Sie sind auf Seite 1von 10

This paper was published in Proceedings of 11th International Congress on Deterioration and Conservation of Stone, eds. J.W.

ukaszewicz, P. Niemcewicz, Wydawnictwo Naukowe Uniwersytetu Mikoaja Kopernika, 2008, 931-938

CONTROL OF THE POROSITY STRUCTURE TO PRODUCE ROMAN CEMENT MORTARS COMPATIBLE WITH THE HISTORIC SUBSTRATE
A. Klisiska-Kopacz1*, R. Tilova2, G. Adamski1, R. Kozowski1
1

Institute of Catalysis and Surface Chemistry, Polish Academy of Sciences, ul. Niezapominajek 8, 30-239 Krakw, Poland

Faculty of Restoration, University of Pardubice, Jiraskova 3, 57001 Litomysl, The Czech Republic

* Corresponding author. E-mail address: ncklisin@cyf-kr.edu.pl (A. Klisiska-Kopacz)

Keywords: Roman cement, mercury porosimetry, hydration, mixture proportioning.

Abstract Two principal categories of pores were distinguished in the Roman cement historic mortars. The finest pores with the diameter below 0.2 m are present within the hardened aged Roman cement matrix. The larger air-pores with the diameters between 0.2 4 m are due to evaporation of water during the hydration process and their volume may be indicative of the w/c ratio used originally. Further, formation of the air pores was enhanced at the surface of the stucco elements exposed to the external environment favouring drying. In the repair mortars, unimodal distribution of pore sizes was observed for low w/c ratios with the threshold pore width decreasing considerably with increasing curing time. With an increase in the w/c ratios a development of a bimodal pore structure was noticed. The formation of larger pores is the result of evaporation of water unbound during the hydration. In this way, Roman cement repair mortars with controlled porosity and strength can be easily obtained by changing the w/c ratio during the preparation process.

1.

Introduction Highly hydraulic binders, known as natural or Roman cements, were key materials for the

economic and easy manufacture of stuccoes for the exterior of buildings during the nineteenth and early twentieth centuries. Roman cements were produced by burning naturally occurring deposits of calcium carbonate rich in clay minerals and grounding the burnt stones to the required fineness. They were distinguished from other hydraulic binders principally by a very fast setting time, agreeable texture and warm yellow-to-brown colour. During the twentieth century Roman cements disappeared from use, displaced by the newer Portland cement, which dominated the market. The lack of appropriate binding materials, matching those available in the nineteenth century, has therefore deprived architects and conservators of the original historic technology for the repair and conservation of these historic renders. Recently, the ROCEM project supported by the European Commission 5th Framework Programme, has extensively investigated historic renders based on Roman cements and has re-established this historic material and technology to the conservation practice (Weber et al. 2007, ROCEM 2008). In the framework of restoration process, an adequate choice of repair mortars is critical so that compatibility between them and the original components of the masonry is assured (Delgado Rodrigues and Grossi 2007). The compatibility can be broadly defined as the capacity of the repair mortar to interact with the original historic substrate without inducing any decay. The pore structure of the mortars is one of the main properties responsible for their compatibility as it greatly influences permeability to water and water vapour, and strength. The mismatch in the pore structure at the substrate-mortar interface may lead to a zone of damage due to ice or salt crystallization affecting the adhesive strength of a repair, and thus compromising its final compatibility. Therefore, a comparative evaluation of porosity structures was carried out for historic Roman cement mortars collected from buildings across Europe and for repair mortars formulated within the ROCEM project. The previous fundamental study of the pore structure of the Roman cement pastes, cured at the ideal moist-air conditions, have revealed a unimodal distribution of pore sizes with the threshold pore width decreasing considerably from up to 0.8 m at early ages to 0.02 m in the matured paste (Vyskocilova et al. 2007). The decrease reflects the evolution of early open pore structure produced by a quick growth of the calcium aluminate hydrate phases (C-A-H) to a more compact structure as the calcium silicate hydrates (C-S-H) gradually block the larger pores.

2.

Materials and methods

2.1 Materials The historic Roman cement mortars were collected from existing historic buildings, rendered and decorated with Roman cement mortars, which were located in various countries and were representative of different European areas in terms of materials, historic periods and geography. The description of selected historic samples studied in this work is given in Table 1. Historic mortars ranged from Wycombe Abbey (1804), an early example of the use of Roman cement in England not long after it was patented there in 1796, to Municipal Trade Academy in Krakow (1904-1906) built shortly before the outbreak of World War I after which the dominance of the newer Portland cement and modern functional architecture brought a quick decline in the production and use of Roman cement. The samples of mortars were representative of different modes of application, from cast ornaments to renders and hand-run elements applied in situ. The Roman cement burnt from the Folwark marl, Poland, was used to produce the repair materials. The characteristics of the original marl feedstock, oxide and mineralogical compositions of the cement, as well as strength and porosity development in its pastes on hydration were described in detail earlier (Hughes et al. 2008). The pastes and mortars were produced at water-to-cement (w/c) ratio varying from 0.6 to 1.0.

2.2 Methods The pore structure of the mortar samples was determined using a Poremaster mercury intrusion porosimeter from Quantachrome, allowing the study of pore sizes in the range 440 0.0035 m. For the measurements of the pore structure of newly prepared materials, the pastes were molded in prismatic 8x2x2cm casts. Next, they were demoulded immediately after setting and cured over distilled water i.e. near 100% relative humidity from 1 day to 3 months until tested. After the predetermined curing period, the specimens were immediately soaked in acetone for 24 hours to stop the hydration. They were placed in a rotary vacuum flask at 20oC for 4 hours to remove acetone and stored in a desiccator.

3.

