You are on page 1of 5

Subnanometer atomic force microscopy of

peptide–mineral interactions links clustering and


competition to acceleration and catastrophe
R. W. Friddlea,b,1, M. L. Weavera,c,1, S. R. Qiua, A. Wierzbickid, W. H. Caseyc, and J. J. De Yoreob,2
a
Physical and Life Sciences Directorate, Lawrence Livermore National Laboratory, Livermore, CA 94551; bMolecular Foundry, Lawrence Berkeley National
Laboratory, Berkeley, CA 94720; cDepartments of Chemistry and Geology, University of California, Davis, CA 95616; and dDepartment of Chemistry,
University of South Alabama, Mobile, AL 36688

Edited by Joanna Aizenberg, Harvard University, and accepted by the Editorial Board November 3, 2009 (received for review July 23, 2009)

In vitro observations have revealed major effects on the structure, sharpness, are too stiff to scan at high rates in contact mode
growth, and composition of biomineral phases, including stabili- without displacing adhered molecules or fracturing the brittle
zation of amorphous precursors, acceleration and inhibition of tip. The common solution is to use cantilever probes fabricated
kinetics, and alteration of impurity signatures. However, decipher- from silicon nitride, which are much softer; but the malleability of
ing the mechanistic sources of these effects has been problematic silicon nitride renders these probes dull and limits their resolving
due to a lack of tools to resolve molecular structures on mineral power. To overcome these limitations, we employed probes com-
surfaces during growth. Here we report atomic force microscopy prised of sharpened Si tips on flexible, low-stress Si3 N4 cantile-
investigations using a system designed to maximize resolution vers (Applied Nanostructures). To augment the relatively weak
while minimizing contact force. By imaging the growth of calcium- laser signal reflected from the thin Si3 N4 cantilever beam
oxalate monohydrate under the influence of aspartic-rich peptides (600 nm thick, uncoated), we utilized a 3-mW diode laser instead
at single-molecule resolution, we reveal how the unique interac- of the ∼1-mW lasers usually employed.
tions of polypeptides with mineral surfaces lead to acceleration, We used this system to investigate peptide interactions with
inhibition, and switching of growth between two distinct states. calcium-oxalate monohydrate (COM, CaC2 O4 · H2 O), the domi-
Interaction with the positively charged face of calcium-oxalate nant mineral phase in human kidney stones (8) (see Methods for
monohydrate leads to formation of a peptide film, but the slow a description of materials and imaging conditions). Aside from
adsorption kinetics and gradual relaxation to a well-bound state clinical relevance, COM is an excellent system for investigating
result in time-dependent effects. These include a positive feedback peptide interactions, because it presents two strongly contrast-
between peptide adsorption and step inhibition described by a

