Скачать книгу

are located in both pore space (Fig. 3a) and in the matrix (Fig. 3b,d) as well as on grain boundaries (Fig. 3c). Smectite has the ability to cause volume increase by swelling, especially when it is located on grain contacts. The alteration of feldspar to clay minerals often leads to the formation of intracrystalline porosity and therefore increases the effective porosity (Fig. 3c).

      The scanning electron microscopy (SEM) images (Fig. 3e–h) reveal very small mordenite needles and clinotilolite laths in CVE, which cause a high specific surface area (Fig. 3e–h).

      The zeolite-rich samples CVO, CVE and CAE show the highest hygroscopic water sorption and hydric expansion of all samples (see Kück et al. 2020a in this issue). During thermal expansion, the zeolite-rich samples (CVO, CVE and CAE) show a large difference between the first and the second heating cycle, with high residual strain (Fig. 4). After the first wet cycle the samples show a decrease in residual strain. CAO and CAE do not recover from contraction after the second drying cycle. The Mitla samples and QB are rather unaffected by both thermal and thermohyric dilatation (see Kück et al. 2020a).

      Figure 3: a–d: Different appearances of swellable clay minerals. a): smectite appears as a ‘spiderweb’ in the pore space as relict of weathered pumice clast in MG, b): kaolinite booklet and smectite in densely packed matrix, c): dissolved feldspar with clay minerals on the grain boundary reveals intracrystalline porosity, d): smectite in the matrix of CVO; e–h): Zeolites in CVE cause a high specific surface area. e): flaky zeolite in the matrix, f): zeolites grow into pore space, g): clinoptilolite laths in the matrix, h): 100 nm thin mordenite needles in a pore.

      123

      Figure 4: Residual strain [mm/m] of CVE in the X- and Z-direction for two heating cycles from 20–90 °C. Noticeable are intense shrinking and non-reversible thermal dilatation.

       Discussion and conclusions

      The cation exchange capacity (CEC) value depends on both clay minerals and zeolites, however, the CEC method used is not suitable for CEC analysis of zeolites because the large colored Cu-trien cations cannot enter the cavities of zeolites (Meier and Kahr 1999). For the analysed samples a clear trend between CEC and hydric expansion could not be observed (Fig. 5a). On the other hand, a linear dependency of hydric expansion and hygroscopic sorption on the specific surface area is observable (Fig. 5b, Fig. 6b). The specific surface area increases when clay minerals and zeolites are present in a volcanic tuff, because they yield a high share of micropores (often in the nanometer-scale) within the rock fabric. The higher the specific surface area is, the more water can interact with the rock and the expansion is resultingly larger. Zeolites can adsorb and release substantial amounts of water and therefore passively lead to hydric expansion, whenever water is released during dehydration.

      The results of this study show, that the presence of both zeolites and swellable clay minerals has a combined effect on the water-related weathering behavior of tuff building stones. The porespace properties (e. g. pore radii, specific surface area) and the water transport and storage properties 124of volcanic tuff rocks are significantly influenced, by either of them. Volcanic tuff rocks with a significant amount of swellable clay minerals as well as zeolites show extremely high hygroscopic water sorption and hydric swelling. Thermal effects like shrinkage and fracturing during drying are particularly high. This study showed that the identification of the clay mineral and zeolite content, as well as their location and shape within the rock fabric is an important measure for the prediction of the weathering behavior of tuff building stones. For the planning of effective conservational treatment of tuff building stones a clay mineral and zeolite analysis should be considered substantial.

      Figure 5: Correlation between a: hydric expansion [mm/m] and CEC [meq/100 g], b: hydric expansion [mm/m] and BET [m2/g].

      Figure 6: Correlation between a) hygroscopic water sorption at 95 % rh [wt.-%] and CEC [meq/100 g], b) hygroscopic water sorption at 95 % rh [wt.-%] and BET [m2/g].

       Acknowledgements

      This work was supported by the German Research Foundation (Si-438/52-1) and the German Federal Environmental Foundation (AZ20017/481). We would like to thank J. L. Noria Sánchez, A. E. Andrade Cuautle and M. Á. González for their support of the onsite work. For the laboratory support and helpful comments we thank K. Wemmer, J. Menningen, W. Wedekind and C. Gross.

       References

      Colella, C., Gennaro, M. d.’ and Aiello, R. (2001), Use of zeolitic tuff in the building industry, Reviews in Mineralogy and Geochemistry, Vol. 45 No. 1, pp. 551–587.

      Di Tchernev (1978), Solar energy application of natural zeolites, Natural zeolites: Occurrence, properties, use, Vol. 474, p. 485.

      Gonzalez, I. J. and Scherer, G. W. (2004), Effect of swelling inhibitors on the swelling and stress relaxation of clay bearing stones, Environmental Geology, Vol. 46 No. 3–4, pp. 364–377.

      Korkuna, O., Leboda, R., Skubiszewska-Zie, J., Vrublevska, T., Gunko, V. M. and Ryczkowski, J. (2006), Structural and physicochemical properties of natural zeolites: clinoptilolite and mordenite, Microporous and Mesoporous Materials, Vol. 87 No. 3, pp. 243–254.

      Kück, A., Pötzl, C., López-Doncel, R. A., Dohrmann, R. and Siegesmund, S. (2020a), Tuffs in pre-Columbian and colonial architecture of Oaxaca, Mexico, in Siegesmund, S. and Middendorf, B. (Eds.), Monument future: Decay and conservation of stone, Göttingen, Kassel, Mitteldeutscher Verlag, Halle.

      Pablo-Galán, L. de (1986), Geochemical trends in the alteration of Miocene vitric tuffs to economic zeolite deposits, Oaxaca, Mexico, Applied geochemistry, Vol. 1 No. 2, pp. 273–285.

      Pötzl, C., Dohrmann, R. and Siegesmund, S. (2018a), Clay swelling mechanism in tuff stones: an example of the Hilbersdorf Tuff from Chemnitz, Germany, Environmental earth sciences, Vol. 77 No. 5, p. 188.

      Pötzl, C., Siegesmund, S., Dohrmann, R., Koning, J. M. and Wedekind, W. (2018b), Deterioration of volcanic tuff rocks from Armenia. Constraints on salt crystallization and hydric expansion, Environmental earth sciences, Vol. 77 No. 19, p. 660.

      Ruedrich, J., Bartelsen, T., Dohrmann, R. and Siegesmund, S. (2011), Moisture expansion as a deterioration factor for sandstone used in buildings, Environmental earth sciences, Vol. 63 No. 7–8, pp. 1545–1564.

      Siegesmund, S. and Dürrast, H. (2011), Physical and mechanical properties of rocks, in Siegesmund, S. and Snethlage, R. (Eds.), Stone in architecture, Springer, pp. 97–225.

      Snethlage, R., Wendler, E. and Klemm, D. D. (1995), Tenside im Gesteinsschutz-bisherige Resultate mit einem neuen Konzept zur Erhaltung von Denkmälern aus Naturstein, Denkmalpflege und Naturwissenschaft-Natursteinkonservierung I. Verlag Ernst & Sohn, Berlin, pp. 127–146.

      Wedekind, W., López-Doncel, R., Dohrmann, R., Kocher, M. and Siegesmund, S. (2013), Weathering of volcanic tuff rocks caused by moisture expansion, Environmental earth sciences, Vol. 69 No. 4, pp. 1203–1224.

      125

       PHYSICO-CHEMICAL CHARACTERIZATION OF THE CARTAGENA WALL AND QUARRY MATERIAL STONE USED FOR ITS RESTORATION