Laser Shock Wave Induced Crystallization - Crystal Growth & Design


Laser Shock Wave Induced Crystallization - Crystal Growth & Design...

1 downloads 70 Views 4MB Size

Subscriber access provided by University of Newcastle, Australia

Article

Laser Shock-Wave Induced Crystallization Nasrin Mirsaleh-Kohan, Andrew Fischer, Bernard Graves, Mehdi Bolorizadeh, Dilip Kendepudi, and Robert N. Compton Cryst. Growth Des., Just Accepted Manuscript • DOI: 10.1021/acs.cgd.6b01437 • Publication Date (Web): 03 Jan 2017 Downloaded from http://pubs.acs.org on January 11, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Crystal Growth & Design is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

Laser Shock-Wave Induced Crystallization Nasrin Mirsaleh-Kohan*,1, Andrew Fischer2, Bernard Graves3, Mehdi Bolorizadeh4, Dilip Kendepudi5 and Robert N. Compton*,3,6 1

Department of Chemistry and Biochemistry, Texas Woman’s University, Denton, TX, 76204 2 Abbott Diagnostics, 1921 Hurd Drive, Irving, TX 75038 3 Department of Physics, University of Tennessee, Knoxville, TN 37996 4 Bloor Solutions, West Orange, NJ 07052 5 Department of Chemistry, Wake Forest University, Winston-Salem, NC 29109 6 Departments of Chemistry, University of Tennessee, Knoxville, TN 37996

Abstract The formations of crystals in saturated and under-saturated solutions are studied as a result of intense sound, or shock waves, produced by high intensity focused laser pulses. Many tiny crystals are created immediately throughout the solution by the laser-generated sound waves. The compression waves can be created by focusing within the liquid or onto the walls of a container. This new method allows for the instantaneous formation of many “seed” crystals, which are then available for further impurity-free crystal growth. More importantly, the tiny “baby” crystals can be harvested before complete growth into “adult” crystals can occur. This method is shown to produce enough “baby” crystals to provide a glimpse into the initial stages of crystal growth using modern microscopy techniques such as SEM. Employing this new method, simple salts such as sodium bromate, sodium chloride, sodium chlorate, and tartaric acid were successfully crystallized. This method of crystal growth may also allow for the generation of crystals, which have previously not been realized or are otherwise difficult to produce.

1 ACS Paragon Plus Environment

Crystal Growth & Design

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 19

Introduction Conventional crystal growth occurs as the result of the ordered assemblage of a species from a supersaturated solution. In general, a highly concentrated solution doesn’t necessarily result in crystal formation. In order for a crystal to grow a crystal must form in an ordered fashion. Very early studies1,2 have shown that mechanical perturbations such as agitation, mechanical shock and pressure gradients can result in rapid crystallization from a supersaturated solution. Khamskii3 also examined the role of external influences such as electromagnetic field effects in crystallization.

Crystallization from supersaturated solutions involves two interconnected

mechanisms: primary and secondary nucleation.

In primary nucleation, the step may be

spontaneous (homogenous) or induced (heterogeneous) nucleation. Our group has explained in some detail the primary and secondary nucleation phenomena in a book chapter4. There are various methods of growing crystals in addition to the traditional evaporative crystallization from a supersaturated solution. For instance, Porkroy et al.5 have grown large, high-quality mercury thiolate single crystals from liquid mercury using sonication. In their experiments, crystals formed within seconds and were two orders of magnitude larger in size when using the sonication technique compared to those produced using the mercury salts. An interesting technique employing a laser was reported by Garetz et al.6 when they were examining second harmonic generation in supersaturated urea solution in water. They discovered urea crystals upon focusing 1.06 µm Q-switched Nd:YAG laser pulses into supersaturated solution of urea in water.

