shape-preserving amorphous-to-crystalline transformation ... · edited by patricia m. dove,...

8
Shape-preserving amorphous-to-crystalline transformation of CaCO 3 revealed by in situ TEM Zhaoming Liu a,b , Zhisen Zhang c,d,e , Zheming Wang b , Biao Jin a,b , Dongsheng Li b , Jinhui Tao b , Ruikang Tang a,1 , and James J. De Yoreo b,f,1 a Department of Chemistry, Zhejiang University, Hangzhou, Zhejiang 310027, China; b Physical Sciences Division, Pacific Northwest National Laboratory, Richland, WA 99352; c Department of Physics, Xiamen University, Xiamen, Fujian 361005, China; d Research Institute for Biomimetics and Soft Matter, Xiamen University, Xiamen, Fujian 361005, China; e Fujian Provincial Key Laboratory for Soft Functional Materials Research, Xiamen University, Xiamen, Fujian 361005, China; and f Department of Materials Science and Engineering, University of Washington, Seattle, WA 98195 Edited by Patricia M. Dove, Virginia Polytechnic Institute and State University, Blacksburg, VA, and approved December 18, 2019 (received for review August 26, 2019) Organisms use inorganic ions and macromolecules to regulate crystallization from amorphous precursors, endowing natural biominerals with complex morphologies and enhanced proper- ties. The mechanisms by which modifiers enable these shape- preserving transformations are poorly understood. We used in situ liquid-phase transmission electron microscopy to follow the evolution from amorphous calcium carbonate to calcite in the presence of additives. A combination of contrast analysis and infrared spectroscopy shows that Mg ions, which are widely present in seawater and biological fluids, alter the transformation pathway in a concentration-dependent manner. The ions bring excess (structural) water into the amorphous bulk so that a direct transformation is triggered by dehydration in the absence of morphological changes. Molecular dynamics simulations suggest Mg-incorporated water induces structural fluctuations, allowing transformation without the need to nucleate a separate crystal. Thus, the obtained calcite re- tains the original morphology of the amorphous state, biomimetically achieving the morphological control of crystals seen in biominerals. crystallization | biomineralization | calcium carbonate | magnesium | morphology C rystallization in nature often follows a multistep pathway (1) involving phase transformation from a transient amorphous precursor (24). Calcium carbonate (CaCO 3 ), which is among the most abundant near-surface minerals and biominerals, is widely used as a model system to investigate fundamental aspects of crystallization in the context of both abiotic and biological mineralization (3, 5). Biomineral formation starting from amorphous CaCO 3 (ACC) results in diverse morphologies including those of sea urchin spicules and teeth (6), nacre layers (7), and crustacean exoskeletons (8), which are strikingly distinct from the growth shapes of abiotic CaCO 3 crystals. In these systems, amorphous nano- particles are deposited within a vesicular space onto the surface of the growing biomineral where they transform to a crystalline material that is shape-preserving with the original ACC precursor and ex- hibits a microstructure reminiscent of aggregated particles (9, 10). The transformation of biogenic amorphous phases is believed to be regulated by inorganic and organic components including Mg, phosphate, and carboxylated macromolecules, such as acidic proteins (8, 1114). In particular, incorporation of Mg into biogenic calcites is common, and its presence is believed essen- tial to stabilization of ACC (8, 13, 15) and inhibition of calcite (12, 16). However, the amorphous to crystalline transformation remains poorly understood (1, 4). Although some laboratory studies reported direct transformation (or solid-state trans- formation) of ACC to crystalline CaCO 3 (1719), most have concluded ACC transforms by dissolutionreprecipitation (1, 1922). Moreover, few studies have been based on direct ob- servation of the transformation and have instead relied upon either ex situ or bulk measurements, which are unable to relate the measured change in properties to the dynamics of individ- ual particles. Consequently, knowledge of how impurities alter transformation pathways and the evolution of particle composition are largely unknown. Here we take advantage of the ability of liquid phase TEM (LP-TEM) to follow mineral transformation processes of indi- vidual particles in real time (22, 23) in order to gain insight into the mechanism by which additives impact the transformation of ACC to crystalline products. LP-TEM has been used to in- vestigate nanoparticle nucleation, growth, dissolution, and as- sembly of many materials including Pt (24, 25), Au (26, 27), Pd (28), Ag (29, 30), Cu (31), iron hydroxide (32), PbS (33), iron oxides (34), and CaCO 3 (22, 23). These studies show that in situ observations by LP-TEM can provide high spatial and temporal resolution to directly observe the progress of phase transforma- tions of individual particles. While we recognize that there are significant differences between the formation of minerals in the vesicles and membranes of organisms and the solution environ- ment of a TEM sample cell, the results provide insights into physiochemically available mechanisms of transformation and the effect of impurities on those mechanisms. Significance Understanding of the shape-preserving crystallization from transient amorphous precursors to crystalline products in bio- mineralization makes it possible to produce topologically complex morphologies that defy crystallographic controls. Here we use in situ liquid-phase TEM, FTIR, and molecular dy- namics simulations to investigate transformation of amor- phous CaCO 3 to crystalline phases in the presence of multiple additives. In situ observations reveal that Mg 2+ , which brings excess water into the bulk, is unique in initiating crystallization within the amorphous phase, leading to a shape-preserving transformation that is accompanied by dehydration. Simula- tion results find that the water accelerates ionic rearrangement within the solid phase, enabling the shape-preserving trans- formation. This in-depth understanding provides an alternative strategy for manufacturing crystals with arbitrary morphol- ogies by design. Author contributions: Z.L., Z.Z., R.T., and J.J.D.Y. designed research; Z.L., Z.Z., Z.W., B.J., D.L., and J.T. performed research; Z.L., Z.Z., Z.W., D.L., J.T., R.T., and J.J.D.Y. analyzed data; and Z.L., Z.Z., R.T., and J.J.D.Y. wrote the paper. The authors declare no competing interest. This article is a PNAS Direct Submission. Published under the PNAS license. Data deposition: In situ LP-TEM files have been deposited in the Zenodo repository (http:// doi.org/10.5281/zenodo.3609354). See online for related content such as Commentaries. 1 To whom correspondence may be addressed. Email: [email protected] or james. [email protected]. This article contains supporting information online at https://www.pnas.org/lookup/suppl/ doi:10.1073/pnas.1914813117/-/DCSupplemental. First published February 3, 2020. www.pnas.org/cgi/doi/10.1073/pnas.1914813117 PNAS | February 18, 2020 | vol. 117 | no. 7 | 33973404 CHEMISTRY Downloaded by guest on February 8, 2021

Upload: others

Post on 02-Oct-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Shape-preserving amorphous-to-crystalline transformation ... · Edited by Patricia M. Dove, Virginia Polytechnic Institute and State University, Blacksburg, VA, and approved December

Shape-preserving amorphous-to-crystallinetransformation of CaCO3 revealed by in situ TEMZhaoming Liua,b, Zhisen Zhangc,d,e, Zheming Wangb, Biao Jina,b, Dongsheng Lib, Jinhui Taob, Ruikang Tanga,1,and James J. De Yoreob,f,1

aDepartment of Chemistry, Zhejiang University, Hangzhou, Zhejiang 310027, China; bPhysical Sciences Division, Pacific Northwest National Laboratory,Richland, WA 99352; cDepartment of Physics, Xiamen University, Xiamen, Fujian 361005, China; dResearch Institute for Biomimetics and Soft Matter, XiamenUniversity, Xiamen, Fujian 361005, China; eFujian Provincial Key Laboratory for Soft Functional Materials Research, Xiamen University, Xiamen, Fujian361005, China; and fDepartment of Materials Science and Engineering, University of Washington, Seattle, WA 98195

Edited by Patricia M. Dove, Virginia Polytechnic Institute and State University, Blacksburg, VA, and approved December 18, 2019 (received for review August26, 2019)