PHYSICS
ing faces. One—ð−101Þ—is calcium-rich and electrostatically
mathematical catastrophe that results in growth hysteresis, charac- positive, whereas another—(010)—is oxalate-rich and strongly
terized by rapid switching from fast to near-zero growth rates for negative (9).
very small reductions in supersaturation. Interactions with the ne- The importance of both specific and nonspecific interactions
gatively charged face result in formation of peptide clusters that between COM surfaces and polypeptides or polyionic molecules
impede step advancement. The result is a competition between ac- has been demonstrated previously. Polypeptides altered step
celerated solute attachment and inhibition due to blocking of the kinetics in a face- and step-specific manner, and their pre-
steps by the clusters. The findings have implications for control of sence in solution reduced adhesion forces between carboxylate-
pathological mineralization and suggest artificial strategies for di-
functionalized AFM tips and COM surfaces (10, 11). In vitro stu-
recting crystallization.
dies of clinical relevance found that COM growth was strongly
inhibited by acidic urinary molecules, particularly citrate and
biomineralization ∣ calcium oxalate monohydrate ∣ crystal growth ∣
osteopontin (OPN) (12, 13). [In this regard, COM is reminiscent
peptide adsorption ∣ molecular resolution
of calcium carbonate in that proteins commonly found in associa-
tion with carbonate biominerals are highly acidic and rich in
T he interaction of proteins with inorganic constituents is a de-
fining feature of biomineral systems. They direct crystalliza-
tion of precursors (1), inhibit pathological mineralization (2, 3),
carboxylic side chains (14).] However, the consequences of intro-
ducing polyanions to COM differ significantly between small
molecules and polypeptides. For example, the effects of citrate
and initiate mineral resorption (4). In vitro observations reveal on COM growth were well explained by classical growth models
major effects on structure, growth, and composition including of step pinning (15), and electrostatic models of citrate binding
stabilization of amorphous phases (5), acceleration (6) and inhi-
provided a clear rationale for the face- and step-specific effects
bition of kinetics (2), and alteration of impurity signatures (7).
(12): Due to repulsive and attractive interactions of the carboxylic
Defining the microscopic mechanism behind these effects has
groups with oxalate and calcium ions, respectively, citrate had
been difficult due to a lack of tools to resolve either proteins
little effect on the negative face while binding strongly to the
or the molecular structure of mineral surfaces during growth.
positive face. In contrast, OPN, which contains peptide segments
In principle, atomic force microscopy (AFM) has the resolving
power to investigate interactions between proteins and crystal
surfaces at the molecular scale. However, in practice this scenario Author contributions: S.R.Q., W.H.C., and J.J.D.Y. designed research; R.W.F., M.L.W., and
presents the difficult challenge of imaging soft biomolecules on a A.W. performed research; R.W.F., M.L.W., S.R.Q., and J.J.D.Y. analyzed data; and R.W.F.,
M.L.W., and J.J.D.Y. wrote the paper.
hard, dynamic surface and requires a delicate balance between
lateral resolution, temporal resolution, and sample disruption. The authors declare no conflict of interest.

To capture the morphological evolution of atomic steps and their This article is a PNAS Direct Submission. J.A. is a guest editor invited by the Editorial Board.
interaction with macromolecules, one must scan rapidly. Thus Freely available online through the PNAS open access option.
cantilever stiffness must be low to minimize the force applied as 1
R.W.F. and M.L.W. contributed equally to this work.
it passes over topographic features such as steps and weakly ad- 2
To whom correspondence should be addressed. E-mail: jjdeyoreo@lbl.gov.
sorbed macromolecules. At the same time, high-resolution imag- This article contains supporting information online at www.pnas.org/cgi/content/full/
ing requires sharp probes. Silicon probes, which offer excellent 0908205107/DCSupplemental.

www.pnas.org/cgi/doi/10.1073/pnas.0908205107 PNAS ∣ January 5, 2010 ∣ vol. 107 ∣ no. 1 ∣ 11–15


consisting of more than 50% aspartic acid (Asp) residues, exhib-
ited face-specific effects that appear to contradict this simple pic-
ture. On the positive face, they formed surface-aggregates having
little effect on growth but strongly inhibited growth of the nega-
tive face (12). Clearly, the simple binding rationale applicable to
small molecules such as citrate is not conserved in their larger,
polypeptide counterparts bearing similar chemical functionality.
The underlying source of these differences is undoubtedly the
unique adsorption behavior of polypeptides at surfaces (16). Poly-
peptide binding to an oppositely charged surface requires cross-
ing a series of kinetic barriers, such as displacement of shielding
counterions, configurational relaxation of the polymer, and coor-
dination with oppositely charged sites on the crystal to reach
the final collapsed state in which the polypeptide residues form
specific bonds to lattice sites (16). In contrast, at like-charged
surfaces, polypeptide solutions containing di- and trivalent coun-
terions exhibit fluctuating polarizations that can diminish the
double-layer repulsion between like-charged objects, allowing
the van der Waals attraction to dominate (16, 17). This effect
can lead to polypeptide clustering (dictated by the balance of
enthalpic gain and entropic cost of aggregation) as well as non-
specific binding to the surface.