Further studies by the same group7 showed that the crystallite

orientation is laser polarization dependent; when the laser light was horizontally polarized, the initial crystallite was found to be oriented horizontally and the laser was vertically polarized, the crystallite was oriented vertically. In a study by Lee et al.8, crystals of hen egg white lysozyme 2 ACS Paragon Plus Environment

Page 3 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

(HEWL) were produced by exposing supersaturated droplets of the HEWL to intense laser pulses with various wavelengths, intensities and pulse duration. Their experiments suggest that laser induced nucleation is more efficient when the stock solution is irradiated by a shorter pulse duration, higher intensity laser operating at 532 nm. Garetz and coworkers11 also reported the growth of the polar γ-polymorph of glycine exposing supersaturated aqueous solutions of glycine to intense pulses of laser light at 1.06 µm. They attributed the crystals produced to the laserinduced nucleation of the supersaturated aqueous glycine. Yoshimura and coworkers9-10 have examined the optimized laser conditions to induce crystallization using femtosecond laser irradiation. In their work, they learned that the efficiency of the nucleation depends on the repletion rate and a high efficiency is realized when they used a nanosecond laser. They also expanded their studies to urea crystals and attributed the crystallization to a photomechanical effect of femtosecond laser irradiation.10 A number of studies11,12 have also discussed crystallization by laser-induced bubble formation and cavitation bubbles in supersaturated aqueous solutions.

Following the reports of crystallization using the laser-induced technique by Garetz and coworkers6,7, our group employed a modification of their technique to produce various crystals. Previously, we have shown that a pulsed Nd:YAG laser focused in a liquid produces a spherical sound wave pulse.13 In that study, we observed formation of a microbubble followed by collapse and the creation of the sound wave. The experimental conditions in the current work is similar to the conditions reported in our previously published paper.13 In this current work, we present the formation of several crystals from saturated and under-saturated solutions when the solutions are exposed to an intense sound, or shock wave, produced by high intensity focused laser pulses.

3 ACS Paragon Plus Environment

Crystal Growth & Design

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 19

When the laser was directed into the solution, many small crystals are created throughout the solution by the laser-generated sound waves. The compression waves can be created by focusing within the liquid or onto the walls of the container. In this technique, many small crystals are produced that can be used as seeds for the growth of larger crystals. Employing this technique we have successfully crystallized simple salts such as sodium chloride, sodium bromate, and tartaric acid.

Experimental Section In the initial studies, supersaturated solutions of the desired compound in water were exposed to 1.06 µm pulses (8-12 ns, 10 Hz) from a Quanta-Ray DCR Nd:YAG laser with an average power of 800 mW. Three or four laser pulses were focused with a 5 cm focal length lens into the supersaturated solution in order to induce immediate crystal growth. Crystals appeared over the extended volume in addition to the laser focus volume. These observations suggested that perhaps compression sound waves were responsible for the crystal growth.

Proof of this

hypothesis came from observing crystal growth by focusing the laser on the container walls or onto a thin metal plate floating on top the liquid. Figure 1 presents an experimental set up for direct irradiation of solutions or irradiating a metal boat on top of the solution. The boat is made up a thin (~0.12mm) sheet of stainless steel. In some experiments the laser beam was split into two beams and focused on opposing metal plates producing converging shock waves in the liquid as depicted in Figure 2. The sample cell in Figure 2 is made from two thin sheets of stainless steel with a thickness of ~ 0.12mm. A separate series of experiments using the set up in Figure 1 was conducted to analyze the effects of the sound waves in the solution. The sound waves were generated by the use of 1.5 × 108 (W/pulse) peak power per pulse then focused with 4 ACS Paragon Plus Environment

Page 5 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

a 5 cm focal length lens into the fluid or onto the “boat”; the power density in those experiments was ~ 1.9×1012 W/cm2. In one experiment using laser irradiation of the boat 20 sodium bromate samples were irradiated for ~ one minute. Twenty nine control samples were also allowed to produce crystals for ~ 72 hours up to one week. These experiments clearly indicate that the crystals are formed as a result of laser induced shock waves (pressure gradients). The crystals were either immediately harvested or allowed to grow for a short period of time.

The crystals were analyzed using an Olympus petrographic

microscope and scanning electron microscope (SEM). The imaging was performed on an Ultra 55, Schottky Field Emitter SEM. The energy of the microscope’s electron beam was chosen in a way to minimize any sample charging, so no sample preparation will be needed and the true structure of the crystals can be observed. The crystal solutions were spin dried on a conductive stub, which was put into the SEM chamber.

