Role of Nuclei in Liesegang Pattern Formation: The Insights from

petri dish (inner diameter: 59 mm) and stored in an incubator at 18 .... as discussed earlier. Fig. 3 shows the RD simulation results with different C...
0 downloads 0 Views 4MB Size
Article Cite This: J. Phys. Chem. C 2018, 122, 3669−3676

pubs.acs.org/JPCC

Role of Nuclei in Liesegang Pattern Formation: Insights from Experiment and Reaction-Diffusion Simulation Masaki Itatani,† Qing Fang,‡ Kei Unoura,§ and Hideki Nabika*,§ †

Graduate School of Science and Engineering, ‡Department of Mathematical Science, Faculty of Science, and §Department of Material and Biological Chemistry, Faculty of Science, Yamagata University, 1-4-12 Kojirakawa, Yamagata 990-8560, Japan S Supporting Information *

ABSTRACT: Many types of periodic patterns can spontaneously form in nature across wide spatiotemporal scales. Construction of a chemical model that mimics these periodic patterns are of considerable interest from both scientific and technological viewpoints. The Liesegang phenomenon is one of the chemical models to form periodic patterns with well-defined periodicity. However, the parameters that influence the mechanism and resultant pattern geometry are not completely known. In this study, we use surface chemistry methods to evaluate the influence of nucleation threshold on the geometry of Liesegang patterns. Cysteine was used as an additional ligand for the precursors (Agn nuclei and/or Ag nanoparticles in the present system) to reduce their surface free energy and thus the nucleation free energy. As a result, the formed Liesegang patterns had smaller spacing coefficient (i.e., finer periodic patterns), a phenomenon that was also reproduced using reaction-diffusion simulation with lowered nucleation threshold. The small spacing coefficient at lowered nucleation threshold was discussed in terms of a slower rate of Ostwald ripening. Similar control of the nucleation threshold through surface chemistry can be applied to various precipitation systems, as well as gaining insight into the comprehensive mechanism underlying various animate and inanimate patterns formed in nature.



INTRODUCTION The formation of spatiotemporal patterns at conditions far from equilibrium is ubiquitous in nature, and it plays a critical role in the dynamic, diverse, flexible, and robust features of both animate and inanimate matters.1 Spatial patterns are often discussed in terms of reaction and diffusion (RD) phenomenon caused by cooperative interaction between reaction and diffusion of substances in the system. There are two types of spatial patterns: traveling patterns and static patterns. Nerve impulse propagation in animals is a typical traveling wave. It is dominated by the activity of ion channels and diffusion of Na+ and K+ ions, and is well explained with the FitzHugh-Nagumo model.2 When the biological activity and diffusivity are replaced by chemical reactivity and diffusivity, respectively, chemical traveling waves are formed in vitro, such as the Belousov− Zhabotinsky (BZ) reaction and proton waves.3−9 Since the key parameters that determine the propagation characteristics can be easily tuned by experimental conditions, these chemical waves are used as models to gain physicochemical insight into traveling waves. Static patterns, on the other hand, can be seen in various fields including chemistry,10,11 biology,12,13 and bacteriology.14,15 Similar to the traveling waves, static patterns formed in various animate and inanimate systems can be modeled by chemical reactions. For example, the stripe pattern of zebrafish can be reproduced by chemical reactions with the © 2018 American Chemical Society

Turing mechanism that produces various patterns from an initially homogeneous medium.16,17 Thus, revealing the mechanism of pattern formation in chemical reaction systems can help us understand the rich features seen in natural patterns. While the Turing patterns are formed in a homogeneous medium, the pattern formed in the presence of a concentration gradient in one reactant is known as the Liesegang pattern.18 In chemical experiments, the gradient can be formed by contact between a gel containing an inner electrolyte and an aqueous solution containing an outer electrolyte. From the solution−gel interface, the outer electrolyte diffuses into the gel media and forms a concentration gradient. When the reaction between the inner and outer electrolytes gives insoluble products, the products will deposit as a precipitate and form periodic precipitation bands in the gel media. There are various reports on chemical Liesegang patterns that are formed by sparingly soluble salts,19−25 chemical reduction,26 and polymerization,27 where the precipitate could be hydroxides, 28−34 chromates,19−25 halides,35,36 carbonates,37 and metal nanoparticles,26 polymers,27 hydrogels,38,39 and so on. Although both Received: December 26, 2017 Revised: January 25, 2018 Published: January 26, 2018 3669