Results
Atomic Resolution of COM Crystal Surfaces During Growth. To attain
Fig. 1. (A–B) AFM height images of steps on the (010) face showing indivi-
high-resolution control images of the surface, we performed in dual kinks and single molecules at the step edge. Also given are schematics
situ AFM on the ð−101Þ and (010) faces of COM at supersatura- showing the relationship between AFM images and underlying COM struc-
tion ∼0.8, in the absence of peptide. (Supersaturation is defined ture. Color scheme: Red—O, green—Ca, and gray—C. (C) AFM deflection im-
 
aðCa2þ Þ aðC O2− Þ age of the (010) face showing step, kinks, and electrostatic map of
Δμ 1
by σ ≡ kB T ¼ 2 ln K sp
2 4
, where Δμ represents the differ- ðDDDSÞ6 DDD for configuration of highest binding energy. (D) AFM deflection
image of the ð−101Þ face and electrostatic map for configuration of highest
ence in chemical potential per COM molecule between solution binding energy for ðDDDSÞ6 DDD as predicted by the in vacuo MD simulations.
and crystal, kB is Boltzmann’s constant, T is the absolute tem- Red is most negative; blue is most positive. In an intuitive and simple electro-
perature, the ai ’s are the activities of the respective components, static picture of binding, the negative (yellow) oxygen atoms of the oxalate
ions that protrude from the (010) terrace are expected to repel the negative
and K sp is the solubility constant. See Methods for details.) As
(yellow-to-red) carboxyl side groups along the peptide. In contrast, on the
Fig. 1 shows, even at room temperature, we obtained true single- ð−101Þ face, the MD simulations predict that the positive (blue) calcium-rich
molecule resolution of faces, atomic steps, kinks, and even what terrace should provide a favorable binding environment for the carboxyl-rich
appear to be single attachment events during growth. In Figs. 1A peptide. Unit cell parameters (9) are a ¼ 9.976 Å, b ¼ 14.588 Å, c ¼ 6.291 Å,
and B, the AFM images of the steps on the (010) face are com- and β ¼ 107.05°. Electrostatic potentials in C and D were rendered by
pared to the interpretations on the basis of the crystal structure. using MOE 2005.06 (Chemical Computing Group). Molecular structures in
Image enhancement by Fourier filtering revealed submolecular Figs. 1 and 3 were rendered by using WebLab ViewerPro 3.7 (Molecular
Simulations).
details such as protruding ends of vertically aligned oxalate
groups (high “peaks”) and recessed, flat-lying oxalate groups
(low “holes”) (Fig. 1C). (See also Fig. S1.) Comparison of Fig. 1C
with Fig. 1D shows the greater vertical relief of the (010) face as revealed starkly contrasting peptide adsorption behavior. As ex-
compared to the ð−101Þ face. The overlying molecular models pected, peptides bound to the ð−101Þ face so that, even at peptide
are based on classical molecular dynamics (MD) simulations concentrations <10 nM, the face was covered by peptides
as described below. (Fig. 3C). In contrast, the (010) atomic structure was still well
resolved in the presence of peptide, showing that individual pep-
Peptide Adsorption and Step Interactions on COM Surfaces. To inves- tides did not bind to that face (Fig. 2C). Instead, despite the like
tigate the consequences of polypeptide adsorption dynamics in a charge of the peptides and the face, the peptides formed surface-
well-controlled system, we synthesized two 27-residue peptides adsorbed clusters about 10 nm in size (Figs. 2B and C). Cluster
consisting of ðDDDXÞ6 DDD (X ¼ S or G) designed to act as number density increased with peptide concentration (Fig. S2),
surrogates for the Asp-rich region of OPN, which contains 19 but cluster size remained nearly constant.
Asp residues occurring primarily in groups of two to three. Not By following the step dynamics at molecular resolution, we
surprisingly, in vacuo simulations of binding supported the simple found that these contrasting adsorption behaviors led to distinct,
electrostatic picture: The negative oxygen atoms of the oxalate face-specific controls on growth (Fig. 4). On the ð−101Þ face, as
ions that protrude from the (010) terrace were predicted to repel steps moved through the adsorbed peptides, they became se-
the negative carboxyl side groups along the peptide (Fig. 1C),
verely roughened due to a high density of pinning sites. But
whereas the positive calcium-rich ð−101Þ terrace provided a
on the (010) face, the steps still consisted of straight segments
favorable binding environment (Fig. 1D). Peptide binding to
the positive face was predicted to be nearly three times greater with visible kinks. When a step encountered a peptide cluster,
than to the negative face (see Methods). Thus, one might naively it formed a hinge point where it was temporarily stopped. Even-
expect the impact of these peptides to mimic that of citrate. How- tually, the steps overcame these obstacles and recovered to their
ever, our in situ observations presented a strikingly different straight morphology, leaving the peptide clusters unperturbed by
picture. the passing step. This process occurred repeatedly without any
As Figs. 2B and 3B show, at low resolution, the morphologies significant change to cluster location or shape (movies of these
of the two faces were similar. But high-resolution imaging growth processes are provided in SI Text).

