Journal of Crystal Growth 125 (1992) 215-222

North-Holland

Ó1992 – Elsevier Science Publishers B.V.  All rights reserved

 

Calcium oxalate monohydrate crystallization: citrate inhibition of nucleation and growth steps

 

Peter A. Antinozzi, Charles M. Brown and Daniel L. Punch

Center for the Study of Urolithiasis and Pathological Calcification and Department of Biochemistry and Molecular Biology, University of Florida College of Medicine, Gainesville, Florida 32610-0247, USA

 

Received 30 March 1992; manuscript received in final form 29 June 1992

 

 

 

The inhibitory action of citrate on calcium oxalate monohydrate (COM) crystallization has been examined in terms of nucleation and crystal growth kinetic properties. Lag-time data for the appearance of crystals and [‘4C] oxalate incorporation under crystal growth conditions allowed us to investigate the influence of citrate at physiological levels (3.5mM). Moreover, through the use of the EQUIL software, we formulated our solutions based on calculations of solute composition such that free calcium concentrations were the same in the absence and presence of this tricarboxylic acid. The presence of citrate had little effect on the apparent interfacial free energy as determined by nucleation kinetic studies, but total particle production was greater in the absence of citrate; this was evident from electron microscopy and was also indicated by corresponding values of pre-exponential terms of the Gibbs—Thomson equation. Crystal growth rates were lowered in the presence of citrate to 30% of the uninhibited value, and distinctive morphological habit modifications were also observed by scanning electron microscopy. Together, these findings suggest that citrate may influence COM crystallization at several stages, and we present a model for face-specific growth inhibition by citrate acting on the (010) COM crystal face.

 

 

1.   Introduction

 

Understanding the factors that influence the course of calcium oxalate monohydrate (COM) crystallization promises to provide insight about processes thought important in urolithiasis. In particular, much attention has been devoted to the analysis of crystallization in terms of dis­cretely defined physical chemical processes (i.e., nucleation, crystal growth, aggregation, and breakup) although such processes may be over­lapping in a temporal sense. Nonetheless, such a physicochemical treatment may afford a means of understanding the action of agents that promote or inhibit COM crystallization. Brown et al. [1] recently described several approaches for distin­guishing between crystal growth and nucleation effects, and they attempted to distinguish be­tween the interfacial free energy term and the “nucleation efficiency” term that are related to the lag-time for crystallization by use of the

Gibbs—Thomson equation. We were motivated by the success of that experimental approach to in­vestigate the action of citrate as an inhibitor of COM crystallization. About half a century ago, Greenwald [2] first recognized the complexation of calcium ion by various organic acids, among them citrate, and he discussed the physiological significance of the sharp increase in the solubility of salts such as calcium sulfate, calcium carbon­ate and calcium phosphate as brought about by malic and fumaric acids. Kissen and Locks [3] established that the urinary citrate levels of pa­tients suffering from urolithiasis were reduced compared with control subjects, and many other investigators have labored to characterize the ba­sis for such a difference and/or its impact on crystallization. The implication that citric acid is a factor in urolithiasis promoted efforts both to understand its ability to dissolve kidney stones and its role in their formation. In 1961, Light and Zinsser [4] examined the rate of formation of calcium oxalate in the presence of various sub­stances found in urine, including citrate, and in fact, they studied nucleation rates by observing lag-times. Subsequent studies of the inhibitory action of citrate focused on determinations of crystal growth viewed in its broadest sense (i.e., there are many steps in the formation of a solid phase from an aqueous solution, and a substance influencing any of these steps may be called an inhibitor). For example, complexation reduces the driving force for crystal growth, and this effect must be distinguished from those that influence incorporation of lattice ions into the crystal. Even some recent crystal growth studies have over­looked this distinction [5] and experimental re­sults represent a convolution of metal-ligand complexation and true growth inhibitory effects.

In this work, we applied nucleation kinetic [1] and crystal growth rate experiments to analyze the action of citrate on COM nucleation and crystal growth. We used the EQUIL speciation software [6] to achieve desired relative supersatu­rations (RS) with respect to calcium oxalate monohydrate (COM) while keeping the RS in control solutions and citrate-containing solutions unchanged. Moreover, to minimize changes in the [Ca2+]free/[OX>]tree ratio, we maintained constant free ionic calcium, [Ca2+ ]free and free ionic oxalate, [OX2~]free, concentrations for the control and experimental solutions of correspond­ing RS values, achieving a virtually constant ratio for all solutions. This ratio also controls surface charging and the zeta potential of the COM surface [7,8], and may therefore influence crystal growth.