Organisms use inorganic ions and macromolecules to regulatecrystallization from amorphous precursors, endowing naturalbiominerals with complex morphologies and enhanced proper-ties. The mechanisms by which modifiers enable these shape-preserving transformations are poorly understood. We used insitu liquid-phase transmission electron microscopy to follow theevolution from amorphous calcium carbonate to calcite in thepresence of additives. A combination of contrast analysis and infraredspectroscopy shows that Mg ions, which are widely present inseawater and biological fluids, alter the transformation pathway ina concentration-dependent manner. The ions bring excess (structural)water into the amorphous bulk so that a direct transformation istriggered by dehydration in the absence of morphological changes.Molecular dynamics simulations suggest Mg-incorporated waterinduces structural fluctuations, allowing transformation without theneed to nucleate a separate crystal. Thus, the obtained calcite re-tains the original morphology of the amorphous state, biomimeticallyachieving the morphological control of crystals seen in biominerals.

crystallization | biomineralization | calcium carbonate | magnesium |morphology

Crystallization in nature often follows a multistep pathway (1)involving phase transformation from a transient amorphous

precursor (2–4). Calcium carbonate (CaCO3), which is amongthe most abundant near-surface minerals and biominerals, iswidely used as a model system to investigate fundamental aspectsof crystallization in the context of both abiotic and biologicalmineralization (3, 5). Biomineral formation starting from amorphousCaCO3 (ACC) results in diverse morphologies including those ofsea urchin spicules and teeth (6), nacre layers (7), and crustaceanexoskeletons (8), which are strikingly distinct from the growthshapes of abiotic CaCO3 crystals. In these systems, amorphous nano-particles are deposited within a vesicular space onto the surface of thegrowing biomineral where they transform to a crystalline materialthat is shape-preserving with the original ACC precursor and ex-hibits a microstructure reminiscent of aggregated particles (9, 10).The transformation of biogenic amorphous phases is believed

to be regulated by inorganic and organic components includingMg, phosphate, and carboxylated macromolecules, such as acidicproteins (8, 11–14). In particular, incorporation of Mg intobiogenic calcites is common, and its presence is believed essen-tial to stabilization of ACC (8, 13, 15) and inhibition of calcite(12, 16). However, the amorphous to crystalline transformationremains poorly understood (1, 4). Although some laboratorystudies reported direct transformation (or solid-state trans-formation) of ACC to crystalline CaCO3 (17–19), most haveconcluded ACC transforms by dissolution–reprecipitation (1,19–22). Moreover, few studies have been based on direct ob-servation of the transformation and have instead relied uponeither ex situ or bulk measurements, which are unable to relatethe measured change in properties to the dynamics of individ-ual particles. Consequently, knowledge of how impurities alter

transformation pathways and the evolution of particle compositionare largely unknown.Here we take advantage of the ability of liquid phase TEM

(LP-TEM) to follow mineral transformation processes of indi-vidual particles in real time (22, 23) in order to gain insight intothe mechanism by which additives impact the transformation ofACC to crystalline products. LP-TEM has been used to in-vestigate nanoparticle nucleation, growth, dissolution, and as-sembly of many materials including Pt (24, 25), Au (26, 27), Pd(28), Ag (29, 30), Cu (31), iron hydroxide (32), PbS (33), ironoxides (34), and CaCO3 (22, 23). These studies show that in situobservations by LP-TEM can provide high spatial and temporalresolution to directly observe the progress of phase transforma-tions of individual particles. While we recognize that there aresignificant differences between the formation of minerals in thevesicles and membranes of organisms and the solution environ-ment of a TEM sample cell, the results provide insights intophysiochemically available mechanisms of transformation andthe effect of impurities on those mechanisms.

Significance

Understanding of the shape-preserving crystallization fromtransient amorphous precursors to crystalline products in bio-mineralization makes it possible to produce topologicallycomplex morphologies that defy crystallographic controls.Here we use in situ liquid-phase TEM, FTIR, and molecular dy-namics simulations to investigate transformation of amor-phous CaCO3 to crystalline phases in the presence of multipleadditives. In situ observations reveal that Mg2+, which bringsexcess water into the bulk, is unique in initiating crystallizationwithin the amorphous phase, leading to a shape-preservingtransformation that is accompanied by dehydration. Simula-tion results find that the water accelerates ionic rearrangementwithin the solid phase, enabling the shape-preserving trans-formation. This in-depth understanding provides an alternativestrategy for manufacturing crystals with arbitrary morphol-ogies by design.

Author contributions: Z.L., Z.Z., R.T., and J.J.D.Y. designed research; Z.L., Z.Z., Z.W., B.J.,D.L., and J.T. performed research; Z.L., Z.Z., Z.W., D.L., J.T., R.T., and J.J.D.Y. analyzed data;and Z.L., Z.Z., R.T., and J.J.D.Y. wrote the paper.

The authors declare no competing interest.

This article is a PNAS Direct Submission.

Published under the PNAS license.

Data deposition: In situ LP-TEM files have been deposited in the Zenodo repository (http://doi.org/10.5281/zenodo.3609354).

See online for related content such as Commentaries.1To whom correspondence may be addressed. Email: [email protected] or [email protected].

This article contains supporting information online at https://www.pnas.org/lookup/suppl/doi:10.1073/pnas.1914813117/-/DCSupplemental.

First published February 3, 2020.

www.pnas.org/cgi/doi/10.1073/pnas.1914813117 PNAS | February 18, 2020 | vol. 117 | no. 7 | 3397–3404

CHEM

ISTR

Y

Dow

nloa

ded

by g

uest

on

Feb

ruar

y 8,

202

1

Page 2: Shape-preserving amorphous-to-crystalline transformation ... · Edited by Patricia M. Dove, Virginia Polytechnic Institute and State University, Blacksburg, VA, and approved December

Results and DiscussionIn Situ Observation of Dissolution–Reprecipitation-Based CaCO3 PhaseTransition. We observed ACC crystallization in the presence ofMg2+, sodium citrate (Na-Cit), and sodium poly-acrylate (PAA),which were chosen to represent inorganic ions, small organicmolecules, and macromolecules found in biomineralizing systems.To initiate the formation of ACC with known concentrations ofthe reagents, solutions containing 50 mM Ca2+ and HCO3

− weredeposited on the top and bottom chip of the TEM fluid cell,respectively. The additives were introduced with the Ca2+ re-agent (see SI Appendix for details). Precipitation of CaCO3 wasinitiated by assembling the chips, and imaging was performed ata calibrated electron dose rate of 3.11 e/nm2s (see SI Appendixfor details on calibration). This dose is low by the standardsof LP-TEM and is considered to exert a limited influence onaqueous environments (35), having, for example, a negligibleeffect on metallic particle nucleation rates, which are highlysensitive to hydrated electron production, for time periods upto nearly 1,000 s (29).For low concentrations of Mg2+ (2.5 mM), Na-Cit (0.5 mM =

1.5 mM of carboxyl side groups), and PAA (250 μg/mL = 2.7 mMof acrylic acid monomers), 100 to 250 nm particles were presentat 0 s (Fig. 1 A–D and SI Appendix, Figs. S1–S3) and, in thesubsequent ∼300 s, grew continuously to ∼500 nm. These par-ticles were identified as ACC via selected area electron diffrac-tion (SAED) (Fig. 1E). Upon cessation of the growth stage, theACC remained stable for times ranging from 400 to 800 s, afterwhich it dissolved over periods of ∼200 to 1,600 s, depending onthe type and concentration of the additive (Fig. 1 F and G and SIAppendix, Figs. S1–S3). The ACC lifetime and timescale fordissolution were quantified using average gray values of theTEM images (Fig. 1H and SI Appendix, Figs. S1B–S3B). DuringACC dissolution, growth of crystalline CaCO3 could also beobserved (Fig. 1 F and G, Insets, and I and SI Appendix, Figs. S4and S5), demonstrating that dissolution–reprecipitation was re-sponsible for the disappearance of ACC. Parallel experimentsperformed without an electron beam to exclude beam effectsgave similar results (SI Appendix, Fig. S6). Relative to additive-

free solutions, all of the additives either increased ACC lifetimes(Mg2+), retarded calcite formation and ACC dissolution (Na-Cit), or both (PAA) (Fig. 1J).