12 ∣ www.pnas.org/cgi/doi/10.1073/pnas.0908205107 Friddle et al.


Fig. 2. The (010) face. (A) A typical dislocation hillock under pure growth
conditions. (B) Morphology during growth in 5 nM ðDDDSÞ6 DDD, at low ionic Fig. 3. The ð−101Þ face. (A) Steps originating at a dislocation hillock under
strength (I ¼ 7 × 10−4 M), showing peptide clusters and pinned steps. pure growth conditions. (B) Morphology during growth in 10 nM
(C) High-resolution image of peptide clusters on the terraces, one in front ðDDDSÞ6 DDD solution, at low ionic strength (7 × 10−4 M), showing extensive
of the approaching steps and one within with the upper step. The crystal step pinning. (C) High-resolution image collected in pure solution showing
lattice is well resolved, and the adsorbed peptides are confined to the clusters lattice structure overlaid with a ball-and-stick model. The color scheme is
(I ¼ 0.15 M). (D) Step speed versus supersaturation σ at various concentra- the same as in Fig. 1A and B. (D) High-resolution image showing that, unlike
tions of ðDDDSÞ6 DDD showing dual functionality—inhibition at low supersa- the (010) face, the ð−101Þ face becomes covered by a uniform film of peptide
turation (or high peptide level) and acceleration at high supersaturation (or preventing visualization of the underlying lattice [ðDDDSÞ6 DDD, I ¼ 0.15 M].