DOI: 10.1021/acs.jpcc.7b12688 J. Phys. Chem. C 2018, 122, 3669−3676

Article

The Journal of Physical Chemistry C

Here, we used a new experimental approach to understand the chemical Liesegang patterns through direct control of C* with the aid of surface chemistry. Surface chemistry has already been used to control the γ value of nanomaterials (such as metal clusters and nanoparticles) that are the dominant chemical species during a nucleation and precipitation growth processes.53 For this purpose, we chose a system based on the chemical reduction of Ag+ that yields Liesegang patterns consisting of Ag nanoparticles26 because the γ value of Agn clusters can be easily tuned by various ligands such as thiols,54−59 phosphate,60,61 and citrate.62−64 Since Agn would be the transition species from Ag atoms to Ag nanoparticles, the control of their γ value may clarify the relationship between Liesegang pattern formation and C* or ΔG. In the present study, we used cysteine as an additional ligand for the stabilization of Agn species because cysteine is easily dissolved in a gelatin medium and expected to strongly ligate Ag species through its sulfide and amino groups. As a result, we found that the presence of cysteine has a strong influence on the geometry of the Liesegang pattern, which was discussed from RD simulation with varying C*. Furthermore, to eliminate the ligation effect of citrate as a reducing agent,26 we also carried out similar experiments with cysteine but no citrate in a photochemical reduction, instead of the chemical reduction with citrate. The results from chemical and photochemical experiments, along with RD simulations, revealed that the ligation of cysteine to Agn decreased C* due to a reduction in γ. As a result, the geometric characteristics of the formed Liesegang pattern were changed. This study, based on surface chemistry, therefore unlocks a new approach to gain deep insight into how chemical conditions control the pattern geometry in the Liesegang mechanism.

Turing and Liesegang patterns are periodic and formed spontaneously, their periodicity is substantially different. In the Turing pattern, the interband spacing is constant, and this distance depends on the diffusivity and the period of the limit cycle.40,41 In contrast, the interband spacing of Liesegang pattern varies as a function of distance from the interface.20,22,26,27,39 Thus, the Turing and Liesegang mechanisms, both appearing in different animate and inanimate systems, are important for understanding pattern formation in nature. The Turing pattern has been comprehensively examined by interdisciplinary studies spanning chemistry, biology, physics, and mathematics. In contrast, for the Liesegang pattern, there is still a large gap between nature and existing chemical models. Several patterns that are similar in geometry to the Liesegang patterns have been reported in bacteria colonies,14 biological tissue,42 rock,43 and the solar system. However, there is no direct evidence to link these natural patterns to the chemical models of the Liesegang pattern. One problem lies in the difficulty of controlling the reaction condition for Liesegang pattern formation in chemical systems. Many proposed models involve a nucleation process followed by precipitate growth, such as in the nucleation and growth44−47 and sol-coagulation models.22,45,48 For example, the nucleation and growth model involves the following three processes: Reaction: A + B → C

(1)

Nucleation: C → N, if [C] > C *

(2)

Precipitate growth: N → P

(3)

where A, B, C, N, and P are the outer electrolyte, inner electrolytes, reaction product, nuclei of particles, and precipitate, respectively. In this model, A diffuses into the gel medium containing B and reacts with B to form C (eq 1). In the gel medium, C nucleates and forms N if the local concentration reaches the nucleation threshold C* (eq 2), which will be followed by spontaneous nuclei growth and precipitation (eq 3). From this model, it is clear that C* controls the progress of precipitate and band formation. Indeed, recent studies based on RD simulation demonstrated the critical importance of C* on the geometry of Liesegang patterns.49−51 C* represents the energy barrier (ΔG) for the transition of the product from the molecular state (C) to solid state (N) with nonzero surface area and volume, as defined by the following equation including two terms corresponding to surface and volume effects:52 ΔG = −4/3πr 3ΔGv + 4πr 2γ



EXPERIMENTAL SECTION Reagents. Gelatin (fine powder) and sodium sulfate were purchased from Nacalai Tesque Industry. Silver nitrate, trisodium citrate dehydrate, L-cysteine, and agarose were purchased from Wako Pure Chemical Industry. All reagents were used without further purification. Chemical Reduction System. A gelatin solution (5.0%) was prepared with Milli-Q water under stirring and heating at 75 °C for 25 min. Sodium sulfate and trisodium citrate dihydrate were added to the gelatin solution to be final concentrations of 130 and 45 mM, respectively. For the system with cysteine, L-cysteine was also added to the gelatin solution at 50 μM. The mixed gelatin solution was stirred and heated at 90 °C for 25 min. Then 1.5 mL of the solution was poured into a glass Petri dish (inner diameter: 59 mm) and stored in an incubator at 18 °C overnight to yield the gelatin gel matrix. An agarose solution (8.0%) was prepared with Milli-Q water, and degassed in vacuum for 5 min. Then the solution was heated by a microwave for 40 s. Immediately after heating, this solution was poured into a silicon tube (inner diameter: 7 mm) and left for 20 min at room temperature. Then agarose stamps were made by slicing the agarose gel into columns with a length of 3 cm and soaked in silver nitrate aqueous solution (1.0 M) for more than 2 weeks. The agarose stamp doped with Ag+ ion was then put on the surface of gelatin gel formed in the Petri dish, and the stacked gels were stored at 18 °C for 6 h. Since gelatin plays an important role in controlling the reduction and nucleation rate of silver species, we used the gelatin gel as the reaction medium. On the other hand, the agarose gel was used as the stamp

(4)

where r, ΔGv, and γ are the radius of nuclei, the free energy per unit volume of nuclei or nanoparticles, and the surface free energy per unit area, respectively. This equation implies that C* can be controlled through ΔG, in which ΔGv and γ would play critical roles. However, there is little knowledge about the relation between Liesegang pattern formation and either ΔGv or γ, except an experimental study that used impurities to control the number of nucleation sites in the reaction medium.22 In that research, the authors clarified that the addition of impurity led to heterogeneous nucleation that was discussed by varying C* in their model. However, the relationship between ΔG and C* is still unknown, even though it provides an alternative and flexible approach to control the precipitation geometry. 3670

DOI: 10.1021/acs.jpcc.7b12688 J. Phys. Chem. C 2018, 122, 3669−3676

Article

The Journal of Physical Chemistry C because of its easy usability compared with the gelatin gel.26 The obtained patterns were observed using an optical microscope BX-43 (Olympus Co., Ltd., Japan). Photochemical Reduction System. The photochemical reduction system was prepared in the same way as for the chemical reduction system, except trisodium citrate dihydrate was not used. The agarose stamp was put on the surface of gelatin gel in Petri dish, which was then transferred to a thermoplate and kept at 24 °C. Then the photochemical reduction was carried out by irradiating UV light (250−400 nm, 45−50 lx; MAX-301, Asahi Spectra Co., Ltd., Japan) for 4 h on the thermoplate. The obtained patterns were observed using the same optical microscope.



