Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2024 Feb 5;121(7):e2320201121. doi: 10.1073/pnas.2320201121

The elementary reactions for incorporation into crystals

Rajshree Chakrabarti a,1, Lakshmanji Verma a,1, Viktor G Hadjiev b, Jeremy C Palmer a, Peter G Vekilov a,c,2
PMCID: PMC10873555  PMID: 38315836

Significance

Crystals are essential structural elements in living organisms and rocks and crucial constituents of the technologies that enable modern civilization. We unravel the mechanism of the chemical reaction between incoming molecules and the unique sites on a crystal surface that receive them, the kinks, which has remained elusive and subject to speculation for over 60 y. The presented paradigm of two-step incorporation may shed light on how minor solution components define the elaborate morphologies of natural crystal formations and guide the search for solvents and additives that stabilize the intermediate state to slow down the growth of, for instance, undesired polymorphs.

Keywords: crystal growth, molecular mechanisms, organic materials, solvent interactions, activation barrier

Abstract

The growth rates of crystals are largely dictated by the chemical reaction between solute and kinks, in which a solute molecule severs its bonds with the solvent and establishes new bonds with the kink. Details on this sequence of bond breaking and rebuilding remain poorly understood. To elucidate the reaction at the kinks we employ four solvents with distinct functionalities as reporters on the microscopic structures and their dynamics along the pathway into a kink. We combine time-resolved in situ atomic force microscopy and x-ray and optical methods with molecular dynamics simulations. We demonstrate that in all four solvents the solute, etioporphyrin I, molecules reach the steps directly from the solution; this finding identifies the measured rate constant for step growth as the rate constant of the reaction between a solute molecule and a kink. We show that the binding of a solute molecule to a kink divides into two elementary reactions. First, the incoming solute molecule sheds a fraction of its solvent shell and attaches to molecules from the kink by bonds distinct from those in its fully incorporated state. In the second step, the solute breaks these initial bonds and relocates to the kink. The strength of the preliminary bonds with the kink determines the free energy barrier for incorporation into a kink. The presence of an intermediate state, whose stability is controlled by solvents and additives, may illuminate how minor solution components guide the construction of elaborate crystal architectures in nature and the search for solution compositions that suppress undesirable or accelerate favored crystallization in industry.


Crystals endow natural and synthetic materials with essential structures and functions (18). The structural and functional utilities of crystals are, to a large extent, defined by their sizes and shapes (9, 10). In turn, the crystal sizes and shapes are determined by the crystal growth rate and its anisotropy related to the distinct structures of the crystal faces (8, 1113). How fast crystal faces grow is largely dictated by the slow chemical reaction between solute molecules and specifically structured crystal surface sites, the kinks (11, 1417), but the respective mechanism, i.e., the characteristic timescales and lengthscales, possible intermediates and their stabilities, remains elusive (18). The incorporation reaction rate is substantially slower than the rate of molecular supply to the kinks [referred to as the diffusion limit (19, 20)] and announces the presence of an activation barrier for incorporation. Since incorporation into kinks from a gas is nearly barrier-free, this barrier was ascribed (11, 17) to solvate shells that coat the kinks and envelope the solutes (7, 2128). This model implies that the reaction for incorporation into kinks comprises breaking of several solute–solvent bonds and instituting new bonds with the molecules forming the kink. Fundamental mechanistic details on the pathway, along which bonds sever and rebuild, the associated energy relief, and potential long-lived intermediates, remain missing.

Indeed, this mechanism could not be probed before as molecular incorporation into kinks is often coupled to the complex kinetics of transport from the solution to the growth sites that may include adsorption on the terraces between steps, diffusion along the crystal surface, and desorption back to the solution (11, 17, 2931). The molecular exchanges between solution, crystal surface, and kinks motivate multiparameter kinetic laws, in which the kinetics of incorporation into kinks convolve with the kinetics of the adjacent processes (32). In the absence of experimental data on the transport to the steps and reaction at the kinks, computer simulations of the growth of ionic crystals from aqueous solutions have suggested that the incorporation of ions into kinks may be a multistep process associated with step-wise dehydration of the ions (33, 34); interestingly, analogous simulations of the same system have yielded divergent intermediate steps (33, 35, 36). These computational studies shared an assumption of direct access of the incoming ions from the solution to the kinks, which was unsupported by experiment and may be unphysical (31). Still, if taken at face value, their results would concur with the classical expectation that the strength of the solute–solvent bonds governs the free energy barrier for incorporation into kinks (28, 37).

To elucidate a potential intermediate state that may be shifted by only fractions of a nanometer from the final molecular incorporation site, exist for times of order of nanoseconds, and thus be inaccessible to the current direct experimental methods, we employed as reporters four solvents with distinct structures and functions. To expose the reaction at the kinks we chose fully organic solute and solvent pairs, in which the solute molecules are expected to reach the steps directly from the solution (31, 38). This strategy is distinct from many solution crystallization systems studied to date, which employed fully or partially aqueous solvents (27, 39, 40) that favor the multistep surface diffusion pathway (31). We explored the growth of etioporphyrin I. Similarly to crystals of other porphyrins, etioporphyrin I crystals have low symmetry, which is conducive of appealing electronic properties (41, 42).

Results

How Etioporphyrin I Crystals Grow.

Etioporphyrin I comprises four pyrrole residues linked by methine groups into a flat porphyrin ring, decorated by alternating four ethyl and four methyl groups (Fig. 1A). In the known unsolvated crystal form with symmetry P1, the etioporphyrin I molecules arrange in columnar stacks supported by π–π bonds and oriented roughly along the [110] direction (Fig. 1B) (42). To illuminate how solute molecules incorporate into crystals, we compare the kinetics of crystallization of etioporphyrin I from three alcohols with increasing alkyl chain length, butyl, hexanyl, and octanyl, and dimethyl sulfoxide (DMSO), a smaller molecule with two short aliphatic residues. This comparison is substantiated by the identical P1 structures of the crystals that grow from the four solvents (SI Appendix, Figs. S1 and S2) (43). The UV–vis absorption spectra of etioporphyrin I in these solvents feature a prominent Soret band, at ca. 400 nm, and four Q bands in the wavelength range 475 to 630 nm (Fig. 1C) attributed to distinct transitions of the conjugated π electron systems of the constituent pyrroles (4446). The Soret and the Q bands are similarly structured in the three alcohols and DMSO (Fig. 1C) and signify that the dominant solute species in the four solvents is a monomer (4749).

Fig. 1.

Fig. 1.

Etioporphyrin I crystals and solutions. (A) The etioporphyrin I molecule. (B) Stacks of etioporphyrin I molecules in the crystal in P1 space group; Cambridge Structural Database REFCODE WOBVUF ref. 42; N is drawn in blue, C, in charcoal, and H, in silver. (C) UV–vis absorption spectra of etioporphyrin I dissolved in DMSO, butanol, hexanol, and octanol. (D) Scanning electron micrograph of an etioporphyrin I crystal; the prominent crystal faces are labeled. (E) In-situ atomic force microscopy (AFM) images of a (010) face of etioporphyrin I growing from solutions in DMSO, butanol, hexanol, and octanol.

(010), (011), and (101) faces are prominent in the habit of etioporphyrin I crystals grown from the three alcohols and DMSO (Fig. 1D). Typically, a (010) face orients parallel to the substrate (SI Appendix, Figs. S1 and S3) and affords a line of sight to atomic force microscopy (AFM). In situ AFM images (Fig. 1E) in supersaturated solutions, where the solute concentration C exceeds the solubility Ce (the concentration at which a crystal and the solution are at equilibrium), reveal that the (010) faces comprise unfinished layers wound into spirals that originate at the outcrops of screw dislocations, similarly to numerous other crystals growing from solution (50, 51).

The Mechanism of Molecular Ingress into Kinks.

The crystals grow as solute molecules associate to kinks, located along the edges of the unfinished layers, the steps (15, 52, 53). The crystal anisotropy enforces distinct velocities of the steps that propagate in different directions such that steps in the [100] or a direction grow fastest and reach furthest from the dislocation (Fig. 1E). In this way, the shape of the growth spiral reveals the ratios between the anisotropic step velocities; these ratios are similar in the four solvents (Fig. 1E and SI Appendix, Fig. S4) and independent of the supersaturation.