PHYSICS
low peptide level). Standard errors in step speed measurements range from (E) Step speed versus supersaturation for various concentrations of
10% at the lowest values of supersaturation to <0.5% at the highest values. ðDDDSÞ6 DDD. Symbols are experimental values taken after introducing
In all cases, the standard errors are smaller than the symbols. Solid curves are the indicated peptide concentration into otherwise pure solution. Red
fits to the model combining step pinning by peptide clusters with reduction squares and blue circles were collected while reducing and increasing super-
in the activation barrier by an amount proportional to the peptide concen- saturation, respectively. Standard errors are as described in Fig. 2. Solid curves
tration (see Supplementary Information for details.) (Figs. S1 and S2 show give the theoretical dependence that takes into account the slow adsorption
dependencies of cluster density and degree of inhibition/acceleration on kinetics of the peptides for decreasing (downward arrow) and increasing
peptide concentration.) (upward arrow) supersaturation. (See Supplementary Information for de-
tails.) The model is in excellent agreement with the data and predicts two
distinct states of growth with a discontinuous drop in step speed to zero
Crystal Growth Kinetics in the Presence of Peptide. Quantification as supersaturation is reduced and a slow steady rise as supersaturation is
of the atomic step speed v revealed unreported consequences increased.
of these distinct adsorption behaviors on growth kinetics as well
as insights into previously unexplained effects. On the ð−101Þ
face, the steps exhibited two distinct states: Upon decreasing Under supersaturated conditions, step growth competes with this
supersaturation σ from high values, v suddenly jumped from that slow peptide binding: When an approaching step arrives at a pep-
of the pure system into a “dead zone”—i.e., a region of finite tide binding site, if the peptide is well-bound, it can block solute
supersaturation in which no growth occurs (Fig. 3E, red, yellow, access to kinks; but if it is weakly adhered, as the peptide bonds
and green curves). But upon increasing σ from near-equilibrium to the crystal fluctuate, step propagation can proceed past the
values, v followed a smooth trajectory with no abrupt changes site. Thus, whether a peptide impedes step motion depends on
(Fig. 3E, blue curve). In contrast, on the (010) face, dual func- the characteristic time scale τ available to relax into the well-
tionality was exhibited: At low supersaturation or high peptide bound state.
concentration, v was reduced, whereas at high σ or low peptide The time window for peptide binding is the terrace exposure
concentration, steps accelerated (Figs. 2D and Fig. S3). This dual time T, which is given by L∕v, where L is the distance between
effect is similar to what was observed for calcite, both with these steps. Analysis of the step speed data on the ð−101Þ face shows
same peptides as well as other Asp-rich peptides (6). Hence this that the characteristic time τ for ðDDDSÞ6 DDD adsorption is
behavior is apparently not an anomalous phenomenon. ∼40 s and, at the higher values of σ used in these experiments,
T ≪ τ (19). However, because L and v decrease and increase with
Discussion supersaturation, respectively, upon reduction of σ, T rapidly in-
Asymmetric Growth Kinetics from Slow Peptide Adsorption Kinetics. creases. When sufficiently low σ is reached, the coverage of well-
We propose that the phenomenon of bistable growth observed on bound peptides is then large enough to appreciably slow the steps.
the ð−101Þ face is a natural consequence of polypeptide adsorp- As they slow, T increases further, resulting in yet higher peptide
tion kinetics at faces to which they strongly bind through specific coverage, which again reduces v, raising the peptide coverage yet
interactions. The series of barriers that slow the adsorption of again, and so on. Because of this positive feedback, step speed
polypeptides (16) results in their passage through multiple con- exhibits the characteristic feature of a mathematical “catas-
figurations before reaching a lowest-energy collapsed state (18). trophe” (20); i.e., there is a point at which v spontaneously jumps

Friddle et al. PNAS ∣ January 5, 2010 ∣ vol. 107 ∣ no. 1 ∣ 13


in Fig. 3E give the predictions of this model for a two-state
adsorption mechanism. (See SI Text for details of the model.)