RESULTS AND DISCUSSION In the chemical reduction system without cysteine, periodic brown bands appeared spontaneously in which multiple band patterns could be observed outside the continuous precipitate (Figure 1a). The brownish precipitates are silver nanoparticles from the reaction between Ag+ and citrate. A previous work showed that two different Ag+ species exist in the presence of gelatin molecules:65 freely diffusing Ag+ ion (free Ag+) and Ag+ ion ligated to amino acid residues in the gelatin matrix (bound Ag+). It has been suggested that only the bound Ag+ is involved in Liesegang pattern formation.26 Since the bound Ag+ and bound Ag0 from reduction were relatively stabilized by the amino acid residues, nucleation commenced when the local concentration of the bound species reached the threshold of C*.65 Afterward, Ag nanoparticles were formed via particle growth and Ostwald ripening, producing the brownish color in Figure 1a owing to surface plasmon resonance. Since Ag+ diffused from the agarose stamp according to Fick’s law, the brownish precipitate appeared and extended from the agarose stamp with time (Figure 1b). Just after placement of the agarose stamp on the gelatin gel (0 min), the gelatin gel was colorless with no visible precipitations. At 60 min, the reaction front can be seen on the left side, forming precipitates as a brownish band. This precipitation zone progressed with time, and the reaction front reached the center of the imaging area at 180 min, at which point continuous precipitation changed to discrete periodic precipitation at the edge of the reaction front. More obvious periodic precipitation was observed at 360 min. The appearance of periodic patterns was also shown as an oscillation of intensity in the line profile (Figure 1c). The interband distance was about 100−200 μm, and this distance increased further away from the edge of agarose stamp. In the presence of cysteine, similar precipitation propagation also occurred, but the periodicity was different. In Figure 2a, it appears that only a continuous precipitation zone was formed, unlike the periodic bands at the reaction front when cysteine was not used (Figure 1a). The enlarged image in Figure 2b also showed no periodic bands with the interband distance of 100− 200 μm. However, this system did yield bands but with much finer periodicity. At 60 min, fine and faint bands started to emerge from the left side, which can be clearly seen in the highcontrast images shown in Supporting Information (Figure S1) and the intensity line profile (Figure 2c). Such fine patterns were not observed in the absence of cysteine, even when highcontrast images were examined. As the reaction proceeded, the dark continuous precipitation band on the left expanded and became visible after 180 min, preceded by the fine bands. The interband spacing of this fine pattern was below 100 μm, meaning that its periodicity was almost half that of the pattern

Figure 1. Liesegang patterns obtained in the chemical reduction system without cysteine. (a) Microscope image of sample 6 h after the agarose stamp (disk at the center of image) was placed on the gelatin gel. (b) Enlarged and successive images of the propagating precipitation reaction front, for the area shown in the white rectangle in (a). (c) Line profiles of the intensity in images shown in (b). The black arrows at 180 and 360 min indicate the positions of individual bands.

formed without cysteine. Therefore, it can be concluded that the presence of cysteine affected the chemical condition for the periodic band formation, presumably through the ligation to Ag species such as Agn nuclei and Ag nanoparticles and thereby altering their surface free energies (γ). Now we consider the mathematical model of reactiondiffusion (RD) simulations. The diffusion part of a chemical molecule u is described by div(D(τ,x) grad u). When D(τ,x) is homogeneous, the diffusion part is given by D∇2u = D div(grad u). The reaction process is modeled by a dynamical system uτ = F(τ,x,u). To discuss the role of cysteine, RD simulations were carried out based on the nucleation and growth model66 as follows: ∂A = ∇2 A − R(A , B) ∂τ 3671

(5) DOI: 10.1021/acs.jpcc.7b12688 J. Phys. Chem. C 2018, 122, 3669−3676

Article

The Journal of Physical Chemistry C

N(C, D) describes the dependence of the precipitation pathway on the presence of pre-existing precipitates at the grid point of interest. The inner electrolyte B is homogeneously distributed in the calculation grid, whereas A is initially present in a circular region at the center of the calculation grid. The parameters used in the present simulation are A0 = 1.0, B0 = 0.01, C0 = D0 = 0, and DA:DB:DC = 1.0:0.75:0.1. The five-point formula with space grid-step Δx = 0.25 is employed on a 1600 × 1600 grid, and the fourth-order Runge−Kutta method with time step Δt = 0.01 is used to integrate the semidiscretized ordinary differential equations. The θ(C − C*) is a step function: θ(C − C*) = 1 when C − C* ≥ 0, while θ(C − C*) = 0 when C − C* < 0. Thus, the value of C* is very important because it controls the progress of precipitate formation (eq 8). To evaluate the effect of cysteine in modifying γ due to the ligation to Agn nuclei, we conducted the RD simulation by changing the value of C* because eq 4 implies that γ alters C* through ΔG, as discussed earlier. Figure 3 shows the RD simulation results

Figure 2. Liesegang patterns obtained in the chemical reduction system with cysteine. (a) Microscope image of sample 6 h after the agarose stamp was placed on the gelatin gel. (b) Enlarged and successive images of the propagating precipitation reaction front, for the white rectangle in (a). (c) Line profiles of the intensity in images shown in (b).

