Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2008 Nov 5;105(45):17362–17366. doi: 10.1073/pnas.0806604105

Transformation mechanism of amorphous calcium carbonate into calcite in the sea urchin larval spicule

Yael Politi a, Rebecca A Metzler b, Mike Abrecht c, Benjamin Gilbert d, Fred H Wilt e, Irit Sagi a, Lia Addadi a,1, Steve Weiner a,1, P U P A Gilbert b,1,2
PMCID: PMC2582271  PMID: 18987314

Abstract

Sea urchin larval spicules transform amorphous calcium carbonate (ACC) into calcite single crystals. The mechanism of transformation is enigmatic: the transforming spicule displays both amorphous and crystalline properties, with no defined crystallization front. Here, we use X-ray photoelectron emission spectromicroscopy with probing size of 40–200 nm. We resolve 3 distinct mineral phases: An initial short-lived, presumably hydrated ACC phase, followed by an intermediate transient form of ACC, and finally the biogenic crystalline calcite phase. The amorphous and crystalline phases are juxtaposed, often appearing in adjacent sites at a scale of tens of nanometers. We propose that the amorphous-crystal transformation propagates in a tortuous path through preexisting 40- to 100-nm amorphous units, via a secondary nucleation mechanism.

Keywords: biomineralization, Ca L-edge X-ray absorption near-edge structure, XANES, X-PEEM, X-ray photoelectron emission spectromicroscopy


A widespread strategy in biomineralization is the initial formation of transient amorphous precursor phases that subsequently transform into one of the more stable crystalline phases (1). This process was first observed in the teeth of chitons where a disordered ferrihydrite precursor transforms into magnetite (2). It has also been observed in different invertebrate phyla (38). Amorphous calcium phosphate was recently identified in the newly deposited fin bones of zebrafish (9). The mechanistic details of these transformations are, however, still poorly understood. Here, we address this fundamental issue by studying the transformation of amorphous calcium carbonate (ACC) to crystalline calcite in the sea urchin larval spicule. Sea urchin larval spicules have long served as a model system for the study of CaCO3 biomineralization processes, and the transient ACC precursor phase was first identified in this system (3). The mature larval spicule is composed of a single crystal of magnesium-bearing calcite (10, 11). Small amounts of organic macromolecules (0.1 wt%) are incorporated within the mineral and are known to play a role in the transient stabilization of the amorphous phase (12).

The spicules are formed inside a syncytium produced by specialized cells (13). The first deposit is a single rhombohedral-shaped calcite crystal. Further growth of the spicule radii follows crystallographic orientations dictated by the initial crystal (10, 14), even though the mineral deposited is mainly in the form of ACC. The rays elongate rapidly for ≈3 days, while the existing rays thicken. ACC is most probably introduced into the mineralization site by the cells in vesicles that fuse with the syncytial membrane (15). The spicule is tightly surrounded by this membrane, with no interstitial water solution detectable at any stage (16). In the polarized light microscope almost the entire spicule behaves as a homogeneously bright birefringent domain, despite being composed mainly of an amorphous phase. The exceptions are the growing tips of the spicule that show no birefringence, suggesting that they are completely amorphous (16). Partial demineralization (etching) of the spicule shows that the spicule is composed of densely packed mineral spherules 40–100 nm in diameter (3, 17). No crystallization front can be detected at the micrometer scale. Extended X-ray absorption fine structure (EXAFS) spectroscopy at the Ca K-edge showed that even at early stages, when the mineral is still predominantly amorphous, it already has a nascent short-range order around the calcium ions similar to that in calcite (18). In contrast to stable biogenic ACC, which contains 1 water molecule per CaCO3, the amorphous phase in the spicules is mostly anhydrous when the spicules are extracted at an advanced developmental stage (12, 19). Macroscopically, therefore, the spicule displays both amorphous and crystalline qualities.