Antagonistic Effects of Enhanced Desolvation and Step Blocking. We


propose that the dual modulation of growth on the (010) face
derives from the unique behavior of polyelectrolytes adhering
to like-charged surfaces as discussed above (16, 17). The clusters,
which result from nonspecific adhesion mediated by van der
Waals attraction, are weakly bound and present only a transient
barrier to solute addition to steps. Their effects should become
significant only at high coverage where they impede step motion.
On the other hand, simulations predict that Asp residues on pep-
tides in solution can assist cation desolvation at mineral surfaces
(21), which is generally recognized to be the rate-limiting step in
mineral growth (22, 23). This phenomenon should lead to step
acceleration. Thus we suggest that Asp-rich peptides near the
(010) face impose two opposing controls on step advancement:
Peptides in solution near the step lower the barrier to solute
incorporation into the step, whereas adsorbed clusters cause
transient blocking of step advancement. At low peptide concen-
trations, when cluster density is low, enhancement by peptides
in solution near the step dominates, whereas at high peptide con-
centration, when the cluster density is high, step blocking by the
clusters overcomes the enhancement. The curves in Fig. 2D were
calculated by including a decrease in the desolvation barrier of a
magnitude proportional to the concentration of peptides in solu-
tion along with pinning by peptide clusters according to the classic
step-pinning model (13, 15). (See SI Text for details.)
Conclusion
The results from this simple system demonstrate that the high-
resolution in situ imaging obtainable with hybrid AFM probes
can provide unique insights into the consequences of interactions
characteristic of macromolecules at inorganic surfaces. More-
Fig. 4. (A–F) Series of sequential images showing interaction of step on (010)
over, those consequences create a versatile palette of controls
face with peptide clusters in a solution containing 10 nM DDDS (I ¼ 0.15 M). over mineral growth: Depending on the strength of the interac-
An individual step approaches two adjacent clusters (A and B), is transiently tion with the faces and steps, the adsorption dynamics and
pinned (C and D), and then recovers (E and F). Frame interval, 4.2 seconds. the tendency towards cluster formation, macromolecules act as
(G–L) Time sequence showing inhibited growth of step on the ð−101Þ face "switches, throttles, or brakes" on crystal formation.
under I ¼ 0.15 M. A single large peptide aggregate near the center of the
image provides a point of reference. The continuous film of peptides on this Methods
face transiently pins subsegments of the oncoming step at many sites along Molecular Modeling. Energy minimizations were performed on the peptide/
the step front. Frame interval, 8.4 seconds. Note the sudden extension of the COM system by using the Compass force field as implemented in MS Model-
step towards the reference cluster in H, followed by inhibition of the same ing version 4.0 software (Accelrys) with the dielectric constant equal to 20 to
step region in I. This behavior of slight acceleration as steps approach clusters attenuate the Coulomb force acting on the peptide at the COM surface.
was often observed on both faces. During minimizations, the COM lattice was held fixed. Binding energies were
obtained by subtracting the optimized energy of the adsorbate–surface sys-
tem from the energies of the surface and adsorbate when separated beyond
from a finite value to zero and does not exhibit intermediate the interaction distance. Flat and stepped surfaces of COM were built by
values. In contrast, when the crystal is first exposed to peptides using CERIUS version 4.2 surface builder (Accelrys) from the unit cell derived
at a value of σ below the jump, step speed is near zero (T ≫ τ) from the crystallography data published by Deganello and Piro (9). Because
and peptide adsorption proceeds unperturbed until equilibrium the reported unit cell by Deganello and Piro did not include hydrogen atoms,
coverage of well-bound peptides is reached. Even as σ increases, hydrogen atom positions were determined computationally by using a com-
mercial density functional theory software package (CASTEP; Accelrys) with
these peptides present a large barrier to desorption that steps
optimization at the Perdew-Burke-Ernzerhof exchange-correlation func-
cannot overcome. As a consequence, the dynamic relation be- tional of the generalized gradient approximation level of theory with 3D per-
tween time scales is lost and v follows a trajectory well-predicted iodic boundary conditions applied to the COM unit cell; the geometric
by the classic pinning model (13, 15) in the absence of time depen- dimensions of the unit cell were fixed at its crystallographically determined
dence (Fig. 3E). values, and all atoms were frozen except the hydrogens, which were free
Thus the slow, multistate binding of peptides to the ð−101Þ to move.
face breaks the symmetry in the step kinetics: Upon decrease
of σ from high levels, weakly bound peptides are displaced by AFM Imaging. All in situ images were collected at 25 °C in contact mode
rapidly advancing steps, but upon increase from equilibrium, (Digital Instruments E scanner; Nanoscope IIIA) on surfaces of COM crystals
anchored inside the enclosed fluid cell. The AFM head was equipped with
the peptides present nearly permanent adsorbates, the symmetry
a 3-mW laser diode, and the probes consisted of hybrid silicon tips on silicon
is broken, and the step exhibits two distinct states—uninhibited nitride cantilevers (HYDRA rectangular lever, k ¼ 35 pN∕nm, and tip radius
growth or arrested growth. Because the surface peptide coverage <8 nm; Applied Nanostructures, Inc, www.appnano.com). Solution flow
will increase with solution peptide concentration, the supersa- was temporarily interrupted during high-resolution image collection. Details
turation at which the jump between states occurs should also of the temperature control and flow systems, image collection, and data
increase, as is observed experimentally. The red and blue curves analysis are provided elsewhere (24, 25).