D ∂B = B ∇2 B − R(A , B) ∂τ DA

(6)

D ∂C = C ∇2 C + R(A , B) − Cθ(C − C*) − CN (C , D) ∂τ DA (7)

∂D = Cθ(C − C*) + CN (C , D) ∂τ

Figure 3. Concentration of precipitate D from the RD simulation at (a) C* = 0.011, (b) C* = 0.0105, and (c) C* = 0.01 in greyscale. The parameters for all simulations are A0 = 1.0, B0 = 0.01, DA = 1.0, DB = 0.75, and DC = 0.1.

(8)

where A, B, C, and D are the concentrations of the outer electrolyte, inner electrolyte, reaction product, and precipitate, respectively. The boundary conditions for A, B, and C are homogeneous Neumann boundary conditions. τ is the dimensionless reaction time (τ = tDA/L2), where L is the characteristic size of the pattern. ∇2 is the Laplacian operator. DA, DB, and DC are the diffusion coefficients of A, B, and C, respectively. R is the reaction kinetics function. The function

with different C*, all of which successfully produced periodic patterns similar to the ones experimentally observed. At C* = 0.011, relatively thick bands formed a concentric pattern. With decreasing of C* to 0.015, each band became thinner, and the concentric patterns became denser. Further decreasing of C* to 3672

DOI: 10.1021/acs.jpcc.7b12688 J. Phys. Chem. C 2018, 122, 3669−3676

Article

The Journal of Physical Chemistry C

different C* appear to converge to the same value, an enlarged plot (inset in Figure 4b) shows a significant difference among them under the observation conditions. Therefore, the simulation result suggested that a reduction in the spacing coefficient, which was observed experimentally by adding cysteine, was induced by a decrease in C*. Thus, it was demonstrated that the addition of cysteine decreased C* through its ligation to Agn nuclei, which led to the formation of fine bands in both experiments and RD simulations. From these results, we concluded that the addition of cysteine produces a similar effect to that of reducing C* in the pattern formation. The added cysteine can reduce the surface free energy γ of Ag clusters by stabilizing their surface atoms. A lowering γ means a reduced energy barrier (ΔG) for the transition from the molecular state (bound Ag) to solid state (Ag nuclei or Ag clusters) with nonzero surface area and volume (eq 4). Then nucleation can proceed under a lower excess concentration C* of the molecular species. Thus, it can be concluded that the presence of cysteine can help to lower C*, as similarly observed in the spacing coefficients from experiments and simulations. In the simulation, the width of each band and interband spacing narrowed when C* was decreased from 0.011 to 0.01, in accordance with our experimental results. In this way, the formed bands shrank and finally became linear ones, and the interband spacing also decreased upon the addition of cysteine. This agreement strongly suggests that the addition of cysteine contributed to the lowering in C*. However, regardless of the addition of cysteine, gelatin gels used in these chemical reduction experiments contained citric acid as a reducing agent. Previous works have reported that citrate ligates to Agn clusters to form Agn−citrate complex,62−64 which can also modulate the ΔG for nucleation and affect the pattern formation process. Therefore, there remains the possibility that citrate and cysteine compete in ligating to Agn clusters, which makes the effect of ligand ligation on ΔG somewhat complicated. To clarify the effect of cysteine, a reducing agent that cannot act as ligand for Agn clusters should be considered. For this purpose, we also conducted photochemical reduction by irradiating UV light to reduce Ag+ ion, instead of using citrate.67,68 Figure 5 shows the formed precipitation bands after UV irradiation for the citrate-free samples without and with cysteine. Both samples formed bright yellow (owing to surface plasmon resonance of Ag nanoparticles) precipitation bands around the agarose stamp, indicating that the reduction reaction proceeded under UV irradiation. The color of the precipitate looks different between the chemical and photochemical reduction (brown vs yellow), and this was attributed to a difference in the duration of light exposure under the microscope to visualize the fine periodic patterns. Without cysteine (Figure 5a), a yellow precipitation zone was formed around the stamp, and an orange band appeared at 2 mm away from the edge of stamp. The enlarged image and intensity line profile (Figure 5c) showed that, in the system without cysteine, the orange band was homogeneous without fine periodic structures. In the presence of cysteine, the orange band contained concentric fine bands according to the enlarged image (Figure 5b). These fine bands were confirmed to satisfy the same spacing law, and the value of 1 + p converged to a constant (Figure 5d). Between the photochemical experiments with and without cysteine, it became even more clear that surface ligation of Agn clusters reduces the interband distance and forms finer Liesegang patterns, which is

0.01 made the bands much thinner and almost linear, and multiple band patterns were aligned in a circular configuration. To compare the geometrical characteristics of experimental and simulated patterns, we analyzed the spacing law for the Liesegang phenomenon in which the ratio between the positions of two consecutive bands (xn+1/xn) tends to converge to a constant known as the spacing coefficient (1 + p) as n becomes large. The experimental spacing coefficients with and without cysteine are shown in Figure 4a, where n = 1 is defined