Results and Discussion

Unraveling the mechanistic complexities of the spatial and temporal interplay between the transforming amorphous and crystalline phases requires the use of high-resolution techniques. Here, we use X-ray photoelectron emission spectromicroscopy (X-PEEM) to study the transformation at high spatial resolution (20, 21). We analyze X-ray absorption near-edge structure (XANES) (22) spectra at the Ca L-edge along the length of spicules at two developmental stages. Ca spectra were acquired by recording 170 images, 0.1 eV apart, and arranging them in stacks in which the energy-dependent intensity of each pixel holds the full spectral information across the Ca L-edge. The pixel size depends on the magnification and is 40–200 nm in this study, while the probing depth is ≈3 nm at the Ca L-edge energy range (23). This technique offers the unique opportunity of characterizing the atomic order of the mineral phase (24) along a single larval spicule with sub-micrometer spatial resolution, providing time and space-resolved snapshots of the crystallization pathway through 2 distinct amorphous phases.

Fig. 1 shows spectra acquired from a 48-h embryo spicule with a pixel size of 200 nm. At this stage the spicule is at the triradiate stage of development and is composed of 70–90% ACC (18). The spectra were extracted from areas near the tip and along 1 of the spicule radii. The spectra near the tip are more heterogeneous than those from the rest of the spicule, revealing that newly-formed regions of the spicule are structurally diverse. Similar results on a different 48-h spicule are presented in supporting information (SI) Fig. S1. For comparison, spectra from synthetic calcite and synthetic ACC are also shown in Fig. 1 D and E, respectively. In calcite, the 2 main peaks (denoted 1 and 3, which are the Ca L2 and L3 peaks, respectively) are split, giving rise to 2 minor peaks [denoted as 2 and 4, the crystal field peaks (25)]. In synthetic ACC, peaks 2 and 4 are less intense and shifted closer to the main peaks, where they appear as shoulders.

Fig. 1.

Fig. 1.

Ca L-edge XANES spectra and an X-PEEM micrograph of a 48-h spicule. (A and B) Ca L-edge XANES spectra extracted from: near the tip (left yellow line in C) (A) and middle part of the spicule (right yellow line in C) (B). (C) X-PEEM micrograph of part of a fresh 48-h spicule. (D) Ca L-edge XANES spectra of synthetic calcite; the L2 peak is split into peaks 1 and 2, and the L3 is split into peaks 3 and 4. The main peaks are 1 and 3, and the crystal field peaks are 2 and 4. (E) Ca L-edge XANES spectra of synthetic ACC. These and all spectra hereafter were extracted from adjacent pixels along a line. The bold spectra at the top of A, B, D, and E are the averages of all spectra below. Blue in A and B highlights a spectrum similar to calcite. Green highlights a spectrum with intense peak 2 and small peak 4. Red, present in A but not in B, highlights a spectrum similar to synthetic hydrated ACC. Each spectrum in A, B, D, and E was extracted from a 200-nm pixel.

Analysis of numerous single pixel spectra revealed 3 independent calcium absorption line-shapes that correspond to 3 different mineral phases (Fig. 1). The first spectrum type (red in Fig. 1) is similar to synthetic ACC, where both peaks 2 and 4 are weak. This type 1 spectrum is found only near the tips of 48-h spicules. The second type (green in Fig. 1) has a pronounced peak 2 but a weak peak 4. The third type of spectrum (blue in Fig. 1) is similar to that of calcite, with both peaks 2 and 4 being intense. This third type of spectrum resembles calcite and is found at locations everywhere on the spicule surface, but with greater frequency and intensity in pixels at the center of the triradiate structure, where the initial rhombohedral calcite crystal was observed (10) (Fig. S1). All other spectra fall between these 3 types. The type 2 spectrum is distinct from those of synthetic ACC, and calcite spectra and cannot be the result of a linear combination of types 1 and 3, because any linear combination of these leads to a parallel change in both peaks 2 and 4 (Fig. S2). The suppression of the L3 crystal field peak 4 in type 2 indicates that this spectrum represents a disordered form of calcium carbonate, although no similar spectrum has been recorded from standard materials. In both 48- and 72-h spicules, the most abundant phase is type 2.