The rate of step growth on the crystal surface is a direct representation of the rate of the chemical reaction at the kinks. We used the step growth rate—or step velocity v—and its responses to thermodynamic and kinetic parameters to explore the mechanism of molecular incorporation into kinks. We measured v by time-resolved in situ AFM (30, 52, 5456). We tracked the locations of steps growing in the [100], or a, direction (Fig. 2A). Step growth was steady in time (Fig. 2B and SI Appendix, Fig. S5), and the slopes of the time correlations of the step displacements present the step velocities v. In all solvents, v increased linearly with the solute concentration C (Fig. 2C), in common with numerous other solution-grown crystals (39). The linear vC correlations indicate that the crystals grow by incorporation of the dominant solution species (30), etioporphyrin I monomers, in the four solvents. Mass preservation prescribes that v equals the product of the rate of the chemical reaction of the solute with the kinks and the contribution of one molecule to step propagation, the solute molecular size a, v=aka(C-Ce). We include in the rate constant ka the dimensionless kink density nk¯-1, which is constant and close to the thermodynamic limit of ca. 0.3 (25, 52, 57, 58) on the rough steps of etioporphyrin I in all solvents (SI Appendix, Fig. S6). Subtracting the solubility Ce accounts for the reversibility of molecular attachment. Alternatively, v has been modeled as the product of the solute flux into a kink and the volume Ω that a solute molecule contributes to the crystal, resulting in v=Ωβ(C-Ce) (11, 17), where β is the step kinetic coefficient. Comparing the two relations informs that aka=Ωβ. Dividing the slopes of the v(C) correlations (Fig. 2C) by a reveals the markedly different rate constants ka for incorporation into kinks from the four solvents (Fig. 2D).

Fig. 2.

Fig. 2.

The chemical reaction at the kinks. (A) Time resolved in situ AFM images of etioporphyrin I (010) face growing from butanol. Arrows trace the growth of a step in the [100] direction. (B) The evolutions of the step displacements during growth from butanol at printed values of C-Ce. Thirty independent measurements were averaged for each data point, and the error bars represent the respective SDs. (C) Step velocity v, determined from the slopes of the time correlations of the step displacement in (B) and SI Appendix, Fig. S5, as a function of the concentration C in the four listed solvents. Error bars represent the SDs of the slopes in (B). Numbers represent the solubilities Ce in the respective solvents in mM; see SI Appendix, Table S1 for the SDs. (D) Bimolecular rate constants ka for the reaction between incoming solute molecules and kinks evaluated from the v(C) correlations in (C); error bars represent the SDs of the slopes in (C). (E) The viscosities η of the four solvents at 28 °C, the temperature of the AFM measurements; error bars represent the SDs from SI Appendix, Table S2. (F) The product kaη for growth form the four solvents. Errors bars represent the propagation of the SDs for ka and η.

The rates of chemical reactions occurring in solution, in parallel with the reactants’ diffusivities (59), scale as the reciprocal viscosity of the solvent η-1 refs. 60 and 61. To expose the molecular behaviors that govern solute incorporation into kinks, we account for the divergent viscosities of the solvents (Fig. 2E) and use the product kaη to quantify the reaction at a kink. The values of kaη are nearly identical for crystallization from solvents in the homologous series butanol–hexanol–octanol and lower by nearly half for crystallization from DMSO (Fig. 2F).

To further illuminate the mechanism of solute incorporation into kinks, we start by examining whether etioporphyrin molecules reach the steps directly from the solution (Fig. 3A) or, as numerous solution-grown crystals (30, 6264), after adsorption on the terraces between steps followed by diffusion toward the steps (Fig. 3B). Two sets of observations affirm that during growth from the four solvents etioporphyrin I, in contrast to the studied water-based systems, prefers the direct incorporation pathway. First, in butanol, hexanol, and octanol, the velocities of steps as high as the b lattice parameter [|b| = 0.991 nm (42)], are close to those of steps of height h=2|b| (Fig. 3 C, D, and F and SI Appendix, Fig. S7). Second, v is independent of the step separation l for ls as short as a few nanometers (Fig. 3E). If solute reaches the steps via the crystal surface, the step supply field is constrained to two dimensions, which stunts the growth of closely spaced steps or steps of double height. Concurrently, analytical models of step growth mediated by surface diffusion predict that v scales with h-1 and sharply slows down at short l (31, 32). Experimental observations reveal that double-height steps that feed via the surface grow slower than single-height steps and may split into two single-height steps owing to unequal supply to the top and bottom layers (63, 65). By contrast, if the steps feed directly from the solution, the supply field is three-dimensional and abundant for closely spaced or twinned steps. Closed-form expressions for this growth mode predict negligible v(h) and weak v(l) correlations (11, 17, 31).

Fig. 3.

Fig. 3.

The pathway of a solute molecule into a kink. (A and B) Schematic representations of two pathways from solution to steps. (A) Solute molecules reach the steps directly from the solution. (B) Solute molecules adsorb on the crystal surface and diffuse toward the steps. (C) Time-resolved AFM images of the growth of single (silver arrows) and double (green arrows) height steps at C = 0.23 mM in hexanol. (D) The evolution of the surface profile along the dotted line in (C). The profiles at 21, 43, and 68 s are shifted vertically for clarity. Double-height steps (green arrows) advance over lengths similar to those of single-height steps (silver arrows). (E) The step velocity in the four solvents does not correlate with the step separation l. C = 0.23 mM in hexanol, 0.09 mM in butanol, 0.33 mM in octanol, and 0.25 mM in DMSO. The averages of 15 measurements for each l interval, represented by horizontal bars, are shown. Vertical bars represent the SDs of the groups of measurements. (F) Comparison of the velocities v of single, h=|b|, and double, h=2|b|, height steps in three solvents. The averages of 10 double and 10 single height steps are shown. Error bars represent the respective SDs. The values of CCe during these measurements are listed in the plots. No double-height steps were observed during growth from DMSO. (G) Schematic of the semi-spherical supply field for direct incorporation of solute into a kink. (H) Schematic of the free energy landscape along the classical one-step reaction pathway of incorporation of an etioporphyrin I molecule from the solution into a kink. The values for the standard free energy of crystallization ΔGo are from ref. 43. (I) Schematic of the free energy landscape along the two-step pathway of incorporation suggested by the free energy barriers for incorporation into kinks ΔG in the four solvents, evaluated from the ratios of the measured rate constants for incorporation into kinks ka (Fig. 2D) to their diffusion limits (SI Appendix, Table S4).

Factors That Determine the Rate Constant for Incorporation into Kinks.

We represent the bimolecular rate constant ka for the reaction between a kink and an incoming solute molecule as ka=k0exp-ΔG/kBT, where k0 is the rate constant of a reaction uninhibited by an activation barrier and proceeding at the diffusion limit, ΔG is the free energy barrier for incorporation of solute molecules into kinks, kB is the Boltzmann constant, and T is temperature (66). Molecules approach kinks only from half the space above the crystal (Fig. 3G), and the kinks can be assumed immobile and separated along the steps, on the average, by n¯k molecules (29). These three constraints modify the Smoluchowski-type expression (67) for k0 to k0=2πDrNA/n¯k (D, etioporphyrin I diffusivity in the respective solvent; r 0.5 nm, reaction radius; and NA, the Avogadro number). In turn, D=kBT/6πηa. The constraints on a reaction at a kink and the elevated viscosities of the organic solvents depress the values of k0 to order 108 M−1s−1 (SI Appendix, Table S4), somewhat lower than the typical values of about 1010 M−1s−1 for diffusion-limited reactions of small molecules dissolved in water (66). The relations for ka, k0, and D expectedly imply that ka scales as η-1. Importantly, given that the kink densities in the four solvents are equal to n¯k’s thermodynamic limit (52, 58) (SI Appendix, Fig. S6), which is solvent-independent (29), the preexponential factor in the product kaη is solvent-independent and the barrier for solute incorporation into kinks ΔG solely regulates kaη.

The found correlation between ΔG and kaη affords an opportunity to apply the kaη data to illuminate the mechanism that governs ΔG. Classical crystal growth theories assign ΔG to a transition state, in which the crucial solute–solvent bonds are broken, but the solute has not yet bound to the kink (11, 17); as the solvent–solute bonds are broken, this complex is nearly identical in all solvents. Within this classical scenario, the activation free energy for incorporation would be instituted by the strength of the solute–solvent bonds and is greater in solvents that bind strongly with the solute (Fig. 3H).