Figure 4. Spacing coefficient as a function of band number (a) obtained from experimental results without cysteine (black circle) and with cysteine (red triangle); (b) obtained from simulation at C* = 0.011 (black), 0.0105 (blue), and 0.01 (red).

as the innermost band that could be resolved under the optical microscope. In other words, the absolute value for n has less meaning than the relative order among the resolvable bands. Experimentally, the spacing coefficient converged to a constant value in each case, but the values are different (1.05 without and 1.02 with cysteine). The smaller spacing coefficient in the latter means the formation of finer periodic bands in the presence of cysteine (Figure 2). A similar tendency was seen in the RD simulations; that is, the spacing coefficient decreased with decreasing C* (Figure 4b). However, unlike the experimental results, the simulated spacing coefficient showed a gradual decrease at small n. A possible reason is that the simulation predicted quite thin precipitation bands that could not be resolved under experimental observation. Thus, the simulated bands in Figure 3 would contain much thinner bands than the experimental ones in Figure 1 and Figure 2. After this gradual decrease, the simulated spacing coefficient became constant at larger n. Although the spacing coefficients with 3673

DOI: 10.1021/acs.jpcc.7b12688 J. Phys. Chem. C 2018, 122, 3669−3676

Article

The Journal of Physical Chemistry C

materials that can interact with the growing precipitate particles can be another key parameter to gain novel insight and comprehensively model the Liesegang phenomenon.



CONCLUSION Our experiments and simulations showed that the surface free energy of precipitate nuclei affects the spacing coefficient in the resulting Liesegang patterns. Because cysteine as a ligand has a high affinity to Agn clusters, adding cysteine in the gelatin gel matrix caused the formation of finer periodic patterns, and the periodicity was confirmed to satisfy the spacing law for Liesegang phenomena. Similar results were observed in the corresponding photochemical reduction system, thereby eliminating the influence of citrate on the surface free energy. While our experiments showed that the addition of cysteine changed the geometry of the Liesegang pattern, the RD simulation suggested that such fine structure could also be observed when C* became lower. All these results indicate that surface ligation of cysteine to Agn clusters reduces the surface free energy and C*, resulting in the appearance of finer periodic patterns due to the slower rate of Ostwald ripening. Such control of C* via γ could be applied to various other precipitation systems. Furthermore, there are many choices of ligands to modulate γ. Thus, this surface chemistry-based approach to explore Liesegang phenomena is useful for understanding the overall mechanism underlying the formation of patterns in various animate and inanimate natural systems.



Figure 5. Liesegang patterns obtained in the photochemical reduction system (a) without and (b) with cysteine. (a) Microscope image of sample without cysteine, taken at 4 h after the agarose stamp was placed on the gelatin gel. (b) Microscope image of sample with cysteine taken after 4 h. Rectangular images at the bottom: enlarged regions indicated by the white rectangle. (c) Line profiles for the sample without and with cysteine. (d) Spacing coefficient for the system with cysteine.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jpcc.7b12688. High contrast images of reaction fronts (PDF)



likely through controlling the values of γ and ΔG for nucleation. From our experiments, we propose that the surface ligation lowers the spacing coefficient. C* represents the concentration at which the nucleation process occurs (eq 2), which can also be defined as the maximum concentration of C that can be used for the nucleation and growth steps. Above C*, nuclei are formed and grow with Ostwald ripening (eq 3), and they can exhibit periodic structure consisting of precipitation and depletion zones. Since the number density of nuclei is proportional to C* (because nucleation cannot start below this concentration), a lower C* means lower densities in the formed nuclei and particles. Furthermore, the rate of Ostwald ripening is proportional to the molar volume and thus the number of particles.69 Above all, lowering C* reduces the Ostwald ripening rate, and as a result thinner precipitation and depletion bands are formed. Therefore, in the suggested mechanism, surface ligation leads to finer Liesegang patterns through lowering the surface free energy. Since there are other possible ligands than citrate or cysteine, an appropriate design of ligands for the nuclei allows well-defined control of the geometry of Liesegang patterns. Furthermore, the relationship between ΔG for nucleation and the spacing coefficient found in the present study may help to reveal the formation mechanism of Liesegang-like patterns with diverse geometries in nature. Finally, our finding indicates that besides the chemical and physical characteristics of the precipitating materials, coexisting

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Masaki Itatani: 0000-0003-1025-0452 Hideki Nabika: 0000-0002-7780-8433 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by JSPS KAKENHI Grant Number 16H04092.