Over time, the amorphous material present in fresh spicules is known to crystallize (3), and we sought to determine whether the ACC phases found here in fresh spicules, 48 h after fertilization, are transient phases. We repeated the measurements on fresh spicules grown for a longer period (72 h after fertilization) and then at precisely the same location on the same spicules after 10 months in storage (Fig. 2 and Fig. S3). The fresh 72-h spicules consist only of type 2 ACC, crystalline biogenic calcite, or combinations thereof, with no type 1 ACC phase (Fig. 2E), whereas the spectra obtained from the 10-month-old spicule are uniformly similar to crystalline biogenic calcite (Fig. 2F). This result implies that the type 2 phase is less stable than calcite. We infer that the crystalline phase grew at the expense of amorphous domains and/or the amorphous phase, having higher solubility, was preferentially removed. The 10-month-old spicule has an etched appearance when imaged in the scanning electron microscope (SEM) (Fig. 2D), resembling an aggregate of spheres of 40–100 nm in diameter. The etched spicule exhibits topographic features that are coarser than nonetched spicules. The homogeneity of the spectra extracted from this sample thus demonstrates that specimen topography cannot produce the spectral diversity shown in Fig. 2E.

Fig. 2.

Fig. 2.

X-PEEM micrographs and Ca L-edge XANES spectra of a fresh 72-h spicule, and the same spicule after 10 months. (A–C) X-PEEM micrographs of the 72-h spicule. (A) Fresh spicule: the dark region immediately below the spicule is its shadow. (B) The same spicule measured 10 months later. The yellow lines indicate the pixels from which the spectra in E and F were extracted. (C) High magnification of the area in A: colored pixels are those from which the corresponding colored spectra in E and F were extracted. (D) SEM micrograph of the same spicule taken after 10 months. (E and F) Ca L-edge spectra extracted from the lines in A and B, respectively. The blue curve in E corresponds to the type 2 ACC phase, which clearly became calcitic in F. Scale bar in A also applies to B, and scale bar in C also applies to D. Pixel size is 100 × 100 nm2. Highlighted in black is the average spectrum.

Selected spectra from the 48- and 72-h spicules in Figs. 1 and 2 and from the reference standards were peak-fitted. The peak-fitting results, presented in Fig. S4, highlight the spectroscopic differences among type 1, 2, and 3 mineral phases, and the similarity of types 1 and 3 with synthetic ACC and calcite.

We further characterized the occurrence of these distinct carbonate phases in 72-h spicules by repeating the measurement at higher magnification (40-nm pixels vs. 200- and 100-nm pixels in Figs. 1 and 2) to gain more insight into the dimensions of the individual domains. Indeed, isolated spectra of amorphous domains can be detected in the middle of crystalline domains. We observed abrupt transitions between calcite and ACC in immediately adjacent pixels, as well as more gradual transitions (Fig. 3). The smallest domains observed are 1 pixel wide, and others are 3 pixels wide (40–120 nm).

Fig. 3.

Fig. 3.

Ca L-edge XANES spectra from a fresh 72-h spicule. The spectra are extracted from individual pixels along a straight line. Each pixel represents 40 × 40 nm2. The 2 series of spectra show different patterns of mineral phase distribution in adjacent pixels. (A) We observe large blocks of 5–10 adjacent spectra of type 3 calcite interspersed with smaller series of spectra of type 2 ACC. (B) The transitions between type 2 and 3 spectra are abrupt between the bottom 2 spectra and gradual for the others. Highlighted in black are the average spectra. Blue highlights 1 of the type 3 calcite spectra, green highlights type 2 ACC.

Because the phases may be distinguished by the crystal field splitting at the L2 and L3 edges, we define an empirical splitting ratio (SR) to be the ratio of the maximum intensity of the crystal field peak (2 or 4) and the intensity of the minimum separating this peak from the corresponding main peak (1 or 3) (Fig. 4 and Fig. S5). For synthetic ACC, both peaks are poorly split and thus both SR values are less than unity. For biogenic and synthetic calcite, both peaks are well resolved, and the SRs are larger than 2 and 1.3, for L2 and L3, respectively. The single-pixel SR values for spicules are varied, and structural trends in the data from the biogenic samples are illustrated by plotting L3 SR vs. L2 SR (Fig. 4B). Each of the 3 phase types identified above falls in a different quadrant of this plot (Fig. 4C): ACC type 1 with both SRs < 1; ACC type 2, with L2 SR > 1 and L3 SR < 1; and calcitic with both SRs > 1.

