Next Article in Journal
Green Synthesis of Pure Superparamagnetic Fe3O4 Nanoparticles Using Shewanella sp. in a Non-Growth Medium
Previous Article in Journal
The Effect of Suppliers’ Green and Traditional Selection Criteria in Supply Chain Management on Purchasing Firms’ Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Modelling the Mechanical Effect of Salt Weathering on Historical Sandstone Blocks through Microdrilling

by
Marco Ludovico-Marques
1,* and
Carlos Chastre
2
1
Polytechnic Institute of Setúbal, ESTBarreiro, RESILIENCE, 2839-001 Lavradio, Portugal
2
CERIS, Department of Civil Engineering, NOVA School of Science and Technology, Universidade NOVA de Lisboa, 2829-516 Caparica, Portugal
*
Author to whom correspondence should be addressed.
Sustainability 2024, 16(15), 6277; https://doi.org/10.3390/su16156277
Submission received: 31 May 2024 / Revised: 8 July 2024 / Accepted: 17 July 2024 / Published: 23 July 2024
(This article belongs to the Special Issue Durability Assessment of Stone-Built Heritage)

Abstract

:
The durability of sandstones of historical building materials and geoheritage landforms is a major issue that requires an assessment methodology to follow salt weathering evolution. The building blocks of monuments support decorative carvings and reliefs that are outstanding testimonies of human activity. An evaluation based on quasi- and non-destructive testing is a reliable and generally accepted way of testing and inspecting historical building materials. Compression tests were performed on specimens of similar building sandstones extracted close to those of from St. Leonard’s Middle Ages Church, and microdrilling tests were carried out on adequate blocks of this monument. The locations of the latter tests were determined using the results of low-pressure water absorption tests, which contributed to finding a link between the sandstone specimens and the building blocks of the monument. This innovative methodology was used to generate simulated stress–strain diagrams of the building blocks of this church based on drilling strength results, avoiding the cutting of specimens from the façades with the sizes needed to ensure the mechanical validity of the results. A good agreement between the predicted and experimental stress–strain curves was achieved. The stress–strain curves of sound stones from historical building blocks and of their weathered envelopes are shown. The evolution of weathering profiles can be followed through the analysis of stress–strain diagrams, allowing an assessment of structural stability, which is essential to the study of the durability of historical building sandstones. This innovative methodology allows the adequate conservation of monuments and is a contribution to the knowledge of sustainable cultural tourism.