REFERENCES

(1) Grzybowski, B. A. Chemistry in Motion: Reaction-Diffusion Systems for Micro-and Nanotechnology; Wiley: New York, 2009. (2) Epstein, I. R.; Pojman, J. A. An Introduction to Nonlinear Chemical Dynamics; Oxford: New York, 1998. (3) Ginn, B. T.; Steinbock, B.; Kahveci, M.; Steinbock, O. Microfluidic Systems for the Belousov−Zhabotinsky Reaction. J. Phys. Chem. A 2004, 108 (8), 1325−1332. (4) Nakata, S.; Matsushita, M.; Sato, T.; Suematsu, N. J.; Kitahata, H.; Amemiya, T.; Mori, Y. Photoexcited Chemical Wave in the Ruthenium-Catalyzed Belousov-Zhabotinsky Reaction. J. Phys. Chem. A 2011, 115 (26), 7406−7412. (5) Masere, J.; Vasquez, D. A.; Edwards, B. F.; Wilder, J. W.; Showalter, K. Nonaxisymmetric and Axisymmetric Convection in 3674

DOI: 10.1021/acs.jpcc.7b12688 J. Phys. Chem. C 2018, 122, 3669−3676

Article

The Journal of Physical Chemistry C Propagating Reaction-Diffusion Fronts. J. Phys. Chem. 1994, 98, 6505− 6508. (6) Szirovicza, L.; Nagypál, I.; Boga, E. An Algorithm for the Design of Propagating Acidity Fronts. J. Am. Chem. Soc. 1989, 111 (8), 2842− 2845. (7) Nabika, H.; Sato, M.; Unoura, K. Microchannel-Induced Change of Chemical Wave Propagation Dynamics: Importance of Ratio between the Inlet and the Channel Sizes. Phys. Chem. Chem. Phys. 2013, 15 (1), 154−158. (8) Arai, M.; Takahashi, K.; Hattori, M.; Hasegawa, T.; Sato, M.; Unoura, K.; Nabika, H. One-Directional Fluidic Flow Induced by Chemical Wave Propagation in a Microchannel. J. Phys. Chem. B 2016, 120 (20), 4654−4660. (9) Nabika, H.; Hasegawa, T.; Unoura, K. Propagation Behaviors of an Acid Wavefront Through a Microchannel Junction. J. Phys. Chem. B 2015, 119 (30), 9874−9882. (10) Nakouzi, E.; Steinbock, O. Self-Organization in Precipitation Reactions Far from the Equilibrium. Sci. Adv. 2016, 2 (8), e1601144− e1601144. (11) Grzybowski, B. A.; Bishop, K. J. M.; Campbell, C. J.; Fialkowski, M.; Smoukov, S. K. Micro- and Nanotechnology via Reaction− Diffusion. Soft Matter 2005, 1 (2), 114. (12) Watanabe, M.; Kondo, S. Is Pigment Patterning in Fish Skin Determined by the Turing Mechanism? Trends Genet. 2015, 31 (2), 88−96. (13) Kondo, S.; Miura, T. Reaction-Diffusion Model as a Framework for Understanding Biological Pattern Formation. Science (Washington, DC, U. S.) 2010, 329 (5999), 1616−1620. (14) Mimura, M.; Sakaguchi, H.; Matsushita, M. Reaction-Diffusion Modelling of Bacterial Colony Patterns. Phys. A 2000, 282 (1), 283− 303. (15) Tasaki, S.; Nakayama, M.; Shoji, W. Morphologies of Bacillus Subtilis Communities Responding to Environmental Variation. Dev. Growth Differ. 2017, 59 (5), 369−378. (16) Nakamasu, A.; Takahashi, G.; Kanbe, A.; Kondo, S. Interactions between Zebrafish Pigment Cells. Proc. Natl. Acad. Sci. U. S. A. 2009, 106 (21), 8429−8434. (17) Asakura, K.; Konishi, R.; Nakatani, T.; Nakano, T.; Kamata, M. Turing Pattern Formation by the CIMA Reaction in a Chemical System Consisting of Quaternary Alkyl Ammonium Cationic Groups. J. Phys. Chem. B 2011, 115 (14), 3959−3963. (18) Liesegang, R. E. Uber Einige Eigenschaften von Gallerten. Naturwiss. Wochenschr. 1896, 11, 353−362. (19) Lagzi, I. Formation of Liesegang Patterns in an Electric Field. Phys. Chem. Chem. Phys. 2002, 4 (8), 1268−1270. (20) Lagzi, I.; Ueyama, D. Pattern Transition between Periodic Liesegang Pattern and Crystal Growth Regime in Reaction-Diffusion Systems. Chem. Phys. Lett. 2009, 468 (4−6), 188−192. (21) Smoukov, S. K.; Lagzi, I.; Grzybowski, B. A. Independence of Primary and Secondary Structures in Periodic Precipitation Patterns. J. Phys. Chem. Lett. 2011, 2 (4), 345−349. (22) Lagzi, I. Controlling and Engineering Precipitation Patterns. Langmuir 2012, 28 (7), 3350−3354. (23) Thomas, S.; Molnár, F.; Rácz, Z.; Lagzi, I. Matalon-Packter Law for Stretched Helicoids Formed in Precipitation Processes. Chem. Phys. Lett. 2013, 577, 38−41. (24) Thomas, S.; Lagzi, I.; Molnár, F.; Rácz, Z. Helices in the Wake of Precipitation Fronts. Phys. Rev. E - Stat. Nonlinear, Soft Matter Phys. 2013, 88 (2), 022141. (25) Karam, T.; El-Rassy, H.; Sultan, R. Mechanism of Revert Spacing in a PbCrO4 Liesegang System. J. Phys. Chem. A 2011, 115 (14), 2994−2998. (26) Nabika, H.; Sato, M.; Unoura, K. Liesegang Patterns Engineered by a Chemical Reaction Assisted by Complex Formation. Langmuir 2014, 30 (18), 5047−5051. (27) Shimizu, Y.; Matsui, J.; Unoura, K.; Nabika, H. Liesegang Mechanism with a Gradual Phase Transition. J. Phys. Chem. B 2017, 121 (11), 2495−2501.