Fig. 4.

Fig. 4.

Peak splitting anaylsis of Ca L-edge XANES spectra from 48-h and 72-h spicules, as well as those from ACC and calcite. (A) Ca L-edge XANES spectra extracted from single pixels of synthetic ACC (bottom red curve) and calcite (top blue curve) and 3 spectra from a 48-h spicule. The 3 spicule spectra are representative of the 3 mineral phases identified in Fig. 1: red is type 1, green is type 2, and blue is type 3. (B) A plot of SR(L3) vs. SR(L2) (see Methods and Fig. S5). The spicule samples indicated by triangles (48-h spicule tip, light blue; middle, purple; 72-h spicule fresh, green; 10 month olds, blue), and the adult sea urchin spine (squares, brown), span 3 quadrants, and synthetic ACC (red circles) is located in the bottom left quadrant, where L2, L3 SR < 1. Calcite (blue diamonds) is located in the top right quadrant where L2, L3 SR > 1. (C) Spicule ratios are shown separately with the relevant quadrants shaded in gray. Color code is as in B.

The spectra are often a mixture of phases. This mixture occurs when a pixel includes, for instance, part of a type 2 and part of a type 3 particle. Thus the corresponding spectrum and SRs are intermediate between types 2 and 3. The spicule SRs tend to become more calcite-like with increasing distance from spicule tip to the middle, and with growth time from 48 to 72 h after fertilization (Fig. 4). All SR values obtained from the 72-h spicule after 10 months fall in the calcite-like top-right quadrant, as expected. However, the values are much smaller than those of synthetic calcite, indicating greater structural disorder. This result might be attributed to the presence in the spicules of ≈5 mole% magnesium and the occluded matrix proteins. To support this hypothesis, we measured the SRs of adult sea urchin spines, considered to be composed of crystalline calcite exclusively. The spine's SRs (Fig. 4B), with Mg concentration similar to the spicules, also do not reach the values of synthetic calcite, although they are shifted toward calcite relative to the spicule. Consistently, spectra of biogenic minerals containing higher amounts of Mg have even lower SRs, but with both SRs > 1 (Yurong Ma, personal communication).

The present data thus show that spicule development involves 2 amorphous precursor phases. The freshly deposited mineral is similar to hydrated synthetic ACC. This phase is short-lived and can be detected only in areas of fast growth (the tip), rather than where slower radial thickening occurs (17). This type 1 ACC rapidly transforms into a second phase that appears amorphous from the XANES data. This type 2 ACC transforms more slowly into biogenic calcite.

Prior studies on spicules extracted at an advanced developmental stage have shown that the amorphous phase in this system is mostly anhydrous (19). However, it has been suggested that water molecules may be present as part of an initial hydrated ACC phase that subsequently transforms into the anhydrous phase (7, 18). Our data are in good agreement with this mechanism, and we suggest that they represent direct observation of the dehydration step for freshly deposited ACC. Because this type 2 phase is the most abundant in fresh spicules it is likely to be the same anhydrous ACC phase observed with bulk methods (18, 19) and now confirmed to be formed from an earlier transient phase. We note that the probing depth in these experiments is only 3 nm. Thus only the fresh material that is deposited on the spicule surface upon thickening is sampled.

A recent EXAFS study showed that the transient ACC phase at this stage has a short-range order that resembles the mature crystalline phase (18), which might be the origin of the prominent peak 2. The type 2 mineral is therefore intermediate between fully disordered, probably hydrated ACC and crystalline calcite with respect to both spicule development stage and crystallinity.