1. Introduction

Several authors have reported on stone degradation patterns in historic buildings [1,2,3]. Crack and deformation, detachment, features induced by material loss, discoloration and deposits, and biological colonization are the five types of stone deterioration patterns referred to in the International Council on Monuments and Sites—International Scientific Committee for Stone (ICOMOS-ISCS) glossary [4].
Material loss at cross-sections, aside fracture and strain damage, is most responsible for decreasing the stability of structural elements such as building blocks in walls, vaults and columns. The weathering of stone reliefs or carved details is very important, and a honeycomb pattern is one of the more visible features, also induced by material loss, in historic buildings.
Some of the most important sandstone monuments in the world with striking reliefs and carvings were studied by Fitzner et al. [5], who focused on Nubian sandstones of the Karnak and Luxor temples in Tebas, quarried in Gebel Silsila (Upper Egypt); Heinrichs [6], whose study object encompassed Cambro–Ordovician sandstone formations in Petra tombs and temples; Wadeson [7], who reported on Nabatean façade tombs in Petra and Hegra; Heinrichs and Fitzner [8], who investigated sandstone stelae of the Nemrud Dag sanctuary in Turkey; Ma [9] and Qiao et al. [10], whose studies concerned the sandstones of the red cliff carved with the Image of Leshan Grand Buddah; Stein [11], who worked with sandstones of the Kanchipuram temples and Khajuraho temples in India; Singh [12], who focused on the weathering of the Kailasanatha Temple at Kanchipuram; Deshpande et al. [13], whose study pertained to reliefs on walls and column carvings on yellow blocks of sandstones of Jaisalmer Fort in Rajasthan in India; Uchida et al. [14], Hosono et al. [15], and Siedal et al. [16], who highlighted the reliefs on the surface of the grey to yellowish-brown sandstones in Phnom Krom monuments (Angkor region) and their detachment; and Ashmore [17], who studied carvings on sandstone major stelae and some masonry structures in the Mayan city of Quiriguá in Guatemala.
The structural stability of medium-scale landforms (Mignón [18]) that are geoheritage, such as sandstone arches, sometimes depends on the integrity of the rock masses remaining after weathering evolution. Significant weathering patterns related to material loss are rendering-like erosion features. Important testimonies are the following: Rainbow Bridge in southern Utah [19,20], sacred to Indigenous Americans in the USA; Delicate Arch [20,21]; Marlong Arch and Looking Glass Arch in Carnarvon National Park in Australia [22]; and the Arch of Afazadjar, south of Akakus Massif in Libya [23].
In order to avoid jeopardizing these artistic or natural details, which are striking testimonies of human or natural activity, the structural stability of their supporting blocks must be assessed and monitored. Monitoring the progress of weathering on sandstones is required to allow the long-lasting existence of reliefs and carvings shown on these building or rock mass blocks, which can only occur if the integrity of the latter is maintained.
Mechanical behaviour can be assessed by compression parameters obtained from stress–strain diagrams; however, minimally intrusive tests and quasi-NDTs (non-destructive tests) are required for the sake of monument integrity, and its weathering progression must be assessed by these tests. The following authors carried out studies comprising microdrilling tests and recorded salt weathering profiles on stone samples and rock mass specimens.
The HardRock research project [24] developed an apparatus to study weathering profiles on historical building stones and the effectiveness of consolidation treatments on impregnation depth by means of measuring the downhole drilling strength and time. Ludovico-Marques and Delgado-Rodrigues [25] performed sodium chloride ageing tests on samples of volcanic tuffs inside a salt mist chamber and assessed the salt weathering effect through drilling resistance profiles on these samples. They also calculated the drilling energy needed to drill the holes downward from the recorded values of the strength and displacement of the drill bit. Further details of this Drilling Resistance Measurement Test (DRMT) are shown in Section 2 in the present study.
Fitzner and Heinrichs, as referenced in Heinrichs [6], studied the salt weathering of the Umm Ishrin Sandstone Formation in the monuments of the Nabataean city of Petra. They determined drilling resistance–depth profiles using a device with drill bits 3 mm in diameter at constant pressure, supply of energy and rotation speed in a representative area with features such as contour scaling and granular disintegration. The drilling resistance was obtained as a ratio between values of drilling time and drilling depth. They found that halite was the representative salt mineral.
Siedal et al. [16] used the Durabo portable drilling system to carry out tests on a scale on the north wall of the Gopura building of Angkor Wat. They recorded, in a chart, the drilling resistance based on the values of the penetration depth of a drill bit with a 3 mm diameter, while the rotation and the pressing force were kept constant and determined the salt weathering profile. These results showed that the zone of detachment behind the scale has a thickness of approximately 1.5 cm.
Bruthans et al. (supplementary information of [21]) measured the drilling resistance as the drilling depth on samples of sandstone exposures in the Bohemian Cretaceous Basin and Colorado Plateau (Navajo, Wingate, Star Point, Kayenta and Hruba Skala sandstones and Crystal Peak Tuff) at constant pressure and rotation speed. This technique allowed them to correlate the drilling resistance with the tensile strength. They also found that the most durable sandstones against salt weathering cycles of NaSO4 showed lower values of drilling resistance.
Dassow et al. [26] assessed artificial salt decay on a sample of Stanton Moor sandstone with ultrasonic-assisted drilling (UAD), recording the force applied and the power required to maintain the oscillation and amplitude while drilling with the drill bit developed for this device.
St. Leonard’s Church is a monument in Atouguia da Baleia, located in the western region of Portugal, about 3 km away from the Atlantic coast, and was built in the Middle Ages. Atouguia da Baleia was near a harbour on the shoreline of an ancient channel between this village and Peniche Island until the 15th century, when it started to be filled with sea and river sediments, and it has since disappeared. The façades of St. Leonard’s Church exhibit cavities in dimension blocks and capitals below the western vault, around 10 cm long and deeply connected, due to sodium chloride crystallization, the major source of these salts being seawater. The salt spray carries sodium chloride to the façades of St. Leonard’s Church and causes deposition and crystallization in the pores of the sandstone. Salt solutions are also absorbed by capillarity [27,28]. The salt effect is widely distributed and damaged the carvings of the portals of St. Leonard’s Church over a period of almost 800 years, inducing the loss of material in the motifs of the structural blocks, as highlighted on the main western façade (Figure 1).
Another important pattern of deterioration is biological colonization, highly developed on stone surfaces, with fungi and lichens being the most visible features. The latter is associated with the blistering and detachment of stone laminae.
The results of quasi-NDTs, such as DRMTs (Drilling Resistance Measurement Tests), carried out on stone monuments can be correlated with data from DTs (Destructive Tests), such as uniaxial compression. Expressions were obtained from a comparison between the DRMT data recorded from representative blocks of the monument St. Leonard’s Church and the results of compression obtained on similar specimens collected from built heritage in the vicinity. The latter specimens were validated by an intrinsic signature through a methodology based on finding similarities by carrying out a visual inspection and petrographic, physical (comprising the absorption of water under low pressure and porosity) and mechanical characterization (compression and DRM tests) [27,29]. Aside from size requirements for mechanical validity and the specifications required by the experimental apparatus, the compression test has no restrictions on the size of the specimens to maintain monument integrity, since it is not performed directly on its materials.
A detailed mineralogical, physical and mechanical characterization is presented concerning the four typologies of sandstone lithotypes found in the built heritage close to the monument in order to contribute to a better understanding of the physical and mechanical behaviour of these sandstone building materials.
An analytical model based on stress–strain diagrams, supported by microdrilling strength values of samples of sound and weathered sandstones, is shown, which allows the assessment of the structural stability of building blocks of monuments as a measure to ensure the long-lasting existence of striking artistic testimonies such as carvings and reliefs.