Direct Transformation of CaCO3 Driven by Mg Incorporation. Uponincreasing Na-Cit to 2.5 mM or PAA to 500 μg/mL, the observedbehavior remained unchanged (Table 1). However, upon in-creasing Mg2+ concentrations to 5.0 or 7.5 mM, which is knownto result in a concomitant increase in the Mg2+ content of theACC (36), a different mechanism of ACC transformation wasobtained: ACC particles initially exhibited low contrast that in-creased along with particle size (Fig. 2 A–G), and dissolution nolonger occurred even after 1 h (SI Appendix, Fig. S7B). SAEDanalysis following disassembly of the TEM cell, as well as in situ,showed the particles had transformed into crystalline calcite withspheroidal morphologies (Fig. 2H and SI Appendix, Figs. S7 Eand F, S8 A and B, and S9). Significantly, as a consequence ofthis direct transformation pathway, the number density of calciteparticles after crystallization was ∼1,000 times higher with Mg(∼20 spheroidal calcite particles per 2 × 2 μm2; SI Appendix, Fig.S9A) than without (one calcite crystal per ∼14 × 14 μm2; SIAppendix, Fig. S4B).Parallel experiments performed without the electron beam

excluded beam effects as the source of the altered trans-formation process (SI Appendix, Fig. S7 A and C). Energy dis-persive X-ray spectroscopy (EDX) on a spheroidal calciteparticle showed that the Mg present in the ACC remained in thesolid after transformation (Fig. 2I and SI Appendix, Fig. S8C),and ∼23.8% Ca (mole ratio) was replaced by Mg. Moreover, thetransformation process observed at high Mg levels was distinctfrom the direct transformation pathways reported previously forMg-free solutions (22), in which nucleation of crystalline phasesoccurred at the ACC particle surface, the spherical morphologyof the original particles was rapidly lost, and the emergingcrystalline particles exhibited the expected morphology for thespecific crystalline phase. Consequently, we conclude that theintroduction of Mg2+ at concentrations of ≥5.0 mM switchesthe ACC-to-calcite transition pathway from dissolution–repre-cipitation to a shape-preserving direct transformation.

Fig. 1. Effect of additives on dissolution–reprecipitation kinetics observed by LP-TEM. (A–D, F, and G) Growth of ACC particles with 2.5 mM Mg2+ and theirsubsequent dissolution. All particles in the main panels are ACC. (Inset) In situ images of growing calcite particles. (Scale bars, Insets of F and G, 2 μm.) (E)Confirming existence of ACC particles on chips disassembled at ∼180 s. Inset in E and I shows SAED labeled with crystal planes. (I) Cuboidal calcite crystalformed by dissolution–reprecipitation on chips disassembled at ∼1,500 s. (H) Change of gray value of image with time showing periods of ACC growth,stability, and dissolution. Each square gives the gray value at the specified time and red lines show the trends in gray values during different time periods (A–D, F, and G). (J) The lifetime (labeled “Stable”) and dissolution period (labeled “Dissolving”) of ACC with the different additives. Additive concentrations wereMg2+, 2.5 mM; sodium citrate, 0.5 mM (1.5 mM of carboxyl groups); and PAA, 250 μg/mL (2.7 mM of acrylic acid monomers). The Student’s t test is used toanalyze the statistical significance, and the P value was 0.016 (*P < 0.05) for lifetime of Mg2+; 0.019 (*P < 0.05) for dissolution period of citrate; 0.003 (**P <0.01) for lifetime of PAA and 0.002 (**P < 0.01) for dissolution period of PAA.

3398 | www.pnas.org/cgi/doi/10.1073/pnas.1914813117 Liu et al.

Dow

nloa

ded

by g

uest

on

Feb

ruar

y 8,

202

1

Page 3: Shape-preserving amorphous-to-crystalline transformation ... · Edited by Patricia M. Dove, Virginia Polytechnic Institute and State University, Blacksburg, VA, and approved December

In Situ Analysis of Structural Evolution During the DirectTransformation. To obtain structural details about the trans-formation, ideally, in situ SAED or lattice fringe analysis wouldbe used (22, 23). However, the electron beam doses required forthese methods were too large to avoid significant beam effects.Thus, we followed the development of electron contrast underthe low-dose conditions used for imaging. While the overallelectron contrast of each particle increases with time, a sharpchange in contrast begins at a specific location in each particleand then spreads throughout (Fig. 3A).The specimen affects both the amplitude and phase of the

electron beam, leading to the formation of amplitude contrast(including mass–thickness contrast and diffraction contrast) andphase contrast, respectively. Although both types of contrastcommonly contribute to an image, specific conditions can causeone type to dominate. As discussed in detail in SI Appendix,mass–thickness contrast was the major contrast observed in ourTEM experiments. The mass–thickness contrast Φ is influencedby mass and density effects on Rutherford scattering (SI Ap-pendix) (37):

Φ=N0QatomρΔD

A, [1]

where N0 is Avogadro’s number, Qatom is cross section for scat-tering per atom, ρ is density, ΔD is thickness, and A isatomic weight.While ρ should increase during the ACC to calcite transition,

particle diameter also increased with time during the transition(Figs. 2 A–F and 3A). We cannot say with certainty whether thethickness increased as well, but to get a conservative estimate ofthe change in contrast, we first assume the particles remainspherical, i.e., they thicken as much as they expand laterally.Thus, the contrast change was influenced by both ΔD and ρ. Toseparate out the effects, we calculated the relative contrast (ΔΦ):

ΔΦðtÞ= ΦðtÞΦð0Þ =

ρðtÞΔDðtÞρð0ÞΔDð0Þ , [2]

where the argument 0 refers to the initial values (Fig. 3B and SIAppendix, Table S1). Consequently, under the assumption thatthe particles are spherical throughout the experiment, the re-sults demonstrate that particle sizes grew linearly throughoutthe transition from ∼350 to 400 nm while the relative contrastexhibited an S-shaped curve with an increase from about 1.17 to2.20 occurring between 400 and 1,300 s. Thus, the increase in

relative contrast is due to increasing density during the directtransformation.From Eq. 2 and the changes in relative contrast and particle

size, which were acquired from the fits to the data in Fig. 3D, weobtain an average density change for all particles of ρ(t)/ρ(0) =1.57 ± 0.08. (The error here arises from averaging over thevalues for multiple particles rather than signal-to-noise seen inthe contrast for any one particle.) If, in fact, the particles are only

Table 1. Dominant transition pathway of ACC with different additives

Additive Concentration Pathway Stable (s) Dissolving (s)

Mg2+ 0.0 mM D-R 424 ± 75 192 ± 212.5 mM D-R 725 ± 54 213 ± 535.0 mM D-T – –

7.5 mM D-T – –

25.0 mM Beam effect* – –

Sodium citrate 0.5 mM (1.5 mM CG) D-R 402 ± 28 357 ± 672.5 mM (7.5 mM CG) D-R 428 ± 66 661 ± 77

10.0 mM (30.0 mM CG) Beam effect* – –

Sodium poly-acrylate 250 μg/mL (2.7 mM CG) D-R 812 ± 113 1,666 ± 165500 μg/mL (5.4 mM CG) D-R 1,260 ± 145 4,431 ± 411750 μg/mL (8.1 mM CG) Beam effect* – –

D-R means dissolution-reprecipitation process, and D-T stands for direct transformation. Stable and dissolvingrepresent the time periods of stability and dissolution for additive concentrations where dissolution–reprecipi-tation is observed. CG is the abbreviation for carboxyl groups.*For Mg, NaCit, and PAA concentrations of 25 and 10 mM and 750 μg/mL, respectively, the data were unreliabledue to electron beam-induced dissolution (see SI Appendix for details).