The present data show the presence of juxtaposed crystalline and amorphous phases in the surface layer of growing spicules, raising the fundamental question of how these heterogeneous mineral domains transform into the single crystal found in the mature spicule. Insight is obtained from the higher-resolution analysis of the size of single domains. In all specimens, we observe discontinuity in mineral phases in immediately adjacent pixels, suggesting that the precursor mineral phase is present in small units. Analysis of 72-h specimens showed that homogeneous domains of type 3 calcite as large as 1 μm are present, interspersed with smaller domains of type 2 ACC. The smallest ACC domains (40–120 nm) observed here with X-PEEM, are consistent in size with the previous observation of 50–100 nm spherules (Fig. 2) (3, 17) and the coherence length of the mature spicule observed in X-ray diffraction (26).

From these observations, we propose that a transformation from type 2 amorphous to type 3 crystalline phase propagates through the spicule via secondary nucleation, in which the crystallization of 1 amorphous unit stimulates the transformation of the domains in contact with it (27, 28). The overall crystallographic orientation is determined by the initial central crystal. Because type 2 ACC is anhydrous, no volume change occurs during the transformation to type 3 calcite, and spicule morphology is unaffected.

The propagation pathway through the spicule is inferred to be complex and tortuous, implying that the rate of transformation depends on the size and interface structure of the amorphous domains. These are probably determined by the presence, location, and concentration of the organic additives. Within larger calcitic regions of 72-h spicules, individual amorphous domains of 40 × 40 nm2 were occasionally identified, indicating that the propagation pathway may leave small domains untransformed. Mapping the distribution of single-phase domains, as demonstrated here, will enable the testing of various hypotheses that may account for the patterns of crystalline-phase propagation that are akin to fractal network percolation (29, 30).

The transformation mechanism presented here may well represent a common strategy in biomineralization, bearing in mind the widespread use of precursor ACC phases in biology and the many cases in which 30- to 50-nm spherulitic structures have been observed in biogenic calcium carbonate minerals from diverse phyla (3135).

Methods

More detailed descriptions of the experimental procedures are provided in SI Text.

Sea Urchin Larval Culture.

Strongylocentrotus purpuratus embryos were grown in artificial sea water containing Gentamycin (20 mg·L−1) at 15 °C, following established methods (36, 37).

Extraction of S. purpuratus Spicules.

Embryos were disrupted in a Polytron homogenizer. The spicules were collected by centrifugation and extracted with SDS and 3.5% NaOCl. The spicules were washed with CaCO3-saturated solution and rinsed with ethanol and acetone, frozen in liquid nitrogen and kept at −80 °C for up to 2 days until the measurements.

Synthetic calcite crystals were grown in Nunc multiwell dishes by diffusion of ammonium carbonate vapor into 10 mM calcium chloride (Merck; A grade) solutions (38). Synthetic ACC was synthesized following Koga et al. (39) by mixing solutions of calcium chloride (0.1 M) with sodium carbonate (0.1 M) and sodium hydroxide (1 M).

X-PEEM Sample Preparation.

Forty-eight- or 72-h spicule samples were resuspended in ethanol. A drop of the suspension was deposited on a 10 mm × 10 mm silicon chip and air-dried. Synthetic ACC and calcite powders were pressed into indium foil. All samples were sputter-coated with 1 nm Pt (40).

X-PEEM Experiments.

We used the spectromicroscope for photoelectron imaging of nanostructures with X-rays (SPHINX), which is an X-PEEM (Elmitec), installed on the VLS-PGM beamline at the Synchrotron Radiation Center. The instrument and its performances are described in detail in ref. 20. Briefly, the sample was mounted vertically and illuminated from a side (16° grazing incidence angle) with monochromatic soft X-rays. The photoelectrons emitted by the sample were accelerated toward an electron optics column and onto a phosphor screen. The chamber was held at ultrahigh vacuum (10−10 Torr). The real-time, sample surface image was acquired by a slow-scanning cooled CCD camera. Movie stacks of 170 images were acquired while scanning the photon energy across the Ca L-absorption edge, so that each pixel in the movie contained a complete XANES spectrum. Images were acquired with fields of view ranging from 20 to 100 μm and corresponding pixel sizes of 40–200 nm. Because the samples are not flat, the spatial resolution may be lower than the pixel size.

Data Analysis.