Results and Discussion Figure 3 shows a photograph of sodium bromate crystals grown over many days from conventional crystal growth procedure from supersaturated solutions (left photo) as compared with the many hundreds of sodium bromate crystals formed from a few laser pulses in laser sound-induced crystallization (right photo). The tiny laser sound-induced crystals were allowed to grow for a few hours to allow for better visualization. Employing the setup in Figure 2, we have also produced various other crystals, the first being sodium chloride crystals. Figure 4 shows an image of NaCl salt crystals grown using the laser shock-wave method. As seen in the figure, the sodium chloride crystals are cubic in structure. Figures 4A-D are taken with magnifications of 2X and 5X. To obtain the actual dimensions of

5 ACS Paragon Plus Environment

Crystal Growth & Design

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 19

these crystals we have used SEM imaging. Figure 5 presents images of laser shock-wave induced crystals of NaCl taken with an SEM at the University of Tennessee, Department of Materials science. Figure 5 (A-D) were taken under different magnifications as noted. An overall image of the crystals is shown in Figure 5A. A three dimensional cube of about 5 × 5 × 3 µm3 is seen in Figure 3 B. Figures 5C and 5D are also presenting NaCl crystals in two dimensional with 6 × 14 µm2 and 2.4 × 4 µm2 sizes, respectively. These samples were also examined using a SEM at the Howard Hughes Medical Institute as shown in Figure 6 (A-D). These pictures are taken randomly across the collection mesh. We have also applied the laser-induced crystallization technique to grow crystals of tartaric acid, C H O . Tartaric acid crystals have long been of interest since the seminal studies of Pasteur due 4

6

6

to their molecular dissymmetry. There have been numerous studies involving the growth of tartaric acid crystals from an aqueous solution of tartaric acid (see Stern and Beevers)14. Figures 7 and 8 exhibit the images of tartaric acid crystals grown using the laser shock wave technique. The images in these figures are taken with a petrographic microscope and a SEM microscope, respectively. As seen in Figure 8, the largest crystal seen has dimensions of 700 × 80 µm2. Smaller crystals in two dimensional with dimensions 80 × 20 µm2 and 200 × 80 µm2 are also observed. Theoretical Considerations The driving force for nucleation is the difference in the solute chemical potential, ∆µ, between the solution and the solid phase. The well-known classical rate of nucleation can be written terms of ∆µ15,16,

J = A.exp[–(16π/3) (γ3Vm2/∆µ2kBT)]

(1)

6 ACS Paragon Plus Environment

Page 7 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

in which A is the pre-exponential factor, γ is the interfacial tension of the solid phase and the solution phase and Vm is the molar volume of the solid phase. One may define super-saturation S=(a/a0), in which a0 is the activity of the solute at saturation and a is the solute activity. In the first approximation, this ratio of activities may be replaced with the corresponding ratio of concentrations, c and c0, which makes S=(c/c0). Now the chemical potential difference, ∆µ, can be written in terms of S: ∆µ = RT lnS = RT ln(c/c0). Noting that R = NAkB, the rate of nucleation can now be expressed as a function of super-saturation S. J = A.exp[–(16π/3) (γ3(Vm/NA)2/ (kBT)3 (lnS)2)]

(2)

Now by identifying (Vm/NA) as the molecular volume v, one arrives at the well-known expression for the nucleation rate: J = A.exp[–(16π/3) (γ3v2/kB3 T3 {lnS}2)]

(3)

Equation (3) is the expression commonly used to describe the rate of primary homogenous nucleation. From this equation, one sees that there are three main variables to describe the rate of nucleation: the temperature, the degree of super-saturation, and the interfacial tension. In deriving the above result, it is assumed that the nucleus is spherical (so that the "size" of the nucleus is its diameter) for which there is a simple relation between the volume and the surface area which is expressed in terms of the radius. If the nucleus is assumed to have a different shape, different geometrical factors relating the surface-to-volume ratio would have to be used, resulting in a different expression. But this difference is only a geometrical factor, not the basic formulation. In the laboratory, nucleation is almost always heterogeneous occurring on small nucleation sites on the container walls or small particles in the solution. The factors that