Fig. 2. In situ observations of direct ACC transformation in presence of5 mM Mg2+. (A–F) Growth and direct transformation of ACC particles. (G)Confirming existence of ACC particles on chips disassembled at ∼180 s. (H)Formation of spheroidal calcite after the direct transformation on chipsdisassembled at ∼1,800 s. Inset in G and H shows SAED labeled with crystalplanes. Typical simulated SAED for calcite is marked in red. (I) EDX analysis ofspheroidal calcite after direct transformation showing the presence of Mg.

Liu et al. PNAS | February 18, 2020 | vol. 117 | no. 7 | 3399

CHEM

ISTR

Y

Dow

nloa

ded

by g

uest

on

Feb

ruar

y 8,

202

1

Page 4: Shape-preserving amorphous-to-crystalline transformation ... · Edited by Patricia M. Dove, Virginia Polytechnic Institute and State University, Blacksburg, VA, and approved December

expanding laterally and are not thickening, then the measureddensity change increases from 1.57 to 1.79. The reported den-sities of hydrated ACC and calcite are 1.62 and 2.75 g/cm3, re-spectively (38). Considering that MgCO3 has a higher densitythan CaCO3, the assumed density of the final calcite particles,which contain ∼23.8% Mg, should be higher than that of calcite.Using a linear interpolation gives a corrected density of 2.8 g/cm3,or ρMg-calcite/ρACC = 1.7. This ratio is within 5 to 8% of the mea-sured value, further supporting the conclusion that Mg2+ inducesa direct transformation from ACC to Mg-calcite.Smeets et al. (39) used cryoTEM to show that under the

conditions used in their experiments, the first particles of calciumcarbonate were, in fact, a dense liquid phase (DLP). Theyquantified the difference in contrast between the DLP, the so-lution, and crystalline CaCO3. The magnitude of the contrastchange seen here is significantly less—2.2 vs. 3.5—indicating thatthe ACC particles, by the time we detected them, were alreadyless hydrated and not of the DLP. However, the ACC particlesare almost certainly more hydrated in solution than in the drystate. Bolze et al. (38) found that the density of ACC in solutionis 1.62 g/cm3, while Fernandez-Martinez et al. (40) obtained adensity for dry ACC powder of 2.18 g/cm3

. As the density mea-sured by Bolze et al. is nearly the same as what we obtain fromthe change in contrast, “a hydrated ACC particle” is the accuratedescription of the Mg-ACC in this system. Whether in this statethe particles exhibit mechanical properties consistent with asolid, a gel, or a dense liquid is unknown.We related the observed density changes to the development

of crystallinity using in situ attenuated total reflection-Fouriertransform infrared spectroscopy (ATR-FTIR) on identical so-lution compositions to follow the evolution of CaCO3 spectralfeatures (Fig. 4A) (12). In-growth of the two primary calcitepeaks at 720 and 880 cm−1, which appear within 2 to 4 min,dominates the time evolution of the spectrum (Fig. 4 B and C).The v4 peak that appears at 720 cm−1 is also indicative of high Mgdoping in calcite (12). However, during the first 4 min we also

observed out-growth of two previously unreported peaks at 3,600to 3,700 cm−1 associated with structural water in minerals (Fig.4D) (41). By ∼5 min, the signals of structural water were no longerobservable while the calcite peaks continued to grow. In com-plementary experiments using 25 mM Mg2+ to amplify the waterloss, thermogravometric analysis (TGA) showed >20% loss bymass, with loss continuing beyond 250 °C (SI Appendix, Fig. S10A).These TGA results reinforce the conclusion from the ATR-FTIR

Fig. 4. In situ ATR-IR analysis of CaCO3 direct transformation in the presenceof 5 mM Mg2+. (A) In situ ATR-FTIR analysis confirming transition from ACC tocalcite. (B–D) Detailed in situ ATR-FTIR spectrum of CaCO3. (B) v2 asymmetricbending. (C) v4 symmetric bending. (D) Structural water. The peak areas beforeand after 4 min at 3,665 and 3,685 cm−1 are compared, respectively. TheStudent’s t test is used to analyze the statistical significance, and the P valuewas 0.006 (**P < 0.01) for 3,665 cm−1 and 0.007 (**P < 0.01) for 3,685 cm−1.

Fig. 3. In situ TEM analysis of direct transformation of CaCO3 in the presence of 5 mM Mg2+. (A) Transformation of a single particle. (B) Gray values of CaCO3

particles in A acquired from area between yellow dashed lines. Red lines are smoothed data. Arrows in A mark locations at which contrast begins to increaseand correspond to arrows in B. (C) Distribution of gray values in B changes with time, indicating increase in area of high contrast region. (D) Size and contrastanalysis for three different particles, which are represented by circles, squares, and triangles, respectively.

3400 | www.pnas.org/cgi/doi/10.1073/pnas.1914813117 Liu et al.

Dow

nloa

ded

by g

uest

on

Feb

ruar

y 8,

202

1

Page 5: Shape-preserving amorphous-to-crystalline transformation ... · Edited by Patricia M. Dove, Virginia Polytechnic Institute and State University, Blacksburg, VA, and approved December

data that the Mg-ACC particles contain structural water. In con-trast, TGA data on ACC prepared without Mg2+ exhibited a lossof <15% with negligible dehydration occurring beyond 250 °C (SIAppendix, Fig. S10B). Notably, previous work by Blue et al. (42)also observed a correlation between Mg2+ content and the waterwithin ACC.

Understanding the Role of Mg in Altering the Mechanism of PhaseTransformation. The above findings show that high Mg2+ incor-poration drives transformation to Mg-calcite within the bulk ofhydrated ACC particles, with the transformation front propa-gating through the particle as the water is expelled and thematerial densifies, leaving the amorphous morphology intact.However, the results do not provide insights into the separateroles of incorporated Mg2+, water, and free Mg2+ in driving thispathway. Previous research showed that Mg2+ ions can inhibitcalcite crystal growth via two main pathways: 1) incorporation ofMg2+, leading to a lower supersaturation of calcite by increasingthe mineral solubility (11), and 2) the adsorption of Mg2+ on step-edges or terraces of calcite, decreasing the velocity of migratingsteps likely due to slower ion desolvation (43, 44). Other studiesshowed that water is a key factor in enabling crystallization andreorientation of ACC (45–48). Indeed, TGA data (SI Appendix,

Fig. S10) prove that Mg2+ incorporation increases the watercontent of ACC. Thus, either Mg2+ or water or the combination ofboth could affect the crystallization pathway.The shape-preserving manner of the direct transformation

enabled us to determine the importance of incorporated Mg,incorporated water, and free Mg2+ in driving that pathway.Spherical Mg-ACC particles were first precipitated in the pres-ence of 25 mM Mg2+ (Fig. 5A). When left in 25 mMMg2+ aqueous solution, spherical calcite was obtained, indicatinga direct transformation (Fig. 5B). In contrast, when the Mg-ACCparticles were precipitated in 25 mM Mg2+ aqueous solution andquickly centrifuged and redispersed in pure water, large facetedcalcite particles were formed in place of the Mg-ACC or sphericalMg-calcite (Fig. 5C), indicating dissolution–reprecipitation domi-nated. Similarly, when pure ACC particles (without Mg doping)were placed into 25 mM Mg2+ solution, faceted calcite particleswith sporadic vaterite were generated (Fig. 5D). Notably, the calcitein Fig. 5C is more spheroidal than is that in Fig. 5D. We assume thatdirect transformation of the ACC begins to occur within the particledue to the Mg2+ doping, but due to a lack of Mg2+ in solution thedissolution–reprecipitation that is usually inhibited when Mg2+ ispresent happens synchronously. Thus, the small morphologicaldifference is observed. Finally, when spherical Mg-ACC was heated