Data processing was done by using routines we designed in National Institutes of Health Image 1.62 and Igor Pro 6.0.3 (WaveMetrics) for Macintosh. From each stack of 170 images Ca L-edge spectra were extracted from single pixels along a straight line on the spicule or standards and divided by a pre-edge linear background. All spectra presented here were normalized to a linear pre-edge fit. The position of peak 1 was set to be 352.6 eV for all spectra, following Benzerara et al. (41). The intensity of the pre-edge was then set to 0 and that of peak 1 to 1. All spectra were smoothed over 5 points and displaced vertically in all figures.

Peak Fitting.

Selected spectra were peak-fitted by using routines we designed in Igor Pro 6.0.3 (WaveMetrics). The most representative peak-fitted spectra are presented in Fig. S4.

Calculations of SRs.

For each peak we divided the intensity value, after approximated-baseline subtraction, of the split peak (spL2, spL3) by the intensity value of the minimum between these peaks and the main peak (dipL2 and dipL3, respectively). Such that:

graphic file with name zpq04508-5482-m01.jpg

See Fig. S5.

Scanning Electron Microscopy.

Samples were coated with 6 nm Cr and viewed with a SEM (Philips; XL30 FESEM FEG), operated at 10 kV.

Supplementary Material

Supporting Information

Acknowledgments.

We thank Prof. Peter Rez for fruitful discussions. This work was supported by National Science Foundation Award CHE&DMR-0613972 (to P.G.), Department of Energy Award DE-FG02-07ER15899 (to P.G. and S.W.), and Israel Ministry of Science Project 777. The experiments were performed at the University of Wisconsin–Synchrotron Radiation Center, which was supported by National Science Foundation Award DMR-0537588. F.H.W. is supported by the National Institutes of Health and National Science Foundation. L.A. is the incumbent of the Dorothy and Patrick Gorman Professorial Chair of Biological Ultrastructure, and S.W. is the incumbent of the Dr. Walter and Dr. Trude Burchardt Professorial Chair of Structural Biology. I.S. is the incumbent of the Pontecorvo Professorial Chair of Cancer Research.

Footnotes

The authors declare no conflict of interest.

This article is a PNAS Direct Submission.

This article contains supporting information online at www.pnas.org/cgi/content/full/0806604105/DCSupplemental.