Results and discussion

3.1 Analysis of original mortars The differential mercury intrusion curves for five selected historical mortars are presented in Figures 1 and 2. The curves show the incremental pore volume intruded as a function of pore diameter. The bimodal distribution of pore sizes was also observed for a significant number of historic mortars, such as ones shown in Figure 2. Such distribution pointed typically to a coexistence of fine pores with the threshold diameter below 0.1 m and of larger pores with the diameters between 0.8 2 m reflecting the presence of well reacted cement matrix (C-SH gel porosity) and of larger pores resulting from the evaporation of excess water not consumed during the process of hydration. The restricted hydration could be due to a number of processes: exposure of the freshly laid surface to dry external environments, high water-tocement ratio in the original mortars, or drawing of water from the stucco mass due to insufficient pre-wetting of the porous masonry or old mortars.

3.2 Pore structure of freshly-prepared Roman cement pastes Figures 4 and 5 compare the differential volume of intruded mercury as a function of pore diameter for the Roman cement pastes prepared at the w/c ratios varying from 0.6 to 1.0, studied at ages of 1 day and 3 months, respectively. The unimodal peak at 0.7 m is observed for the specimen prepared at the water-to-cement ratio of 0.6. With the increase in the water content, the initial peak shifts to higher pore diameters, 1.2 m and 1.9 m for the w/c ratios of 0.8 and 1.0, respectively. Further, the second porosity region at a pore diameter of 0.3 m can be noticed in all specimens, especially well visible for higher w/c ratios. For samples cured for 3 months, the threshold pore diameters are reduced to 0.01-0.03 m. This shift is the result of a progress of hydration and closing the broad pore diameters by the growth of the C-S-H gel. However, the mercury intrusion curves point to the presence of larger pores of 0.60 and 1.20 m in diameter, for samples prepared at w/c ratios of 0.8 and 1.0, respectively. The observation can be interpreted in terms of an insufficient amount of CS-H gel to fill large pores present in the structure due to an excess amount of water. The evolution of the pore structure of the Roman cement pastes is confirmed by the measurements of the total porosities. For all the samples, the total porosities were higher for

samples prepared at higher w/c ratios. The total porosities decreased with the increase of the hydration time (Fig. 5).

4.

Conclusions Mercury porosimetry has revealed two principal categories of pores in the freshly prepared

Roman cement repair materials. The finest pores with the pore diameter below 0.2 m were present within very well-hydrated mature Roman cement matrix. Larger pores with the diameters of around 1 m were characteristic of materials in which the hydration process was interrupted by the evaporation of water. The historic Roman cement mortars have rarely been found to develop a dense microstructure characteristic of the C-S-H gel abundantly formed under ideal condition of moist-curing of Roman cements. The presence of much larger air pores, a result of the evaporation or drawing of the excess unbound water and restricted hydration was in turn almost universally observed. The above considerations lead to a conclusion that Roman cement repair mortars are in broad terms compatible with the historic stuccoes. However, the porosity and strength of the Roman cement repair mortars may be easily controlled by changing the water-to-cement ratio during the preparation process.

5.

Acknowledgements A part of this research was carried out within the ROCEM project (contract EVK4-CT-

2002-00084), supported financially by the European Commission 5th Framework Programme, Thematic Priority: Environment and Sustainable Development, Key Action 4: City of Tomorrow and Cultural Heritage.

6.

References Delgado Rodrigues J., Grossi A. 2007. Indicators and ratings for the compatibility

assessment of conservation actions. Journal of Culture Heritage 8, 32-43. Hughes D.C., Jaglin D., Kozowski R., Mucha D. 2008. Roman cements - belite cements calcined at low temperature, Cement and Concrete Research, submitted for publication. ROCEM Roman cements to restore built heritage effectively,

http://www.heritage.xtd.pl/roman_cement/index.html (accessed 9 March 2008) Vyskocilova R., Schwarz W., Mucha D., Hughes D., Kozlowski R., Weber J. 2007. Hydration processes in pastes of Roman and natural American cements. Journal of the American Society for Testing Materials 4(2), Paper ID JAI100669. Weber J., Mayr N., Bayer K., Hughes D.C., Kozlowski R., Stillhammerova M., Ullrich D., Vyskocilova R. 2007. Roman Cement Mortars in Europes Architectural Heritage of the 19th Century. Journal of the American Society for Testing Materials 4(8), Paper ID JAI100667.

Table 1: Description of historic samples Sample symbol UK Location High Wycombe, Buckinghamshire, UK Hlboka Cesta 4, Bratislava, Slovakia Palackho 126, Brno, The Czech Republic Lehargasse 6-8, Vienna, Austria Kapucyska 2-4, Krakow, Poland Date Building Element

1804

Wycombe Abbey cornice run in situ

SK

1900 2nd half of 19th C.

school Semilasso Palace former warehouse

cornice run in situ

CZ

cast ornament

AT

1873

of the Court Theatres

cast ornament

PL

1904-1906

former Trade Academy

plain window frame run in situ

Figure 1: Differential volume of intruded mercury vs pore diameter for the historic Roman cement mortars with unimodal pore size distribution.

Figure 2: Differential volume of intruded mercury vs pore diameter for the historic Roman cement mortars with bimodal pore size distribution.

Figure 3: Differential volume of intruded mercury vs pore diameter for the pastes prepared at various w/c ratio changing from 0.6 to 1.0, cured for 1 day.

Figure 4: Differential volume of intruded mercury vs pore diameter for the pastes prepared at various w/c ratio changing from 0.6 to 1.0, cured for 3 months.

Figure 5: Correlation between total porosity and w/c ratio for Roman cement pastes at various ages: 1 day, 14 days and 3 months.

Das könnte Ihnen auch gefallen