To test the correlation between ΔG and the strength of the solute-solvents bonds, we use that etioporphyrin I crystallizes as identical P1 crystals (SI Appendix, Figs. S1 and S2) from all four solvents (43). In consequence, the equilibrium free energies of crystallization ΔGo, evaluated from the respective solubilities (43), characterize the relative strengths of the solute–solvent interactions in each solvent. A correlation between ΔGo and ΔG would exemplify the venerable Berthollet rule, according to which crystals grow faster from solvents, in which their solubility is lower (68). Etioporphyrin I, however, defies both the Berthollet rule and the classical understanding of activation barriers for solution crystallization. It binds weakly to butanol, and its interactions with DMSO, hexanol, and octanol are stronger and nearly equal (43). The classical principles concertedly predict lowest activation barrier and fastest growth from butanol solutions and comparable barriers and growth rates from DMSO, hexanol and octanol (Fig. 3H). The AFM results on the step velocities belie both parts of this prediction after accounting for the divergent viscosities of the solvents (Fig. 2F).

The discrepancy between the strength of the interactions of etioporphyrin I with the solvents, manifest as ΔGo, and the activation barriers for crystallization ΔG advocates for two sequential elementary reactions of incorporation into kinks, separated by a metastable state that a solute molecule occupies in reaction one, before it invades the kink in reaction two. This intermediate complex locates along the direct access route of a molecule from the solution to a kink (Fig. 3A) and likely represents a state in which the solute molecule is bound, but only partially, to the molecules comprising the kink. The stability of the intermediate state, regulated by the strengths of the preliminary bonds with the kink and interactions with the solvent, dictates the height of the barrier that the solute molecule overcomes as it relocates to the kink (Fig. 3I). The values of ΔG in the four solvents imply that DMSO strongly stabilizes the intermediate state, and the contributions of the three alcohols are weaker and close (Fig. 3I).

A Microscopic View of the Reaction at the Kinks.

For a microscopic view of the pathway from the solution into a kink, we performed all-atom molecular dynamics (MD) simulations. We monitored the ingress of an etioporphyrin I molecule into a kink in contact with each of the four solvents (Fig. 4 AC and Movie S1). Before an incoming molecule approaches a kink, it freely rotates. Further along the reaction pathway, the rotations are constrained by interactions with the molecules in the kink and not by the model assumptions (Fig. 4 A–C and Movie S1). We examined the normal distance of the molecule’s center of mass from the surface of the kink, zCOM, and the number of solvent molecules in the kink cavity, Nsolv. The respective MD simulation trajectories reveal that incoming molecules stall at an intermediate state at ca. zCOM1.1nm (Fig. 4 B and D), where they associate with more solvent molecules than in the kink (Fig. 4E). In all solvents, the potential of mean force FNsolv,zCOM along the kink access route features two minima: a global minimum at zCOM0.7nm, which corresponds to a molecule in the kink, and a metastable minimum at zCOM1.1nm (Fig. 4F and SI Appendix, Fig. S9), corresponding to the transient intermediate state observed in the MD trajectories (Fig. 4 B, D, and E).

Fig. 4.

Fig. 4.

Microscopic view of the ingress of an etioporphyrin I molecule into a kink. (AC) Representative snapshots of an etioporphyrin I molecule in the solution near a kink (A), at the intermediate state (B), and at its final location in the kink (C). Classical MD simulation results. The green sphere in (B) defines our choice of the kink cavity, i.e., the volume where the dynamics of the solute molecule that partake in the reaction between a solute molecule and a kink are evaluated. Solvent molecules are omitted for clarity. The kink is viewed along the unfinished molecular row at the step edge. The direction of step growth is from Left to Right. Etioporphyrin I molecules in the crystal lattice are shown in gray. The lattice planes of etioporphyrin I molecules in front and behind the plane, which hosts the kink, are omitted. In the incoming etioporphyrin I molecule, C atoms are shown in teal, N in blue, and H in silver. (D and E) A representative MD simulation trajectory of an incoming etioporphyrin I molecule displayed in terms of the normal distance zCOM from the center of mass of an incoming molecule to that of the molecule at the bottom surface of a kink in (D) and the number of solvent molecules Nsolv in the kink cavity [shown in green in (B)] in (E). (F) Two-dimensional potential of mean force profiles F(Nsolv,zCOM) for incorporation of a solute molecule into a kink from the four solvents. F was computed from well-tempered metadynamics simulations. The locations of the three states shown in (AC) are indicated in the DMSO plot.

The intermediate complex represents the stepping stone from which a molecule launches to surmount the activation barrier and relocate to the kink. The saddle point between the two minima hosts the transition state for transfer from the intermediate state into a kink and the energy difference between the intermediate state and the saddle point represents the activation barrier for molecular incorporation into kinks ΔG (Fig. 4F and SI Appendix, Fig. S9). In the four solvents, ΔG is between 11 and 14 kJ mol−1 (Fig. 4F and SI Appendix, Fig. S9 and Table S5), close to the experimental estimates (Fig. 3I and SI Appendix, Table S5), although the statistical uncertainty in the computational values (ca. 2 kJ mol−1, SI Appendix, Fig. S8 and Table S5) prevents precise quantitative comparisons.

How the Solvent Stabilizes the Intermediate State.

To rationalize the relative stabilities of the intermediate states in the four solvents, we turn to the dynamics of the solvent molecules occupying the ca. 0.4-nm gap between an incoming etioporphyrin I molecule and the lower molecule of the kink (Fig. 4B). Entropy maximization requires that this gap fills with solvent. This gap is populated by about two DMSO, one butanol, 2/3 hexanol, and 1/2 octanol molecules (Fig. 4F), which consistently comprise four -CH3 or -CH2- residues. The distributions of the centers of mass of the solvent molecules along the zCOM coordinate reveal that similar ranges of zCOM values are occupied, suggesting that the translational freedoms of the four solvents are similar. A substantial population of the smaller DMSO molecules retain significant orientational freedom (Fig. 5A), whereas the larger molecules of butanol, hexanol, and octanol mostly rest parallel to the lower kink surface (Fig. 5 A and B). We propose that the elevated orientational freedom and the greater number of DMSO molecules entropically stabilize the intermediate state in that solvent and also bolster the search for stronger bonds of the solvent with the etioporphyrin I molecules that frame the gap. The enhanced stability of the intermediate state during growth from DMSO increases ΔG and dictates the slower rate of growth from this solvent (Figs. 2F and 3I). The similar dynamics of the three alcohols in the gap (Fig. 5A) and their similar chemistries underlie the close stabilities of the intermediate complexes for incorporation from these solvents and the near identical values of the respective ΔG and kaη products (Fig. 2F).

Fig. 5.

Fig. 5.

Solvent dynamics in the intermediate state. (A) Distributions P(cos θ, zCOM) of the molecules of the four solvents trapped in the gap between an incoming solute and the base of a kink, Fig. 4B, along the vertical coordinate zCOM and the angle θ defined in (B). P(cos θ, zCOM) is normalized to unity far from the crystal surface. (B) Definition of the angle θ between the vertical axis z and a vector, shown in red, that characterizes the orientation of each solvent. cosθ=0 when θ=π/2 and a solvent molecule rests parallel to the crystal surface. Etioporphyrin I crystal surface is shown in cyan. In the solvents, C is shown in gray, O, in red, S, in yellow, and H, in white.

Discussion

Crystallization of organic, inorganic, biomolecular, and colloid materials commonly involves the formation of several contacts between a solute and a kink. It is likely that two-step incorporation with an intermediate state will be a part of the respective mechanisms of access to kinks. For solutes that reach the kinks by the surface diffusion pathway, adsorption on the crystal surface leaves most of the solute valences saturated with solvent and allows the formation of a similar intermediate state as the solute interacts with a kink. It is feasible that for compounds with structures more complex than those of their respective solvents more than one intermediate state may exist, as suggested for the association of Ba2+ ions to steps on barite crystals (69). For other compounds, a long-living intermediate state may recruit more than one of the incoming solute molecules or ions (34). Rare instances of one-step solute incorporation into kinks are likely limited to high-symmetry solutes that are comparable in size to the solvent.