(28) Tinsley, M. R.; Collison, D.; Showalter, K. Propagating Precipitation Waves: Experiments and Modeling. J. Phys. Chem. A 2013, 117 (48), 12719−12725. (29) Karam, T.; Sultan, R. Effect of an Alternating Current Electric Field on Co(OH)2 Periodic Precipitation. Chem. Phys. 2013, 412, 7− 12. (30) Batlouni, H.; Al-Ghoul, M. Experimental Study of the Dynamics of Front Propagation in the Co(OH)2/NH4OH Liesegang System Using Spectrophotometry. J. Phys. Chem. A 2008, 112 (35), 8038− 8045. (31) Badr, L.; Sultan, R. Ring Morphology and pH Effects in 2D and ID Co(OH)2 Liesegang Systems. J. Phys. Chem. A 2009, 113 (24), 6581−6586. (32) Badr, L.; El-Rassy, H.; El-Joubeily, S.; Sultan, R. Morphology of a 2D Mg2+/NH4OH Liesegang Pattern in Zero, Positive and Negative Radial Electric Field. Chem. Phys. Lett. 2010, 492 (1−3), 35−39. (33) Mansour, A. A.; Al-Ghoul, M. Band Propagation, Scaling Laws, and Phase Transition in a Precipitate System. 2. Computational Study. J. Phys. Chem. A 2015, 119 (35), 9201−9209. (34) Kuz’min, V. I.; Gadzaov, A. F.; Tytik, D. L.; Busev, S. A.; Revina, A. A.; Vysotskii, V. V. Kinetics of the Formation of Liesegang Rings. J. Struct. Chem. 2013, 54 (S2), 363−378. (35) Ayass, M. M.; Al-Ghoul, M. Superdiffusive Cusp-like Waves in the Mercuric Iodide Precipitate System and Their Transition to Regular Reaction Bands. J. Phys. Chem. A 2014, 118 (22), 3857−3865. (36) Ripszám, M.; Nagy, Á .; Volford, A.; Izsák, F.; Lagzi, I. The Liesegang Eyes Phenomenon. Chem. Phys. Lett. 2005, 414 (4−6), 384−388. (37) Makki, R.; Al-Ghoul, M.; Sultan, R. Propagating Fronts in Thin Tubes: Concentration, Electric, and pH Effects in a Two-Dimensional Precipitation Pulse System. J. Phys. Chem. A 2009, 113 (21), 6049− 6057. (38) Li, B.; Gao, Y.; Li, X.; Feng, Y.; Zhou, Y. Chitosan Hydrogels with 3D Liesegang Ring Structure for Rifampicin Release. J. Controlled Release 2011, 152 (Suppl), e47−9. (39) Narita, T.; Tokita, M. Liesegang Pattern Formation in κCarrageenan Gel. Langmuir 2006, 22 (1), 349−352. (40) Ouyang, Q.; Li, R.; Li, G.; Swinney, H. L. Dependence of Turing Pattern Wavelength on Diffusion Rate. J. Chem. Phys. 1995, 102 (6), 2551−2555. (41) Maini, P. K.; Painter, K. J.; Chau, H. N. P. Spatial Pattern Formation in Chemical and Biological Systems. J. Chem. Soc., Faraday Trans. 1997, 93 (20), 3601−3610. (42) Khonsari, R. H.; Calvez, V. The Origins of Concentric Demyelination: Self-Organization in the Human Brain. PLoS One 2007, 2 (1), e150. (43) Kettler, R. M.; Loope, D. B.; Weber, K. a.; Niles, P. B. Life and Liesegang: Outcrop-Scale Microbially Induced Diagenetic Structures and Geochemical Self-Organization Phenomena Produced by Oxidation of Reduced Iron. Astrobiology 2015, 15 (8), 616−636. (44) Ayass, M. M.; Abi Mansour, A.; Al-Ghoul, M. Alternating Metastable/stable Pattern in the Mercuric Iodide Crystal Formation Outside the Ostwald Rule of Stages. J. Phys. Chem. A 2014, 118 (36), 7725−7731. (45) Antal, T.; Droz, M.; Magnin, J.; Rácz, Z.; Zrinyi, M. Derivation of the Matalon-Packter Law for Liesegang Patterns. J. Chem. Phys. 1998, 109 (21), 9479−9486. (46) Hantz, P. Regular Microscopic Patterns Produced by Simple Reaction−diffusion Systems. Phys. Chem. Chem. Phys. 2002, 4 (8), 1262−1267. (47) Bensemann, I. T.; Fialkowski, M.; Grzybowski, B. A. Wet Stamping of Microscale Periodic Precipitation Patterns. J. Phys. Chem. B 2005, 109 (7), 2774−2778. (48) Shinohara, S. A Theory of One-Dimensional Liesegang Phenomena. J. Phys. Soc. Jpn. 1970, 29 (4), 1073−1087. (49) Büki, A.; Kárpáti-Smidróczki, E.; Zrínyi, M. Computer Simulation of Regular Liesegang Structures. J. Chem. Phys. 1995, 103 (23), 10387. 3675

