research papers\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoJOURNAL OF
APPLIED
CRYSTALLOGRAPHY
ISSN: 1600-5767

Qu­antitative phase analysis and microstructural characterization of urinary tract calculi with X-ray diffraction Rietveld analysis on a Caribbean island

crossmark logo

aUniversity of the West Indies, St Augustine, Trinidad and Tobago, and bEric Williams Medical Sciences Complex, Mount Hope, Trinidad and Tobago
*Correspondence e-mail: jaimie.greasley@gmail.com

Edited by J. Hajdu, Uppsala University, Sweden, and The European Extreme Light Infrastucture, Czech Republic (Received 3 August 2021; accepted 3 November 2021)

In the twin-island state of Trinidad and Tobago, urinary stone analysis is not routinely performed. This study investigates, via powder X-ray diffraction, 52 urinary tract calculi collected from hospitals in Trinidad. Of these, 46 stones were analysed with Rietveld refinement for quantitative analysis and materials characterization. Refined unit-cell, microstructural and weight fraction parameters were obtained, with the last being used for stone classification. The results revealed seven distinct mineralogical phases of varying frequency: calcium oxalate monohydrate (COM, 58%), calcium oxalate dihydrate (COD, 23%), carbonated apatite (APA, 48%), brushite (BRU, 6%), struvite (STR, 42%), uric acid (UA, 23%) and ammonium acid urate (AAU, 19%). The average refined crystallite sizes were 1352 ± 90 Å (COM), 1921 ± 285 Å (COD), 83 ± 5 Å (APA), 1172 ± 9 Å (BRU), 1843 ± 138 Å (STR), 981 ± 87 Å (UA) and 292 ± 83 Å (AAU). Subsequently, 36.5% of stones were categorized as phosphates, 34.6% as oxalates, 13.5% as uric acid/urates and 15.4% as mixed compositions. The study findings highlight the importance of stone analysis as a necessary step towards disease management of local patients, and endorse the application of Rietveld refinement as a natural extension to diffraction-based kidney stone investigations.

1. Introduction

Urolithiasis, the pathological formation of concretions in the urinary tract, is an affliction suffered by many around the globe. It is the third most common urological disorder (Prezioso et al., 2014[Prezioso, D., Illiano, E., Piccinocchi, G., Cricelli, C., Piccinocchi, R., Saita, A., Micheli, C. & Trinchieri, A. (2014). Arch. Ital. Urol. Androl. 86, 99-102.]), with a risk of 1–19% for Asian populations, 5–9% for Europeans and 12–15% for North Americans (Liu et al., 2018[Liu, Y., Chen, Y., Liao, B., Luo, D., Wang, K., Li, H. & Zeng, G. (2018). Asia. J. Urol. 5, 205-214.]; Ramello et al., 2001[Ramello, A., Vitale, C. & Marangella, M. (2001). J. Nephrol. 13, S45-S50.]). Epidemiological data depict a globally increased incidence of urolithiasis over the past few decades (Yoshida et al., 1999[Yoshida, O., Terai, A., Ohkawa, T. & Okada, Y. (1999). Kidney Int. 56, 1899-1904.]; Hesse et al., 2003[Hesse, A., Brändle, E., Wilbert, D., Köhrmann, K.-U. & Alken, P. (2003). Eur. Urol. 44, 709-713.]; Stamatelou et al., 2003[Stamatelou, K. K., Francis, M. E., Jones, C. A., Nyberg, L. M. Jr & Curhan, G. C. (2003). Kidney Int. 63, 1817-1823.]; Sharma & Filler, 2010[Sharma, A. P. & Filler, G. (2010). Indian J. Urol. 26, 516.]). Despite a patient being rendered stone free after medical and surgical interventions, there is also the issue of recurrence (Williams, 1963[Williams, R. E. (1963). Br. J. Urol. 35, 416-437.]). Urolithiasis may bring about acute renal failure as a complication of urinary obstruction and/or inflammation of the kidney (Jamal & Ramzan, 2004[Jamal, A. & Ramzan, A. (2004). J. Coll. Phys. Surg. Pak. 14, 411-415.]; Keddis & Rule, 2013[Keddis, M. T. & Rule, A. D. (2013). Curr. Opin. Nephrol. Hypertens. 22, 390-396.]; Tang & Lieske, 2014[Tang, X. & Lieske, J. C. (2014). Curr. Opin. Nephrol. Hypertens. 23, 385-390.]). Recurrent stone formers are thus vulnerable to decreased renal function (Gillen et al., 2005[Gillen, D. L., Worcester, E. M. & Coe, F. L. (2005). Kidney Int. 67, 685-690.]), the development of chronic kidney disease (CKD) and end-stage renal failure (ESRF) (Rule et al., 2011[Rule, A. D., Krambeck, A. E. & Lieske, J. C. (2011). Clin. J. Am. Soc. Nephrol. 6, 2069-2075.]; Kartha et al., 2012[Kartha, G., Calle, J. C., Marchini, G. S. & Monga, M. (2012). Urol. Clin. North Am. 40, 135-147.]; Keddis & Rule, 2013[Keddis, M. T. & Rule, A. D. (2013). Curr. Opin. Nephrol. Hypertens. 22, 390-396.]). Other concomitant associations include hypertension (Madore et al., 1998[Madore, F., Stampfer, M. J., Rimm, E. B. & Curhan, G. C. (1998). Am. J. Hypertens. 11, 46-53.]), sepsis (Al-Mamari, 2017[Al-Mamari, S. A. (2017). Complications of Urolithiasis, pp. 121-129. Heidelberg: Springer.]), osteoporosis (Pfau & Knauf, 2016[Pfau, A. & Knauf, F. (2016). Am. J. Kidney Dis. 68, 973-985.]) and urothelial carcinoma (Sun et al., 2013[Sun, L.-M., Lin, C.-L., Chang, Y.-J., Liang, J.-A., Liu, S.-H., Sung, F.-C. & Kao, C.-H. (2013). BJU Int. 112, 1150-1155.]).

Urinary tract calculi are due to the emergence, growth and clustering of mineralogical crystals in urine. Kidney stones are mostly crystalline, but are held together by a complex organic matrix comprising proteins and lipids (Khan & Kok, 2004[Khan, S. R. & Kok, D. J. (2004). Front. Biosci. 9, 1450-1482.]; Khan et al., 2016[Khan, S. R., Pearle, M. S., Robertson, W. G., Gambaro, G., Canales, B. K., Doizi, S., Traxer, O. & Tiselius, H.-G. (2016). Nat. Rev. Dis. Primers, 2, 1-23.]). Although it is reported that more than 100 different chemical species have been detected in kidney stones (Daudon & Jungers, 2012[Daudon, M. & Jungers, P. (2012). Urolithiasis: Basic Science and Clinical Practice, edited by J. T. Talati, H.-G. Tiselius, D. M. Albala & Z. Ye, pp. 113-140. Heidelberg: Springer.]), only about a dozen of these are found in greater than 1% of cases. The most common minerals are calcium-based oxalates (CaOx) and phosphates (CaPh) (Daudon et al., 2009[Daudon, M., Bazin, D., André, G., Jungers, P., Cousson, A., Chevallier, P., Véron, E. & Matzen, G. (2009). J. Appl. Cryst. 42, 109-115.]). Non-calcareous minerals include magnesium phosphates, uric acid/urates, and, exceptionally, rare protein and drug-related compositions (Daudon et al., 2016[Daudon, M., Dessombz, A., Frochot, V., Letavernier, E., Haymann, J.-P., Jungers, P. & Bazin, D. (2016). C. R. Chim. 19, 1470-1491.]).

Calcium oxalate monohydrate (COM, CaC2O4·H2O) or whewellite is by far the most frequent composition (Schubert, 2006[Schubert, G. (2006). Urol. Res. 34, 146-150.]). Its structure has been found in the monoclinic space group P21/c (a = 6.316, b = 14.541, c = 10.116 Å, β = 109°). Calcium oxalate dihydrate (COD, CaC2O4·2H2O), called weddellite, is the less chemically stable oxalate form and converts to COM over time (He et al., 2010[He, J.-Y., Deng, S.-P. & Ouyang, J.-M. (2010). IEEE Trans. Nanobiosci. 9, 156-163.]; Izatulina et al., 2018[Izatulina, A. R., Gurzhiy, V. V., Krzhizhanovskaya, M. G., Kuz'mina, M. A., Leoni, M. & Frank-Kamenetskaya, O. V. (2018). Cryst. Growth Des. 18, 5465-5478.]). It crystallizes in the tetragonal system with space group I4/m (a = 12.371, c = 7.357 Å). CaOx stones tend to be ambiguous with regard to their aetiology. Elevated levels of oxalate and calcium ions in the urine, referred to medically as hyperoxaluria and hypercalciuria, respectively, are associated with their formation. However, these conditions in turn require further assessment as they arise from a multitude of possible metabolic, dietary or genetic influences (Pak, 1998[Pak, C. Y. C. (1998). Lancet, 351, 1797-1801.]; Moe, 2006[Moe, O. W. (2006). Lancet, 367, 333-344.]).

Calcium phosphates include carbonated hydroxyapatite (APA, Ca5[PO4,CO3]3OH) and calcium hydrogen phosphate dihydrate or brushite (BRU, CaHPO4·2H2O). Hydroxy­apatite's structure belongs to the hexagonal P63/m space group (a = 9.424, c = 6.879 Å), whereas brushite has been assigned to monoclinic space group Cc (a = 5.8105, b = 15.1758, c = 6.2337 Å, β = 116.405°). The latter phase is rare, but its presence in calculi signals severe stone disease (Klee et al., 1991[Klee, L. W., Brito, C. G. & Lingeman, J. E. (1991). J. Urol. 145, 715-718.]; Evan et al., 2005[Evan, A. P., Lingeman, J. E., Coe, F. L., Shao, Y., Parks, J. H., Bledsoe, S. B., Phillips, C. L., Bonsib, S., Worcester, E. M., Sommer, A. J., Kim, S. A. M. C., Tinmouth, W. W. & Grynpas, M. (2005). Kidney Int. 67, 576-591.]).

Magnesium ammonium phosphate hexahydrate or struvite (STR, MgNH4PO4·6H2O) is more prevalent than the monohydrate form. It is orthorhombic, belonging to space group Pmn21 (a = 6.941, b = 6.137, c = 11.199 Å). Struvite stones arise from bacterial infection of urine. The production of urease, by specific strains of bacteria, facilitates the breakdown of urea into ammonia and carbon dioxide. This reaction alkalizes the urine and promotes the precipitation of struvite and carbonated apatite (Hess, 1990[Hess, B. (1990). Urol. Res. 18, S45-S48.]; Rahman et al., 2003[Rahman, N. U., Meng, M. V. & Stoller, M. L. (2003). Curr. Pharm. Des. 9, 975-981.]). Consequently, struvite and sometimes carbapatite stones are labelled as infection stones (Prywer & Torzewska, 2010[Prywer, J. & Torzewska, A. (2010). Cryst. Res. Technol. 45, 1283-1289.]).

Phases of uric acid and derivatives include anhydrous uric acid or uricite (UA, C5H4N4O3), uric acid dihydrate (C5H4N4O3·2H2O), ammonium acid urate (AAU, C5H7N5O3) and sodium urate (C5H3N4NaO3). Uricite belongs to space group P21/a (a = 14.464, b = 7.403, c = 6.208 Å, β = 65°), and there is one hypothesized structural model for AAU which puts it in the triclinic space group [P {\overline 1}] (a = 3.65, b = 10.215, c = 10.597 Å, α = 113.9, β = 91.1, γ = 92.3°) (Friedel et al., 2015[Friedel, P., Bergmann, J., Kleeberg, R. & Schubert, G. (2015). Z. Cryst. Suppl. 23, 517-522.]). Uric acid requires consistently low pH urine for precipitation (Shekarriz & Stoller, 2002[Shekarriz, B. & Stoller, M. L. (2002). J. Urol. 168, 1307-1314.]). Risk factors include gout or a family history of gout and a diet high in protein (Breslau et al., 1988[Breslau, N. A., Brinkley, L., Hill, K. D. & Pak, C. Y. C. (1988). J. Clin. Endocrinol. Metab. 66, 140-146.]). Persons who are overweight, obese or diabetic have also been shown to be at risk for uric acid urolithiasis (Sakhaee et al., 2002[Sakhaee, K., Adams-Huet, B., Moe, O. W. & Pak, C. Y. C. (2002). Kidney Int. 62, 971-979.]; Taylor et al., 2005[Taylor, E. N., Stampfer, M. J. & Curhan, G. C. (2005). JAMA, 293, 455-462.]; Mosli et al., 2013[Mosli, H. A., Mosli, H. H. & Kamal, W. K. (2013). Res. Rep. Urol. 5, 11-15.]; Lieske et al., 2006[Lieske, J. C., de la Vega, L. S. P., Gettman, M. T., Slezak, J. M., Bergstralh, E. J., Melton, L. J. III & Leibson, C. L. (2006). Am. J. Kidney Dis. 48, 897-904.]).

Multiple factors are implicated in stone formation. These may be classified as anatomical, genetic, metabolic, dietary or environmental in origin. Regardless of the aetiopathogenic forces at play, urinary stones are brought on by a specific set of physicochemical conditions and events. These are (i) the persistent supersaturation of the urine which leads to (ii) crystal nucleation and ultimately (iii) crystal growth and agglomeration (Finlayson, 1978[Finlayson, B. (1978). Kidney Int. 13, 344-360.]). The first two circumstances are termed `pre-requisites' as they produce crystals but do not compel the formation of a macroscopic stone (Rodgers, 2017[Rodgers, A. L. (2017). Urolithiasis, 45, 27-32.]). Urine is normally supersaturated with certain solutes and the appearance of crystals is not uncommon in non-stone formers (Grases et al., 2000[Grases, F., Costa-Bauzá, A., Königsberger, E. & Königsberger, L.-C. (2000). Int. Urol. Nephrol. 32, 19-27.]). What distinguishes lithogenic from normal urine is the frequency, size, morphology and extent of aggregation of the crystals (Kok et al., 1990[Kok, D. J., Papapoulos, S. E. & Bijvoet, O. L. M. (1990). Kidney Int. 37, 51-56.]; He et al., 2010[He, J.-Y., Deng, S.-P. & Ouyang, J.-M. (2010). IEEE Trans. Nanobiosci. 9, 156-163.]).

1.1. Stone investigations

Kidney stones are proof of an atypical urinary environment, with appearance and crystalline composition being direct clues to their origin and development (Daudon et al., 2008[Daudon, M., Jungers, P. & Bazin, D. (2008). AIP Conf. Proc. 1049, 199-215.], 2016[Daudon, M., Dessombz, A., Frochot, V., Letavernier, E., Haymann, J.-P., Jungers, P. & Bazin, D. (2016). C. R. Chim. 19, 1470-1491.]; Cloutier et al., 2015[Cloutier, J., Villa, L., Traxer, O. & Daudon, M. (2015). World J. Urol. 33, 157-169.]). Analysis of stones should be a central component of diagnostic evaluation for all urolithiasis patients (Coe et al., 1992[Coe, F. L., Parks, J. H. & Asplin, J. R. (1992). N. Engl. J. Med. 327, 1141-1152.]; Grases et al., 1998[Grases, F., Costa-Bauzá, A. & García-Ferragut, L. (1998). Adv. Colloid Interface Sci. 74, 169-194.]; Tiselius, 2000[Tiselius, H.-G. (2000). Braz. J. Urol. 26, 452-462.]; Kourambas et al., 2001[Kourambas, J., Aslan, P., Teh, C. L., Mathias, B. J. & Preminger, G. M. (2001). J. Endourol. 15, 181-186.]). Knowledge of molecular composition is key to patients' risk evaluation for recurrent stone disease or more grave developments like CKD and ESRF. For this purpose, popular stone analysis techniques include wet chemical analysis, infrared spectroscopy (IR), scanning electron microscopy (SEM) and powder X-ray diffraction (PXRD) (Basiri et al., 2012[Basiri, A., Taheri, M. & Taheri, F. (2012). Urol. J. 9, 445-454.]). PXRD and IR are the standard and recommended methods (Turk et al., 2020[Turk, C., Neisius, A., Petřík, A., Seitz, C., Thomas, K. & Skolarikos, A. (2020). European Association of Urology Guidelines, 2020 ed., presented at the EAU Annual Congress, Amsterdam, 2020. https://uroweb.org/guideline/urolithiasis/. Arnhem: The European Association of Urology Guidelines Office.]).

Powder X-ray diffractometry permits the direct identification of crystalline phases due to the acquisition of characteristic diffraction patterns upon sample interaction with a monochromatic X-ray beam. Kidney stones are polycrystalline and may be effectively assessed via this technique. The inclusion of the Rietveld method for analysis of diffractograms helps to extract as much information as possible, other than simply stone composition (Le Bail et al., 2008[Le Bail, A., Cranswick, L. M. D. & Madsen, I. (2008). Powder Diffraction: Theory and Practice. Cambridge: Royal Society of Chemistry.]). A Rietveld procedure involves the iterative refinement of structural, microstructural and sample parameters such that a calculated profile becomes well fitted to the measured diffraction pattern. For kidney stone analysis, Rietveld studies have the potential to reveal structural details related to pathology-specific physicochemical conditions and stone growth mechanisms. In terms of microstructure, there are several indicators for further investigation into crystallite sizes of kidney stone phases (Shapur et al., 2012[Shapur, N. K., Uvarov, V., Popov, I., Katz, R., Gofrit, O. N., Landau, E. H., Pode, D. & Duvdevani, M. (2012). Urology, 80, 980-985.]; Daudon et al., 2016[Daudon, M., Letavernier, E., Weil, R., Véron, E., Matzen, G., André, G. & Bazin, D. (2016). C. R. Chim. 19, 1527-1534.]; Bazin et al., 2021[Bazin, D., Frochot, V., Haymann, J.-P., Letavernier, E. & Daudon, M. (2021). C. R. Chim. 24(S2), 96.]).

This paper reports the analysis of urinary tract calculi in Trinidad using the powder X-ray diffraction method with Rietveld analysis. Our results are submitted in the form of weighted proportions of detected crystalline components for each stone, the refined unit-cell dimensions and angles, and the crystallite size for each phase. Stones were classified on the basis of their quantitative composition. The data were analysed and compared with other batch stone studies conducted globally. The benefits of Rietveld studies for stone analysis are also discussed.

2. Experimental

2.1. Sample collection and preparation

Urinary tract calculi were collected from two public hospitals. The stones were rinsed with a saline solution post-removal and then stored in plastic or glass containers. Prior to analysis, the stones were rinsed again with de-ionized water and allowed to air-dry for up to 48 h. A ceramic mortar and pestle were used to grind the stones into a fine powder for XRD scanning.

2.2. Instrumentation and scanning

Diffraction scans were conducted with the Bruker D2 Phaser Tabletop X-ray diffractometer for the angular range 2 < 2θ < 55° with a 0.02° step size. The diffractometer setup was the standard Bragg–Brentano geometry with a primary and secondary goniometer radius of 141.4 mm. The radiation source was a ceramic X-ray tube with a Cu anode target. The wavelengths of Kα1, Kα2 and Kβ were 1.5406, 1.5444 and 1.3922 Å, respectively, with a Kα2/Kα1 ratio of approximately 0.5. The optical system comprised 2.5° Soller modules, a 1 mm (0.6°) fixed divergence slit and an Ni Kβ filter. The detector was a LynxEye linear position-sensitive detector.

2.3. Phase identification

Scan files were imported into the DIFFRAC.EVA software (Version 4.2; Bruker AXS GmbH, Karlsruhe, Germany). Crystalline phases were identified with the aid of the program's Search/Match operation using the ICDD PDF-2 (2011 version; International Centre for Diffraction Data, https://www.icdd.com) reference database. Phase presence was subsequently confirmed with Rietveld analysis.

2.4. Crystallite size and crystallinity

In DIFFRAC.EVA, the percent crystallinity of the samples was calculated with the Crystallinity function. This operation automatically determines the crystallinity according to equation (1[link]),

[{\rm Crystallinity} = \left (1 - {{G - R} \over {G}} \right) \,100\% , \eqno(1)]

where G is the global area under the diffraction profile and R is the reduced area (area of Bragg peaks) after background subtraction.

Crystallite size calculations via the Scherrer method using both integral breadth (IB) and FWHM were performed on selected samples. The Create Area tool in DIFFRAC.EVA allows the user to demarcate the angular range of a single peak for crystallite size determination. Several single peaks were chosen from the scans and used for the DIFFRAC.EVA calculation. The instrumental contribution was set as the averaged FWHM of corundum peaks scanned using the same settings as the samples. The Scherrer constant k was provided as 1.

2.5. Rietveld refinement

The Rietveld method was first introduced in the 1960s (Rietveld, 1967[Rietveld, H. M. (1967). Acta Cryst. 22, 151-152.], 1969[Rietveld, H. M. (1969). J. Appl. Cryst. 2, 65-71.]) and is today adopted into numerous software programs (Le Bail et al., 2008[Le Bail, A., Cranswick, L. M. D. & Madsen, I. (2008). Powder Diffraction: Theory and Practice. Cambridge: Royal Society of Chemistry.]). The structure refinement routine involves the extraction of data from a calculated profile (Icalc) which has been fitted as well as possible to the observed data (Iexp) via minimization of the weighted sum of squares, WSS,

[{\rm WSS} = \textstyle\sum\limits_{i=1}^N w_i^2 \left ( I_i^{\rm exp} - I_i^{\rm calc} \right )^2 . \eqno(2)]

The calculated intensity at the ith point of a pattern ([I_i^{\rm calc}]) is a sum over the contributions of all the phases, peaks and background at that point,

[\eqalignno{I_i^{\rm calc} = & \, \underbrace{I\sum_{j=1}^{N_{\rm phases}} {{f_j} \over {V_j^2}}}_{\rm scale \, factor} \left [ \sum_{k=1}^{N_{\rm peaks}} L_k |F_{k,j}|^2 \underbrace{G_j (2 \theta_i - 2\theta_{k,j})}_{\rm PSF} P_{k,j} A_j \right ] \cr & \, + \underbrace{\sum_{n=0}^{N_{\rm b}} a_n (2 \theta_i)^n}_{\rm bkg} . &(3)}]

How well a calculated pattern matches the observed profile is ordinarily judged by means of the weighted profile R factor:

[R_{\rm wp} = \left [ {{\rm WSS} \over {\sum_{i=1}^N \left ( w_i I_i^{\rm exp} \right )^2}} \right ]^{1/2} . \eqno(4)]

The lower the value of Rwp, the better the fit is presumed to be.

In these equations, wi = [1/(I_i^{\rm exp})^{1/2}], I is the beam intensity, fj is the phase volume fraction for the jth phase, Vj is the unit-cell volume of phase j, Lk is the Lorentz polarization factor of the kth peak, |Fk, j|2 is the squared structure factor for the kth peak of the jth phase, Gj is the peak shape function (PSF) for phase j, Pk,j is the preferred orientation correction for peak k in phase j, Aj is the X-ray absorption correction for phase j, an is the background (bkg) coefficient for the nth polynomial and Nb is the degree of polynomial for background modelling.

2.5.1. Phase quantification

For a multiphase sample, the weight fraction of the pth phase is computed by the simple relation in equation (5[link]) (Hill & Howard, 1987[Hill, R. J. & Howard, C. J. (1987). J. Appl. Cryst. 20, 467-474.]). In Rietveld software, the scale factors and structural parameter values from the refinement are used to calculate these fractions.

[W_p = {{S_p Z_p M_p V_p} \over {\sum_i S_i Z_i M_i V_i}} , \eqno(5)]

where S is the Rietveld scale factor, Z is the number of formula units per unit cell, M is the mass per unit cell and V is the unit-cell volume.

2.5.2. MAUD analysis

The refinements were performed in Materials Analysis Using Diffraction (MAUD) (Lutterotti et al., 1999[Lutterotti, L., Matthies, S. & Wenk, H. R. (1999). IUCr Commission on Powder Diffraction Newsletter, No. 21, pp. 14-15.]). The software features classic optimization algorithms for quantitative phase analysis and microstructural and texture analysis for materials characterization from X-ray, neutron and electron diffraction data. Structural data for detected phases were imported as CIFs from the Crystallography Open Database (Gražulis et al., 2009[Gražulis, S., Chateigner, D., Downs, R. T., Yokochi, A. F. T., Quirós, M., Lutterotti, L., Manakova, E., Butkus, J., Moeck, P. & Le Bail, A. (2009). J. Appl. Cryst. 42, 726-729.]) for COM (ID 2300210; Daudon et al., 2009[Daudon, M., Bazin, D., André, G., Jungers, P., Cousson, A., Chevallier, P., Véron, E. & Matzen, G. (2009). J. Appl. Cryst. 42, 109-115.]), COD (ID 9000764; Tazzoli & Domeneghetti, 1980[Tazzoli, V. & Domeneghetti, C. (1980). Am. Mineral. 65, 327-334.]), APA (ID 9011094; Sudarsanan & Young, 1969[Sudarsanan, K. & Young, R. A. (1969). Acta Cryst. B25, 1534-1543.]), BRU (ID 9007305; Schofield et al., 2004[Schofield, P. F., Knight, K. S., Houwen, J. A. M. & Valsami-Jones, E. (2004). Phys. Chem. Miner. 31, 606-624.]), STR (ID 2106462; Whitaker & Jeffery, 1970[Whitaker, A. & Jeffery, J. W. (1970). Acta Cryst. B26, 1429-1440.]) and UA (ID 9011061; Ringertz, 1966[Ringertz, H. (1966). Acta Cryst. 20, 397-403.]). The structural data for AAU (Friedel et al., 2015[Friedel, P., Bergmann, J., Kleeberg, R. & Schubert, G. (2015). Z. Cryst. Suppl. 23, 517-522.]) were registered manually into MAUD.

A typical procedure involves step-wise refinement of the following parameters: (i) background and scale, (ii) zero offset, unit-cell lengths and angles, and displacement parameters, (iii) crystallite size and microstrain, and (iv) preferred orientation. For crystallite size modelling in MAUD, anisotropic line broadening was chosen. Crystallite sizes were started at 1000 Å for most phases, except APA and AAU which were started at 100 and 500 Å, respectively. The general spherical harmonics model was applied to accommodate preferred orientation.

3. Results and discussion

3.1. Phase composition

Phase identification revealed seven distinct crystalline phases, which were COM, COD, hydroxyapatite/carbapatite (APA), BRU, STR, UA and AAU. There was no attempt made to distinguish hydroxyapatite from carbapatite as biological apatites are always carbonated to some degree (Maurice-Estepa et al., 1999[Maurice-Estepa, L., Levillain, P., Lacour, B. & Daudon, M. (1999). Scand. J. Urol. Nephrol. 33, 299-305.]; Bazin et al., 2009[Bazin, D., Chappard, C., Combes, C., Carpentier, X., Rouzière, S., André, G., Matzen, G., Allix, M., Thiaudière, D., Reguer, S., Jungers, P. & Daudon, M. (2009). Osteoporos. Int. 20, 1065-1075.]). There were between one and four phases in each stone, with a modal value of two phases per stone. Thirteen were monophasic, 23 were bi-phasic, nine had three phases and seven had four phases. It was more likely that a stone comprised more than one phase, which is in accordance with the findings of Schubert (2006[Schubert, G. (2006). Urol. Res. 34, 146-150.]).

The bar chart in Fig. 1[link] depicts the number of calculi found to contain the seven phases. COM was the most abundant mineral, having been discovered in 30 of 52 stones. This was followed by APA (N = 25), STR (N = 22), UA (N = 12), COD (N = 12) and AAU (N = 10), and the least common was BRU (N = 3). All phases, with the exception of the last, were detected as majority constituents in at least one stone (Fig. 2[link]). Here, we define `majority' as containing a weight fraction more than or equal to half. Pure (monophasic) stones were observed for COM, APA, STR and UA, but no pure AAU or COD stones were found.

[Figure 1]
Figure 1
The frequency of crystalline phases across 52 stones.
[Figure 2]
Figure 2
The relative abundance of the phase majorities.

3.2. Phase correlation

With respect to the co-existence of two given phases within a stone, Pearson correlation co-efficients were computed. The colour matrix is depicted in Fig. 3[link], with shades of blue indicating positive correlation (r > 0) and red indicating a negative correlation (r < 0) between a phase pair. Correlations deemed significant (p ≤ 0.01) are labelled within colour blocks. A significant positive covariance was found between COM and COD (r = 0.469, p < 0.001), APA and STR (r = 0.500, p < 0.001), and UA and AAU (r = 0.428, p = 0.002). Significant negative covariances were found for COM and STR (r = −0.448, p < 0.001), COD and STR (r = −0.377, p = 0.006), and APA and UA (r = −0.527, p < 0.001). Overall, these values indicate that oxalate, phosphate and uric acid phases are likely to be present alongside other phases of the same group, but the co-existence of two phases of different groups is unlikely. In the chemical context, this is logical as the conditions giving rise to specific phases may preclude others. It should be noteworthy when unlikely phases appear together, as this may signal crucial changes in urinary conditions and lithogenic contributors.

[Figure 3]
Figure 3
The relative abundance of the phase majorities.

3.3. Stone classification

Fifty-two urinary tract stones were analysed qualitatively for their composition. The majority, 46, were submitted to a complete Rietveld analysis. The remainder were only partially assessed because of difficulties in the refinement of multiple phases or an unsatisfactory Rwp factor due to inadequate modelling of strong preferred orientation. Nevertheless, preliminary quantitative results facilitated their classification. The stones fell into four major categories: oxalates, phosphates, uric acid/urates and mixed stones. There were no discovered protein or drug stones. A sample was categorized as an `oxalate', `phosphate' or 'uric acid/urate' if about 70% of its weight constituted phases belonging to the specified class. Calculi were deemed `mixed' if there was a weight ratio of at least 3:2 of phases belonging to two distinct classes.

Our results show two major stone categories (phosphate and oxalate) and two minor categories (uric acid/urate and mixed). The relative abundance of each category is depicted in the pie chart in Fig. 4[link]. The numbers of stones classified as oxalate, phosphate, uric acid/urate and mixed were 18, 19, seven and eight, respectively. Among oxalate stones, 59% consisted of solely oxalate phases, that is whewellite and weddellite. The rest were majority oxalate with some phosphate or uric acid/urate content. Pure whewellite stones represented 41% of all oxalates, but no pure weddellite stones were found. Within the phosphate group, 80% consisted of solely phosphate phases, mainly apatite and struvite, and the remainder contained trace amounts of whewellite. From the seven samples classified as uric acid/urate stones, two were pure uric acid stones, three were uric acid with ammonium acid urate and the rest contained small amounts of oxalate or phosphate phases. For the mixed stones category, half were a combination of oxalate and uric acid/urate phases, 38% were a combination of phosphate and uric acid/urate phases, and a single stone was a mixed oxalate/phosphate stone.

[Figure 4]
Figure 4
The abundance of stone categories and sub-categories.

Chatterjee et al. (2018[Chatterjee, P., Chakraborty, A. & Mukherjee, A. K. (2018). Spectrochim. Acta A Mol. Biomol. Spectrosc. 200, 33-42.]) reported a dominance of oxalate stones for eastern India on the basis of XRD data. From a nearly identical sample size of 50 stones to our 52, 82% were classified as oxalates, which is different from what we have reported. It is obvious from our data that whewellite was the most frequent phase (30/52). However, oxalate phases were the majority (wt% ≥ 50%) for just 42.3% (22/52), half the number reported by Chatterjee and co-workers. Moreover, our classification criterion for mixed calculi led to only 34.6% (18/52) being definitely `oxalate' stones as substantial amounts of non-oxalate phases were also present.

A Japan-based study also showed a high dominance of stones consisting of oxalate, comparable to the reports of Chatterjee et al. (2018[Chatterjee, P., Chakraborty, A. & Mukherjee, A. K. (2018). Spectrochim. Acta A Mol. Biomol. Spectrosc. 200, 33-42.]). Hossain et al. (2003[Hossain, R. Z., Ogawa, Y., Hokama, S., Morozumi, M. & Hatano, T. (2003). Int. J. Urol. 10, 411-415.]) recorded 81.6% stones with CaOx, 15.8% uric acid/urate-containing stones and just 3.7% struvite stones via semi-quantitative IR spectroscopic analysis. An advantage of their analysis is a large sample size of more than 1800 stones, but the lack of a fully quantitative method and rigid classification protocol makes direct comparison difficult. In the current work, UA and AAU were also detected at a high frequency in 23.1 and 19.2% of stones, respectively.

Uvarov et al. (2011[Uvarov, V., Popov, I., Shapur, N., Abdin, T., Gofrit, O. N., Pode, D. & Duvdevani, M. (2011). Environ. Geochem. Health, 33, 613-622.]) gave figures of 43.2% oxalates, 35.9% mixed stones, 10.3% urates and 7.7% phosphates from a reference intensity ratio XRD assessment of 278 stones in Jerusalem. A high proportion of mixed stones is prominent from their results, but is most likely attributable to the absence of a quantitative boundary in the classification method. Our results give 15.4% for mixed calculi, though following the same approach would yield 40.4%, scaling similarly to their findings.

Giannossi et al. (2012[Giannossi, M. L., Mongelli, G., Tateo, F. & Summa, V. (2012). J. X-ray Sci. Technol. 20, 175-186.]) documented 59% oxalate stones followed by 18% uric acid from a batch of 80 stones in southern Italy using qualitative analysis with SEM, PXRD and optical microscopy. Ma et al. (2017[Ma, R., Luo, X., Li, Q. & Zhong, H. (2017). Int. J. Surg. 41, 150-161.]) conducted an extensive qualitative analysis of 2437 stones with Fourier transform IR spectroscopy, of which 720 were imaged with SEM. They reported 53% oxalates, 18% uric acid and 6% total phosphates (APA, BRU, STR) for their hospital in Guangzhou. Keshavarzi et al. (2016[Keshavarzi, B., Yavar Ashayeri, N., Moore, F., Irani, D., Asadi, S., Zarasvandi, A. & Salari, M. (2016). Minerals, 6, 131.]) noted an abundance of whewellite and uricite from XRD Rietveld analysis of 39 stones for Iran. Most of their data set were sole oxalates (28%) and uric acid (21%) or mixed oxalate–uric acid compositions (41%). The remainder were split up into isolated or double cases of cystine or pure or mixed oxalate–phosphates.

Some consistency is seen from the above reports: oxalates are most common, followed by uric acid and then phosphates. Our data deviate from this, as phosphates and oxalates are equally dominant. Similar findings of co-dominant oxalate and phosphate stones, 32 and 37%, respectively, were cited in a very recent study on a Mayan population in Mexico (Cruz-May et al., 2021[Cruz-May, T. N., Herrera, A., Rodríguez-Hernández, J., Basulto-Martínez, M., Flores-Tapia, J. P. & Quintana, P. (2021). J. Mol. Struct. 1235, 130267.]).

One variation amongst the previous reports is the wide gap between the dominant oxalate and uric acid groups outlined by Hossain et al. (2003[Hossain, R. Z., Ogawa, Y., Hokama, S., Morozumi, M. & Hatano, T. (2003). Int. J. Urol. 10, 411-415.]) and Chatterjee et al. (2018[Chatterjee, P., Chakraborty, A. & Mukherjee, A. K. (2018). Spectrochim. Acta A Mol. Biomol. Spectrosc. 200, 33-42.]) but not by the others, who report a lesser abundance of oxalates, greater frequency of other groups and a `mixed' category. Our report also follows the trend of the latter.

Caution must be taken in making and interpreting these comparisons. Uvarov et al. (2011[Uvarov, V., Popov, I., Shapur, N., Abdin, T., Gofrit, O. N., Pode, D. & Duvdevani, M. (2011). Environ. Geochem. Health, 33, 613-622.]) highlighted the lack of a standard in classifying stones amongst studies, which is especially apparent in our discussion for `mixed' calculi. The greatest uncertainty lies in differences in the analytical procedures used by stone researchers. Some surveys employed XRD as their principal technique, others IR. Some analyses are quantitative or semi-quantitative, with others relying on a qualitative assessment alone. Only Chatterjee et al. (2018[Chatterjee, P., Chakraborty, A. & Mukherjee, A. K. (2018). Spectrochim. Acta A Mol. Biomol. Spectrosc. 200, 33-42.]) and Keshavarzi et al. (2016[Keshavarzi, B., Yavar Ashayeri, N., Moore, F., Irani, D., Asadi, S., Zarasvandi, A. & Salari, M. (2016). Minerals, 6, 131.]) employed Rietveld analysis in the above studies. Additionally, the sample size is quite varied, ranging from dozens of calculi to thousands. Whilst using more advanced methods, smaller studies like ours may not accurately represent the entire study population.

3.4. Rietveld analysis

Forty-six calculi underwent a complete Rietveld analysis with MAUD. The refined values of unit-cell parameters, percentage weight (Wt%) and crystallite size (Crys.) for each identified phase alongside the weighted R factor (Rwp) are listed in Table 1[link]. The final stone classifications are also labelled in the table as oxalate, phosphate, uric acid/urate and mixed. Observed and calculated intensities for samples with one, two, three and four phases co-present are provided in Figs. 5[link]–7[link][link]. The entire collection of Rietveld plots is available as supporting information.

Table 1
Composition, classification (Class.) and Rietveld refinement data for 46 urinary tract calculi

The table shows refined values of unit-cell parameters, percentage weight (Wt%) and crystallite size (Crys.) for each identified phase alongside the weighted R factor (Rwp). Final stone classifications are labelled as oxalate (Ox), phosphate (Ph), uric acid/urate (Ur) and mixed (Mx).

          Cell dimensions (Å) Cell angles (°)  
Label Rwp Class. Phase Wt% a b c α β γ Crys. (Å)
C1 0.0859 Ur UA 100 14.502 7.469 6.22   65.08   657
                       
C2 0.0812 Ur UA 74.7 14.497 7.457 6.215   64.99   729
COD 22.2 12.399   7.37       1006
COM 3.1 6.301 14.621 10.149   109.55   1323
                       
C3 0.1045 Ph APA 56.3 9.437   6.899       77
STR 43.7 6.958 6.144 11.223       1371
                       
C4 0.0986 Ox COD 60.9 12.378   7.359       1704
COM 39.1 6.294 14.595 10.124   109.42   1215
                       
C5 0.0788 Ur UA 61.4 14.482 7.459 6.214   65.05   539
AAU 33.2 3.659 10.16 10.618 113.61 91.02 92.7 174
STR 5.4 6.945 6.153 11.237       818
                       
C6 0.0944 Ph STR 51.3 6.949 6.139 11.217       1345
APA 48.7 9.456   6.884       81
                       
C7 0.0944 Ur UA 100 14.511 7.473 6.226   65.06   781
                       
C8 0.0959 Ph STR 86.5 6.951 6.142 11.221       2307
APA 13.5 9.401   6.91       67
                       
C9 0.0998 Ox COD 57.7 12.368   7.36       1244
APA 21 9.437   6.875       67
BRU 10.8 5.817 15.171 6.253   116.41   1163
COM 10.5 6.211 14.564 10.164   107.08   1724
                       
C10 0.083 Ox COM 82.6 6.298 14.606 10.123   109.47   1303
APA 15 9.416   6.89       94
COD 2.5 12.36   7.36       1260
                       
C11 0.0934 Ox COM 80.8 6.298 14.604 10.124   109.46   2225
APA 15.4 9.443   6.88       122
COD 3.8 12.378   7.365       3027
                       
C12 0.0549 Ph APA 84.9 9.44   6.887       69
STR 15.1 6.954 6.143 11.219       2420
                       
C13 0.1031 Ox COM 69.3 6.297 14.607 10.127   109.45   2075
COD 30.7 12.378   7.361       1949
                       
C14 0.0644 Ph APA 58.9 9.48   6.887       60
STR 41.1 6.957 6.14 11.216       2650
                       
C15 0.078 Ox COM 72.8 6.297 14.603 10.125   109.47   2364
APA 25 9.446   6.885       147
COD 2.2 12.368   7.364       1175
                       
C16 0.0725 Ph APA 71.3 9.447   6.883       67
COM 14.1 6.302 14.61 10.129   109.47   1246
STR 8.5 6.988 6.133 11.22       1832
BRU 6.1 5.85 15.183 6.31   117.04   1180
                       
C17 0.112 Ox COM 100 6.296 14.607 10.126   109.46   5328
                       
C18 0.0574 Ph APA 81.3 9.431   6.885       84
STR 18.7 6.947 6.136 11.22       2221
                       
C19 0.075 Mx (Ph/Ox) APA 48.5 9.476   6.887       83
COM 26.2 6.3 14.612 10.129   109.47   1933
STR 13.6 6.955 6.14 11.225       1998
COD 11.8 12.377   7.364       2728
                       
C20 0.0595 Ph APA 71 9.428   6.885       89
COM 29 6.297 14.608 10.127   109.44   988
                       
C21 0.075 Ox COM 81.1 6.297 14.605 10.128   109.47   1184
UA 18.9 14.472 7.473 6.205   65.04   1142
                       
C22 0.0521 Ph APA 88.5 9.436   6.875       76
STR 11.5 6.941 6.131 11.211       1492
                       
C23 0.1073 Ox COM 100 6.299 14.613 10.13   109.48   1655
                       
C24 0.0758 Ox COM 100 6.299 14.608 10.124   109.45   1211
                       
C25 0.0925 Ox COM 100 6.301 14.612 10.131   109.47   1015
                       
C26 0.081 Ox COM 100 6.303 14.619 10.129   109.45   920
                       
C27 0.0449 Ph APA 64.1 9.426   6.895       80
STR 35.9 6.951 6.14 11.214       1966
                       
C28 0.0684 Ox COM 74.3 6.3 14.608 10.124   109.45   1005
COD 25.7 12.373   7.358       3199
                       
C29 0.0578 Ph APA 89.4 9.446   6.871       64
STR 10.6 6.936 6.135 11.214       1013
                       
C30 0.0743 Ur UA 88.2 14.493 7.463 6.214   65.04   947
AAU 11.8 3.69 10.12 10.671 113.6 90.15 91.93 117
                       
C31 0.0669 Mx (Ox/UA) COM 53.1 6.297 14.604 10.126   109.45   897
UA 37.6 14.473 7.455 6.204   65.01   1022
STR 6.3 6.959 6.138 11.207       1604
AAU 2.9 3.675 10.171 10.608 113.76 90.7 92.78 595
                       
C32 0.0875 Ph STR 100 6.952 6.14 11.223       2276
                       
C33 0.051 Ph APA 100 9.433   6.872       81
                       
C34 0.0577 Ph APA 100 9.452   6.864       69
                       
C35 0.0731 Ur AAU 58.7 3.669 10.094 10.656 113.46 90.81 93.02 478
UA 41.3 14.53 7.444 6.252   64.6   880
                       
C36 0.0544 Ph APA 71.6 9.428   6.879       127
COM 28.4 6.298 14.6 10.121   109.47   1350
                       
C37 0.0866 Mx (Ph/UA) AAU 50 3.675 10.076 10.661 113.52 90.75 92.54 280
STR 37.3 6.957 6.141 11.226       1208
APA 12.7 9.51   6.89       25
                       
C38 0.0666 Ph APA 72 9.437   6.887       70
STR 18.7 6.953 6.14 11.22       2763
COM 9.3 6.296 14.608 10.125   109.47   1693
                       
C39 0.0554 Ph APA 61.5 9.439   6.884       75
STR 38.5 6.949 6.138 11.215       2844
                       
C40 0.0607 Ox COM 100 6.301 14.61 10.127   109.45   908
                       
C41 0.0815 Ox COM 100 6.3 14.614 10.13   109.46   1316
                       
C42 0.1175 Ph STR 100 6.955 6.139 11.221       1253
                       
C43 0.0689 Mx (Ox/UA) COM 60.8 6.299 14.609 10.126   109.46   1080
UA 39.2 14.479 7.457 6.206   65.01   1539
                       
C44 0.1088 Ph STR 82.7 6.957 6.142 11.227       2297
APA 17.3 9.428   6.87       124
                       
C45 0.0691 Mx (Ox/UA) COM 42.5 6.301 14.614 10.13   109.47   957
UA 38.3 14.486 7.458 6.206   65.01   957
AAU 15.9 3.666 10.175 10.599 113.85 90.92 93.53 105
STR 3.3 6.97 6.141 11.213       1185
                       
C46 0.0516 Mx (Ox/UA) COM 55 6.301 14.61 10.127   109.45   854
UA 45 14.48 7.459 6.206   65.02   1112
[Figure 5]
Figure 5
Rietveld refinement of whewellite sample C24 (top) and apatite sample C33 (bottom).
[Figure 6]
Figure 6
Rietveld refinement of monophasic sample C1 (top) and biphasic sample C46 (bottom).
[Figure 7]
Figure 7
Rietveld refinement of triphasic sample C2 (top) and four-phase sample C9 (bottom).

The final unit-cell parameters were generally quite similar to the starting values ([\Delta \,\lt\, 0.5\%]) with a few exceptions ([0.5 \,\lt] [ \Delta \,\lt\, 2\%]). The greatest variations were observed for AAU. However, only six refinements were performed for this phase since its triclinic structure made refinement challenging. Deviations were also observed for some samples with minor phases, e.g. C9 for COM, C37 for APA, and C5, C16 and C45 for STR.

The crystallite size data are statistically summarized for each phase in Table 2[link], with the crystallite size distributions for the three most frequent phases illustrated in Fig. 8[link]. Box plots for most phases are provided in Fig. 9[link]. For COM, the distribution is right skewed. One outlier was omitted from the statistical calculations for a pure COM stone (C17). The diffractogram displayed extreme texture which was difficult to model, and the final crystallite size was 5328 Å. Apatite (APA) showed the smallest crystallite sizes and a positively skewed distribution. The lower bound of 25 Å (C37) is a possible outlier, as may be seen from Fig. 8[link]. The second smallest value was 60 Å. Struvite revealed quite large crystallite sizes with the highest median value of 1899 Å (Table 2[link]). The distribution appears bimodal at 1250 and 2250 Å. For other phases, the number of refinements performed was significantly lower. COD crystallites presented the widest range of values from N = 9 refinements and the largest mean size at 1921 ± 285 Å. The uric acid crystallites had mean and median sizes below 1000 Å. Ammonium acid urate had the second smallest sizes following apatite. Brushite was only refined twice and took values of 1160 and 1183 Å.

Table 2
Summary of Rietveld refinement crystallite size data

Phase No. of refinements Range (Å) Median (Å) Mean (Å)
COM 24 854–2364 1231 1352 ± 90
COD 9 1006–3199 1704 1921 ± 285
APA 23 25–147 77 83 ± 5
STR 20 818–2844 1899 1843 ± 138
BRU 2 1160–1183    
UA 12 539–1539 952 981 ± 87
AAU 6 105–595 227 292 ± 83
[Figure 8]
Figure 8
Histogram distributions of crystallite sizes for whewellite, weddellite and struvite.
[Figure 9]
Figure 9
Box plots of crystallite sizes for whewellite, weddellite, struvite, uric acid and ammonium acid urate.

The sizes for COM, UA and AAU are quite similar to what has been published by Chatterjee et al. (2015[Chatterjee, P., Pramanik, S. & Mukherjee, A. K. (2015). J. Appl. Cryst. 48, 1794-1804.], 2018[Chatterjee, P., Chakraborty, A. & Mukherjee, A. K. (2018). Spectrochim. Acta A Mol. Biomol. Spectrosc. 200, 33-42.]). The APA values are smaller in our study as most fell below 10 nm, whereas Chatterjee and co-workers reported a range of 10–35 nm. Conversely, a broader and larger range of crystallite sizes was obtained in our study for COD (101–320 nm) than theirs (42–167 nm) for the same number of refinements. There is a difference in methodology, however, as the anisotropic model was implemented in our refinements as opposed to the isotropic model for crystallite size modelling. Bazin et al. (2009[Bazin, D., Chappard, C., Combes, C., Carpentier, X., Rouzière, S., André, G., Matzen, G., Allix, M., Thiaudière, D., Reguer, S., Jungers, P. & Daudon, M. (2009). Osteoporos. Int. 20, 1065-1075.]) wrote that biological apatite nanocrystals in bone were about 10 nm, which matches well with our results. Bazin et al. (2012[Bazin, D., André, G., Weil, R., Matzen, G., Emmanuel, V., Carpentier, X. & Daudon, M. (2012). Urology, 79, 786-790.]) measured a mean value of 250 nm for struvite crystals via powder neutron diffraction. In agreement, 40% of our values reflect a size greater than 200 nm and 60% above 150 nm for STR.

3.5. DIFFRAC.EVA analysis

3.5.1. Scherrer crystallite size

A handful of samples were analysed where possible with the traditional IB and FWHM approaches included in the DIFFRAC.EVA software (Table 3[link]). The values were lower than the Rietveld-refined crystallite sizes by 33–70%, but the general order of the phases is maintained. The UA sizes were smaller than those of COM, which in turn were smaller than those of STR. Uvarov et al. (2011[Uvarov, V., Popov, I., Shapur, N., Abdin, T., Gofrit, O. N., Pode, D. & Duvdevani, M. (2011). Environ. Geochem. Health, 33, 613-622.]) also employed the FWHM method to evaluate crystallite sizes for hundreds of samples. Our findings correlate well with their modal values for COM and UA of 70 and 45 nm, respectively. Limitations of this method are the multi-phasic compositions of some samples and overlapping peaks, especially with APA, which make single peak area and width measurements difficult or even impossible. Only distinctly separated low-angle peaks were considered, which was remarkably limiting and not representative of all hkl peaks.

Table 3
Integral breadth (IB) and FWHM averaged crystallite sizes for selected samples

Label Phase IB (Å) FWHM (Å)
C1 UA 386 428
C7 UA 413 443
C8 STR 749 829
C12 STR 914 1109
C40 COM 636 690
C41 COM 710 799
C44 STR 657 734
3.5.2. Crystallinity

The percent crystallinity averaged 56.3% across all samples. In ranking order of increasing crystallinity, the stone categories were phosphates (48.5% crystalline), mixed (54%), oxalates (61.9%) and uric acid stones (65.2%). These data are similar to the results of Mirković et al. (2020[Mirković, M., Dosen, A., Erić, S., Vulić, P., Matović, B. & Rosić, A. (2020). Microchem. J. 152, 104429.]), who calculated high crystallinities for COM and COD of 68% and for UA of 65%, although our averages were based on stone type and not individual phases as in their study. The phosphate category ranked the lowest due to the low crystallinity calculated for hydroxyapatite, which Mirković et al. (2020[Mirković, M., Dosen, A., Erić, S., Vulić, P., Matović, B. & Rosić, A. (2020). Microchem. J. 152, 104429.]) recorded to be around 18%. The high percent crystallinity calculated for struvite and brushite, however, acts as the buffer for the phosphate group.

In reference to biological hydroxyapatite, Londoño-Restrepo et al. (2019[Londoño-Restrepo, S. M., Jeronimo-Cruz, R., Millán-Malo, B. M., Rivera-Muñoz, E. M. & Rodriguez-García, M. E. (2019). Sci. Rep. 9, 5915.]) highlighted issues regarding reports of low crystallinity by the scientific community. In fact, their high-resolution transmission electron microscopy study revealed high atomic structural order for APA crystals from human, bovine and porcine bones. Londoño-Restrepo and co-workers measured bi-dimensional APA crystals to be 21 ± 8 nm long and 6 ± 2 nm wide for human bone and even smaller for the bovine and porcine samples. The ability to evaluate crystalline quality accurately using traditional XRD methods for apatite is tied to the difficulty in isolating the effects of inelastic scattering for nano-scaled crystallites. Whilst the present investigation has found `low crystallinity' for apatite, we have also provided results of nanometric APA crystals in kidney stones based on 23 Rietveld refinements.

3.6. Clinical implications

Schubert (2006[Schubert, G. (2006). Urol. Res. 34, 146-150.]) assessed the compositions of more than 110 000 stones and published the frequencies of mineral components as 78% for COM, 43% COD, 33% APA, 10% UA, 6% STR, 1–2% BRU and 1% AAU. Here, we have reported lower frequencies of COM and COD and a significantly higher prevalence of all other phases amongst a small sample of 52 stones.

An excessive 42% of calculi contained struvite, which speaks of infection either as the origin of the stone or occurring at some point thereafter. Struvite stones are regarded as high risk for recurrence and sepsis (Gao et al., 2020[Gao, X., Lu, C., Xie, F., Li, L., Liu, M., Fang, Z., Wang, Z., Ming, S., Dong, H., Shen, R., Sun, Y., Peng, Y. & Gao, X. (2020). World J. Urol. 38, 219-229.]; Turk et al., 2020[Turk, C., Neisius, A., Petřík, A., Seitz, C., Thomas, K. & Skolarikos, A. (2020). European Association of Urology Guidelines, 2020 ed., presented at the EAU Annual Congress, Amsterdam, 2020. https://uroweb.org/guideline/urolithiasis/. Arnhem: The European Association of Urology Guidelines Office.]). The higher the struvite content, the greater the recurrence risk (Nevo et al., 2019[Nevo, A., Shahait, M., Shah, A., Jackman, S. & Averch, T. (2019). Int. Urol. Nephrol. 51, 585-591.]). One study reported a strong correlation between mixed CaOx–struvite calculi and hypercalciuria in patients (Kristensen et al., 1987[Kristensen, C., Parks, J. H., Lindheimer, M. & Coe, F. L. (1987). Kidney Int. 32, 749-753.]), implicating infection as a secondary event to metabolically triggered CaOx urolithiasis. This is significant, as our phase correlation statistics have shown the unlikelihood of COM content in a struvite stone. Nevertheless, a few cases (C16, C19, C31, C38 and C45) detailed in Table 1[link] present a COM–STR combination. According to the literature, infection may be the secondary cause and a clinical investigation ought to include metabolic evaluation of the patient.

A considerable number of stones (19%) were revealed to comprise AAU. Although this phase is predominantly linked with infection, similarly to struvite (Tiselius, 2000[Tiselius, H.-G. (2000). Braz. J. Urol. 26, 452-462.]; Chou et al., 2012[Chou, Y.-H., Huang, C.-N., Li, W.-M., Huang, S.-P., Wu, W.-J., Tsai, C.-C., Chang, A.-W., Chen, S.-M., Lin, Y.-L. & Lin, Y.-P. (2012). Kaohsiung J. Med. Sci. 28, 259-264.]), associations have also been made with morbid obesity, recurrent uric acid stones, irritable bowel syndrome and laxative abuse (Soble et al., 1999[Soble, J. J., Hamilton, B. D. & Streem, S. B. (1999). J. Urol. 161, 869-873.]; Kuruma et al., 2006[Kuruma, H., Arakawa, T., Kubo, S., Hyodo, T., Matsumoto, K., Satoh, T., Egawa, S. & Baba, S. (2006). Int. J. Urol. 13, 498-501.]; Lomas et al., 2017[Lomas, D. J., Jaeger, C. D. & Krambeck, A. E. (2017). Urology, 102, 43-47.]). Kuruma et al. (2006[Kuruma, H., Arakawa, T., Kubo, S., Hyodo, T., Matsumoto, K., Satoh, T., Egawa, S. & Baba, S. (2006). Int. J. Urol. 13, 498-501.]) contemplated a need to distinguish between pure and mixed AAU stones as there were perceptible clinical differences between the two groups. They reported that 70% of the pure AAU group had history with laxative abuse, whilst mixed AAU stone formers were principally older men deemed as overweight. In the present work, AAU-containing stones were never pure but most often discovered with struvite (60%) and UA (60%), or with both phases 30% of the time. Infection seems innately related to AAU crystallization but it is doubtful that this is what actually initiated most stones. The key to this is the UA content, which suggests a metabolic origin, with a possible infection as the secondary promoter. A common thread of AAU surveys, though, is the regular appearance of serious co-morbidities in study participants. Lomas et al. (2017[Lomas, D. J., Jaeger, C. D. & Krambeck, A. E. (2017). Urology, 102, 43-47.]) found diabetes in 9% and chronic kidney disease in 11% of patients. Chou et al. (2012[Chou, Y.-H., Huang, C.-N., Li, W.-M., Huang, S.-P., Wu, W.-J., Tsai, C.-C., Chang, A.-W., Chen, S.-M., Lin, Y.-L. & Lin, Y.-P. (2012). Kaohsiung J. Med. Sci. 28, 259-264.]) gave figures of 60% for CKD and 12% for urothelial carcinoma and noted an elevated recurrence risk. In the light of this, AAU lithiasis represents a critical class of stone formers from this study that must be evaluated further.

Infection-related, uric acid and brushite-containing stones are all classed as high risk for recurrence by the European Association of Urology (Turk et al., 2020[Turk, C., Neisius, A., Petřík, A., Seitz, C., Thomas, K. & Skolarikos, A. (2020). European Association of Urology Guidelines, 2020 ed., presented at the EAU Annual Congress, Amsterdam, 2020. https://uroweb.org/guideline/urolithiasis/. Arnhem: The European Association of Urology Guidelines Office.]). Furthermore, non-calcium stones are associated with reduced renal function (Chou et al., 2011[Chou, Y.-H., Li, C.-C., Hsu, H., Chang, W.-C., Liu, C.-C., Li, W.-M., Ke, H.-L., Lee, M.-H., Liu, M.-E., Pan, S.-C. & Wang, H. S. (2011). Kaohsiung J. Med. Sci. 27, 264-267.]). This investigation has highlighted a substantial proportion of these high-risk constituent phases amongst a sample of stones from hospitals in Trinidad. Due to the limited sample size, additional research is required to confirm whether the trend persists for the larger population of stone patients in the country. Should this be the case, an investigation into specific risk factors for the local population would be necessary for preventative care.

3.7. Benefits and outlook

Analysis of powder XRD data with a Rietveld-based approach not only provides a means for quantitative estimation of crystalline phases but often facilitates a qualitative assessment. In practice, minority yet critical phases had been missed in the phase identification step, with small peaks being overlooked as `impurity' peaks prior to Rietveld analysis. The calculation of an entire profile pattern from the already-known phases would then allow us to correct such oversights. For example, minor amounts of apatite were overshadowed by sharp peaks of other phases as in samples C9–C11, or the reverse scenario might occur whereby peak overlap of phases with small crystallite sizes masks normally well defined peaks of other minerals. Additionally, crowded diffractograms with three to four minerals sometimes concealed one of the components, like for struvite in C31 or whewellite in C16. The subsequent classification of stones according to phase fraction estimates is made more accurate by whole-powder-pattern fitting.

A further benefit of the Rietveld method for kidney stone analysis is the deeper insight granted by the characterization of structure and microstructure of mineral components. Shapur et al. (2012[Shapur, N. K., Uvarov, V., Popov, I., Katz, R., Gofrit, O. N., Landau, E. H., Pode, D. & Duvdevani, M. (2012). Urology, 80, 980-985.]) proposed that, for whewellite and apatite, crystallite size is related to the eventual volume of the macroscopic stone. In their study, smaller crystallites were associated with higher stone burdens, whereas larger crystallites were associated with smaller stones. Shapur and co-workers suggested that this may be useful for predicting the potential for an obstructive stone in a patient. In the biological context, crystallite size is a parameter for renal cytotoxicity. It has been shown that smaller whewellite, weddellite and apatite crystals result in higher renal cell death (Sun, Gan & Ouyang, 2015[Sun, X. Y., Gan, Q. Z. & Ouyang, J. M. (2015). Cell Death Discovery, 1, 1-8.]; Sun, Ouyang et al., 2015[Sun, X.-Y., Ouyang, J.-M., Zhu, W.-Y., Li, Y.-B. & Gan, Q.-Z. (2015). J. Mater. Chem. B, 3, 1864-1878.]; Sun et al., 2020[Sun, X.-Y., Chen, J.-Y., Rao, C.-Y. & Ouyang, J.-M. (2020). Int. J. Nanomed. 15, 5043-5060.]; Bazin et al., 2021[Bazin, D., Frochot, V., Haymann, J.-P., Letavernier, E. & Daudon, M. (2021). C. R. Chim. 24(S2), 96.]). Daudon et al. (2016[Daudon, M., Letavernier, E., Weil, R., Véron, E., Matzen, G., André, G. & Bazin, D. (2016). C. R. Chim. 19, 1527-1534.]) carried out neutron powder diffraction analysis and discovered significant differences in uric acid crystallite size for non-diabetic males and females, but no variation in size between diabetics. From findings such as these, it is clear that the role of crystallite size should be more thoroughly assessed for its biological and medical implications.

Significant progress has been made in urolithiasis research within the past four decades, yet there are still a few fundamental gaps in knowledge, for example with regard to mechanistic theories of crystal growth and aggregation, modulator macromolecules, and the role of trace elements in stone formation (Khan & Kok, 2004[Khan, S. R. & Kok, D. J. (2004). Front. Biosci. 9, 1450-1482.]; Aggarwal et al., 2013[Aggarwal, K. P., Narula, S., Kakkar, M. & Tandon, C. (2013). Biomed. Res. Int. 2013, 292953. ]; Giannossi et al., 2013[Giannossi, M. L., Summa, V. & Mongelli, G. (2013). J. Trace Elem. Med. Biol. 27, 91-97.]; Singh & Rai, 2014[Singh, V. K. & Rai, P. K. (2014). Biophys. Rev. 6, 291-310.]; Ramaswamy et al., 2015[Ramaswamy, K., Killilea, D. W., Kapahi, P., Kahn, A. J., Chi, T. & Stoller, M. L. (2015). Nat. Rev. Urol. 12, 543-557.]; Rodgers, 2017[Rodgers, A. L. (2017). Urolithiasis, 45, 27-32.], 2019[Rodgers, A., Trinchieri, A., Ather, M. H. & Buchholz, N. (2019). Urolithiasis, 47, 401-413.]). Increased structural knowledge of biogenic crystalline materials could prove helpful to our understanding (Izatulina & Yelnikov, 2008[Izatulina, A. R. & Yelnikov, V. Y. (2008). Minerals as Advanced Materials I, edited by S. V. Krivovichev, pp. 231-239. Heidelberg: Springer.]). Variation in unit-cell parameters signals changes at the atomic and/or microscopic level indicative of the crystal growth and stone formation process. For instance, a high degree of carbonation in hydroxyapatite stones is affiliated to bacterial origin (Carpentier et al., 2009[Carpentier, X., Daudon, M., Traxer, O., Jungers, P., Mazouyes, A., Matzen, G., Véron, E. & Bazin, D. (2009). Urology, 73, 968-975.]). As apatite is well studied for its applications, the relationship between incorporation of carbonate ions in the structure and the distance parameters a and c is established (Ren et al., 2013[Ren, F., Lu, X. & Leng, Y. (2013). J. Mech. Behav. Biomed. Mater. 26, 59-67.]). For weddellite, Izatulina et al. (2014[Izatulina, A., Gurzhiy, V. & Frank-Kamenetskaya, O. (2014). Am. Mineral. 99, 2-7.]) found a linear relationship between the unit-cell parameter a and the zeolitic water contained in its structure. More revelations such as these may be revealed with larger-scaled crystallographic investigations into kidney stone materials.

The current investigation has generated unit-cell parameter and crystallite size data for the most common kidney stone phases from PXRD Rietveld data. There are too few studies which have stepped in this direction (Izatulina & Yelnikov, 2008[Izatulina, A. R. & Yelnikov, V. Y. (2008). Minerals as Advanced Materials I, edited by S. V. Krivovichev, pp. 231-239. Heidelberg: Springer.]; Ghosh et al., 2009[Ghosh, S., Basu, S., Chakraborty, S. & Mukherjee, A. K. (2009). J. Appl. Cryst. 42, 629-635.], 2014[Ghosh, S., Bhattacharya, A., Chatterjee, P. & Mukherjee, A. K. (2014). Z. Kristallogr. Cryst. Mater. 229, 451-458.]; Mukherjee, 2014[Mukherjee, A. K. (2014). J. Indian Inst. Sci. 94, 35-44.]; Chatterjee et al., 2015[Chatterjee, P., Pramanik, S. & Mukherjee, A. K. (2015). J. Appl. Cryst. 48, 1794-1804.], 2018[Chatterjee, P., Chakraborty, A. & Mukherjee, A. K. (2018). Spectrochim. Acta A Mol. Biomol. Spectrosc. 200, 33-42.]; Cruz-May et al., 2021[Cruz-May, T. N., Herrera, A., Rodríguez-Hernández, J., Basulto-Martínez, M., Flores-Tapia, J. P. & Quintana, P. (2021). J. Mol. Struct. 1235, 130267.]). Considering the challenges of multiphasic refinement, only one other study thus far has published data for more than 30 stones (Chatterjee et al., 2018[Chatterjee, P., Chakraborty, A. & Mukherjee, A. K. (2018). Spectrochim. Acta A Mol. Biomol. Spectrosc. 200, 33-42.]). More studies would be vital for affirmation of structural and microstructural data for classic kidney stones and finding any trends which may prove clinically relevant.

4. Conclusions

Stone analysis for the determination of crystalline constituents is a crucial step in risk assessment for recurrence prevention of stone disease. A powder X-ray diffraction study with Rietveld analysis was employed for a quantitative, structural and microstructural assessment of the compositional crystalline phases in 46 urinary tract calculi.

The refined crystallite sizes ranged from 85 to 236 nm for COM (N = 24), 101 to 320 nm for COD (N = 9), 3 to 15 nm for APA (N = 23), 82 to 284 nm for STR (N = 20), 116 to 118 nm for BRU (N = 2), 54 to 154 nm for UA (N = 12) and 11 to 60 nm for AAU (N = 6). The phase weight fractions allowed the classification of the sample set as 36.5% phosphates, 34.6% oxalates, 15.4% mixed stones and 13.5% uric acid/urates.

The study has found an elevated frequency in the appearance of high-risk phases such as struvite (42%), uric acid (23%), ammonium acid urate (19%) and brushite (6%), indicating the need for prophylactic intervention in study patients.

The application of the Rietveld method is beneficial for enhanced accuracy through whole-pattern fitting, but also for establishing structural values for crystalline phases which may be helpful for understanding stone growth processes.

Supporting information


Acknowledgements

The authors thank Mr Adrian Gayah and Ms Sadira Khan of the Materials Sciences Laboratory at the University of the West Indies for their technical assistance.

References

First citationAggarwal, K. P., Narula, S., Kakkar, M. & Tandon, C. (2013). Biomed. Res. Int. 2013, 292953.   Google Scholar
First citationAl-Mamari, S. A. (2017). Complications of Urolithiasis, pp. 121–129. Heidelberg: Springer.  Google Scholar
First citationBasiri, A., Taheri, M. & Taheri, F. (2012). Urol. J. 9, 445–454.  Web of Science PubMed Google Scholar
First citationBazin, D., André, G., Weil, R., Matzen, G., Emmanuel, V., Carpentier, X. & Daudon, M. (2012). Urology, 79, 786–790.  Web of Science CrossRef PubMed Google Scholar
First citationBazin, D., Chappard, C., Combes, C., Carpentier, X., Rouzière, S., André, G., Matzen, G., Allix, M., Thiaudière, D., Reguer, S., Jungers, P. & Daudon, M. (2009). Osteoporos. Int. 20, 1065–1075.  Web of Science CrossRef PubMed CAS Google Scholar
First citationBazin, D., Frochot, V., Haymann, J.-P., Letavernier, E. & Daudon, M. (2021). C. R. Chim. 24(S2), 96.  Google Scholar
First citationBreslau, N. A., Brinkley, L., Hill, K. D. & Pak, C. Y. C. (1988). J. Clin. Endocrinol. Metab. 66, 140–146.  CrossRef CAS PubMed Web of Science Google Scholar
First citationCarpentier, X., Daudon, M., Traxer, O., Jungers, P., Mazouyes, A., Matzen, G., Véron, E. & Bazin, D. (2009). Urology, 73, 968–975.  Web of Science CrossRef PubMed Google Scholar
First citationChatterjee, P., Chakraborty, A. & Mukherjee, A. K. (2018). Spectrochim. Acta A Mol. Biomol. Spectrosc. 200, 33–42.  Web of Science CrossRef CAS PubMed Google Scholar
First citationChatterjee, P., Pramanik, S. & Mukherjee, A. K. (2015). J. Appl. Cryst. 48, 1794–1804.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationChou, Y.-H., Huang, C.-N., Li, W.-M., Huang, S.-P., Wu, W.-J., Tsai, C.-C., Chang, A.-W., Chen, S.-M., Lin, Y.-L. & Lin, Y.-P. (2012). Kaohsiung J. Med. Sci. 28, 259–264.  Web of Science CrossRef CAS PubMed Google Scholar
First citationChou, Y.-H., Li, C.-C., Hsu, H., Chang, W.-C., Liu, C.-C., Li, W.-M., Ke, H.-L., Lee, M.-H., Liu, M.-E., Pan, S.-C. & Wang, H. S. (2011). Kaohsiung J. Med. Sci. 27, 264–267.  Web of Science CrossRef PubMed Google Scholar
First citationCloutier, J., Villa, L., Traxer, O. & Daudon, M. (2015). World J. Urol. 33, 157–169.  Web of Science CrossRef PubMed Google Scholar
First citationCoe, F. L., Parks, J. H. & Asplin, J. R. (1992). N. Engl. J. Med. 327, 1141–1152.  CrossRef PubMed CAS Web of Science Google Scholar
First citationCruz-May, T. N., Herrera, A., Rodríguez-Hernández, J., Basulto-Martínez, M., Flores-Tapia, J. P. & Quintana, P. (2021). J. Mol. Struct. 1235, 130267.  Google Scholar
First citationDaudon, M., Bazin, D., André, G., Jungers, P., Cousson, A., Chevallier, P., Véron, E. & Matzen, G. (2009). J. Appl. Cryst. 42, 109–115.  Web of Science CSD CrossRef CAS IUCr Journals Google Scholar
First citationDaudon, M., Dessombz, A., Frochot, V., Letavernier, E., Haymann, J.-P., Jungers, P. & Bazin, D. (2016). C. R. Chim. 19, 1470–1491.  Web of Science CrossRef CAS Google Scholar
First citationDaudon, M. & Jungers, P. (2012). Urolithiasis: Basic Science and Clinical Practice, edited by J. T. Talati, H.-G. Tiselius, D. M. Albala & Z. Ye, pp. 113–140. Heidelberg: Springer.  Google Scholar
First citationDaudon, M., Jungers, P. & Bazin, D. (2008). AIP Conf. Proc. 1049, 199–215.  CrossRef CAS Google Scholar
First citationDaudon, M., Letavernier, E., Weil, R., Véron, E., Matzen, G., André, G. & Bazin, D. (2016). C. R. Chim. 19, 1527–1534.  Web of Science CrossRef CAS Google Scholar
First citationEvan, A. P., Lingeman, J. E., Coe, F. L., Shao, Y., Parks, J. H., Bledsoe, S. B., Phillips, C. L., Bonsib, S., Worcester, E. M., Sommer, A. J., Kim, S. A. M. C., Tinmouth, W. W. & Grynpas, M. (2005). Kidney Int. 67, 576–591.  Web of Science CrossRef PubMed CAS Google Scholar
First citationFinlayson, B. (1978). Kidney Int. 13, 344–360.  CrossRef CAS PubMed Web of Science Google Scholar
First citationFriedel, P., Bergmann, J., Kleeberg, R. & Schubert, G. (2015). Z. Cryst. Suppl. 23, 517–522.  Google Scholar
First citationGao, X., Lu, C., Xie, F., Li, L., Liu, M., Fang, Z., Wang, Z., Ming, S., Dong, H., Shen, R., Sun, Y., Peng, Y. & Gao, X. (2020). World J. Urol. 38, 219–229.  Web of Science CrossRef CAS PubMed Google Scholar
First citationGhosh, S., Basu, S., Chakraborty, S. & Mukherjee, A. K. (2009). J. Appl. Cryst. 42, 629–635.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationGhosh, S., Bhattacharya, A., Chatterjee, P. & Mukherjee, A. K. (2014). Z. Kristallogr. Cryst. Mater. 229, 451–458.  Web of Science CrossRef CAS Google Scholar
First citationGiannossi, M. L., Mongelli, G., Tateo, F. & Summa, V. (2012). J. X-ray Sci. Technol. 20, 175–186.  Web of Science CAS PubMed Google Scholar
First citationGiannossi, M. L., Summa, V. & Mongelli, G. (2013). J. Trace Elem. Med. Biol. 27, 91–97.  Web of Science CrossRef CAS PubMed Google Scholar
First citationGillen, D. L., Worcester, E. M. & Coe, F. L. (2005). Kidney Int. 67, 685–690.  Web of Science CrossRef PubMed Google Scholar
First citationGrases, F., Costa-Bauzá, A. & García-Ferragut, L. (1998). Adv. Colloid Interface Sci. 74, 169–194.  Web of Science CrossRef CAS PubMed Google Scholar
First citationGrases, F., Costa-Bauzá, A., Königsberger, E. & Königsberger, L.-C. (2000). Int. Urol. Nephrol. 32, 19–27.  CrossRef PubMed CAS Google Scholar
First citationGražulis, S., Chateigner, D., Downs, R. T., Yokochi, A. F. T., Quirós, M., Lutterotti, L., Manakova, E., Butkus, J., Moeck, P. & Le Bail, A. (2009). J. Appl. Cryst. 42, 726–729.  Web of Science CrossRef IUCr Journals Google Scholar
First citationHe, J.-Y., Deng, S.-P. & Ouyang, J.-M. (2010). IEEE Trans. Nanobiosci. 9, 156–163.  Google Scholar
First citationHess, B. (1990). Urol. Res. 18, S45–S48.  CrossRef PubMed Web of Science Google Scholar
First citationHesse, A., Brändle, E., Wilbert, D., Köhrmann, K.-U. & Alken, P. (2003). Eur. Urol. 44, 709–713.  Web of Science CrossRef PubMed CAS Google Scholar
First citationHill, R. J. & Howard, C. J. (1987). J. Appl. Cryst. 20, 467–474.  CrossRef CAS Web of Science IUCr Journals Google Scholar
First citationHossain, R. Z., Ogawa, Y., Hokama, S., Morozumi, M. & Hatano, T. (2003). Int. J. Urol. 10, 411–415.  Web of Science CrossRef PubMed Google Scholar
First citationIzatulina, A., Gurzhiy, V. & Frank-Kamenetskaya, O. (2014). Am. Mineral. 99, 2–7.  Web of Science CrossRef ICSD CAS Google Scholar
First citationIzatulina, A. R., Gurzhiy, V. V., Krzhizhanovskaya, M. G., Kuz'mina, M. A., Leoni, M. & Frank-Kamenetskaya, O. V. (2018). Cryst. Growth Des. 18, 5465–5478.  Web of Science CSD CrossRef CAS Google Scholar
First citationIzatulina, A. R. & Yelnikov, V. Y. (2008). Minerals as Advanced Materials I, edited by S. V. Krivovichev, pp. 231–239. Heidelberg: Springer.  Google Scholar
First citationJamal, A. & Ramzan, A. (2004). J. Coll. Phys. Surg. Pak. 14, 411–415.  Google Scholar
First citationKartha, G., Calle, J. C., Marchini, G. S. & Monga, M. (2012). Urol. Clin. North Am. 40, 135–147.  Web of Science CrossRef PubMed Google Scholar
First citationKeddis, M. T. & Rule, A. D. (2013). Curr. Opin. Nephrol. Hypertens. 22, 390–396.  Web of Science CrossRef PubMed Google Scholar
First citationKeshavarzi, B., Yavar Ashayeri, N., Moore, F., Irani, D., Asadi, S., Zarasvandi, A. & Salari, M. (2016). Minerals, 6, 131.  Web of Science CrossRef Google Scholar
First citationKhan, S. R. & Kok, D. J. (2004). Front. Biosci. 9, 1450–1482.  Web of Science CrossRef PubMed CAS Google Scholar
First citationKhan, S. R., Pearle, M. S., Robertson, W. G., Gambaro, G., Canales, B. K., Doizi, S., Traxer, O. & Tiselius, H.-G. (2016). Nat. Rev. Dis. Primers, 2, 1–23.  Web of Science CrossRef Google Scholar
First citationKlee, L. W., Brito, C. G. & Lingeman, J. E. (1991). J. Urol. 145, 715–718.  CrossRef CAS PubMed Web of Science Google Scholar
First citationKok, D. J., Papapoulos, S. E. & Bijvoet, O. L. M. (1990). Kidney Int. 37, 51–56.  CrossRef CAS PubMed Web of Science Google Scholar
First citationKourambas, J., Aslan, P., Teh, C. L., Mathias, B. J. & Preminger, G. M. (2001). J. Endourol. 15, 181–186.  Web of Science CrossRef PubMed CAS Google Scholar
First citationKristensen, C., Parks, J. H., Lindheimer, M. & Coe, F. L. (1987). Kidney Int. 32, 749–753.  CrossRef CAS PubMed Web of Science Google Scholar
First citationKuruma, H., Arakawa, T., Kubo, S., Hyodo, T., Matsumoto, K., Satoh, T., Egawa, S. & Baba, S. (2006). Int. J. Urol. 13, 498–501.  Web of Science CrossRef PubMed CAS Google Scholar
First citationLe Bail, A., Cranswick, L. M. D. & Madsen, I. (2008). Powder Diffraction: Theory and Practice. Cambridge: Royal Society of Chemistry.  Google Scholar
First citationLieske, J. C., de la Vega, L. S. P., Gettman, M. T., Slezak, J. M., Bergstralh, E. J., Melton, L. J. III & Leibson, C. L. (2006). Am. J. Kidney Dis. 48, 897–904.  Web of Science CrossRef PubMed Google Scholar
First citationLiu, Y., Chen, Y., Liao, B., Luo, D., Wang, K., Li, H. & Zeng, G. (2018). Asia. J. Urol. 5, 205–214.  Web of Science CrossRef Google Scholar
First citationLomas, D. J., Jaeger, C. D. & Krambeck, A. E. (2017). Urology, 102, 43–47.  Web of Science CrossRef PubMed Google Scholar
First citationLondoño-Restrepo, S. M., Jeronimo-Cruz, R., Millán-Malo, B. M., Rivera-Muñoz, E. M. & Rodriguez-García, M. E. (2019). Sci. Rep. 9, 5915.  Web of Science PubMed Google Scholar
First citationLutterotti, L., Matthies, S. & Wenk, H. R. (1999). IUCr Commission on Powder Diffraction Newsletter, No. 21, pp. 14–15.  Google Scholar
First citationMa, R., Luo, X., Li, Q. & Zhong, H. (2017). Int. J. Surg. 41, 150–161.  Web of Science CrossRef PubMed Google Scholar
First citationMadore, F., Stampfer, M. J., Rimm, E. B. & Curhan, G. C. (1998). Am. J. Hypertens. 11, 46–53.  Web of Science CrossRef CAS PubMed Google Scholar
First citationMaurice-Estepa, L., Levillain, P., Lacour, B. & Daudon, M. (1999). Scand. J. Urol. Nephrol. 33, 299–305.  Web of Science PubMed CAS Google Scholar
First citationMirković, M., Dosen, A., Erić, S., Vulić, P., Matović, B. & Rosić, A. (2020). Microchem. J. 152, 104429.  Google Scholar
First citationMoe, O. W. (2006). Lancet, 367, 333–344.  Web of Science CrossRef PubMed CAS Google Scholar
First citationMosli, H. A., Mosli, H. H. & Kamal, W. K. (2013). Res. Rep. Urol. 5, 11–15.  PubMed Google Scholar
First citationMukherjee, A. K. (2014). J. Indian Inst. Sci. 94, 35–44.  CAS Google Scholar
First citationNevo, A., Shahait, M., Shah, A., Jackman, S. & Averch, T. (2019). Int. Urol. Nephrol. 51, 585–591.  Web of Science CrossRef PubMed Google Scholar
First citationPak, C. Y. C. (1998). Lancet, 351, 1797–1801.  Web of Science CrossRef CAS PubMed Google Scholar
First citationPfau, A. & Knauf, F. (2016). Am. J. Kidney Dis. 68, 973–985.  Web of Science CrossRef PubMed Google Scholar
First citationPrezioso, D., Illiano, E., Piccinocchi, G., Cricelli, C., Piccinocchi, R., Saita, A., Micheli, C. & Trinchieri, A. (2014). Arch. Ital. Urol. Androl. 86, 99–102.  CrossRef PubMed Google Scholar
First citationPrywer, J. & Torzewska, A. (2010). Cryst. Res. Technol. 45, 1283–1289.  Web of Science CrossRef CAS Google Scholar
First citationRahman, N. U., Meng, M. V. & Stoller, M. L. (2003). Curr. Pharm. Des. 9, 975–981.  Web of Science CrossRef PubMed CAS Google Scholar
First citationRamaswamy, K., Killilea, D. W., Kapahi, P., Kahn, A. J., Chi, T. & Stoller, M. L. (2015). Nat. Rev. Urol. 12, 543–557.  Web of Science CrossRef CAS PubMed Google Scholar
First citationRamello, A., Vitale, C. & Marangella, M. (2001). J. Nephrol. 13, S45–S50.  Google Scholar
First citationRen, F., Lu, X. & Leng, Y. (2013). J. Mech. Behav. Biomed. Mater. 26, 59–67.  Web of Science CrossRef CAS PubMed Google Scholar
First citationRietveld, H. M. (1967). Acta Cryst. 22, 151–152.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationRietveld, H. M. (1969). J. Appl. Cryst. 2, 65–71.  CrossRef CAS IUCr Journals Web of Science Google Scholar
First citationRingertz, H. (1966). Acta Cryst. 20, 397–403.  CSD CrossRef ICSD CAS IUCr Journals Web of Science Google Scholar
First citationRodgers, A. L. (2017). Urolithiasis, 45, 27–32.  Web of Science CrossRef PubMed Google Scholar
First citationRodgers, A., Trinchieri, A., Ather, M. H. & Buchholz, N. (2019). Urolithiasis, 47, 401–413.  Web of Science CrossRef CAS PubMed Google Scholar
First citationRule, A. D., Krambeck, A. E. & Lieske, J. C. (2011). Clin. J. Am. Soc. Nephrol. 6, 2069–2075.  Web of Science CrossRef PubMed Google Scholar
First citationSakhaee, K., Adams-Huet, B., Moe, O. W. & Pak, C. Y. C. (2002). Kidney Int. 62, 971–979.  Web of Science CrossRef PubMed CAS Google Scholar
First citationSchofield, P. F., Knight, K. S., Houwen, J. A. M. & Valsami-Jones, E. (2004). Phys. Chem. Miner. 31, 606–624.  Web of Science CrossRef ICSD CAS Google Scholar
First citationSchubert, G. (2006). Urol. Res. 34, 146–150.  Web of Science CrossRef PubMed Google Scholar
First citationShapur, N. K., Uvarov, V., Popov, I., Katz, R., Gofrit, O. N., Landau, E. H., Pode, D. & Duvdevani, M. (2012). Urology, 80, 980–985.  Web of Science CrossRef PubMed Google Scholar
First citationSharma, A. P. & Filler, G. (2010). Indian J. Urol. 26, 516.  CrossRef PubMed Google Scholar
First citationShekarriz, B. & Stoller, M. L. (2002). J. Urol. 168, 1307–1314.  Web of Science CrossRef PubMed CAS Google Scholar
First citationSingh, V. K. & Rai, P. K. (2014). Biophys. Rev. 6, 291–310.  CrossRef CAS PubMed Google Scholar
First citationSoble, J. J., Hamilton, B. D. & Streem, S. B. (1999). J. Urol. 161, 869–873.  Web of Science CrossRef PubMed CAS Google Scholar
First citationStamatelou, K. K., Francis, M. E., Jones, C. A., Nyberg, L. M. Jr & Curhan, G. C. (2003). Kidney Int. 63, 1817–1823.  Web of Science CrossRef PubMed Google Scholar
First citationSudarsanan, K. & Young, R. A. (1969). Acta Cryst. B25, 1534–1543.  CrossRef ICSD IUCr Journals Web of Science Google Scholar
First citationSun, L.-M., Lin, C.-L., Chang, Y.-J., Liang, J.-A., Liu, S.-H., Sung, F.-C. & Kao, C.-H. (2013). BJU Int. 112, 1150–1155.  Web of Science CrossRef PubMed Google Scholar
First citationSun, X.-Y., Chen, J.-Y., Rao, C.-Y. & Ouyang, J.-M. (2020). Int. J. Nanomed. 15, 5043–5060.  Web of Science CrossRef CAS Google Scholar
First citationSun, X. Y., Gan, Q. Z. & Ouyang, J. M. (2015). Cell Death Discovery, 1, 1–8.  Web of Science CrossRef Google Scholar
First citationSun, X.-Y., Ouyang, J.-M., Zhu, W.-Y., Li, Y.-B. & Gan, Q.-Z. (2015). J. Mater. Chem. B, 3, 1864–1878.  Web of Science CrossRef CAS PubMed Google Scholar
First citationTang, X. & Lieske, J. C. (2014). Curr. Opin. Nephrol. Hypertens. 23, 385–390.  Web of Science CrossRef CAS PubMed Google Scholar
First citationTaylor, E. N., Stampfer, M. J. & Curhan, G. C. (2005). JAMA, 293, 455–462.  Web of Science CrossRef PubMed CAS Google Scholar
First citationTazzoli, V. & Domeneghetti, C. (1980). Am. Mineral. 65, 327–334.  CAS Google Scholar
First citationTiselius, H.-G. (2000). Braz. J. Urol. 26, 452–462.  Google Scholar
First citationTurk, C., Neisius, A., Petřík, A., Seitz, C., Thomas, K. & Skolarikos, A. (2020). European Association of Urology Guidelines, 2020 ed., presented at the EAU Annual Congress, Amsterdam, 2020. https://uroweb.org/guideline/urolithiasis/. Arnhem: The European Association of Urology Guidelines Office.  Google Scholar
First citationUvarov, V., Popov, I., Shapur, N., Abdin, T., Gofrit, O. N., Pode, D. & Duvdevani, M. (2011). Environ. Geochem. Health, 33, 613–622.  Web of Science CrossRef CAS PubMed Google Scholar
First citationWhitaker, A. & Jeffery, J. W. (1970). Acta Cryst. B26, 1429–1440.  CrossRef ICSD CAS IUCr Journals Web of Science Google Scholar
First citationWilliams, R. E. (1963). Br. J. Urol. 35, 416–437.  CrossRef PubMed CAS Google Scholar
First citationYoshida, O., Terai, A., Ohkawa, T. & Okada, Y. (1999). Kidney Int. 56, 1899–1904.  Web of Science CrossRef PubMed CAS Google Scholar

This article is published by the International Union of Crystallography. Prior permission is not required to reproduce short quotations, tables and figures from this article, provided the original authors and source are cited. For more information, click here.

Journal logoJOURNAL OF
APPLIED
CRYSTALLOGRAPHY
ISSN: 1600-5767
Follow J. Appl. Cryst.
Sign up for e-alerts
Follow J. Appl. Cryst. on Twitter
Follow us on facebook
Sign up for RSS feeds