The breaking of the preliminary bonds between a growing crystal and an incoming molecule in the second stage of the reaction is necessary to accomplish incorporation into kinks but is highly counterintuitive. On the other hand, it is akin to backtracking, observed for amyloid fibril growth, in which an incoming peptide chain docks to the fibril tip by contacts that are distinct from those in the fibril bulk (7076). The shared features of molecular incorporation of the relatively rigid (excepting the methyl and ethyl side groups) etioporphyrin I molecules and the flexible peptide chains suggest that backtracking may be a common feature of the mechanism of aggregation in solution.

The concept of an intermediate state for molecular incorporation may illuminate poorly understood examples of additive-driven faster crystal growth (7, 26, 77). Expectations, based on myriad prior studies, state that the additives would delay growth by binding to specific crystal surface sites (17). The acceleration of calcite growth by a line of polypeptides was attributed to reduced activation barriers due to “perturbation of the local water structure” by the peptides (7). The low concentration of polypeptides, from 10−7 to 10−5 M, needed to invoke growth acceleration, suggests that the peptides do not perturb the solvation of the calcium and carbonate ions in the solution, which would require additive amounts comparable to the solute concentrations of order of 10−4 M (50). On the other hand, the peptide effect on kinetics is consistent with destabilization of a potential intermediate state at the kinks, which exclusively localize at the crystal–solution interface at much lower concentrations. In another example, acetone, added to crystallization solutions of the peptide hormone insulin at concentrations up to 20 vol.%, increased multiple-fold the crystal solubility (78), suggesting that acetone weakens the crystal contacts. Instead of suppressing crystallization, however, acetone substantially boosted the rate of growth of the insulin crystals (26), another violation of the Berthollet rule (68). A possible explanation of the kinetic effect of acetone is that this additive weakens the contact that support an intermediate state for incorporation into kinks, which destabilizes it and lowers the activation barrier for association to the kinks.

Conclusions

The kinetics of etioporphyrin I crystal growth in four solvents disconnect from the respective crystallization thermodynamics and indicate that the incorporation of molecules into crystal growth sites is mediated by an intermediate state. MD simulations support the presence of an intermediate state, in which an incoming solute molecule forms preliminary contacts with the molecules from the kink. The breaking of the preliminary bonds of the solute with the solvent and the kink that stabilize this intermediate state constitutes the free energy barrier for incorporation into kinks.

The model of an intermediate state for molecular incorporation, in concert with the ubiquitous two-step crystal nucleation (7983), suggests that two-step transitions may be common for phase transformations defined by multiple order parameters (84). The power of solvents and additives that uphold or undermine the intermediate state may illuminate how complex environments define the elaborate morphologies of natural crystal formations and guide the search for solvents and additives that suppress or accelerate crystallization during the synthesis of pharmaceuticals, fine chemicals, and advanced functional and structural materials.

Materials and Methods

Solution Preparation.

Etioporphyrin I was purchased from Santa Cruz Biotechnology Ltd. The solvents, octanol (anhydrous, ≥99%), butanol (anhydrous, ≥99%), hexanol (anhydrous, ≥99%), and DMSO (anhydrous, ≥99%) were bought from Sigma Aldrich. Deionized (DI) water for cleaning glass vials was produced by a Millipore reverse osmosis–ion exchange system (Rios-8 Proguard 2—MilliQ Q-guard). Solutions of etioporphyrin I in each solvent were prepared by dissolving etioporphyrin I in respective solvents at 70 °C in glass vials. The supersaturated solutions were filtered, and concentrations were measured using UV–vis spectroscopy. Solutions of desired concentration for bulk crystallization and in-situ AFM studies were made by subsequent dilution with respective solvents.

Viscosity.

The viscosity of organic solvents varies non-linearly with temperature. Solvent viscosities at 28 °C were determined by linear interpolation between two bracketing temperatures 25 °C and 30 °C or 35 °C (8590). The solvent viscosities at 28 °C are reported in SI Appendix, Table S1.

Scanning Electron Microscopy.

For bulk crystal characterization with SEM, supersaturated solutions of etioporphyrin I (C = 0.5mM) in different solvents were passed through a 0.22-µm PTFE filter and kept in clean 20-mL capped borosilicate glass vials in an incubator. In addition, 15-µm diameter clean glass coverslips scratched at the center were kept at the bottom of the glass vials in contact with respective supersaturated solutions. Crystals of etioporphyrin I were made by slow cooling at the rate of 5 °C/15 min, upon reaching room temperature, the vials were kept undisturbed overnight. After 1 to 2 d, etioporphyrin I crystals nucleated on glass coverslips were taken out, rinsed with octanol to remove loose crystals, air dried, mounted on metal stubs with carbon tape and coated with 10 to 20 nm gold. SEM images (as in SI Appendix, Fig. S1) were collected at a system vacuum of 2 × 10−5 mbar pressure and EHT 3 kV.

UV–Vis Spectroscopy.

Weighted amounts of etioporphyrin I was completely dissolved in known volumes of solvents in borosilicate glass vials. The glass vials were left at 70 °C for several days with occasional stirring until complete dissolution. The stock solutions were filtered and diluted serially to achieve the desired concentrations. UV–vis absorbance spectra were recorded with DU Spectrophotometer 800 (Beckman) in the wavelength range 200 to 800 nm (Fig. 1C). For each porphyrin, distinct Soret and Q bands were clearly identifiable. The correlations between optical density at the wavelength of maximum absorbance in the respective Q bands and the concentration are linear (91). For experimental statistics, all measurements were carried out in triplicate, in independently prepared solutions. The extinction coefficients for each porphyrin–solvent pair were determined from the respective slopes of absorbance vs. concentration at a given wavelength (91).

Powder X-ray Diffraction.

Powder X-ray diffraction (PXRD) patterns (SI Appendix, Fig. S2) were obtained using BB Rigaku Samartlab X-ray diffractometer with primary monochromatic radiation CuKαI of wavelength 1.54056 Å. Powder patterns were collected by laying finely ground crystals on a quartz sample holder. Data were collected in the 2θ range 5 to 40°, with a 2θ step of 0.015˚ and 2s/step speed. A reference powder pattern for etioporphyrin I was computed using Mercury 3.6 (CCDC) software using single crystal data from the Cambridge Structural Database (42).

Crystal Face Indexing.

The crystal faces (SI Appendix, Fig. S3) were indexed on a Bruker D8 Venture diffractometer equipped with a charge-coupled device (CCD) detector using graphite-monochromated Cu Kα radiation (λ = 1.54056 Å) and an APEX-3 face indexing plug-in.

In Situ Monitoring of the Etioporphyrin I Crystal Evolution.

We used the multimode atomic force microscope (Nanoscope VIII and IV) from Bruker for all step velocity measurements. AFM images were collected in tapping mode using Olympus TR800PSA probes (Silicon nitride, Cr/Au coated 5/30, 0.15 N/m spring constant) with a tapping frequency of 32 kHz. Image sizes ranged from 2 μm to 20 μm. Due to the fast growth of etioporphyrin I layers on the (010) face, the scan rates were between 2.50 and 6.10 s−1 depending on the scan size. Height, amplitude, and phase imaging modes were employed. The captured images contained 128 or 256 scan lines based on the scan aspect ratio, at angles depending on the orientation of the monitored crystal. The temperature in the fluid cell reached a steady value of 28 ± 0.1 °C within 15 min of imaging (92). This value was higher than room temperature (ca. 22 °C) owing to heating by the AFM scanner and laser.

Pure etioporphyrin I crystals were grown on glass disks as described above. The density of attached etioporphyrin I crystals was monitored under an optical microscope. We ensured similar crystal density for all samples to minimize potential depletion of solutes due to a high crystal number. The glass slides were mounted on AFM sample disks (Ted Pella Inc.), and the samples were placed on the AFM scanner. Etioporphyrin I solutions in respective solvents with desired concentrations were prepared less than 2 h in advance. This solution was loaded into the AFM liquid cell using 1-mL disposable polypropylene syringes (Henck Sass Wolf), tolerant of organic solvents. After loading, the system was left at rest for 10 min to equilibrate thermally. The crystal edges and surface morphology were used to identify the (010) crystal surface. Etioporphyrin I crystals are very dynamic hence the surface features changed rapidly in the presence of the growth solution. We set the scan direction approximately parallel to the [100] crystallographic direction and AFM images were collected for 30 min. The solution in the AFM fluid cell was refreshed every 10 min to maintain constant concentration. The evolution of the etioporphyrin I crystal surface was characterized by the velocity of growing steps v. To determine v, we monitored the displacements of 8 to 15 individual steps with a measured step height h = 1.17 ± 0.07 nm (Fig. 2A and SI Appendix, Fig. S4). Between 5 and 10 measurements were taken for each individual step with time. The step displacements were plotted as functions of time (Fig. 2B and SI Appendix, Fig. S5). The slope of these plots was used to determine the step velocity v (Fig. 2C) at a given supersaturation.