14 ∣ www.pnas.org/cgi/doi/10.1073/pnas.0908205107 Friddle et al.


Peptide Synthesis. Peptides were synthesized according to standard proce- and oxalate activities and the respective solution concentrations can be
dures (26) by sequential addition of Fmoc amino acids. After synthesis, the found in ref. 13.
peptides were purified and molecular weights verified by mass spectrometry Ionic activities were estimated through speciation calculations in which
as described previously (27). the Davies equation was used to approximate activity coefficient corrections.
Equilibrium activities were taken to be those at which the speed of atomic
Materials
steps on the COM crystal surface was measured to be zero. In a previous
The COM crystals used for these experiments were grown in vitro with
a gel method (28). Aqueous solutions of calcium oxalate with calcium-to- study, for calcium-oxalate solutions at ionic strength 0.05 M, we determined
oxalate ratio ¼ 1 were prepared by using reagent-grade K2 C2 O4 and CaCl2 · the value of K sp to be 1.56 × 10−9 M2 , which was close to the reported value
2H2 O dissolved in distilled deionized (18 MΩ >) water. The ionic strength was of 1.66 × 10−9 M2 (29).
fixed by using reagent-grade KCl.
Inductively coupled plasma–atomic emission spectroscopy was used to ACKNOWLEDGMENTS. The authors are grateful to Dr. John R. Hoyer of the
confirm the purity of all reactants. Solution pH was adjusted to 7.0  0.05 University of Delaware, who performed the peptide synthesis and character-
before each experiment by using a calibrated glass electrode and potas- ization, to Dr. E. Alan Salter for his assistance in molecular modeling and
sium hydroxide. The calcium and oxalate concentrations were varied from preparation of figures, and to Prof. Pupa Gilbert for valuable discussion.
0.1 to 0.35 mM, respectively, leading to a range of supersaturation Development of subnanometer AFM was supported by Office of Science,
a a

ðCa2þ Þ ðC2 O2− Þ Office of Basic Energy Sciences of the US Department of Energy under
σ ¼ 12 ln K sp
4
between 0 and 1.2. Here we ignore the activity of water,
Contract DE-AC52-07NA27344. Theoretical analysis was supported by Office
which is assumed to be close to unity and unchanged during the experiments. of Science, Office of Basic Energy Sciences of the US Department of Energy
Similarly, the activity of the pure calcium-oxalate monohydrate solid is unity under Contract DE-AC02-05CH1123. AFM investigations of COM growth
and is independent of peptide level. The relationship between the calcium were supported by Grant DK61673 from the National Institutes of Health.

1. Politi Y, et al. (2008) Transformation mechanism of amorphous calcium carbonate into 15. Cabrera N, Vermilyea DA (1958) Growth of crystals from solution. Growth and Perfec-
calcite in the sea urchin larval spicule. Proc Natl Acad Sci USA, 105:17362–17366. tion of Crystals; Proceedings, eds Doremus RH (Wiley, New York), pp 393–410.
2. Shiraga H, et al. (1992) Inhibition of calcium-oxalate crystal-growth invitro by uropon- 16. Harding JH, et al. (2008) Computational techniques at the organic-inorganic interface
tin—Another member of the aspartic acid-rich protein superfamily. Proc Natl Acad Sci in biomineralization. Chem Rev, 108:4823–4854.
USA, 89:426–430. 17. Pastre D, et al. (2003) Adsorption of DNA to mica mediated by divalent counterions: A
3. Graether SP (2000) Beta-helix structure and ice-binding properties of a hyperactive theoretical and experimental study. Biophys J, 85:2507–2518.
antifreeze protein from an insect. Nature, 406:325–328. 18. Bachmann M, Janke W (2005) Conformational transitions of nongrafted polymers
4. Lange PF, Wartosch L, Jentsch TJ, Fuhrmann JC (2006) ClC-7 requires Ostm1 as a beta- near an absorbing substrate. Phys Rev Lett, 95:058102.
subunit to support bone resorption and lysosomal function. Nature, 440:220–223. 19. De Yoreo JJ, Wierzbicki A, Dove PM (2007) New insights into mechanisms of biomo-
5. Gower LB, Odom DJ (2000) Deposition of calcium carbonate films by a polymer- lecular control on growth of inorganic crystals. CrystEngComm, 9:1144–1152.
induced liquid-precursor (PILP) process. J Cryst Growth, 210:719–734. 20. Arnol’d VI (1986) Catastrophe Theory (Springer, New York).
6. Elhadj S, De Yoreo JJ, Hoyer JR, Dove PM (2006) Role of molecular charge and 21. Piana S, Jones F, Gale JD (2007) Aspartic acid as a crystal growth catalyst. CrystEng-
hydrophilicity in regulating the kinetics of crystal growth. Proc Natl Acad Sci USA, 103: Comm, 9:1187–1191.
19237–19242. 22. Piana S, Jones F, Gale JD (2006) Assisted desolvation as a key kinetic step for crystal
7. Stephenson AE, et al. (2008) Peptides enhance magnesium signature in calcite: Insights growth. J Am Chem Soc, 128:13568–13574.