DOI: 10.1021/acs.jpcc.7b12688 J. Phys. Chem. C 2018, 122, 3669−3676

Article

The Journal of Physical Chemistry C (50) Chen, L.; Kang, Q.; He, Y. L.; Tao, W. Q. Mesoscopic Study of the Effects of Gel Concentration and Materials on the Formation of Precipitation Patterns. Langmuir 2012, 28 (32), 11745−11754. (51) Molnár, F., Jr.; Izsák, F.; Lagzi, I. Design of Equidistant and Revert Type Precipitation Patterns in Reaction-Diffusion Systems. Phys. Chem. Chem. Phys. 2008, 10 (17), 2368−2373. (52) Thanh, N. T. K.; Maclean, N.; Mahiddine, S. Mechanisms of Nucleation and Growth of Nanoparticles in Solution. Chem. Rev. 2014, 114 (15), 7610−7630. (53) Aikens, C. M. Electronic Structure of Ligand-Passivated Gold and Silver Nanoclusters. J. Phys. Chem. Lett. 2011, 2 (2), 99−104. (54) Yang, H.; Wang, Y.; Huang, H.; Gell, L.; Lehtovaara, L.; Malola, S.; Häkkinen, H.; Zheng, N. All-Thiol-Stabilized Ag44 and Au12 Ag32 Nanoparticles with Single-Crystal Structures. Nat. Commun. 2013, 4, 2422. (55) Desireddy, A.; Conn, B. E.; Guo, J.; Yoon, B.; Barnett, R. N.; Monahan, B. M.; Kirschbaum, K.; Griffith, W. P.; Whetten, R. L.; Landman, U.; et al. Ultrastable Silver Nanoparticles. Nature 2013, 501 (7467), 399−402. (56) Kumar, S.; Bolan, M. D.; Bigioni, T. P. Glutathione-Stabilized Magic-Number Silver Cluster Compounds. J. Am. Chem. Soc. 2010, 132 (38), 13141−13143. (57) Wu, Z.; Lanni, E.; Chen, W.; Bier, M. E.; Ly, D.; Jin, R. High Yield, Large Scale Synthesis of Thiolate-Protected Ag 7 Clusters. J. Am. Chem. Soc. 2009, 131, 7−9. (58) Cathcart, N.; Mistry, P.; Makra, C.; Pietrobon, B.; Coombs, N.; Jelokhani-Niaraki, M.; Kitaev, V. Chiral Thiol-Stabilized Silver Nanoclusters with Well-Resolved Optical Transitions Synthesized by a Facile Etching Procedure in Aqueous Solutions. Langmuir 2009, 25 (10), 5840−5846. (59) Yang, H.; Wang, Y.; Zheng, N. Stabilizing Subnanometer Ag(0) Nanoclusters by Thiolate and Diphosphine Ligands and Their Crystal Structures. Nanoscale 2013, 5 (7), 2674. (60) Catalano, V. J.; Kar, H. M.; Garnas, J. A Highly Luminescent Tetranuclear Silver(I) Cluster and Its Ligation-Induced Core Rearrangement. Angew. Chem., Int. Ed. 1999, 38 (13), 1979−1982. (61) Teo, B. K.; Keating, K. Novel Triicosahedral Structure of the Largest Metal Alloy Cluster: [(Ph3P)12Au13Ag12Cl6]m+. J. Am. Chem. Soc. 1984, 106 (7), 2224−2226. (62) Pillai, Z. S.; Kamat, P. V. What Factors Control the Size and Shape of Silver Nanoparticles in the Citrate Ion Reduction Method? J. Phys. Chem. B 2004, 108 (3), 945−951. (63) Jiang, X. C.; Chen, C. Y.; Chen, W. M.; Yu, A. B. Role of Citric Acid in the Formation of Silver Nanoplates through a Synergistic Reduction Approach. Langmuir 2010, 26 (6), 4400−4408. (64) Henglein, A.; Giersig, M. Formation of Colloidal Silver Nanoparticles: Capping Action of Citrate. J. Phys. Chem. B 1999, 103 (44), 9533−9539. (65) Kapoor, S.; Lawless, D.; Kennepohl, P.; Meisel, D.; Serpone, N. Reduction and Aggregation of Silver Ions in Aqueous Gelatin Solutions. Langmuir 1994, 10 (9), 3018−3022. (66) Pan, C.; Gao, Q.; Xie, J.; Xia, Y.; Epstein, I. R. Precipitation Patterns with Polygonal Boundaries between Electrolytes. Phys. Chem. Chem. Phys. 2009, 11 (46), 11033. (67) Darroudi, M.; Ahmad, M. B.; Zak, A. K.; Zamiri, R.; Hakimi, M. Fabrication and Characterization of Gelatin Stabilized Silver Nanoparticles under UV-Light. Int. J. Mol. Sci. 2011, 12 (9), 6346−6356. (68) Marignier, J. L.; Belloni, J.; Delcourt, M. O.; Chevalier, J. P. Microaggregates of Non-Noble Metals and Bimetallic Alloys Prepared by Radiation-Induced Reduction. Nature 1985, 317 (6035), 344−345. (69) Hoang, T. K. N.; La, V. B.; Deriemaeker, L.; Finsy, R. Ostwald Ripening of Alkane in Water Emulsions Stabilized by Polyoxyethylene (20) Sorbitan Monolaurate. Langmuir 2002, 18 (5), 1485−1489.

3676

DOI: 10.1021/acs.jpcc.7b12688 J. Phys. Chem. C 2018, 122, 3669−3676