Observation of Kinks along the Steps.

The kink densities on the (010) face of etioporphyrin I were determined using high-resolution images obtained by Cypher ES environmental AFM (Asylum Research), with Bruker SNL 10 AFM tips. In situ monitoring of small step segments on (010) face of etioporphyrin I in the four solvents at their respective solubilities show abundant kink densities in all the four solvents chosen for the study. The lateral resolution of these images, ca. 2 nm, is insufficient to resolve individual kinks (50, 52, 57, 93, 94). The lack of straight step segments, however, suggests that the average separation between kinks is comparable to the resolution limit, ca. 3 molecules, equivalent to kink density n¯k-1 close to the thermodynamic limit of 0.3 (SI Appendix, Fig. S6).

Determinations of Activation Free Energies.

The activation energy of solute molecule incorporation into the growth site was determined from the step velocities. The rate constant for solute incorporation may be presented as ka=koexp(ΔG/kBT), where k0=2πDrNA/n¯k combines the rate constant for reaction between a kink and a solute molecule only limited by diffusion with the constant kink density n¯k-1. The ka values were determined from the slopes of the step velocity correlation with solute concentration. The k0, values are determined from the known values of solvent viscosity η, kink density n¯k-1, and assuming r = 0.5 nm. The resulting values of k0 are listed in SI Appendix, Table S4. The values of activation free energy ΔG are listed in SI Appendix, Table S5.

Molecular Dynamics Simulations.

To investigate the attachment/detachment of etioporphyrin I molecules to etioporphyrin I crystals in DMSO, octanol, hexanol, and butanol, we carried out all-atom molecular dynamics (MD) simulations. We employed GROMACS 5.1.5 (95). Etioporphyrin I and the solvent species were modeled with the generalized AMBER force field (96, 97) (GAFF), using the ACPYPE package (98) to generate topology files for GROMACS. Partial charges were derived from B3LYP/6-31G(d) density functional theory calculations using the restrained electrostatic potential (RESP) method implemented in the PyRED program (99). The equations of motion were propagated using the leapfrog integration algorithm with a 2-fs timestep. All the bond lengths were constrained using the LINCS (100) algorithm to remove fast degrees of freedom associated with bond vibrations and facilitate the use of a large time step. Van der Waals interactions were truncated using a 1-nm cutoff, and periodic boundary conditions were applied in all three dimensions. Coulomb interactions were handled with the particle mesh Ewald method (101), using a cutoff of 1 nm for the real-space contributions and an error tolerance of 10−5.

Unit cell information for etioporphyrin I was obtained from the Cambridge Structural Database REFCODE WOBVUF (42). The unit cell parameters predicted by the GAFF model were evaluated by performing a 10-ns MD simulation of a 4 × 4 × 4 supercell of etioporphyrin I and computing averages over configurations sampled every 10 ps over the last 5 ns. During the MD simulation, the temperature was maintained at 300 K and isostress conditions of 1 bar were imposed using a Bussi–Donadio–Parrinello thermostat (102) (0.1-ps time constant) and an anisotropic Parrinello–Rahman barostat (103) (2-ps time constant), respectively. Unit cell parameters from the GAFF model were found to deviate from the literature data by less than 3% (SI Appendix, Table S6).

The kink site on the (010) face of etioporphyrin I was modeled by replicating the unit cell to create a 7×2×5 supercell, which was then rotated to align the (010) face with the z-axis of the simulation box, preserving the triclinic periodicity of the crystal. Next, the z-dimension of the simulation box was enlarged to create a crystal slab with top and bottom (010) faces exposed to vacuum; the z-dimension of the simulation box was chosen to be at least 3 times that of the crystal slab thickness to minimize finite size effects in the simulations. A portion of the top layer of the crystal slab was then cleaved to expose an obtuse kink in the [100] direction on a [001] facing step. The final kink model consisted of 52 etioporphyrin molecules in the crystal and one tagged etioporphyrin occupying the kink site. Last, the space above and below the slab was filled with molecules of the desired solvent (SI Appendix, Table S7).

Energy minimization of the solvated kink models was performed using the steepest descent algorithm with a stopping criterion of 1,000 kJ mol−1nm−1 for the maximum force on an atom. Each model system was then equilibrated by performing a 20-ns MD simulation in the canonical (NVT) ensemble at 300 K using a Bussi–Donadio–Parrinello thermostat (102) with a 0.1-ps time constant. The NVT simulation was followed by an additional 20 ns of equilibration in the isothermal–isostress (NPzT) ensemble at 300 K and 1 bar, coupling the z-dimension of the simulation box normal to the surface to a Parrinello–Rahman barostat (103) with a 2-ps time constant while fixing the x- and y-dimensions.

The equilibrated systems were used to initialize well-tempered metadynamics (WTmetaD) (104) simulations to study the attachment/detachment of a tagged etioporphyrin molecule at the kink site. The WTmetaD simulations were performed in the NPzT ensemble at 300 K and 1 bar using the PLUMED 2.4.3 free energy plug-in for GROMACS. The choice of collective variables (CVs) was motivated by a study, which showed that both slow solvent and solute degrees of freedom must be sampled to properly converge potential of mean force (PMF) calculations for the detachment/attachment of Cl and Na+ at kink sites on NaCl crystals (36). Accordingly, we used two CVs: the normal distance between the kink site and the center of mass (COM) of the tagged etioporphyrin molecule, zCOM, and the number of solvent molecules in the kink cavity, (Nsol). The first CV was defined as the normal distance from a reference carbon in the etioporphyrin I crystal located on the bottom surface of the kink. The second CV was defined using a stretched and shifted rational switching function to obtain a CV suitable for biasing in WTmetaD:

Nsolv=1nhjsrij-sdmaxs0-sdmaxHdmax-rij,

where

sr=1,r-d0r001-r-d0r0m1-r-d0r02m,r-d0r0<0,

rij is the distance between solvent heavy atom j and the reference carbon atom on the kink’s surface, nh is the number of heavy atoms in a single solvent molecule, H is the Heaviside function, r0=0.9 nm, d0=0.85 nm, dmax=0.90 nm, and m=12. This CV was implemented using the DENSITY and INSPHERE functions in PLUMED and provides a continuous approximation to the number of solvent molecules within a 0.9-nm radius of the kink site.

The WTmetaD simulations were performed using a bias factor of 15 and a 1-ps interval between the deposition of repulsive Gaussian potentials. The initial height of the Gaussian potentials was 1 kJ/mol, and Gaussian widths of 0.02 nm and 0.15 molecules were chosen for zCOM and Nsolv, respectively. To limit the explored CV region, upper quadratic walls were placed at zCOM=1.7 nm and Nsolv=11 molecules using spring constants of 5,000 kJ/(mol−1 nm−2) and 5,000 kJ mol−1nm−2, respectively. To prevent lateral drift away from the kink site, an auxiliary upper quadratic wall was also placed on the in-plane (x-y) distance between the reference carbon atom on the kink’s surface and the center of mass of the tagged etioporphyrin molecule at 0.4 nm using a spring constant of 2,000 kJ mol−1nm−2. For each solvent, three independent WTmetaD simulations were run for 5.5 to 9.5 µs (SI Appendix, Table S6) to obtain converged estimates of the PMF characterized by the two CVs. The PMF from each WTmetaD simulation was calculated from the generated time series data using the reweighting method proposed in ref. 105. The final PMF reported for each solvent was obtained by averaging over the results from the three independent simulations, using the SD to estimate statistical uncertainties.

Supplementary Material

Appendix 01 (PDF)

Dataset S01 (XLSX)

Movie S1.

The ingress of an etioporphyrin I molecules into a kink. Successive snapshots of an etioporphyrin I molecule in the solution near a kink (ass seen in Fig. 4a), at the intermediate state (as in Fig. 4b), and at its final location in the kink (as in Fig. 4c). Classical MD simulation results. Solvent molecules are omitted for clarity. The kink is viewed along the unfinished molecular row at the step edge. The direction of step growth is from left to right. Etioporphyrin I molecules in the crystal lattice are shown in grey. The lattice planes of etioporphyrin I molecules in front and behind the plane, which hosts the kink, are omitted. In the incoming etioporphyrin I molecule, C atoms are shown in teal, N in blue, and H in silver.