 

 

2.  Materials and methods

 

Solution preparation. Reagent grade chemicals were used without further purification, and water of 10 M12 conductivity was produced with a Milli­Q high purity water system. All solutions were filtered through 0.22 ~tm Millipore GS filters (4.7 cm diameter) and cation concentrations were de­termined with a Perkin-Elmer atomic absorption spectrophotometer. Calcium and oxalate concen­trations were adjusted to achieve desired relative

supersaturation (RS) values, with RS defined as the calcium-oxalate ion activity product divided by its equilibrium value. At each RS, two reactant solutions were prepared, one containing potas­sium oxalate and the other calcium chloride dihy­drate. Typically, the buffered solution consisted of 0.1M sodium chloride, 0.O1M HEPES, and determined levels of either potassium oxalate or calcium chloride dihydrate. The solutions were maintained at 370C and pH was adjusted to 6.5. The total calcium and oxalate levels in each were chosen using EQUIL to ensure that uncomplexed Ca2~ was the same for samples with and without citrate. The second constraint in the EQUIL computations was to maintain the relative super-saturations of each pair of samples (i.e., with and without citrate). Calculated concentrations of free ionic and complex species for calcium oxalate monohydrate solutions in the presence and ab­sence (values in parentheses) of citrate at a rela­tive supersaturation of 19.7 (pH 6.5): total citrate, 3.5mM (none); sodium ion, 99.8mM (99.8mM); potassium ion, 1.60mM (1.88mM); calcium ion, 0.73mM (0.72mM); chloride, 106mM (102mM); oxalate, 0.51mM (0.49mM); citrate, 1.17mM (none); HEPES (unprotonated), 1.34mM (1.33 mM); HEPES (protonated), 8.66mM (8.67mM); potassium chloride, 19.8mM (22.7mM); monohy­drogen oxalate ion, 1.51~tM (1.48~tM); mono­sodium oxalate ion, 1751LM (230j.tM); mono-potassium oxalate ion, 0.044mM (0.052mM); cal­cium oxalate, 121mM (121mM); dicalcium oxalate ion, 6.27mM (6.18mM); calcium dioxalate ion, 1.06mM (1.02mM); calcium hydrogen oxalate ion, 0.021mM (0.021mM); monohydrogen citrate, 0.28mM (none); dihydnogen citrate, 2.O5btM (none); monopotassium citrate, 6.65mM (none); calcium citrate anion, 2.02mM (none); calcium hydnogen cjtrate, 16.6mM (none). Between cit­rate-containing and control solutions, uncom­plexed Ca2 + levels agreed within 1%, and uncom­plexed Ox2 - within 4%; correspondingly, surface charge effects on COM due to variation in [Ca2 + ]tree were minimized.

Nucleation. A typical run began by rapidly mixing 2.5 mL each of the two reactant solutions by manually pushing the fluids through an in-line helical mixer into a 1 cm pathlength polystyrene

Fig. 1. Typical turbidity plot used for lag time measurements. RS 37 with 3.5mM citrate. r = 100 s.

 

 

 

cuvette. We chose polystyrene because glass, quartz, and acrylic cuvettes produced appreciable growth on their surfaces. Turbidity was measured for 10 mm using a Perkin-Elmer 559A UV/VIS spectrometer in absorbance mode at 530 nm. Lag-times, r, were determined from plots of ab­sorbance versus time (fig. 1). The selected RS range was based on the behavior of turbidity measurements with respect to increasing RS. Be­low RS 20, the turbidity increase did not exceed 0.05 absorbance units, and lag-times resulting from small deviations above the baseline were considerably less certain. Above RS 37, lag-times were under 30 s. A typical experimental run is shown in fig. 1 where the dashed line indicates how the lag-time was evaluated by extrapolating to a turbidity value of zero; for example, in the case of citrate-containing systems at RS 19.7, the lag-time was 86 + 17 s. This method gave repro­ducible estimates of the apparent nucleation lag­-time. Therefore, we examined the dependence of r on changes in the initial relative supersatura­tion of calcium oxalate. Apparent interfacial free energy, s, was evaluated by plotting ln(1/r) ver­sus (ln RS)2 at six different relative supersatu­ration values to produce a linear plot (fig. 2); s was obtained from the slope of the line as de­fined by the Gibbs-Thomson equation [9]:

                                                        

J=A exp[(-16ps3v2)/(3k3T3m2[ln(RS)]2)],   (1)

 

where J is the nucleation rate which is propor­tional to 1/t (s-1), A the pre-exponential factor, s the apparent interfacial surface energy (erg cm-2, v the molecular volume (for COM, 1.10 x 10-22 cm3), k the Boltzmann constant (1.38 x 10-16 erg K-1), T the absolute temperature (in these experiments, 310 K), m the number of growth units represented by v, and RS the rela­tive supersaturation.

      Crystal growth. COM seeds were produced us­ing the dimethyl oxalate method [10]. First, a 2.5mM calcium chloride solution was prepared, and the pH adjusted to 4.7 with dilute ammonium hydroxide. Calcium chloride solution (150 mL) was added to 100 mL of ammonium acetate— acetic acid buffer (2.5 M with respect to each) into a 500 mL polymethylpentene plastic flask; then dimethyl oxalate (10 g) was added. The flask was tightly closed and heated in an oven at 900C for 2.5 h, followed by rapid cooling to room temperature. Crystals were collected by centrifu­gation, washed with a RS 1 solution in NaCl— HEPES buffer, and diluted to 0.311 mg mL-1. Crystals produced were monoclinic, with an aver­age length of 3.5 mm (fig. 4c). The advantage of this method is that a substantial quantity of large morphologicaily well described crystals were produced by slowly generating oxalate in situ under zero-order kinetic conditions. Surface area of 1700 cm2 g-1 was determined using the BET surface area analysis (Porous Materials, Inc.).

Equal volumes of calcium and oxalate reactant solutions were added to a 50 mL polymethylpen­tene flask maintained at 370C in a water bath (14C-oxalate was added as tracer to oxalate solu­tions). COM seed slurry (0.062 mg mLt final concentration) was added to initiate crystalliza­tion. At 5 mm intervals, aliquots were removed and filtered through a 0.22 ~.tm Nucleopore filter (25 mm). The filtrate was dispensed into a scintil­lation vial with 100 ~.tL normal HCI and scintilla­tion cocktail (Scintiverse II). Samples were counted using a Beckman LS 3801 liquid scintilla­tion counter. The crystal-laden filters were also counted after rinses with 3 mL of RS 1 solution. Oxalate concentrations were determined, and RS values were calculated assuming a 1 : 1 calcium oxalate precipitate. We should note that under these conditions [‘4C] oxalate exchange with seeds should be negligible. To estimate the influence of citrate on the crystal growth rate of calcium oxalate monohydrate, we used the parabolic growth rate law: (-d RS/dt =  kst[RSi - RS¥2). Integrating this equation gives

 

kstt =  (RSt - RS¥)-1 - (RSt - RS¥)-1,

 

where t is the time interval from the beginning of crystallization (in seconds), RS~ the relative su­persaturation at time t, RSi the initial relative supersaturation, RS¥. the relative supersaturation at equilibrium (defined as 1), K the crystal growth rate constant (s-1), and st the total crystal sur­face area at time t. We calculated st by matching fractional changes in total seed mass as deter­mined from the growth experiments to fractional changes in volume and correlated these to frac­tional changes in surface area using the surface area of the seeds (1700 cm2 g1) as s0. Volume and surface were related to each other based on the geometry of hexagonal prisms closely similar in morphology and dimension to the actual seeds as observed by SEM.

Morphology. Samples for microscopic analysis were taken at specified intervals after beginning

the experiment; typically, a 0.5 mL aliquot was removed and filtered through a 0.22 mm Nucleo­pore filter (13 mm). In the nucleation experi­ments, crystals were fixed after five and ten mm for each of the twelve solutions. For the crystal growth experiments, crystals were filtered after 0, 3, and 24 h for RS 20 solutions with and without citrate. Crystals were then gold-coated and exam­ined by scanning electron microscopy. Surface analysis was performed using a KEVEX X-ray spectrometer; no surface contaminants were found.

Particle characterization. Experimental systems the same as those used for lag-phase measure­ments were employed for purposes of particle characterization. Aliquots of these solutions were taken 20 mm after mixing [11]. Total particle number and mode particle diameter (equivalent spherical diameter) were measured using an Elzone 80 XY (Particle Data, Inc.).

 

 

 

3.  Results

 

To understand the action of citrate, we first applied a lag-phase kinetic analysis in which the appearance of crystals was evaluated turbidimet­rically using a spectrophotometer at a non-ab­sorbing wavelength (530 nm). When data col­lected in these experiments were analyzed using the Gibbs-Thomson equation as discussed above, plots of ln(1/r) versus (ln RS)-2 gave slopes of -62.5 + 1.92 for control and -49.5 + 6.26 for the experiment with citrate; intercepts were 1.03 + 0.2 and —0.72 ± 0.6 respectively. The slopes were further analyzed by converting them into values for the apparent interfacial energy for nucleation. This was done by solving the equation

 

slope = (—16ps3v2/3k3T3),             (2)

 

taken from the linearized form of the Gibbs-Thomson equation (cf. eq. (1)). Apparent interfa­cial energy for the control was 28.9 erg cm2, and for the citrate system it was 26.8 erg cm2. These values may be compared with those derived for systems very similar to our control system. In earlier work [1], we found a value of 27.3 erg

 


Fig. 2. Gibbs—Thomson nucleation plot: (●) control slope = -63±6.3, intercept = 1.0±0.60, r2 = 0.96; (▲) 3.5mM citrate slope = -50± 1.9, intercept=0.71±0.18, r2 = 0.99.

 

 

 


cm2and Finlayson [11] gave the value 31.1 erg-2 cm.

As another way of looking at nucleation, we examined particle production in our nucleating systems. Based on the observations of Finlayson [11], we expected total particles, N, to be a rather flat and somewhat noisy function of relative su­persaturation in this range of RS. This proved to be the case, however, we did detect statistically significant differences in N and in equivalent spherical diameter between the control and cit­rate systems. Without citrate, the nucleating sys­tems produced an average of 3.46 (± 1.94) x iO~ particles per liter, with an average mode diameter of 12.4 + 3.8 ~tm. With citrate, N was 7.14 (± 2.78) x i0~ particles per liter, with an average mode diameter of 8.2 ± 1.3 ~.tm.

Although there was a relatively small but sig­nificant difference between the apparent interfa­cial energies of the control and experimental systems, there was a considerable difference be­tween the lag-times in the two sets of solutions at corresponding relative supersaturations. Table 1 shows that lag-times in citrate systems were in­variably longer than the control systems by an average of 75%. We believed that the interfacial energies were an accurate reflection of nucle­ation in our experiments, and that the discrep­ancy in lag-times could be explained by growth inhibition due to the presence of citrate.

Crystal growth studies did indeed show a sig­nificant difference in growth rates at a citrate

 

 

 

Table 1

Kinetics of calcium oxalate monohydrate nucleation in the

absence and presence of 3.5mM citratea)

Relative                        Observed lag times (s)

supersaturation             Control             3.5mM citrate

20                                       400±44             560±26
22                                       250±14             350±80
24                                       170+66             290±47
28                                       130±57             170±61
33                                         70±19             140±33
37                                         30±12               90±17

a)Note that the uncomplexed calcium ion concentrations for each pair of control and citrate samples were the same, based on calculations with EQUIL.


Fig. 3. Determination of crystal growth rates: (e) control, K = 4.7±0.33X iO~ cm2 ~ intercept = 0.093+0 052 r2 =

0.98,  N= 7; (A) 3.5mM citrate K= 11+0 10x105 cm2 ~, intercept = 0.068±0.018, r2 = 0.96, N= 7.

 

 

 

concentration of 3.5mM (see fig. 3). Again, we sought to control the solution very closely so that differences in crystal growth rates could be clearly attributed to citrate. The surface normalized crys­tal growth rate in the control was 2.36 (± 0.16) x 106 51 cm2, whereas for the citrate system it was 0.66 (±0.03) x 106 51 cm2. The reduced growth rate of COM in the presence of citrate allowed us to explain the differences in particle counts and sizes we had observed. It was clear that growth inhibition would cause smaller parti­cles as we had seen with citrate, but because growth was delayed, the relative supersaturation, and therefore the nucleation rate, did not fall as quickly, consequently, more particles were pro­duced. Taken by itself, the doubling of the lag-times with citrate might have been interpreted as inhibition of nucleation while the doubling of the particle counts might have been interpreted as promotion of nucleation. By bringing several techniques to bear on this closely controlled ex­perimental design, a self-consistent view of the action of citrate emerged which explained these apparently conflicting results.

 

Fig. 4. Photographs of COM using scanning electron microscopy. Nucleation experiments at 10 mm: (a) RS 20 control; (b) RS 20 with 3.5mM citrate. Crystal growth experiments: (c) COM seeds at time zero; (d) COM seeds grown for 3 h in RS 20 control solution; (e) COM seeds grown for 3 h in an RS 20 containing 3.5mM citrate solution (magnification: 4000 x); (f) COM seeds

grown for 3 h in RS 20 containing 3.5mM citrate solution (magnification: 7800 x).


 


As in our previous report on nucleation, we attempted to model these results with our crystal­lization simulation program, PSD [1,12]. These efforts are still in the preliminary stages, but early results are encouraging because the simulations do reflect the general trends of the experiments.

The effect of citrate was seen in a striking way in electron micrographs of crystals from the nu­cleating systems. Fig. 4a shows crystals of COM nucleated in the control system. These crystals had a characteristic morphology for COM; they were twinned (as in fig. 4c) and had a prominent elongated hexagonal face. In the presence of cit­rate, however, the crystals were broader and flat­ter with an aspect similar to regular hexagons (fig. 4b). Seed crystals used in seeded growth experiments (fig. 4c) were not altered much in shape by either control systems (fig. 4d) or sys­tems containing citrate (figs. 4e and 4f).

 

4.  Discussion

 

In evaluating the inhibitory action of citrate, we found that compensation for the complexation of cations made a very considerable difference in the observed nucleation and growth rate behav­ior. To achieve equal uncomplexed calcium ion concentrations in the absence and presence of citrate, the total calcium concentration had to be raised significantly. In the control samples, un­complexed calcium ion corresponded to about 85% of the total calcium ion concentration, whereas it was only 25% of total calcium in the presence of citrate. Failure to account for the importance of complexation would have resulted

 

                                                   


Fig. 5. Computer simulation of COM crystal growth. (a) Morphology of control crystal as predicted by maturing each crystal face. Initial “nucleus” has an identical morphology. (b) Morphology of crystal computer grown in a 3.5mM citrate solution by restricting growth on (010) face. See fig. 4b.

 

 


in misleading inferences about apparent differ­ences in the kinetics of nucleation and crystal growth. By using EQUIL, however, we could correct for these chelation effects by citrate, and our data show that the presence of calcium cit­rate complex and citrate ions resulted in about a 70% decrease in crystal growth rate. This obser­vation indicates that uncomplexed citrate and/or calcium citrate must reduce the efficiency of adding calcium ion, oxalate ion, on calcium ox­alate complex to crystal growth sites.

From our kinetic and morphological studies of calcium oxalate monohydrate crystals, we now propose that citrate adsorbs preferentially to one crystal face, thereby altering the morphology of the crystals during the further accretion of cal­cium oxalate into the crystal. As shown in fig. 5, we could start with a common initial crystal mor­phology and allow crystals to “grow” using a computer program that represented changes in crystal mass as a change in total volume of the geometrically defined crystal. Without any change in apparent rates of addition to the various crys­tal faces, morphology was maintained; however, by restricting growth on the (010) face, increased addition to the other faces resulted in hexagonal plates shown in this figure. Using the X-ray crys­tallographic data of Deganello and Piro [13] we developed a specific proposal regarding this bind­ing behavior. Of the three major planes defining the calcium oxalate monohydrate crystal (i.e., the (010), (101), and the (001) planes), only the first

two have oxalate groups parallel to the face, and citrate would most probably replace oxalate ion by binding on the (101) face. In any case, the morphological changes due to the adsorption of citrate may be significant in urolithiasis, because crystal—cell interactions and crystal aggregation processes are both likely to be influenced by changes in crystal morphology. For example, Wiessner et al. [14] reported that COM crystals exhibit a much higher capacity to cause red blood cell membranolysis than do dihydrate crystals. Likewise, the adhesion of COM crystals to papil­lary cells are likely to be affected by the contour and nature of the various faces of calcium oxalate crystals. Crystal aggregation may also depend upon morphology, and the hexagonal plates formed in the presence of citrate may aggregate more readily. All of these considerations have led us to initiate a longer term study using stereo-chemical considerations as the basis for molecu­lar recognition at crystal interfaces [15]. Such efforts have already provided many valuable in­ferences regarding changes in crystal morphology arising from face-specific interactions of crystals and growth inhibitors [16,17]. While clearly be­yond the scope of our present work on citrate, we are attracted by the potential of the molecular recognition approach which may reveal how low-molecular-weight inhibitors such as citrate and pyrophosphate, as well as macromolecules (e.g., nephrocalcin and Tamm—Horsfall proteins) affect crystal growth processes.

Finally, although nucleation and growth proc­esses in the case of calcium oxalate monohydrate appear to overlap during the initial lag-phase of precipitation, our earlier studies [1] as well as those of Söhnel and Mullin [18] indicate that such lag-phase kinetics can be treated phenomenologi­cally in terms of the Gibbs—Thomson formula­tion. Nevertheless, as we probe further into the details of the early steps in COM crystallization using the tools of molecular recognition theory to establish a structural perspective, we recognize that more advanced theoretical treatments of precipitation could be helpful. Such models might, for example, allow for the deconvolution of the overlapping time domains of nucleation and growth steps.

 

References

 

[1]   C.M. Brown, D.K. Ackermann, D.L. Punch and B. Fin­layson, J. Crystal Growth 108 (1991) 445.

[2] I. Greenwald, J. Biol. Chem. 124 (1938) 437.

[3]   B. Kissen and MO. Locks, Soc. Expt. Biol. Med. Proc. 46 (1941) 216.

[4]   IS. Light and H.H. Zinsser, Arch. Biochem. Biophys. 92 (1961) 487.

[5]   H. Sidhu, R. Gupta, S. Thind and R. Nath, Urol. Res. 14 (1986) 299.

[6]   P. Werness, C. Brown, L. Smith and B. Finlayson, J. Urol. 134 (1985) 1242.

[7]   P. Curreri, G.Y. Onoda and B. Finlayson, J. Colloid Interface Sci. 69 (1978) 170.

[8]   J.H. Adair, Coagulation of Calcium Oxalate Monohy­drate Suspensions, PhD Dissertation, University of Florida, Gainesville, FL (1981).

[9]   AG. Walton, The Formation and Properties of Precipi­tates (Kreiger, Huntington, NY, 1979) pp. 3—7.

[10] L. Gordon, M.L. Salutsky and H.H. Willard, Precipita­

tion from Homogeneous Solution (Wiley, New York, 1959) pp. 57—58.

[11] B. Fiolayson, Kidney Intern. 13 (1978) 344.

[12] B. Finlayson, Comment on the round table discussion on theoretical models related to urolithiasis, in: Urolithiasis and Related Clinical Research, Eds. P.O. Schwille, L.H. Smith, W.G. Robertson and W. Vahlensieck (Plenum, New York, 1985) pp. 923—933.

[13] S. Deganello and O.E. Piro, Neues Jahrb. Mineral. Monatsh. H 2 (1981) 81.

[14] J.H. Wiessner, G.S. Mandel and N.S. Mandel, J. Urol. 135 (1986) 835.

[15] I. Weissbuch, L. Addadi, M. Lahav and L. Leiserowitz, Science 253 (1991) 637.

[16] J.D. Foot and E.A: Colbourn, J. Mol. Graphics 6 (1988) 93.

[17] I. Weissbuch, L. Addadi, L. Leiserowitz and M. Lahav, J. Am. Chem. Soc. 110 (1988) 561.

[18] 0. Sdhnel and J. Mullin, J. Colloid Interface Sci. 123 (1988) 43.