Fig. 5. Understanding the role of free Mg2+, doped Mg, and water on crystallization of ACC by SEM and FTIR. (A) Precipitation of spherical Mg-ACC in thepresence of 25 mM Mg2+. (B) Crystallization of Mg-ACC to spherical calcite in the presence of 25 mM free Mg2+. (C) Crystallization of Mg-ACC to calcitewithout Mg2+ in solution. (D) Precipitation of ACC without Mg2+. In the subsequent phase transition stage, Mg2+ was added to maintain 25 mM free Mg2+ insolution. (E) Heating the Mg-ACC in A at 200 °C for 20 min. The dried CaCO3 was still amorphous with a spherical shape. (F) The dried Mg-ACC in E wascrystallized to aragonite (Ara) and calcite (Cal) in the presence of 25 mM free Mg2+ in solution.

Liu et al. PNAS | February 18, 2020 | vol. 117 | no. 7 | 3401

CHEM

ISTR

Y

Dow

nloa

ded

by g

uest

on

Feb

ruar

y 8,

202

1

Page 6: Shape-preserving amorphous-to-crystalline transformation ... · Edited by Patricia M. Dove, Virginia Polytechnic Institute and State University, Blacksburg, VA, and approved December

to 200 °C to remove the ∼16% incorporated water (Fig. 5E) andthen placed in 25 mMMg2+ solution, only needle-like aragonite wasobtained (Fig. 5F), again implying a dissolution–reprecipitationpathway. Thus, we conclude that all three components are necessaryfor the direct transformation to take place: incorporated Mg2+,incorporated water, and free Mg2+.From the data on ACC lifetime (Fig. 1H), the known in-

hibition of calcite crystallization by free Mg2+ (12, 45), and theobservation that in the absence of free Mg2+, Mg-ACC isreplaced by faceted calcite, the role of Mg2+ is clear. First, in itsabsence, ACC is kinetically stabilized due to the long incubationtime for nucleation of calcite in solution. However, once calciteforms in solution, it consumes the dissolved ions as it grows,reducing the activity product toward the solubility of calcite.Once the activity product drops below the solubility limit ofACC, the ACC begins to dissolve. In contrast, when Mg2+ ispresent, two changes take place. First, Mg2+ inhibits the nucle-ation of calcite in solution thus increasing the lifetime of theACC, even at low Mg2+ concentrations (Fig. 1H). This longerlifetime then combines with the second effect of Mg2+, which isto destabilize the ionic network and enhance ionic reconfigura-tions, to drive calcite nucleation within the Mg-ACC particleinstead of the surrounding solution. Then the dehydration–transformation front must reach the surface of the Mg-ACCparticle and come in contact with surrounding solution beforeit can begin to deplete the dissolved ions through growth. As longas the dehydration–transformation front moves through the en-tire particle too rapidly for significant dissolution to occur fol-lowing first contact of the front with the solution, the particle willremain largely intact. The specific and relative roles of in-corporated Mg and incorporated water are less clear. Moreover,whether the water must be part of the ACC structure or only theamount of water in the ACC matters is uncertain. However,based on the observations above, we hypothesize that Mg2+,which is a strongly hydrated ion, brings additional water into theACC (45), and the water provides the molecular mobility neededto nucleate calcite within the Mg ACC particles.To test the hypothesized function of Mg2+ and water in the

transformation process and understand whether the water plays astructural role, we employed molecular dynamics (MD) simulationsto examine the structural dynamics of ACC and Mg-ACC clustersfor a range of Mg concentrations. ACC clusters were generatedfrom Ca2+/Mg2+ and CO3

2− solutions by a procedure which wasused to minimize the artificial influence on the initial structure ofthe ACC clusters (refer to simulation details of SI Appendix, Fig.S11). Specifically, we placed the solutions between two walls whichonly allowed water molecules to pass. The free water and weak-binding water passed through the walls, leaving behind the ions andstrongly binding water, which constitute the ACC/Mg-ACC clusters.These clusters were subsequently used as the initial configurationsfor the final production simulations. Both sets of simulations werecarried out using previously developed potentials (49–51).This approach produced clusters that could be reasonably as-

sumed as representative of those formed in the experiments. Forexample, according to the results from the TGA experiments,ACC has ∼15 wt % water content, and Mg-ACC with 23.8% Mgdoping has ∼20 wt % water content. In our simulations, the Mg-free ACC clusters contained ∼16.1 wt % water, while the Mg-ACC clusters into which Mg was doped at 20 wt % had a watercontent of ∼19.6 wt %, in good agreement with experiments.Moreover, the atomic configurations obtained in the simulationscould be reasonably taken as reflective of isolated regions withinlarger ACC particles because the Mg-ACC cluster containedabout 3 times as many ions and molecules (800 Ca2+ plus Mg2+,800 CO3

2−, and 840 to 1,400 water molecules) than used in pre-vious published simulations (800 ions and molecules in total) (51).The structural dynamics of the resulting ACC clusters were

evaluated by examining the radial distribution function (RDF)

profiles, residence times, and coordination numbers between theion pairs in ACC clusters. The MD results predict that, due to itshigh dehydration barrier, Mg2+ brings additional water into ACC(Fig. 6 A and B and SI Appendix, Tables S2 and S3). The resultsalso predict that the water molecules, which form hydrogenbonds to CO3

2−, are relatively unstable sites in the ionic-bindingnetwork. This is seen by comparing the sharpness of the peaks inthe RDF profiles, which depends on the binding energies (SIAppendix). The peaks representing cation–OC binding, thoughdifferent in position due to the smaller size of Mg2+ ion thanCa2+, are both sharp, implying strong binding affinities betweenCO3

2− and both Ca2+ and Mg2+ (Fig. 6C). In contrast, the RDFprofile of H2O–OC exhibits two broad overlapping peaks, in-dicating weak binding between water and CO3

2−, which estab-lishes a weak point in the ionic-binding network of the ACCcluster that results in stronger ionic rearrangements for Mg-ACCthan ACC (Fig. 6D). This greater propensity for rearrangementis reflected in the residence times of water molecules aroundCO3

2− (HW–CO32−), Ca2+ (OW–Ca2+), and Mg2+ (OW–Mg2+),

which rank in the order of Mg2+ >> Ca2+ >> CO32− (SI Ap-

pendix, Table S4). The fact that strong cation–anion bonds arepartially replaced by weak, unstable H bonds between H2O andCO3

2− indicates that some of the water has a role in the structureof the ACC at least some of the time (Fig. 6 C and D and SIAppendix, Fig. S12). More importantly, the weak H2O–OCbinding and high rate of HW–CO3

2− rearrangement in thewater-rich Mg-ACC implies an enhanced probability for nucle-ation of a crystalline phase within the particle.These conclusions are consistent with recent results based on

coherent X-ray and incoherent neutron scattering demonstratingthat the presence of Mg2+ in ACC can increase the content ofboth mobile and rigid water, leading to a higher frequency ofstructural rearrangements within the mineral (52), although the

Fig. 6. Mechanism of Mg2+ regulation of ACC phase transition pathway.Atomic configuration details around (A) Ca2+ and (B) Mg2+ in MD simulation(Ca2+, green sphere; Mg2+, purple sphere; H2O, red/white; CO3, gray/red). (C)Radial distribution function (RDF, G(r)) between positively and negativelycharged atoms showing broadened peak for H atoms of water and O atomsof CO3 (Ca, Ca2+; Mg, Mg2+; OC, O atoms in CO3

2−; HW, H atoms of water).(D) Local binding network of water molecules in simulated ACC. The Hbonds between water molecules and CO3

2− are unstable, easily breakingand regenerating spontaneously. (E) Two pathways of ACC phase trans-formation regulated by Mg2+. Without Mg2+, ACC typically follows a dis-solution–recrystallization pathway to form rhombohedral calcite. However,in the presence of Mg2+, nucleation of calcite occurs within the ACC and isaccompanied by dehydration. As a result, a direct transformation is achievedin the absence of morphological changes.