Download video file (7.7MB, mp4)

Acknowledgments

We thank Monika Warzecha and Jeffrey Rimer for valuable discussions and suggestions; the former also for help with X-ray characterization of the crystals and the latter for access to Asylum Cypher ES AFM. This work was supported by the NSF (Award No. DMR-2128121) and the Welch Foundation (Grants E-2170, E-1882 and the Center for Advanced Bioactive Materials Crystallization, Award V-E-0001). Computational resources were provided by the Hewlett Packard Enterprise Data Science Institute at the University of Houston and the Texas Advanced Computing Center at the University of Texas at Austin.

Author contributions

J.C.P. and P.G.V. designed research; R.C., L.V., V.G.H., and J.C.P. performed research; R.C., L.V., V.G.H., J.C.P., and P.G.V. analyzed data; and R.C., L.V., J.C.P., and P.G.V. wrote the paper.

Competing interests

The authors declare no competing interest.

Footnotes

This article is a PNAS Direct Submission.

Data, Materials, and Software Availability

The datasets generated and analyzed during the current study are provided in SI Appendix. No custom-made computer code was used in this work. All other data are included in the manuscript and/or supporting information.

Supporting Information

References

  • 1.Ackerson M. R., Mysen B. O., Tailby N. D., Watson E. B., Low-temperature crystallization of granites and the implications for crustal magmatism. Nature 559, 94–97 (2018). [DOI] [PubMed] [Google Scholar]
  • 2.Ni P., Chabot N. L., Ryan C. J., Shahar A., Heavy iron isotope composition of iron meteorites explained by core crystallization. Nat. Geosci. 13, 611–615 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Daniels M. J. D., et al. , Fenamate NSAIDs inhibit the NLRP3 inflammasome and protect against Alzheimer’s disease in rodent models. Nat. Commun. 7, 12504 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Wang Y., et al. , Two-photon excited deep-red and near-infrared emissive organic co-crystals. Nat. Commun. 11, 4633 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Pagola S., Stephens P. W., Bohle D. S., Kosar A. D., Madsen S. K., The structure of malaria pigment β-haematin. Nature 404, 307–310 (2000). [DOI] [PubMed] [Google Scholar]
  • 6.Weiner S., Dove P. M., An Overview of Biomineralization Processes and the Problem of the Vital Effect. Rev. Mineral. Geochem. 54, 1–29 (2003). [Google Scholar]
  • 7.Elhadj S., De Yoreo J. J., Hoyer J. R., Dove P. M., Role of molecular charge and hydrophilicity in regulating the kinetics of crystal growth. Proc. Natl. Acad. Sci. U.S.A. 103, 19237–19242 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Dandekar P., Kuvadia Z. B., Doherty M. F., Engineering crystal morphology. Ann. Rev. Mater. Res. 43, 359–386 (2013). [Google Scholar]
  • 9.Myerson A. S., Handbook of Industrial Crystallization (Cambrifge University Press, Cambridge, ed. 3, 2019). [Google Scholar]
  • 10.Buckley H. E., Crystal Growth (John Wiley & Sons Inc, New York, 1951). [Google Scholar]
  • 11.Chernov A. A., The spiral growth of crystals. Sov. Phys. Uspekhi 4, 116–148 (1961). [Google Scholar]
  • 12.Schlichtkrull J., Insulin crystals: II. Shape of rhombohedral zinc-insulin crystals in relation to species and crystallization media. Acta Chem. Scand. 10, 1459–1464 (1956). [Google Scholar]
  • 13.Rimer J. D., Inorganic ions regulate amorphous-to-crystal shape preservation in biomineralization. Proc. Natl. Acad. Sci. U.S.A. 117, 3360–3362 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Kossel W., Zur Theorie des Kristallwachstums. Nachr. Ges. Wiss. Götingen 135–138 (1928). [Google Scholar]
  • 15.Stranski I. N., Zur Theorie des Kristallwachstums. Z. Phys. Chem. 136, 259–278 (1928). [Google Scholar]
  • 16.Burton R. C., Ferrari E. S., Davey R. J., Finney J. L., Bowron D. T., The relationship between solution structure and crystal nucleation: A neutron scattering study of supersaturated methanolic solutions of benzoic acid. J. Phys. Chem. B 114, 8807–8816 (2010). [DOI] [PubMed] [Google Scholar]
  • 17.Chernov A. A., Modern Crystallography III, Crystal Growth (Springer, Berlin, 1984). [Google Scholar]
  • 18.Chung J., et al. , Molecular modifiers reveal a mechanism of pathological crystal growth inhibition. Nature 536, 446–450 (2016). [DOI] [PubMed] [Google Scholar]
  • 19.Smoluchowski M., Drei Vorträge uber Diffusion, Brownsche Bewegung und Koagulation von Kolloidteilchen. Physik Z. 17, 557–585 (1916). [Google Scholar]
  • 20.Hänggi P., Talkner P., Borkovec M., Reaction-rate theory: Fifty years after Kramers. Rev. Mod. Phys. 62, 251–341 (1990). [Google Scholar]
  • 21.Pal S. K., Zewail A. H., Dynamics of water in biological recognition. Chem. Rev. 104, 2099–2124 (2004). [DOI] [PubMed] [Google Scholar]
  • 22.Israelachvili J. N., Intermolecular and Surface Forces (Academic Press, New York, 1995). [Google Scholar]
  • 23.Araki Y., et al. , Direct observation of the influence of additives on calcite hydration by frequency modulation atomic force microscopy. Cryst. Growth Design 14, 6254–6260 (2014). [Google Scholar]
  • 24.Nakouzi E., et al. , Moving beyond the solvent-tip approximation to determine site-specific variations of interfacial water structure through 3D force microscopy. J. Phys. Chem. C 125, 1282–1291 (2021). [Google Scholar]
  • 25.Petsev D. N., Chen K., Gliko O., Vekilov P. G., Diffusion-limited kinetics of the solution-solid phase transition of molecular substances. Proc. Natl. Acad. Sci. U.S.A. 100, 792–796 (2003). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Reviakine I., Georgiou D. K., Vekilov P. G., Capillarity effects on the crystallization kinetics: Insulin. J. Am. Chem. Soc. 125, 11684–11693 (2003). [DOI] [PubMed] [Google Scholar]
  • 27.Vekilov P. G., What determines the rate of growth of crystals from solution? Cryst. Growth Design 7, 2796–2810 (2007). [Google Scholar]
  • 28.Kowacz M., Putnis C., Putnis A., The effect of cation: Anion ratio in solution on the mechanism of barite growth at constant supersaturation: Role of the desolvation process on the growth kinetics. Geochim. Cosmochim. Acta 71, 5168–5179 (2007). [Google Scholar]
  • 29.Burton W. K., Cabrera N., Frank F. C., The growth of crystals and equilibrium structure of their surfaces. Phil. Trans. Roy. Soc. London Ser. A 243, 299–360 (1951). [Google Scholar]
  • 30.Warzecha M., et al. , Olanzapine crystal symmetry originates in preformed centrosymmetric solute dimers. Nat. Chem. 12, 914–920 (2020). [DOI] [PubMed] [Google Scholar]
  • 31.Vekilov P. G., Verma L., Palmer J. C., Chakrabarti R., Warzecha M., The pathway from the solution to the steps. J. Cryst. Growth 599, 126870 (2022). [Google Scholar]
  • 32.Gilmer G. H., Ghez R., Cabrera N., An analysis of combined volume and surface diffusion processes in crystal growth. J. Cryst. Growth 8, 79–93 (1971). [Google Scholar]
  • 33.Silvestri A., Raiteri P., Gale J. D., Obtaining consistent free energies for ion binding at surfaces from solution: Pathways versus alchemy for determining kink site stability. J. Chem. Theory Comput. 18, 5901–5919 (2022). [DOI] [PubMed] [Google Scholar]
  • 34.Broad A., Darkins R., Duffy D. M., Ford I. J., Calcite kinks grow via a multistep mechanism. J. Phys. Chem. C 126, 15980–15985 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Joswiak M. N., Peters B., Doherty M. F., In silico crystal growth rate prediction for NaCl from aqueous solution. Cryst. Growth Design 18, 6302–6306 (2018). [Google Scholar]
