shown in Figures 7 and S7. In aqueous status, H2PO4− groups are interacted with H2O by hydrogen bonding, and each H2PO4− unit is surrounded by four hydrogen bonds. In such a case, four IR absorption bands exist for H2PO4− groups, respectively; the asymmetric and symmetric stretching vibrations of PO2 in H2PO4−, i.e., vas(PO2) at 1150 cm−1 and vs(PO2) at 1075 cm−1, and the asymmetric and symmetric stretching vibrations of P(OH)2 in H2PO4−, i.e., vas(P(OH)2) at 940 cm−1 and vs(P(OH)2) at 874 cm−1 (Figure 7A, Table S1). According to the group theory, the H2PO4− possesses C2v symmetry in aqueous status. In aggregation status, the slight appearance of P−O···H−O−P vibration band at ∼1242 cm−1 indicates the aggregation of H2PO4− groups. On the basis of IR spectral characteristics, the H2PO4− group keeps C2v symmetry, but the relative intensity between asymmetric and symmetric stretching bands changes, indicating the structural variation in H2PO4− group. In the prenucleation status, KH2PO4 molecules are formed via hydrogen bonding between H2PO4− groups and the K−O between K+ and H2PO4−. The dramatic increase of δ(P− O···H−O−P) denotes the formation of nucleation clusters. After these metastable nucleation clusters, the relative positions of H2PO4− groups are engineered via P−O···H−O−P hydrogen bonding. In nucleation status, the intensity of δ(P−O···H− O−P) enhances and the absorption band shifts from 1260 to 1271 cm−1, demonstrating the increased hydrogen bonding between H2PO4−. For H2PO4− groups, the vibration modes transform from nondegenerate state (i.e., 2v1(A1) at 874 and 1075 cm−1, v3(B1) at 1150 cm−1, and v3(B2) at 940 cm−1) to degenerate state (v3(B2 + E) at 1050 cm−1 and v1(A1) at 874 cm−1), indicating the symmetry of H2PO4− group increases from C2v to S4 owing to the intermediate position of H between two H2PO4− groups. The hydrogen bonding guides the formation of crystalline KDP with D2d symmetry (Figure S8). The chemical bonding characteristics of crystalline KDP are shown in Figure 7B, which guide the mesoscale structural evolution in the formation of crystalline KDP from aqueous solution.

stronger in such a case, leading to its sharp shift toward higher wavenumbers. At this stage, KDP prenucleation clusters are formed, and then they enter the structural evolution period. According to Figure 5B, KDP prenucleation clusters undergo a relative longer metastable stage, about 400 s. Finally, the P− O···H−O−P hydrogen bonding increases again and guides the rotation and twisting of H2PO4− as well as K+. This results in the formation of crystalline KDP nuclei at 520 s. In molecular vibration spectroscopy, the spectral characteristics, such as the relative intensity, position, and number of vibration bands correspond to the symmetry of constituent groups. The variation of spectral characteristics indicates the evolution of structural symmetry, which is induced by the phase transition. Therefore, the relative height of two different bands can also be used to reflect the symmetry of mesoscale structures at different stages in the formation crystalline KDP. In order to further testify this structural evolution stage, we plot the relative height between δ(P−O···H−O−P) and vs(PO2), H(δ(P−O··· H−O−P))/H(vs(PO2)) as well as the relative height between vas(PO2) and vs(PO2), H(vas(PO2))/H(vs(PO2)) against the measurement time (Figure 6). We can deduce the same results about the period partition in the nucleation process of KDP from aqueous solution with the concentration of 2.33 mol/L. Those are aqueous status before 80 s, KDP prenucleation clusters at 80 s, the metastable status of prenucleation status



CONCLUSION In this work, both in situ Raman and ATR-IR spectroscopes are carried out in order to clarify the hydrogen bonding paradigm in the nucleation of KDP from aqueous solution. On the basis of the evolution of hydrogen bonding vibrations, the period partition in the formation of crystalline KDP from aqueous solution becomes clear. The Raman observations confirm the formation of P−O···H−O−P hydrogen bonding between H2PO4− groups in solution system by breaking the hydrogen bonding between H2O and H2PO4−. On the basis of in situ ATR-IR spectra, the formation process of crystalline KDP includes the aqueous status, aggregation status, prenucleation status, and nucleation status. It is notable that there is a relatively longer metastable status of prenucleation clusters, which favors to enhance the P−O···H−O−P hydrogen bonding to promote the nucleation. Finally, we can deduce two roles of P−O···H−O−P hydrogen bonding in the formation of

Figure 6. Time-dependent ratio between the relative height of vas(PO2), vs(PO2), and δ(P−O···H−O−P) vibration bands. (A) Timedependent ratio between the relative height of δ(P−O···H−O−P) and vs(PO2). (B) Time-dependent ratio between the relative height of vas(PO2) and vs(PO2). Red dotted lines are drawn to guide the partition of formation process of crystalline KDP. 3182

DOI: 10.1021/acs.cgd.7b00139 Cryst. Growth Des. 2017, 17, 3178−3184

Crystal Growth & Design

Article

Figure 7. Intermolecular hydrogen bonding guided the structural evolution in the formation of crystalline KDP. (A) In situ ATR-IR observation of H2PO4− groups and the P−O···H−O−P interaction between H2PO4− groups. (B) Schematics of mesoscale structural evolution in the formation crystalline KDP from aqueous solution. (6) Suponitsky, K. Y.; Masunov, A. E. J. Chem. Phys. 2013, 139, 094310. (7) Seidler, T.; Stadnicka, K.; Champagne, B. J. Chem. Phys. 2014, 141, 104109. (8) Latajka, Z.; Gajewski, G.; Barnes, A. J.; Xue, D.; Ratajczak, H.; Latajka, Z.; Gajewski, G.; Barnes, A. J.; Xue, D.; Ratajczak, H. J. Mol. Struct. 2009, 928, 121−124. (9) Plachinda, P. A.; Dolgikh, V. A.; Stefanovich, S. Y.; Berdonosov, P. S. Solid State Sci. 2005, 7, 1194−1200. (10) Xue, D.; Betzler, K.; Hesse, H.; Ratajczak, H. Bull. Polish Acad. Sci. Chem. 2001, 49, 289−298. (11) Cazade, P.-A.; Cazade, H.; Bereau, T.; Das, A. K.; Kläsi, F.; Hamm, P.; Meuwly, M. J. Chem. Phys. 2015, 142, 212415. (12) Inoue, K. I.; Nihonyanagi, S.; Singh, P. C.; Yamaguchi, S.; Tahara, T. J. Chem. Phys. 2015, 142, 212431. (13) Xue, D.; Ratajczak, H. J. Mol. Struct.: THEOCHEM 2005, 716, 207−210. (14) Zhang, F.; Li, K.; Ratajczak, H.; Xue, D. J. Mol. Struct. 2010, 976, 69−72. (15) Black, H. T.; Perepichka, D. F. Angew. Chem. 2014, 126, 2170− 2174. (16) Xue, D.; Zhang, S. Chem. Phys. Lett. 1999, 301, 449−452. (17) Xue, D.; Zhang, S. J. Phys. Chem. Solids 1996, 57, 1321−1328. (18) Xue, D.; Zhang, S. Phys. B 1999, 262, 78−83. (19) Ranganathan, M.; Weeks, J. D. Phys. Rev. Lett. 2013, 110, 055503. (20) Sun, C.; Xue, D. Phys. Chem. Chem. Phys. 2013, 15, 14414− 14419. (21) Sun, C.; Xue, D. J. Phys. Chem. C 2013, 117, 5505−5511. (22) Sun, C.; Xue, D. CrystEngComm 2014, 16, 2129−2135. (23) Shultz, M. J.; Bisson, P.; Vu, T. H. J. Chem. Phys. 2014, 141, 18C521. (24) Sun, C.; Xu, D.; Xue, D. CrystEngComm 2013, 15, 7783−7791. (25) Sun, C.; Xue, D. Cryst. Growth Des. 2015, 15, 2867−2873. (26) Rajbanshi, A.; Wan, S.; Custelcean, R. Cryst. Growth Des. 2013, 13, 2233−2237. (27) Kawahata, Y.; Tominaga, Y. Solid State Commun. 2008, 145, 218−222. (28) Feng, X.; Zhu, L.; Li, F.; Wang, F.; Han, W.; Wang, Z.; Zhu, Q.; Sun, X. RSC Adv. 2016, 6, 33983−33989. (29) Cai, W.; Katrusiak, A. Dalton Trans. 2013, 42, 863−866. (30) Goods, J. B.; Sydlik, S. A.; Walish, J. J.; Swager, T. M. Adv. Mater. 2014, 26, 718−723. (31) Asha, K. S.; Bhattacharjee, R.; Mandal, S. Angew. Chem. 2016, 128, 11700−11704. (32) Levinson, N. M.; Bolte, E. E.; Miller, C. S.; Corcelli, S. A.; Boxer, S. G. J. Am. Chem. Soc. 2011, 133, 13236−13239.

crystalline KDP. At the initial stage of aggregation formation, P−O bond in H2PO4− group guides the P−O···H−O−P hydrogen bonding, keeping H2PO4− with C2v symmetry. In prenucleation clusters, P−O···H−O−P hydrogen bonding guides the twist and rotation of H2PO4− groups, allowing the structural evolution from C2v to D2d, producing the crystalline KH2PO4 nuclei. The present work deepens the hydrogen bonding effect that can warrant much space to adjust the chemical bonding environment in constructing crystallographic frames.



ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.cgd.7b00139. Additional Figures S1−S8, Table S1, and references depicting experimental results (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Congting Sun: 0000-0002-6949-6417 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS Financial support from the National Natural Science Foundation of China (grant nos. 51125009, 91434118, and 21401185), National Natural Science Foundation for Creative Research Group (grant no. 21521092), Hundred Talents Program of Chinese Academy of Sciences, and Jilin Province Science and Technology Development Project (grant nos. 20170101092JC and 20160520006JH) is acknowledged.



REFERENCES

(1) Lasave, J.; Koval, S.; Dalal, N. S.; Migoni, R. L. Phys. Rev. Lett. 2007, 98, 267601. (2) Xue, D.; Ratajczak, H. Chem. Phys. Lett. 2003, 371, 601−607. (3) Mastalerz, M. Nat. Chem. 2013, 5, 810−811. (4) Desiraju, G. R. J. Am. Chem. Soc. 2013, 135, 9952−9967. (5) Xue, D.; Zhang, S. J. Phys. Chem. A 1997, 101, 5547−5550. 3183

DOI: 10.1021/acs.cgd.7b00139 Cryst. Growth Des. 2017, 17, 3178−3184

Crystal Growth & Design

Article

(33) Chandran, S.; Paulraj, R.; Ramasamy, P. Mater. Chem. Phys. 2017, 186, 365−371. (34) Gavette, J. V.; Mills, N. S.; Zakharov, L. N.; Johnson, C. A.; Johnson, D. W.; Haley, M. M. Angew. Chem. 2013, 125, 10460−10464. (35) Van Mooy, B. A. S.; Krupke, A.; Dyhrman, S. T.; Fredricks, H. F.; Frischkorn, K. R.; Ossolinski, J. E.; Repeta, D. J.; Rouco, M.; Seewald, J. D.; Sylva, S. P. Science 2015, 348, 783−785. (36) Kababya, S.; Gal, A.; Kahil, K.; Weiner, S.; Addadi, L.; Schmidt, A. J. Am. Chem. Soc. 2015, 137, 990−998. (37) Lee, S.; Wi, H. S.; Jo, W.; Cho, Y. C.; Lee, H. H.; Jeong, S. Y.; Kim, Y.; Lee, G. W. Proc. Natl. Acad. Sci. U. S. A. 2016, 113, 13618− 13623. (38) Pandian, T. S.; Cho, S. J.; Kang, J. J. Org. Chem. 2013, 78, 12121−12127. (39) Tang, E.; Di Tommaso, D.; de Leeuw, N. H. J. Chem. Phys. 2009, 130, 234502. (40) Sun, S. T.; Jiang, L.; Liu, J. W.; Heine, N.; Yacovitch, T. I.; Wende, T.; Asmis, K. R.; Neumark, D. K.; Liu, Z. F. Phys. Chem. Chem. Phys. 2015, 17, 25714−25724.

3184

DOI: 10.1021/acs.cgd.7b00139 Cryst. Growth Des. 2017, 17, 3178−3184