PHYSICS
into origins of vital effects. Science, 322:724–727. 23. Petsev DN, Chen K, Gliko O, Vekilov PG (2003) Diffusion-limited kinetics of the
8. Hoyer JR, Otvos L, Urge L (1995) Role of osteopontin in urinary stone formation. Ann solution-solid phase transition of molecular substances. Proc Natl Acad Sci USA,
NY Acad Sci, 760:257–265. 100:792–796.
9. Deganello S, Piro OE (1981) The crystal structure of calcium oxalate monohydrate 24. DeYoreo JJ, Orme CA, Land TA (2001) Using atomic force microscopy to investigate
(whewellite). N Jb Miner Mh, 2:81–88 (translated from German). solution crystal growth. Advances in Crystal Growth Research, eds Sato K, Nakajima
10. Sheng XX, Jung TS, Wesson JA, Ward MD (2005) Adhesion at calcium oxalate crystal K, Furukawa Y (Elsevier Science, Amsterdam), pp 361–380.
surfaces and the effect of urinary constituents. Proc Natl Acad Sci USA, 102:267–272. 25. Qiu SR, et al. (2005) Modulation of calcium oxalate monohydrate crystallization by
11. Jung T, et al. (2004) Probing crystallization of calcium oxalate monohydrate and the citrate through selective binding to atomic steps. J Am Chem Soc, 127:9036–9044.
role of macromolecule additives with in situ atomic force microscopy. Langmuir, 20: 26. Fields GB, Noble RL (1990) Solid-phase peptide-synthesis utilizing 9-fluorenylmethox-
8587–8596. ycarbonyl amino-acids. Int J Peptide Protein Res, 35:161–214.
12. Qiu SR, et al. (2004) Molecular modulation of calcium oxalate crystallization by 27. Hoyer JR, Asplin JR, Otvos L (2001) Phosphorylated osteopontin peptides suppress crys-
osteopontin and citrate. Proc Natl Acad Sci USA, 101:1811–1815. tallization by inhibiting the growth of calcium oxalate crystals. Kidney Int, 60:77–82.
13. Weaver ML, et al. (2007) Inhibition of calcium oxalate monohydrate growth by citrate 28. Cody AM, Horner HT, Cody RD (1982) SEM study of the fine surface features of
and the effect of the background electrolyte. J Cryst Growth, 306:135–145. synthetic calcium oxalate monohydrate crystals. Scan Electron Microsc, 1:185–197.
14. Mann S (2001) Biomineralization: Principles and Concepts in Bioinorganic Materials 29. Tomazic BB, Nancollas GH (1979) A study of the phase transformation of calcium
Chemistry (Oxford Univ Press, New York). oxalate trihydrate monohydrate. Invest Urol, 16:329–335.

Friddle et al. PNAS ∣ January 5, 2010 ∣ vol. 107 ∣ no. 1 ∣ 15

You might also like