  • 36.Joswiak M. N., Doherty M. F., Peters B., Ion dissolution mechanism and kinetics at kink sites on NaCl surfaces. Proc. Natl. Acad. Sci. U.S.A. 115, 656–661 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.De La Pierre M., Raiteri P., Gale J. D., Structure and dynamics of water at step edges on the calcite 1014 surface. Cryst. Growth Design 16, 5907–5914 (2016). [Google Scholar]
  • 38.Warzecha M., et al. , Precrystallization solute assemblies and crystal symmetry. Faraday Discussions 235, 307–321 (2022). [DOI] [PubMed] [Google Scholar]
  • 39.De Yoreo J. J., Vekilov P. G., “Principles of crystal nucleation and growth” in Reviews in Mineralogy & Geochemistry, P. M. Dove, J. J. De Yoreo, S. Weiner, Eds. (Mineralogical Society of America Geochemical Society, 2003), vol. 54, pp. 57–93. [Google Scholar]
  • 40.Vekilov P. G., “Incorporation at kinks: Kink density and activation barriers, perspectives on inorganic, organic and biological crystal growth: From fundamentals to applications” in AIP Conference Proceedings, AIP Conference Series, Skowronski M., DeYoreo J. J., Wang C. A., Eds. (AIP, Melville, NY), 2007), vol. 916, pp. 235–267. [Google Scholar]
  • 41.Hoang M. H., et al. , Unusually high-performing organic field-effect transistors based on π-extended semiconducting porphyrins. Adv. Mater. 24, 5363–5367 (2012). [DOI] [PubMed] [Google Scholar]
  • 42.Che C.-M., et al. , A high-performance organic field-effect transistor based on platinum(II) porphyrin: Peripheral substituents on porphyrin ligand significantly affect film structure and charge mobility. Chem. Asian J. 3, 1092–1103 (2008). [DOI] [PubMed] [Google Scholar]
  • 43.Chakrabarti R. G., Vekilov P. G., Attraction between permanent dipoles and london dispersion forces dominate the thermodynamics of organic crystallization. Cryst. Growth Design 20, 7429–7438 (2020). [Google Scholar]
  • 44.Giovannetti R., “The use of spectrophotometry UV-vis for the study of porphyrins” in Macro To Nano Spectroscopy, Uddin J., Ed. (IntechOpen, Rijeka, 2012). [Google Scholar]
  • 45.Bruhn T., Brückner C., The origin of the absorption spectra of porphyrin N- and dithiaporphyrin S-oxides in their neutral and protonated states. Phys. Chem. Chem. Phys. 17, 3560–3569 (2015). [DOI] [PubMed] [Google Scholar]
  • 46.Verma L., et al. , How to identify the crystal growth unit. Israel J. Chem. 61, 1–11 (2021). [Google Scholar]
  • 47.Kano K., Minamizono H., Kitae T., Negi S., Self-aggregation of cationic porphyrins in water. Can π−π stacking interaction overcome electrostatic repulsive force? J. Phys. Chem. A 101, 6118–6124 (1997). [Google Scholar]
  • 48.Dvornikov S. S., Solov’ev K. N., Tsvirko M. P., Spectroscopic manifestations of dimerization of etioporphyrin I in hydrocarbon solvents. J. Appl. Spectr. 38, 581–585 (1983). [Google Scholar]
  • 49.Paine J. B. III, Dolphin D., Gouterman M., Exciton and electron interaction in covalently-linked dimeric porphyrins. Canadian J. Chem. 56, 1712–1715 (1978). [Google Scholar]
  • 50.Teng H. H., Dove P. M., Orme C. A., De Yoreo J. J., Thermodynamics of calcite growth: Baseline for understanding biomineral formation. Science 282, 724–727 (1998). [DOI] [PubMed] [Google Scholar]
  • 51.Gliko O., Reviakine I., Vekilov P. G., Stable equidistant step trains during crystallization of insulin. Phys. Rev. Lett. 90, 225503 (2003). [DOI] [PubMed] [Google Scholar]
  • 52.Yau S.-T., Thomas B. R., Vekilov P. G., Molecular mechanisms of crystallization and defect formation. Phys. Rev. Lett. 85, 353–356 (2000). [DOI] [PubMed] [Google Scholar]
  • 53.Orme C. A., et al. , Formation of chiral morphologies through selective binding of amino acids to calcite surface steps. Nature 411, 775–779 (2001). [DOI] [PubMed] [Google Scholar]
  • 54.Malkin A. J., Land T. A., Yu G., Kuznetsov A., McPherson J. J. D., Investigation of virus crystal growth mechanism by in situ atomic force microscopy. Phys. Rev. Lett. 75, 2778–2781 (1995). [DOI] [PubMed] [Google Scholar]
  • 55.Yau S.-T., Vekilov P. G., Quasi-planar nucleus structure in apoferritin crystallisation. Nature 406, 494–497 (2000). [DOI] [PubMed] [Google Scholar]
  • 56.Ma W., Lutsko J. F., Rimer J. D., Vekilov P. G., Antagonistic cooperativity between crystal growth modifiers. Nature 577, 497–501 (2020). [DOI] [PubMed] [Google Scholar]
  • 57.Yau S.-T., Petsev D. N., Thomas B. R., Vekilov P. G., Molecular-level thermodynamic and kinetic parameters for the self-assembly of apoferritin molecules into crystals. J. Mol. Biol. 303, 667–678 (2000). [DOI] [PubMed] [Google Scholar]
  • 58.Lovette M. A., Doherty M. F., Multisite models to determine the distribution of kink sites adjacent to low-energy edges. Phys. Rev. E 85, 021604 (2012). [DOI] [PubMed] [Google Scholar]
  • 59.Einstein A., Zur theorie der brownschen bewegung. Annalen der Physik 19, 371–381 (1906). [Google Scholar]
  • 60.Kramers H. A., Brownian motion in a field of force and the diffusion model of chemical reactions. Phys. A 7, 284–304 (1940). [Google Scholar]
  • 61.Frauenfelder H., Wolynes P. G., Rate theories and puzzles of hemeprotein kinetics. Science 229, 337–345 (1985). [DOI] [PubMed] [Google Scholar]
  • 62.Vekilov P. G., Kuznetsov Y. G., Chernov A. A., Interstep interaction in solution growth; (101) ADP face. J. Cryst. Growth 121, 643–655 (1992). [Google Scholar]
  • 63.Land T. A., DeYoreo J. J., Lee J. D., An in-situ AFM investigation of canavalin crystallization kinetics. Surf. Sci. 384, 136–155 (1997). [Google Scholar]
  • 64.Olafson K. N., Rimer J. D., Vekilov P. G., Early onset of kinetic roughening due to a finite step width in hematin crystallization. Phys. Rev. Lett. 119, 198101 (2017). [DOI] [PubMed] [Google Scholar]
  • 65.Olafson K. N., Ketchum M. A., Rimer J. D., Vekilov P. G., Mechanisms of hematin crystallization and inhibition by the antimalarial drug chloroquine. Proc. Natl. Acad. Sci. U.S.A. 112, 4946–4951 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Atkins P., DePaula J., Physical Chemistry (Freeman, ed. 7, New York, 2002). [Google Scholar]
  • 67.von Smoluchowski M., Versuch einer matematischen theorie der koagulationskinetik kolloider losungen. Z. Phys. Chem. 92, 129–135 (1918). [Google Scholar]
  • 68.Zambelli S., “Chemical kinetics, an historical introduction” in Chemical Kinetics, Patel V., Ed. (InTech Rijeka, Croatia, 2012), pp. 3–28. [Google Scholar]
  • 69.Stack A. G., Raiteri P., Gale J. D., Accurate rates of the complex mechanisms for growth and dissolution of minerals using a combination of rare-event theories. J. Am. Chem. Soc. 134, 11–14 (2012). [DOI] [PubMed] [Google Scholar]
  • 70.Bryngelson J. D., Wolynes P. G., Spin glasses and the statistical mechanics of protein folding. Proc. Natl. Acad. Sci. U.S.A. 84, 7524–7528 (1987). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Onuchic J. N., Socci N. D., Luthey-Schulten Z., Wolynes P. G., Protein folding funnels: The nature of the transition state ensemble. Folding Design 1, 441–450 (1996). [DOI] [PubMed] [Google Scholar]