3402 | www.pnas.org/cgi/doi/10.1073/pnas.1914813117 Liu et al.

Dow

nloa

ded

by g

uest

on

Feb

ruar

y 8,

202

1

Page 7: Shape-preserving amorphous-to-crystalline transformation ... · Edited by Patricia M. Dove, Virginia Polytechnic Institute and State University, Blacksburg, VA, and approved December

authors attributed the behavior to stress effects. Yang et al. (53)found that Mg-ACC actually consists of two types of carbonates,amorphous CaCO3 and amorphous MgCO3. Using quasi-elasticneutron scattering, Jensen et al. (54) recently observed thatwater diffuses in amorphous MgCO3 with similar dynamics tothat found for water in clays, although they did not report higherwater mobility in Mg-ACC than in pure ACC. All these resultsare consistent with the conclusion that incorporation of Mg2+

into ACC introduces relatively unstable sites in the ACC structure,increasing the possibility of ion rearrangement, thus enhancingthe probability of structural transformations. With regard to thefinding that a direct transformation process dominates in thepresence of high Mg content, most previous studies indeedfound that free Mg2+ inhibits nucleation of calcite in solution,thereby decreasing the crystallization rate during the dissolution-reprecipitation process, but the applicability of the findings tothe direct transformation process investigated here is unclear.Moreover, previous work has shown that extremely high Mg2+

concentrations can initiate transformation of ACC to aragoniteeither in simulated sea water (55) or in solution containing ahigh content of polymer (56). Whether the mechanism reportedhere still dominates is unknown.Dissolution–reprecipitation and direct transformation can be

understood as two competitive pathways, with the dominantpathway dependent on the site of nucleation: nucleation of calcitein solution leads to dissolution–reprecipitation, while nucleationwithin ACC particles initiates a direct transformation. Commonly,the mobility of ions in solution is much higher than in the solid;thus, nucleation in solution is much more rapid. However, thepresence of Mg2+ in CaCO3 solutions lowers the supersaturationwith respect to calcite by increasing the mineral solubility (11). Inaddition, the adsorption of Mg2+ onto the surfaces of calciteblocks its growth (43, 44). In contrast, Mg2+ brings more waterinto ACC, increasing the nucleation rate in (or on) the solid.Moreover, the additional structural water that Mg2+ introducesinto Mg-ACC likely makes it less soluble than pure ACC (52).Consequently, once the concentration of Mg2+ reaches a certainlevel (5 mM for our experimental conditions), nucleation fromACC happens before nucleation from solution, causing the directtransformation to dominate (Fig. 6E).

ConclusionsThe findings reported above reveal that, while for pure solutionsand most additives investigated here, ACC transformation intocalcite is dominated by dissolution–reprecipitation, doping ofMg2+ into ACC increases the water content, weakens the ionic-binding network, and leads to a direct transformation in theabsence of morphological changes (Fig. 6E). In addition, Mginhibits the nucleation of calcite within the bulk solution, leadingto an increase in the period of stability (11, 43, 44). Thus, Mg2+

plays two roles: 1) It decreases the supersaturation of the solu-tion with respect to calcite, as well as the crystal growth rate,leading to the inhibition of the dissolution-reprecipitationpathway and increasing the time available for direct trans-formation. 2) It acts as a carrier to bring excess structural waterinto ACC, generating relatively unstable sites that promote thetransformation, as proposed previously (46, 57). The results alsoshow that direct transformation of ACC is a condensation pro-cess accompanied by dehydration. As a result, the Mg is retainedwithin the particle, leading to formation of Mg-calcite. Thisoutcome stands in stark contrast to previous findings showingMg2+ is lost during ACC transformation via dissolution–repre-cipitation (36) and suggests a plausible reason why Mg is such acommon impurity in biominerals (8, 11, 13, 15, 58). Moreover, thedirect transformation is shape-preserving, implying that Mg dopingis potentially exploited by organisms to assist in formation of curved,noncrystalline forms commonly seen in natural biominerals (2, 6, 12,15, 58). Because the morphological characteristics of materials

endow them with specific functions (59, 60), the ability to controlmaterial morphology is an important goal in material science.Previous work has shown that amorphous precursors are moldable(61). The findings reported here demonstrate a potential methodfor initiating a shape-preserving direct transformation in such pre-cursors. Consequently, our results point toward a process that di-rectly transforms an amorphous precursor to a crystalline phasewithout morphological changes, providing a viable route tomanufacturing crystals with arbitrary morphologies—especiallywith curved surfaces—by design.

MethodsTEM Liquid Cells. Liquid cell chips (Hummingbird Scientific) similar to thosedescribed previously were used for LC-TEM observation (22). Briefly, the liquidcell consists of two square 4 mm2 silicon chips with 50 nm thick silicon nitridemembranes in 30 × 200 μm2 windows for imaging. Each chip has eight Au padswith a thickness of 500 nm, and the chips were aligned with the pads were onthe internal faces to create a space for particle growth. All of the chips wereplasma cleaned in a Plasma Cleaner (Harrick Plasma) for 30 s prior to use bybleeding in 200 to 400 mTorr of ambient atmosphere.Sample preparation. Different from previous LC-TEM experiments on CaCO3

crystallization, the dual flow method was not used here owing to the un-certainty of the resulting ion concentrations, which was an important param-eter to fix in investigating the influence of additives. Instead, a static methodwas applied by placing 100 mM CaCl2 (with or without additives) and NaHCO3

solution separately on different liquid cell chips and assembling the chips tostart the precipitation of ACC. Although this method can miss the earliestmoments of ACC nucleation, the phase transition from ACC to crystallinepolymorph could be easily captured at fixed and known solution conditions.

Typically, after the chips were cleaned, 0.6 μL of the abovementionedCaCl2 solution (with or without additives) was placed on one chip, while0.6 μL of 100 mM NaHCO3 aqueous solution was placed on the other one.Then the chips were quickly assembled inside the liquid cell holder (Hum-mingbird Scientific) to check the vacuum seal. Posttest, the samples wereimmediately put into TEM for observation, with the time between chip as-sembling and observation being within 5 min.

ElectronMicroscopy. Electronmicroscopywas conducted in a field emission FEITecnai G2 F20 (Thermo Fisher Scientific) operated at 200 kV. Snapshots werecaptured by an Eagle CCD (1,024 × 1,024 pixels), which was controlled fromthe Tecnai User Interface (TUI), and snapshots were further processed usingthe TEM Imaging & Analysis (TIA) interface.

Different from previous LC-TEM observation on particle nucleation (22,23), the crystallization from ACC to crystal required a longer time, especiallywhen additives stabilized the ACC phase. To minimize beam effects a lowelectron dose rate was used to observe the phase transition.