References

  • 1.Weiner S, Sagi I, Addadi L. Choosing the crystallization path less traveled. Science. 2005;309:1027–1028. doi: 10.1126/science.1114920. [DOI] [PubMed] [Google Scholar]
  • 2.Towe KM, Lowenstam HA. Ultrastructure and development of iron mineralization in the radular teeth of Cryptochiton stelleri (Mollusca) J Ultrastruct Res. 1967;17:1–13. doi: 10.1016/s0022-5320(67)80015-7. [DOI] [PubMed] [Google Scholar]
  • 3.Beniash E, Aizenberg J, Addadi L, Weiner S. Amorphous calcium carbonate transforms into calcite during sea-urchin larval spicule growth. Proc R Soc London Ser B. 1997;264:461–465. [Google Scholar]
  • 4.Weiss IM, Tuross N, Addadi L, Weiner S. Mollusk larval shell formation: Amorphous calcium carbonate is a precursor for aragonite. J Exp Zool. 2002;293:478–491. doi: 10.1002/jez.90004. [DOI] [PubMed] [Google Scholar]
  • 5.Dillaman R, Hequembourg S, Gay M. Early pattern of calcification in the dorsal carapace of the blue crab, Callinectes sapidus. J Morphol. 2005;263:356–374. doi: 10.1002/jmor.10311. [DOI] [PubMed] [Google Scholar]
  • 6.Marxen JC, Becker W, Finke D, Hasse B, Epple M. Early mineralization in Biomphalaria glabrata: Microscopic and structural results. J Molluscan Studies. 2003;69:113–121. [Google Scholar]
  • 7.Politi Y, Klein E, Arad T, Weiner S, Addadi L. Sea urchin spine calcite forms via a transient amorphous calcium carbonate phase. Science. 2004;306:1161–1164. doi: 10.1126/science.1102289. [DOI] [PubMed] [Google Scholar]
  • 8.Meibom A, et al. Distribution of magnesium in coral skeleton. Geophys Res Let. 2004;31:L23306–L23310. [Google Scholar]
  • 9.Mahamid J, Sharir A, Addadi L, Weiner S. Amorphous calcium phosphate is a major component of the forming fin bones of zebrafish: Indications for an amorphous precursor phase. Proc Natl Acad Sci USA. 2008;105:12748–12753. doi: 10.1073/pnas.0803354105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Okazaki K, Dillaman RM, Wilbur KM. Crystalline axes of the spine and test of the sea urchin Strongylocentrotus purpuratus: Determination by crystal etching and decoration. Biol Bull. 1981;161:402–415. [Google Scholar]
  • 11.Theel H. On the development of Ehinocyamus pusillus. Nova Acta Res Soc Sci Upsala. 1892;15:1–57. [Google Scholar]
  • 12.Raz S, Hamilton PC, Wilt FH, Weiner S, Addadi L. The transient phase of amorphous calcium carbonate in sea urchin larval spicules: The involvement of proteins and magnesium ions in its formation and stabilization. Adv Funct Mat. 2003;13:480–486. [Google Scholar]
  • 13.Gibbins JR, Tilney LG, Porter KR. Microtubules in the formation and development of the mesenchime in Arbacia punctulata. I. Distribution of microtubules. J Cell Biol. 1969;41:201–226. doi: 10.1083/jcb.41.1.201. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Okazaki K. Skeleton formation of sea urchin larvae. IV. Correlation of the shape of spicule and matrix. Embryologia. 1962;7:21–38. [Google Scholar]
  • 15.Wilt F, Killian CE, Hamilton PC, Croker L. The dynamics of secretion during sea urchin embryonic skeleton formation. Exp Cell Res. 2008;314:1744–1752. doi: 10.1016/j.yexcr.2008.01.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Beniash E, Addadi L, Weiner S. Cellular control over spicule formation in sea urchin embryos: A structural approach. J Struct Biol. 1999;125:50–62. doi: 10.1006/jsbi.1998.4081. [DOI] [PubMed] [Google Scholar]
  • 17.Wilt FH, Ettensohn CA. In: Handbook of Biomineralization. Bauerlein E, Behrens P, editors. Vol 2. Germany: Wiley, Weinheim; 2007. pp. 183–210. [Google Scholar]
  • 18.Politi Y, et al. Strucutural characterization of the transient calcium carbonate amorphous precursor phase in sea urchin embryos. Adv Funct Mat. 2006;16:1289–1298. [Google Scholar]
  • 19.Addadi L, Raz S, Weiner S. Taking advantage of disorder: Amorphous calcium carbonate and its roles in biomineralization. Adv Mat. 2003;15:959–970. [Google Scholar]
  • 20.Frazer BH, Girasole M, Wiese LM, Franz T, De Stasio G. Spectromicroscope for the photoelectron imaging of nanostructures with X-rays (SPHINX): Performance in biology, medicine, and geology. Ultramicroscopy. 2004;99:87–94. doi: 10.1016/j.ultramic.2003.10.001. [DOI] [PubMed] [Google Scholar]
  • 21.Gilbert P, Frazer BH, Abrecht M. In: Reviews in Mineralogy and Geochemistry. Banfield JF, Nealson KH, Cervini-Silva J, editors. Vol 59. Washington DC: Mineralogical Soc Am; 2005. pp. 157–185. [Google Scholar]
  • 22.Stohr J. NEXAFS Spectroscopy. Berlin: Spriger; 1992. [Google Scholar]
  • 23.Metzler RA, et al. Polarization-dependent imaging contrast in abalone shells. Phys Rev B. 2008;77 064110. [Google Scholar]
  • 24.Chan CS, et al. Microbial polysaccharides template assembly of nanocrystal fibers. Science. 2004;303:1656–1658. doi: 10.1126/science.1092098. [DOI] [PubMed] [Google Scholar]
  • 25.Himpsel FJ, et al. Fine-structure of the Ca 2p X-ray-absorption edge for bulk compounds, surfaces, and interfaces. Phys Rev B. 1991;43:6899–6907. doi: 10.1103/physrevb.43.6899. [DOI] [PubMed] [Google Scholar]
  • 26.Berman A, et al. Biological control of crystal texture: A widespread strategy for adapting crystal properties to function. Science. 1993;259:776–779. doi: 10.1126/science.259.5096.776. [DOI] [PubMed] [Google Scholar]
  • 27.Xu AW, Dong WF, Antonietti M, Colfen H. Polymorph switching of calcium carbonate crystals by polymer-controlled crystallization. Adv Funct Mat. 2008;18:1307–1313. [Google Scholar]
  • 28.Wang T, Antonietti M, Colfen H. Calcite mesocrystals: “Morphing” crystals by polyelectrolyte. Chem Eur J. 2006;12:5722–5730. doi: 10.1002/chem.200501019. [DOI] [PubMed] [Google Scholar]
  • 29.Stauffer D, Aharony A. Introduction to Percolation Theory. London: Taylor and Francis; 1992. [Google Scholar]
  • 30.Jensen P, et al. Direct observation of the infinte precolation cluster in thin films: Evidence for a double percolation process. Phys Rev B. 1993;47:5008–5012. doi: 10.1103/physrevb.47.5008. [DOI] [PubMed] [Google Scholar]
  • 31.Sethmann I, Putnis A, Grassmann O, Lobmann P. Observation of nano-clustered calcite growth via a transient phase mediated by organic polyanions: A close match for biomineralization. Am Mineral. 2005;90:1213–1217. [Google Scholar]
  • 32.Sethmann I, Worheide G. Structure and composition of calcareous sponge spicules: A review and comparison to structurally related biominerals. Micron. 2008;39:209–228. doi: 10.1016/j.micron.2007.01.006. [DOI] [PubMed] [Google Scholar]
  • 33.Dauphin Y, Guszman N, Denis A, Cuif JP, Ortlieb L. Microstructure, nanostructure, and composition of the shell of Concholepas concholepas (Gastropoda, Muricidae) Aquat Living Resour. 2003;16:95–103. [Google Scholar]
  • 34.Przenioslo R, Stolariski J, Mazur M, Berunelli M. Hierarchically structured scleractinian coral biocrystals. J Struct Biol. 2008;161:74–82. doi: 10.1016/j.jsb.2007.09.020. [DOI] [PubMed] [Google Scholar]
  • 35.Li X, Xu ZH, Wang R. In situ observation of nanograin rotation and deformation in nacre. Nano Lett. 2006;6:2301–2304. doi: 10.1021/nl061775u. [DOI] [PubMed] [Google Scholar]
  • 36.Hinegradner R. In: Methods in Developmental Biology. Wilt F, Wessells N, editors. New York: Crowell; 1967. pp. 139–156. [Google Scholar]
  • 37.Killian CE, Wilt FH. Characterization of the proteins comprising the integral matrix of Strongylocentrotus purpuratus embryonic spicules. J Biol Chem. 1996;271:9150–9159. doi: 10.1074/jbc.271.15.9150. [DOI] [PubMed] [Google Scholar]
  • 38.Falini G, Albeck S, Weiner S, Addadi L. Control of aragonite or calcite polymorphism by mollusk shell macromolecules. Science. 1996;271:67–69. [Google Scholar]
  • 39.Koga N, Nakagoe Y, Tanaka H. Crystallization of amorphous calcium carbonate. Thermochim Acta. 1998;318:239–244. [Google Scholar]
  • 40.De Stasio G, et al. Compensation of charging in X-PEEM: A successful test on mineral inclusions in 4.4-Ga-old zircon. Ultramicroscopy. 2003;98:57–62. doi: 10.1016/S0304-3991(03)00088-3. [DOI] [PubMed] [Google Scholar]
  • 41.Benzerara K, et al. Scanning transmission X-ray microscopy study of microbial calcification. Geobiology. 2004;2:249–259. [Google Scholar]

Associated Data

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

Supplementary Materials

Supporting Information

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

RESOURCES