  • 72.Chen M., et al. , Surveying biomolecular frustration at atomic resolution. Nat. Commun. 11, 5944 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Zheng W., Tsai M.-Y., Chen M., Wolynes P. G., Exploring the aggregation free energy landscape of the amyloid-β protein (1–40). Proc. Natl. Acad. Sci. U.S.A. 113, 11835–11840 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Chen M., Wolynes P. G., Aggregation landscapes of Huntingtin exon 1 protein fragments and the critical repeat length for the onset of Huntington’s disease. Proc. Natl. Acad. Sci. U.S.A. 114, 4406–4411 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Chen M., Schafer N. P., Wolynes P. G., Surveying the energy landscapes of Aβ fibril polymorphism. J. Phys. Chem. B 122, 11414–11430 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Xu Y., et al. , Frustrated peptide chains at the fibril tip control the kinetics of growth of amyloid-β fibrils. Proc. Natl. Acad. Sci. U.S.A. 118, e2110995118 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Elhadj S., et al. , Peptide controls on calcite mineralization: Polyaspartate chain length affects growth kinetics and acts as a stereochemical switch on morphology. Cryst. Growth Des. 6, 197–201 (2006). [Google Scholar]
  • 78.Bergeron L., Filobelo L., Galkin O., Vekilov P. G., Thermodynamics of the hydrophobicity in crystallization of insulin. Biophys. J. 85, 3935–3942 (2003). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Houben L., Weissman H., Wolf S. G., Rybtchinski B., A mechanism of ferritin crystallization revealed by cryo-STEM tomography. Nature 579, 540–543 (2020). [DOI] [PubMed] [Google Scholar]
  • 80.Yamazaki T., et al. , Two types of amorphous protein particles facilitate crystal nucleation. Proc. Natl. Acad. Sci. U.S.A. 114, 2154–2159 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Hu Y.-C., Tanaka H., Revealing the role of liquid preordering in crystallisation of supercooled liquids. Nat. Commun. 13, 4519 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Nicolis G., Nicolis C., Enhancement of the nucleation of protein crystals by the presence of an intermediate phase: A kinetic model. Phys. A 323, 139–154 (2003). [Google Scholar]
  • 83.Nicolis G., Maes D., Eds., Kinetics and Thermodynamics of Multistep Nucleation and Self-Assembly in Nanoscale Materials (John Wiley & Sons Inc, NY, 2012). [Google Scholar]
  • 84.ten Wolde P. R., Frenkel D., Enhancement of protein crystal nucleation by critical density fluctuations. Science 277, 1975–1978 (1997). [DOI] [PubMed] [Google Scholar]
  • 85.Bhattacharjee A., Roy M. N., Density, viscosity, and speed of sound of (1-Octanol + 2-Methoxyethanol), (1-Octanol + N, N-Dimethylacetamide), and (1-Octanol + Acetophenone) at temperatures of (298.15, 308.15, and 318.15) K. J. Chem. Eng. Data 55, 5914–5920 (2010). [Google Scholar]
  • 86.Sastry N. V., Valand M. K., Viscosities and densities for Heptane + 1-Pentanol, +1-Hexanol, +1-Heptanol, +1-Octanol, +1-Decanol, and +1-Dodecanol at 298.15 K and 308.15 K. J. Chem. Eng. Data 41, 1426–1428 (1996). [Google Scholar]
  • 87.Matsuo S., Makita T., Viscosities of six 1-alkanols at temperatures in the range 298–348 K and pressures up to 200 MPa. Int. J. Thermophys. 10, 833–843 (1989). [Google Scholar]
  • 88.Assael M. J., Polimatidou S. K., Measurements of the viscosity of alcohols in the temperature range 290–340 K at pressures up to 30 MPa. Int. J. Thermophys. 15, 95–107 (1994). [Google Scholar]
  • 89.Dubey G. P., Sharma M., Dubey N., Study of densities, viscosities, and speeds of sound of binary liquid mixtures of butan-1-ol with n-alkanes (C6, C8, and C10) at T=(298.15, 303.15, and 308.15)K. J. Chem. Thermodyn. 40, 309–320 (2008). [Google Scholar]
  • 90.Pan I. C., Tang M., Chen Y.-P., Densities and viscosities of binary liquid mixtures of vinyl acetate, diethyl oxalate, and dibutyl phthalate with normal alkanols at 303.15 k. J. Chem. Eng. Data 45, 1012–1015 (2000). [Google Scholar]
  • 91.Chakrabarti R., Vekilov P. G., Attraction between permanent dipoles and london dispersion forces dominate the thermodynamics of organic crystallization. Cryst. Growth Design 20, 7429–7438 (2020). [Google Scholar]
  • 92.Olafson K. N., Ketchum M. A., Rimer J. D., Vekilov P. G., Molecular mechanisms of hematin crystallization from organic solvent. Cryst. Growth Design 15, 5535–5542 (2015). [Google Scholar]
  • 93.Georgiou D. K., Vekilov P. G., A fast response mechanism for insulin storage in crystals may involve kink generation by association of 2D clusters. Proc. Natl. Acad. Sci. U.S.A. 103, 1681–1686 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Chernov A. A., Rashkovich L. N., Vekilov P. G., Steps in solution growth: Dynamics of kinks, bunching and turbulence. J. Cryst. Growth 275, 1–18 (2005). [Google Scholar]
  • 95.Abraham M. J., et al. , GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX 1–2, 19–25 (2015). [Google Scholar]
  • 96.Wang J., Wang W., Kollman P. A., Case D. A., Automatic atom type and bond type perception in molecular mechanical calculations. J. Mol. Graph. Modell. 25, 247–260 (2006). [DOI] [PubMed] [Google Scholar]
  • 97.Wang J., Wolf R. M., Caldwell J. W., Kollman P. A., Case D. A., Development and testing of a general amber force field. J. Comput. Chem. 25, 1157–1174 (2004). [DOI] [PubMed] [Google Scholar]
  • 98.Sousa da Silva A. W., Vranken W. F., ACPYPE - AnteChamber PYthon Parser interfacE. BMC Res. Notes 5, 367 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Vanquelef E., et al. , Server: A web service for deriving RESP and ESP charges and building force field libraries for new molecules and molecular fragments. Nucleic Acids Res. 39, W511–517 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Hess B., Bekker H., Berendsen H. J. C., Fraaije J. G. E. M., LINCS: A linear constraint solver for molecular simulations. J. Comput. Chem. 18, 1463–1472 (1997). [Google Scholar]
  • 101.Essmann U., et al. , A smooth particle mesh Ewald method. J. Chem. Phys. 103, 8577–8593 (1995). [Google Scholar]
  • 102.Bussi G., Donadio D., Parrinello M., Canonical sampling through velocity rescaling. J. Chem. Phys. 126, 014101 (2007). [DOI] [PubMed] [Google Scholar]
  • 103.Parrinello M., Rahman A., Polymorphic transitions in single crystals: A new molecular dynamics method. J. Appl. Phys. 52, 7182–7190 (1981). [Google Scholar]
  • 104.Barducci A., Bussi G., Parrinello M., Well-tempered metadynamics: A smoothly converging and tunable free-energy method. Phys. Res. Lett. 100, 020603 (2008). [DOI] [PubMed] [Google Scholar]
  • 105.Tiwary P., Parrinello M., A time-independent free energy estimator for metadynamics. J. Phys. Chem. B 119, 736–742 (2015). [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Appendix 01 (PDF)

Dataset S01 (XLSX)

Movie S1.

The ingress of an etioporphyrin I molecules into a kink. Successive snapshots of an etioporphyrin I molecule in the solution near a kink (ass seen in Fig. 4a), at the intermediate state (as in Fig. 4b), and at its final location in the kink (as in Fig. 4c). Classical MD simulation results. Solvent molecules are omitted for clarity. The kink is viewed along the unfinished molecular row at the step edge. The direction of step growth is from left to right. Etioporphyrin I molecules in the crystal lattice are shown in grey. The lattice planes of etioporphyrin I molecules in front and behind the plane, which hosts the kink, are omitted. In the incoming etioporphyrin I molecule, C atoms are shown in teal, N in blue, and H in silver.

Download video file (7.7MB, mp4)

Data Availability Statement

The datasets generated and analyzed during the current study are provided in SI Appendix. No custom-made computer code was used in this work. All other data are included in the manuscript and/or supporting information.


Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES