Vous êtes sur la page 1sur 8

Materials Research Bulletin, Vol. 34, Nos. 12/13, pp.

1959 1966, 1999 Copyright 2000 Elsevier Science Ltd Printed in the USA. All rights reserved 0025-5408/99/$see front matter

PII S0025-5408(99)00206-8

SYNTHESIS AND CHARACTERIZATION OF -HOPEITE, Zn3(PO4)2 4H2O

O. Pawlig* and R. Trettin Johannes Gutenberg Universitat, Institut fur Geowissenschaften (Angewandte und Technische Mineralogie), D-55099 Mainz, Germany (Refereed) (Received November 30, 1998; Accepted December 28, 1998)

ABSTRACT Orthorhombic -Zn3(PO4)2 4H2O, -hopeite, was synthesized from an aqueous solution of zinc acetate and orthophosphoric acid. The synthesis from these starting materials yielded best results with respect to purity and crystallinity. Structural characteristics of the compound were investigated by X-ray diffraction (XRD), Fourier transform infrared (FT-IR), scanning electron microscopy (SEM), and thermogravimetric/differential thermal analysis (TG/DTA). Supplemental data came from surface area, particle size distribution, and density measurements. The obtained compound proved to be highly crystalline and stoichiometric. Dehydration of -hopeite is accompanied by at least two endothermic effects and already starts at temperatures 373 K. This observation has important implications for the technical zinc phosphate coating process. The dehydration behavior can be explained by results from low-temperature diffuse reectance IR spectroscopy (DRIFT). 2000 Elsevier
Science Ltd

KEYWORDS: A. inorganic compounds, B. chemical synthesis, C. thermogravimetric analysis, C. infrared spectroscopy, D. crystal structure INTRODUCTION Zinc phosphate coatings on steel consist mainly of crystalline -hopeite, orthorhombic -Zn3(PO4)2 4H2O [1 4]. When the substrate is pretreated by phosphating or other conversion coating processes, organic coatings (e.g., polyurethane) on metal products are improved.

*To whom correspondence should be addressed. 1959

1960

O. PAWLIG and R. TRETTIN

Vol. 34, Nos. 12/13

Phosphating leaves the surface clean and microrough, which improves the adhesion of these coatings to steel by increasing the number of sites for mechanical locking of the topcoating to the phosphate crystals [59]. Further, the application of the phosphate coatings facilitates cold formation of metals [10]. Among various phosphate coatings of technical interest, zinc phosphate coatings are desirable for superior performance in retarding the lateral creep of corrosion between the organic topcoating and the metal. By potentiodynamic current and capacity measurements as well as by impedance spectroscopy, it can be shown that these coatings exhibit electrical insulating properties. For zinc phosphate coatings, after oil posttreatment, the breakdown voltage is more than 700 V [11]. Thus, zinc phosphate coatings have found widespread application in electric motors and transformers and in the automotive industry. In addition to its application in important metal products, -hopeite is the crystalline reaction product of biomedical zinc phosphate cement, which is used to attach prefabricated crowns and inlays to teeth, to line cavities for the protection of dental pulp against chemical attack and thermal stresses, and even to ll cavities. Crystals of -hopeite grow on the outermost surface of the cement, in direct contact to human dentine and enamel [1214]. In nature, Zn3(PO4)2 4H2O exists in two structures, orthorhombic hopeite and parahopeite, its triclinic polymorph. Although for a long time most studies of the P2O5ZnOH2O system supported the existence of only two orthorhombic modications, - and -hopeite [1517], in newer studies [18,19] evidence of a third modication, -hopeite, was found from isoperibol solution-reaction calorimetry. The structure of -hopeite is dominated by the presence of puckered sheets of corner-sharing tetrahedra of O atoms perpendicular to the b axis, separated by sheets of face-sharing O octahedra. The Zn atoms are divided in the ratio 2:1 between one-half of the tetrahedral sites (ZnO4) and the octahedral sites (ZnO2(H2O)4). The two O atoms of the ZnO2(H2O)4 octahedra are arranged in a cis-conguration. The remaining tetrahedral positions contain P [20]. From X-ray diffraction, it is known that the nonhydrogen atom positions of the various modications of hopeite are essentially identical, whereas the hydrogen atoms are arranged in different positions within the crystal lattice. A distinction between these modications is possible by means of thermal analysis and infrared spectroscopy. There is need for synthesis of -hopeite as a standard material for the analysis of phosphate coatings and biomedical cements in quality control and industrial research or development. However, problems arise because the seemingly trivial synthesis from plain zinc oxide and orthophosphoric acid leads to the formation of a mixture of , -, and -hopeite in many cases [16,18,19]. The conventional synthesis method of hopeite consists in its preparation from an aqueous solution using zinc sulfate and sodium dihydrogen phosphate according to the following reaction: 3ZnSO 4 7H 2O 2Na 2SO 4 2Na 2HPO 4 2H 2O 3 Zn 3(PO 4) 2 4H 2O 21H 2O H 2SO 4

This method is time-consuming and suffers from a low efciency of reaction. It has been reported [21] that in most cases a maximum of about 60% efciency could not be exceeded after 48 h. Further, it is known from the literature [15] that this synthesis may lead to a mixture of - and -hopeite.

Vol. 34, Nos. 12/13

ZINC PHOSPHATE HYDRATE

1961

EXPERIMENTAL Synthesis. Deionized water and stoichiometric amounts of high purity zinc acetate and 85 wt% orthophosphoric acid (both ACS grade, Fluka, Germany) were used to prepare the required aqueous solution: 3Zn(CH 3COO) 2 2H 2O 2H 2O The pH was adjusted to 4.0 by the addition of 1 M KOH solution (prepared from ACS grade KOH pellets, Fluka, Germany), and the solution was heated to 323 K for 2 h under magnetic stirring. The mixture was cooled and ltered through a 0.22 m millipore lter. The white precipitate was washed 4 times, twice with the acid solution and twice with deionized water, and subsequently dried in an oven at 303 K. Approximately 4 h were needed for a complete synthesis procedure. Characterization. XRD spectra were recorded using a Seiffert XRD 3000 TT BraggBrentano diffractometer operating with Cu K radiation at 40kV/30mA. Automatic divergence slit, step size of 0.01, and a time per step of 2 s were used. Infrared transmission spectra were recorded with Perkin-Elmer equipment (model 1760) from 400 to 4000 cm 1 and 4 cm 1 resolution using KBr pellets. Low-temperature diffuse reectance IR (DRIFT) spectra in the range 2900 to 3700 cm 1 were recorded using a Bruker infrared microscope connected to a Bruker IFS 66. The spectrometer was equipped with a nitrogen-cooled mercury cadmium telluride (MCT) detector. Spectra were recorded at 2 cm 1 resolution and the sample cell was purged with nitrogen. TG/DTA experiments were performed on a Setaram TG-DTA 92-16 thermobalance, in a Pt crucible at heating rates of 10, 5, and 1 K/min, from 298 to 673 K under argon atmosphere. SEM pictures were taken on a Zeiss DSM 962. BET surface area measurements were performed using an ASAP 2010 Micromeritics instrument. Nitrogen was used as an analysis adsorptive. Before each measurement, the sample was outgazed for 6 h at 313 K. Grain size distributions were measured on a Galai CIS 1 particle size analyzer. Density measurements were performed with a Micromeritics AccuPyc 1330 Pycnometer by measuring the amount of displaced gas. RESULTS AND DISCUSSION The synthesis of -hopeite was tried from various starting materials (zinc carbonate hydroxide, zinc chloride, zinc citrate, zinc nitrate, and zinc acetate). An aqueous solution of zinc acetate and orthophosphoric acid yielded best results with respect to purity and crystallinity. When the pH of the mixture containing zinc acetate and orthophosphoric acid was adjusted to 4.0 using 1 M KOH, the efciency of the reaction was between 78 and 91%. A typical XRD pattern of the product obtained from the reaction between these starting materials is shown in Figure 1. The six strongest peaks were indexed. The bimodal grain size distribution of the platelike crystals (Fig. 2) shows two maxima at 5.07 and 31.42 m. Peak intensities in the X-ray diffraction spectrum and the measured BET surface area of 0.6 m2g 1 indicate a high crystallinity. The X-ray diffraction pattern matches those reported for -hopeite [12]. A renement of the lattice constants using the program TREOR [22] yielded a 10.591(3), b 18.312(4), and c 5.027(2) . 2H 3PO 4 3 Zn 3(PO 4) 2 4H 2O 6CH 3COOH

1962

O. PAWLIG and R. TRETTIN

Vol. 34, Nos. 12/13

FIG. 1 X-ray diffraction pattern of synthetic -hopeite.

In the literature, a certain level of nonstoichiometry of -hopeite is suggested from X-ray renement. It has been proposed that nonstoichiometry affects natural as well as synthetic material prepared in the laboratory [20]. Moreover, an incorporation of manganese and nickel from trication phosphating bath into the phosphate layer has been shown by XPS/EDX measurements, and it has been concluded that these cations may occupy vacant zinc sites

FIG. 2 Scanning electron micrograph of synthetic -hopeite.

Vol. 34, Nos. 12/13

ZINC PHOSPHATE HYDRATE

1963

FIG. 3 FT-IR transmission spectrum of synthetic -hopeite.

[23]. In our study, density of -hopeite was calculated from rened lattice constants to be 3.122(4) g cm 3. Since the measured density is 3.128(5) g cm 3, there is no indication for nonstoichiometry in our case. Wavenumbers of bands in the infrared spectrum (Fig. 3) were taken from contrast enhancing D2 derivative spectra. The spectra show characteristic absorptions due to the PO3 group ( 1: 952 and 929 cm 1; 2: 472, 455, 431, 409, and 388 cm 1; 3: 1156, 1130, 4 1105, 1066, 1026, and 1000 cm 1, 4: 635, 600, 576, 561, 524, and 502 cm 1). In addition, water bending (1639 cm 1) and OH stretching (3546 and a broad band centered at 3300 3400 cm 1) vibrations can be observed. Their positions are consistent with the presence of a hydrogen-bonding network within the crystal lattice. However, the sharp band at 3546 cm 1 indicates that some OH groups do not participate in hydrogen bonding. The numerous bands in the phosphate region can be explained by the low C1 site symmetry of the phosphate ion in the hopeite structure. Thus, the full set of nine infrared modes becomes active. Additional bands arise from correlation eld splitting. The DTA curves in Figure 4 show the thermal dehydration of -hopeite. When a heating rate of 5 or 10 K/min was used, the dehydration was accompanied by three endothermic effects, in keeping with the literature reports [1517,20]. The onset temperatures were 374, 411, and 542 K for a heating rate of 10 K/min and 363, 410, and 535 K for a heating rate of 5 K/min. It is well known that -hopeite loses all structural water at once at 421 K [15,16]. The dehydration pattern of -hopeite, however, is not yet reported in the literature. According to the mass-loss curves (Fig. 5), heating to 433 K was accompanied by the removal of two water molecules, indicating that Zn3(PO4)2 2H2O is established at this temperature. The third stage of the dehydration gave an anhydrous Zn3(PO4)2 phase. When a heating rate of 1 K/min was used, the dehydration pattern changed (Fig. 4). A rst broad endothermic peak was located at 330 K, followed by a less intense peak at 347 K. These endothermic effects are due

1964

O. PAWLIG and R. TRETTIN

Vol. 34, Nos. 12/13

FIG. 4 DTA analysis curves of synthetic -hopeite using different heating rates.

to adsorptive water in the precipitate. Compared to the DTA measurements performed with heating rates of 10 and 5 K/min, the rst dehydration peak of structural water shows an onset temperature of 354 K. From the mass-loss curve (Fig. 5), it was calculated that Zn3(PO4)2

FIG. 5 TG analysis curves of synthetic -hopeite using different heating rates.

Vol. 34, Nos. 12/13

ZINC PHOSPHATE HYDRATE

1965

FIG. 6 Diffuse reectance infrared spectra (DRIFT) of synthetic -hopeite at room temperature and upon cooling. 2H2O was already formed at 379 K, using a heating rate of 1 K/min. Complete dehydration of this phase occurred at 538 K under these conditions. The loss of water molecules from -hopeite at temperatures 373 K has some important implications for the technical zinc phosphate coating process. From XPS analysis it is well known that Zn3(PO4)2 2H2O is formed on the surface of phosphated sheets, whereas the bulk composition consists of -hopeite, as shown by XRD analysis [23,24]. The formation of the dihydrate phase is undesirable due to insufcient anticorrosive properties and a lower breakdown voltage [23,24]. This observation was not well understood, since the dehydration of -hopeite was said to occur above 373 K, to yield a dihydrate structure by loss of two H2O molecules. According to our study, the dehydration of -hopeite already starts at about 354 K; therefore, it may be possible that the dehydration occurs in the outermost layer due to the procedure of hot-air drying after phosphating, producing an amorphous dihydrate structure. In an attempt to assign the DTA dehydration peaks of -hopeite to energetically different bound water, we performed a low-temperature diffuse reectance IR (DRIFT) spectroscopic investigation on the H2O stretching region (Fig. 6). At 298 K this region is dominated by a broad featureless band centered at 3300 3400 cm 1. This result is consistent with the presence of a large variety of structurally unique H2O groups, each with different stretching frequencies and different hydrogen bond properties. Upon cooling to 123 K, the broad band splits into two groups of water stretching frequencies, which nally become resolved into seven bands (3531, 3489, 3442, 3344, 3292, 3172, and 3072 cm 1) arranged in two main groups at 93 K. For heating rates of 10 and 5 K/min, the rst and second endothermic peaks in the DTA curve can be assigned to the group of weaker bonded or free OH (Zn2 OH2 OH2), giving rise to higher stretching frequencies (32923531 cm 1); whereas, the third endothermic peak correlates with the group of stretching frequencies at lower wavenumbers

1966

O. PAWLIG and R. TRETTIN

Vol. 34, Nos. 12/13

(30723172 cm 1), indicating stronger bonded OH (Zn2 OH2OPO3). When a heating rate of 1 K/min was used, only two endothermic peaks, due to dehydration of structural water, appeared in the DTA curve. The assignment to the mentioned groups of OH stretching frequencies is obvious. Thus, low-temperature diffuse reectance IR spectroscopical (DRIFT) and thermoanalytical results are consistent in the case of synthetic -hopeite. ACKNOWLEDGMENT The authors thank the Stiftung Rheinland-Pfalz fur Innovation for nancial support of this work. REFERENCES
1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. T. Sugama and T. Takahashi, J. Mater. Sci. 30, 809 (1995). U.B. Nair and M. Subbaiyan, J. Mater. Sci. 30, 2108 (1995). K. Molt, M. Pohl, R. Seidel, and B. Mayer, Mikrochim. Acta 116, 101 (1994). K. Molt, D. Behmer, and M. Pohl, Fresenius J. Anal. Chem. 358, 36 (1997). G. Lorin, Phosphating of Metals: Constitution, Physical Chemistry and Technical Applications of Phosphating Solutions (Engl. transl. by F.H. Reid), Finishing Publications Ltd., London (1974). D.B. Freeman, Phosphating and Metal Pre-treatment: A Guide to Modern Processes and Practice, Woodhead-Faulkner, London (1986). W. Rausch, The Phosphating of Metals, ASM International, Metals Park, OH (1990). N.C. Debnath and P.K. Roy, Trans. IMF 74, 17 (1996). W.F. Heung, P.C. Wong, K.A. Mitchell, and T. Foster, J. Mater. Sci. Lett. 14, 1461 (1995). V. Burokas, A. Martusiene, and G. Bikulcius, Surf. Coat. Technol. 102, 233 (1998). D. Weng, P. Jokiel, A. Uebleis, and H. Boehni, Surf. Coat. Technol. 88, 147 (1996). J. Margerit, B. Cluzel, J.M. Leloup, J. Nurit, B. Pauvert, and A. Terol, J. Mater. Sci.Mater. Med. 7, 623 (1996). C.K. Park, M.R. Silsbee, and D.M. Roy, Cem. Concr. Res. 28, 141 (1998). A.D. Wilson and J.W. Nicholson, Acid-Base Reaction Cements; Their Biomedical and Industrial Applications, Cambridge University Press, New York (1993). M.V. Goloshapov and T.N. Filatova, Russ. J. Inorg. Chem. 14, 424 (1969). E.A. Nikonenko, I.I. Olikov, I.N. Marenkova, L.N. Margolin, and L.A. Reznikova, Russ. J. Inorg. Chem. 30, 25 (1985). E.A. Nikonenko and I.N. Marenkova, Russ. J. Inorg. Chem. 31, 397 (1986). H.M.A. Al-Maydama, P.J. Gardner, and I.W. McAra, Thermochim. Acta 194, 117 (1992). H.M.A. Al-Maydama, P.J. Gardner, and I.W. McAra, Thermochim. Acta 196, 117 (1992). R.J. Hill and J.B. Jones, Am. Mineral. 61, 987 (1976). M. Pohl, Doctoral thesis, University of Duisburg, Germany (1992). P.E. Werner, L. Erikson, and M. Westdahl, J. Appl. Cryst. 18, 367 (1985). E. Klusmann, U. Konig, and J.W. Schultze, Mater. Corros. 46, 83 (1995). S. Maeda and M. Yamamoto, Progr. Org. Coat. 33, 83 (1998).

Vous aimerez peut-être aussi