7 ACS Paragon Plus Environment

Crystal Growth & Design

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 19

affect nucleation in this case are the interfacial tension and geometry which can enhance the nucleation rate. From the above expression it is clear that any change in interfacial tension, γ, or the chemical potential difference, ∆µ, or the temperature T can alter the nucleation rate. Any external factor that changes these parameters will change the nucleation rate. In the case of a sound wave or a shock wave, there is an increase in temperature and pressure. Changes in pressure and temperature can have a major influence on the chemical potential. An expression for the change in chemical potential due to a change, δp, in pressure and a change, δT, temperature can be obtained as follows:

(4)

Noting that the derivative (∂µ/∂p)T = Vm, the molar volume, and that (∂µ/∂T)p = –Sm, the above expression can be written in terms of the corresponding changes in the molar volume and entropy of the solute between the solution and the solid phase, ∆Vm , –∆Sm, respectively. Thus:

∆µ(p 0 + δp,T0 + δT) = ∆µ(p 0 ,T0 ) + ∆Vmδp − ∆SmδT

(5)

During crystallization from solution, ∆Vm > 0, because the molar volume of the solute in solution phase is generally larger than that of the solid phase. Similarly, the molar entropy change ∆Sm is also positive. This implies increases in p and T have opposite effects on ∆µ; one increases ∆µ while the other causes it to decrease. The overall effect of these changes in the nucleation rate can be calculated by substituting these terms in the expression for the nucleation rate in eq.:

8 ACS Paragon Plus Environment

Page 9 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

(6)

In this expression, the change in γ due to change in T is implicit. As we have already noted, one could use the approximation ∆µ(p0,T0) =RTln(S), in which S is the super-saturation parameter. Considering the effects of sound waves, since change in pressure and T are periodic functions we may use the root-mean-square values of δp (which in turn is related to the sound intensity) and δT. Using the subscripts "rms" for these quantities, we obtain the nucleation rate for sound or shock-wave induced crystallization:

 16π  γ 3Vm2 J = AExp−   3 [RTln(S) + ∆Vmδp rms − ∆S mδTrms ] 2 k B (T0 + δTrms ) 

(7)

From this expression, it is clear that an increase in nucleation rate due to sound waves will result only when (∆Vmδprms– ∆SmδTrms) > 0; the fact that the pressure wave generated by the laser increases nucleation rate implies that this term is positive. The above expression can now be written as: (8) in which the nucleation rate and the exponent are expressed as a function of δp and δT. The increase in the nucleation rate due to the shockwave can now be expressed as the ratio (9) Conditions under which ∆f >0 will result in an increase in the nucleation rate due to laser induced shockwave. 9 ACS Paragon Plus Environment

Crystal Growth & Design

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 19

The generation and propagation of a shock-wave is a very complex process involving rapid changes in pressure and temperature17. As the laser-pulse induced shockwave propagates a few millimeters, pressure over 20 MPa have been observed; this pressure decreases as the shockwave propagates out. The temperature rises to well above the boiling point of water and fall quickly to the ambient temperature.

With the data on the pressures generated by a

shockwave we can make numerical estimates for NaCl. For these estimates, we shall use the following data for NaCl: the crystal-solution interfacial tension is about 60×10-3 N/m 18. Using the solution density data, we estimate ∆Vm ≈ 20×10-6 m3; from the heat of solution of NaCl, we estimate ∆Sm ≈ 10 J/K. Figure 9 shows the regions of δP and δT where ∆f is positive for NaCl. It clearly shows that in the range of pressures that the laser pulse generates, ∆f >0 when δT decreases to values below 10K, i.e., as the temperature of the shockwave decreases to 10K above the ambient T, the shockwave pressure drives the nucleation and crystallization taking place at an enhanced rate. For a given δT, the value of ∆f increases very rapidly with δP: for instance, when δT = 4, at δP=1.92, ∆f = –3.0 and when δP=1.93, ∆f = +16.6, indicating the rapid increase in nucleation rate due to the pressure of the shockwave. Thus we see under the conditions that our experiments were performed, the above theoretical estimates show considerable enhancement in homogeneous nucleation rate. Surely, heterogeneous nucleation will also be contributing to the increase in nucleation rate but its rate depends on the effective γ of the nucleation sites and other factors.