In order to control the electron dose rate, we used the same TEM ob-servation conditions (including condenser aperture, spot size, magnification,C2 lens, etc.) for every experiment.We chose a diameter of 50 μm for the secondcondenser aperture. The spot size was 6. In order to increase the mass–thicknesscontrast, an objective aperture of 35 μm was used. The measurements weremade with a blank holder under same observation condition, as well as withthe liquid cell chips in the holder for comparison. The dose rate after correctingfor the measured calibration factor was 3.11 e/nm2s with a blank holder; thus,the electron dose in our experiments was no more than 3.11 e/nm2s, which was100 times smaller than previous work on the observation of calcium carbonate.As a result, the exposure time for each acquisition was prolonged to 2 s toobtain adequate contrast while keeping the total electron dose at low values.We confirmed that this was a suitable dose for observing the phase transitionwith negligible impact on the timescale of transformation events, as discussedin SI Appendix.

Scanning electronmicroscopy images were acquired by a field emission SU-8010 (Hitachi) operated at 5 kV.

ATR-IR Measurements. The precipitation of Mg-ACC was monitored over timeusing a Bruker Alpha Fourier transform infrared (FTIR) spectrometerequipped with a Harrick ConcentratIR sampling accessory with a diamondattenuated total reflectance (ATR) sampling plate. For each FTIR spectrumrecorded, 16 scans were made at a resolution of 2 cm−1. All samples weremade by mixing equal volumes of calcium solution (100 mM CaCl2, 10 mMMgCl2) and bicarbonate solution (100 mM NaHCO3). The backgrounds weredetermined using 50 mM NaHCO3 solution. The first spectrum was recorded∼10 s after mixing the solutions and the rest of the spectra were recorded atthe defined time intervals.

Liu et al. PNAS | February 18, 2020 | vol. 117 | no. 7 | 3403

CHEM

ISTR

Y

Dow

nloa

ded

by g

uest

on

Feb

ruar

y 8,

202

1

Page 8: Shape-preserving amorphous-to-crystalline transformation ... · Edited by Patricia M. Dove, Virginia Polytechnic Institute and State University, Blacksburg, VA, and approved December

Data Availability. The detailed description of the experiment and experimentaldata are available in the manuscript and SI Appendix. In situ LP-TEM files areavailable in the Zenodo repository (http://doi.org/10.5281/zenodo.3609354).

ACKNOWLEDGMENTS. We thank Jennifer Soltis and Guomin Zhu for TEMassistance and Yueqi Zhao for SEM assistance. We also thank Fan Yang,Jiajun Chen, and Boyan Li for help. Materials synthesis, ATR-IR spectroscopyand electron microscopy were supported by the US Department of Energy(DOE), Office of Basic Energy Sciences (BES), Division of Materials Science andEngineering, Synthesis and Processing Sciences Program at Pacific NorthwestNational Laboratory (PNNL). Analysis of TEM and TGA data and MD simulationswere supported by the International Exchange Program of Zhejiang University,

National Natural Science Foundation of China (Grants 21625105, 21805241,and 11904300), the Fujian Provincial Department of Science & Technology(Grant 2017J05028), the Fundamental Research Funds for the Central Uni-versities (Grant 20720180018), and the China Postdoctoral Science Founda-tion (Grants 2017M621909 and 2018T110585). ATR-IR measurements wereperformed at the Environmental and Molecular Sciences Laboratory, a DOEOffice of Science User Facility at PNNL sponsored by the Office of Biologicaland Environmental Research. In situ TEM capability development was sup-ported by the Materials Synthesis and Simulation Across Scales Initiativethrough the Laboratory Directed Research and Development Program atPNNL. PNNL is a multiprogram national laboratory operated for Departmentof Energy by Battelle under Contract DE-AC05-76RL01830.

1. J. J. De Yoreo et al., CRYSTAL GROWTH. Crystallization by particle attachment insynthetic, biogenic, and geologic environments. Science 349, aaa6760 (2015).

2. L. Addadi, S. Raz, S. Weiner, Taking advantage of disorder: Amorphous calcium car-bonate and its roles in biomineralization. Adv. Mater. 15, 959–970 (2003).

3. L. B. Gower, Biomimetic model systems for investigating the amorphous precursorpathway and its role in biomineralization. Chem. Rev. 108, 4551–4627 (2008).

4. J. H. Cartwright, A. G. Checa, J. D. Gale, D. Gebauer, C. I. Sainz-Díaz, Calcium car-bonate polyamorphism and its role in biomineralization: How many amorphous cal-cium carbonates are there? Angew. Chem. Int. Ed. Engl. 51, 11960–11970 (2012).

5. J. W. Morse, R. S. Arvidson, A. Lüttge, Calcium carbonate formation and dissolution.Chem. Rev. 107, 342–381 (2007).

6. Y. Politi, T. Arad, E. Klein, S. Weiner, L. Addadi, Sea urchin spine calcite forms via atransient amorphous calcium carbonate phase. Science 306, 1161–1164 (2004).

7. N. Nassif et al., Amorphous layer around aragonite platelets in nacre. Proc. Natl. Acad.Sci. U.S.A. 102, 12653–12655 (2005).

8. J. Tao, D. Zhou, Z. Zhang, X. Xu, R. Tang, Magnesium-aspartate-based crystallizationswitch inspired from shell molt of crustacean. Proc. Natl. Acad. Sci. U.S.A. 106, 22096–22101 (2009).

9. Y. Politi et al., Transformation mechanism of amorphous calcium carbonate into calcitein the sea urchin larval spicule. Proc. Natl. Acad. Sci. U.S.A. 105, 17362–17366 (2008).

10. C. E. Killian et al., Mechanism of calcite co-orientation in the sea urchin tooth. J. Am.Chem. Soc. 131, 18404–18409 (2009).

11. K. J. Davis, P. M. Dove, J. J. De Yoreo, The role of Mg2+ as an impurity in calcitegrowth. Science 290, 1134–1137 (2000).

12. E. Loste, R. M. Wilson, R. Seshadri, F. C. Meldrum, The role of magnesium in stabilisingamorphous calcium carbonate and controlling calcite morphologies. J. Cryst. Growth254, 206–218 (2003).

13. D. Wang, A. F. Wallace, J. J. De Yoreo, P. M. Dove, Carboxylated molecules regulatemagnesium content of amorphous calcium carbonates during calcification. Proc. Natl.Acad. Sci. U.S.A. 106, 21511–21516 (2009).

14. Y. U. Gong et al., Phase transitions in biogenic amorphous calcium carbonate. Proc.Natl. Acad. Sci. U.S.A. 109, 6088–6093 (2012).

15. S. Raz, S. Weiner, L. Addadi, Formation of high‐magnesian calcites via an amorphousprecursor phase: Possible biological implications. Adv. Mater. 12, 38–42 (2000).

16. F. C. Meldrum, S. T. Hyde, Morphological influence of magnesium and organic ad-ditives on the precipitation of calcite. J. Cryst. Growth 231, 544–558 (2001).

17. E. M. Pouget et al., The development of morphology and structure in hexagonalvaterite. J. Am. Chem. Soc. 132, 11560–11565 (2010).

18. J. Ihli et al., Dehydration and crystallization of amorphous calcium carbonate in so-lution and in air. Nat. Commun. 5, 3169 (2014).

19. A. Gal et al., Calcite crystal growth by a solid-state transformation of stabilizedamorphous calcium carbonate nanospheres in a hydrogel. Angew. Chem. Int. Ed.Engl. 52, 4867–4870 (2013).

20. G. Assaf et al., Particle accretion mechanism underlies biological crystal growth froman amorphous precursor phase. Adv. Funct. Mater. 24, 5420–5426 (2014).

21. Q. Hu et al., The thermodynamics of calcite nucleation at organic interfaces: Classicalvs. non-classical pathways. Faraday Discuss. 159, 509 (2012).

22. M. H. Nielsen, S. Aloni, J. J. De Yoreo, In situ TEM imaging of CaCO3 nucleation revealscoexistence of direct and indirect pathways. Science 345, 1158–1162 (2014).

23. P. J. Smeets, K. R. Cho, R. G. Kempen, N. A. Sommerdijk, J. J. De Yoreo, Calciumcarbonate nucleation driven by ion binding in a biomimetic matrix revealed by in situelectron microscopy. Nat. Mater. 14, 394–399 (2015).

24. H. Zheng et al., Observation of single colloidal platinum nanocrystal growth trajec-tories. Science 324, 1309–1312 (2009).

25. J. M. Yuk et al., High-resolution EM of colloidal nanocrystal growth using grapheneliquid cells. Science 336, 61–64 (2012).

26. X. Ye et al., Single-particle mapping of nonequilibrium nanocrystal transformations.Science 354, 874–877 (2016).

27. N. D. Loh et al., Multistep nucleation of nanocrystals in aqueous solution. Nat. Chem.9, 77–82 (2017).

28. B. Jin et al., Revealing the cluster-cloud and its role in nanocrystallization. Adv. Mater.31, e1808225 (2019).

29. T. J. Woehl, J. E. Evans, I. Arslan, W. D. Ristenpart, N. D. Browning, Direct in situdetermination of the mechanisms controlling nanoparticle nucleation and growth.ACS Nano 6, 8599–8610 (2012).

30. E. Sutter et al., In situ liquid-cell electron microscopy of silver-palladium galvanic re-placement reactions on silver nanoparticles. Nat. Commun. 5, 4946 (2014).

31. M. J. Williamson, R. M. Tromp, P. M. Vereecken, R. Hull, F. M. Ross, Dynamic mi-croscopy of nanoscale cluster growth at the solid-liquid interface. Nat. Mater. 2, 532–536 (2003).

32. K. Niu et al., Bubble nucleation and migration in a lead-iron hydr(oxide) core-shellnanoparticle. Proc. Natl. Acad. Sci. U.S.A. 112, 12928–12932 (2015).

33. J. E. Evans, K. L. Jungjohann, N. D. Browning, I. Arslan, Controlled growth of nano-particles from solution with in situ liquid transmission electron microscopy. Nano Lett.11, 2809–2813 (2011).

34. D. Li et al., Direction-specific interactions control crystal growth by oriented attach-ment. Science 336, 1014–1018 (2012).

35. N. M. Schneider et al., Electron-water interactions and implications for liquid cellelectron microscopy. J. Phys. Chem. C 118, 22373–22382 (2014).

36. C. R. Blue et al., Chemical and physical controls on the transformation of amorphouscalcium carbonate into crystalline CaCO3 polymorphs. Geochim. Cosmochim. Acta196, 179–196 (2017).

37. C. B. Carter, D. B. Williams, Transmission Electron Microscopy (Springer-Verlag, 2009).38. J. Bolze et al., Formation and growth of amorphous colloidal CaCO3 precursor par-

ticles as detected by time-resolved SAXS. Langmuir 18, 8364–8369 (2002).39. P. J. M. Smeets et al., A classical view on nonclassical nucleation. Proc. Natl. Acad. Sci.

U.S.A. 114, E7882–E7890 (2017).40. A. Fernandez-Martinez, B. Kalkan, S. M. Clark, G. A. Waychunas, Pressure-induced

polyamorphism and formation of ‘aragonitic’ amorphous calcium carbonate.Angew. Chem. Int. Ed. Engl. 52, 8354–8357 (2013).

41. J. Madejová, FTIR techniques in clay mineral studies. Vib. Spectrosc. 31, 1–10 (2003).42. C. Blue, P. Dove, Chemical controls on the magnesium content of amorphous calcium

carbonate. Geochim. Cosmochim. Acta 148, 23–33 (2015).43. A. Gutjahr, H. Dabringhaus, R. Lacmann, Studies of the growth and dissolution ki-

netics of the CaCO3 polymorphs calcite and aragonite. 2. The influence of divalentcation additives on the growth and dissolution rates. J. Cryst. Growth 158, 310–315(1996).

44. N. Kanzaki, K. Onuma, G. Treboux, S. Tsutsumi, A. Ito, Inhibitory effect of magnesiumand zinc on crystallization kinetics of hydroxyapatite (0001) face. J. Phys. Chem. B 104,4189–4194 (2000).

45. Y. Politi et al., Role of magnesium ion in the stabilization of biogenic amorphouscalcium carbonate: A structure−Function investigation. Chem. Mater. 22, 161–166(2009).

46. J. Ihli, A. N. Kulak, F. C. Meldrum, Freeze-drying yields stable and pure amorphouscalcium carbonate (ACC). Chem. Commun. (Camb.) 49, 3134–3136 (2013).

47. A. Jensen et al., Hydrogen bonding in amorphous calcium carbonate and molecularreorientation induced by dehydration. J. Phys. Chem. C 122, 3591–3598 (2018).

48. M. Albéric et al., The crystallization of amorphous calcium carbonate is kineticallygoverned by ion impurities and water. Adv. Sci. (Weinh.) 5, 1701000 (2018).

49. J. W. Shen, C. Li, N. F. Van Der Vegt, C. Peter, Understanding the control of miner-alization by polyelectrolyte additives: Simulation of preferential binding to calcitesurfaces. J. Phys. Chem. C 117, 6904–6913 (2013).

50. P. Raiteri, J. D. Gale, D. Quigley, P. M. Rodger, Derivation of an accurate force-field forsimulating the growth of calcium carbonate from aqueous solution: A new model forthe calcite−Water interface. J. Phys. Chem. C 114, 5997–6010 (2010).

51. H. Tomono et al., Effects of magnesium ions and water molecules on the structure ofamorphous calcium carbonate: A molecular dynamics study. J. Phys. Chem. B 117,14849–14856 (2013).

52. A. Koishi et al., Role of impurities in the kinetic persistence of amorphous calciumcarbonate: A nanoscopic dynamics view. J. Phys. Chem. C 122, 16983–16991 (2018).

53. S. Y. Yang, H. H. Chang, C. J. Lin, S. J. Huang, J. C. Chan, Is Mg-stabilized amorphouscalcium carbonate a homogeneous mixture of amorphous magnesium carbonate andamorphous calcium carbonate? Chem. Commun. (Camb.) 52, 11527–11530 (2016).

54. A. C. S. Jensen, I. Rodriguez, W. J. E. M. Habraken, P. Fratzl, L. Bertinetti, Mobility ofhydrous species in amorphous calcium/magnesium carbonates. Phys. Chem. Chem.Phys. 20, 19682–19688 (2018).

55. W. Sun, S. Jayaraman, W. Chen, K. A. Persson, G. Ceder, Nucleation of metastablearagonite CaCO3 in seawater. Proc. Natl. Acad. Sci. U.S.A. 112, 3199–3204 (2015).

56. X. Cheng, P. L. Varona, M. J. Olszta, L. B. Gower, Biomimetic synthesis of calcite filmsby a polymer-induced liquid-precursor (PILP) process: 1. Influence and incorporationof magnesium. J. Cryst. Growth 307, 395–404 (2007).

57. P. Raiteri, J. D. Gale, Water is the key to nonclassical nucleation of amorphous calciumcarbonate. J. Am. Chem. Soc. 132, 17623–17634 (2010).

58. I. Polishchuk et al., Coherently aligned nanoparticles within a biogenic single crystal:A biological prestressing strategy. Science 358, 1294–1298 (2017).

59. S. Mann, P. Behrens, E. Bäuerlein, Handbook of Biomineralization: Biomimetic andBioinspired Chemistry (Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, 2007).

60. B. Yeom et al., Abiotic tooth enamel. Nature 543, 95–98 (2017).61. H. Cölfen, M. Antonietti, Mesocrystals and Nonclassical Crystallization (John Wiley &

Sons Ltd., West Sussex, 2008).

3404 | www.pnas.org/cgi/doi/10.1073/pnas.1914813117 Liu et al.

Dow

nloa

ded

by g

uest

on

Feb

ruar

y 8,

202

1