10 ACS Paragon Plus Environment

Page 11 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

Conclusions In this work, we have described a simple laser-induced shock wave technique to produce crystals of various compounds, specifically sodium chloride, tartaric acid, and sodium bromate. This method produces many very tiny crystals. For example, NaCl crystals were determined to be about 6 × 14 µm2, and 2.4 × 4 µm2. These examples exhibit the potential of this technique to make smaller crystals, perhaps on the order of "nano crystals" using under-saturated solutions. Further experiments would be important to investigate the impact of basic experimental parameters (e.g., temperature, solvent, exposure time, etc.,) on the laser induced crystallization technique. This method of crystal growth may also allow for the generation of crystals, which have previously not been realized or are otherwise difficult to produce such as protein crystals. Also, the passive nature of this technique would be favorable to crystal growth under extreme environments.

11 ACS Paragon Plus Environment

Crystal Growth & Design

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 19

Figures

Figure 1. The Experimental setup for the irradiation of the sodium bromate solutions is shown. For some of the sound experiments, the metal boat was on top of the solution while irradiating.

Figure 2. A picture of one experimental arrangement (left) along with a schematic presentation (right)

12 ACS Paragon Plus Environment

Page 13 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

Figure 3. Photograph of sodium bromate crystals grown over many days from conventional crystal growth procedure from supersaturated solutions (left photo) as compared with the hundreds of sodium bromate crystals formed from laser sound-induced crystallization (right photo). The tiny laser sound-induced crystals were allowed to grow for a few hours to allow for better visualization.

Figure 4. Laser shock-wave induced crystals of NaCl, pictures were taken with a petrographic microscope. Pictures are taken with magnifications of 20X and 50X. For the actual size of the crystals see Figure 5.

13 ACS Paragon Plus Environment

Crystal Growth & Design

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 19

Figure 5. SEM image of laser shock-wave induced crystals of NaCl. A three dimensional cube of about 5 × 5 × 3 µm3 is seen in Figure 3B. Figures 5C and 5D are also presenting crystals in two dimensional with 6 × 14 µm2 and 2.4 × 4 µm2 sizes, respectively.

Figure 6. SEM image of laser shock-wave induced crystals of NaCl, taken at the Howard Hughes Medical Institute. These pictures are taken randomly across the mesh.

14 ACS Paragon Plus Environment

Page 15 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

Figure 7. Laser shock-wave induced crystals of Tartaric acid, C4H6O6. Images were taken with a petrographic microscope. Pictures are taken with magnification of 20X. For the actual size of the crystals see Figure 8.

Figure 8. SEM image of laser shock-wave induced crystals of Tartaric acid, C4H6O6, taken at the Howard Hughes Medical Institute. The largest crystal seen has dimensions of 700 × 80 µm2. Smaller crystals in two dimensional with dimensions 80 × 20 µm2 and 200 × 80 µm2 are also observed.

15 ACS Paragon Plus Environment

Crystal Growth & Design

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 19

Figure 9. Graph showing regions of the pressure δP and δT of the shockwave at which δf is positive for NaCl (In regions where δf > 0, the nucleation rate is enhanced). The ratio of nucleation rates J(δP, δT) /J0(0,0) = Exp[∆f]

AUTHOR INFORMATION Corresponding Author * E-mail: [email protected] * E-mail: [email protected] Author Contributions The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.

16 ACS Paragon Plus Environment

Page 17 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

Acknowledgments This work was supported in part by the National Science Foundation (CHE-0848487). One of the authors (N.M.K.) also would like to thank the Robert H. Welch Foundation, Research Enhancement Program, and the Chancellor's Research Fellows program at Texas Woman’s University. We thank Dr. David C. Joy for useful discussions and for allowing us to perform imaging using the SEM system in his laboratory.

References: (1) Young, S. Mechanical stimulus to crystallization in super-cooled liquids. J. Am. Chem. Soc. 1911, 33, 148-162. (2) Berkeley E. Solubility and supersolubility from the osmotic standpoint. Philo. Mag. 1912, 24, 254-268. (3) Khamskii, E. V. Crystallization from solutions. New York: Consultants Bureau; 1969 (4) Nano-clusters-A bridge across disciplines, Volume 1, (2010) Publisher: Elsevier; first edition Edited by Purusottam Jena and A. Welford Castleman, Chapter 8, Laser Induced Crystallization, Andrew Fischer, R.M. Pagni, R.N. Compton, D. Kondepudi. Pages 343-364 (5) Pokroy, B. ; Aichmayer, B.; Schenk, A. S.; Haimov, B.; Kang, S. H.; Fratzl, P.; Aizenberg, J. Sonication-assisted synthesis of large, high-quality mercury thiolate single crystals directly from liquid mercury. J. Am. Chem. Soc., 2010, 132, 14355-14357. (6) Garetz, B. A.; Aber, J. E.; Goddard, N. L.; Young, R. G.; Myerson, A. S. Nonphotochemical, polarization-dependent, laser-induced nucleation in supersaturated aqueous urea solutions. Phys. Rev. Lett. 1996, 77, 3475-3476. (7) Zaccaro, J.; Matic, J.; Myerson, A. S.; Garetz, B. A. Nonphotochemical, laser-induced nucleation of supersaturated aqueous glycine produces unexpected γ-polymorph. Crystal Growth & Design, 2001, 1, 5-8.

17 ACS Paragon Plus Environment

Crystal Growth & Design

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 19

(8) Sung Lee, I.; Evans, J. M. B.; Erdemir, D.; Lee, A. Y.; Garetz, B. A.; Myerson, A. S. Nonphotochemical laser induced nucleation of hen egg white lysozyme crystals. Crystal Growth & Design, 2008, 8, 4255-4261. (9) Hosokawa, Y.; Adachi, H.; Yoshimura, M.; Mori, Y.; Sasaki, T.; Masuhara, H. Femtosecond laser-induced crystallization of 4-(dimethylamino)-N-methyl-4-stilbazolium tosylate. Crystal Growth & Design, 2005, 5, 861-863. (10)

Yoshikawa, H. Y.; Hosokawa, Y.; Masuhara, H. Spatial control of urea crystal growth by

focused femtosecond laser irradiation. Crystal Growth & Design, 2006, 6, 302-305. (11) Nakamura, K.; Hosokawa, Y.; Masuhara, H. Anthracene crystallization induced by singleshot femtosecond laser irradiation: experimental evidence for the important role of bubbles. Crystal Growth & Design, 2007, 7, 885-889. (12) Soare, A.; Dijkink, R.; Pascual, M. R.; Sun, C.; Cains, P. W.; Lohse, D.; Stankiewicz, A. I.; Kramer, H. J. M. Crystal nucleation by laser-induced cavitation. Crystal Growth & Design, 2011, 11, 2311-2316. (13) Fischer, A. T; Compton, R. N. Laser-based speed of sound measurements in water and aqueous D, L,and DL alanine solutions. Rev. Sci. Instrument, 2003, 4, 3730-3734. (14) Stern, F.; Beevers, C. A. The crystal structure of tartaric acid. Acta Cryst. 1950, 3, 341-346. (15) Randolp, A.; Larson, M. Theory of Particulate Processes, Academic Press, New York, 1971. (16) Kondepudi, D. K.; Prigogine, I. Modern Thermodynamics, 2nd Ed. John Wiley, 2015. (17) Sankin, G.N.; Zhou, Y.; Zhong, P. Focusing of shockwaves induced by optical breakdown in water, J. Acoust. Soc. Am. 2008, 123, 4071-4081. (18) Bahadur,R.; Russell, L.M.; Alavi, S. Surface tensions in NaCl−water−air systems from MD simulations. J. Phys. Chem. B, 2007, 111, 11989–11996.

18 ACS Paragon Plus Environment

Page 19 of 19

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

For Table of Contents Use Only Synopsis: A schematic diagram of formation crystals from saturated and under-saturated solutions employing intense sound or shock waves produced by pulsed lasers.

Title: Laser Shock-Wave Induced Crystallization Authors: Nasrin Mirsaleh-Kohan, Andrew Fischer, Bernard Graves, Mehdi Bolorizadeh, Dilip Kendepudi and Robert N. Compton

19 ACS Paragon Plus Environment