2. Experimental Work

More than 240 samples were extracted from different varieties of sandstone blocks of the Lourinhã Formation with 5 cm long cross-sections, corresponding to height-to-length ratios of 1 and 2. Three varieties of sandstones were detected on the façades of St. Leonard’s Church, identified as B, C and M. Variety A was not found in this monument, but it is present in the masonry walls located close to it. Varieties A and B belong to lithotype A + B, and varieties C + M constitute lithotype C + M.
Lithotype A + B’s macroscopic laminations and lineations were aligned parallel to the major axis. The prismatic specimens were randomly cut, as no well-defined macroscopic laminations were found in the C and M varieties of sandstone [27].
Archimedes’ principle was used to determine the porosity and the bulk and real densities of the sandstone specimens, according to the Recommendations of [30,31]; after saturation under vacuum at a minimum value of 15 mbar for 24 h, air voids were filled with trapped deionized water, lasting 24 h, and atmospheric pressure was reintroduced after 24 h. The drying of the specimens was carried out in an oven at a temperature of 105 °C and lasted 72 h.
Mercury intrusion porosimetry (MIP) was carried out in order to obtain the pore size distribution (PSD) of B, C and M sandstone varieties [27]. Mercury capillary pressures of up to 200 MPa were applied by an AutoPore IV9500 from Micromeritics Instrument Corporation. The Katz and Thompson method [32] was used to infer the permeability values.
The water absorption was measured in a low-pressure test [30] carried out using a Karsten tube fixed to removable putty on the top face of the sandstone samples. The top opening of the vertical pipe was filled with water dropped towards the tube top, with the graduation “0”. The slope of the linear part of the test curve was given by water absorption graphs of the mass of water per unit area as a function of the square root of time. The values of the water absorption coefficient, k, were then obtained. Figure 2 shows Karsten pipes puttied on a portal of the south façade.
A covered tray containing deionized water was used to carry out the capillarity tests on prismatic samples of sandstones [30]. These samples were introduced into a tray with a water-level thickness of 2 mm, which was kept constant during the test. The absorbed amounts of water were weighed at different times, and the values were divided per unit of cross area of the specimens. The values of this parameter were plotted as a function of the square root of time. The capillarity water absorption coefficient was obtained through the slope of the linear part of the test curve.
Drying tests were carried out according to the Recommendations of NORMAL 29/88 (1991) and RILEM (1980) [30,33] on cubic sandstone samples under laboratory environment conditions of a 20 ± 2 °C temperature (T) and 55 ± 10% relative humidity (RH). Drying occurred on the six faces of the cubic specimens of sandstone lithotype A + B and only on the top face of specimens of sandstone lithotype C + M. It was not possible on the other five faces of the latter samples because they were sealed with an epoxy resin impermeable to water.
The drying index (DI) of NORMAL 29/88 (1991) [33] expresses the drying behaviour of experimental curves and is given in Equation (1):
D I = ( t 0 t f f ( Q i ) ) / ( Q m a x t f )
with f(Qi) as the percentage of dry mass towards the end of drying (tf), Qmax as the initial water content and t0 as the initial time.
A Seidner servo-controlled press, model 3000D, with a load capacity of up to 3000 kN and a piston stroke of 50 mm was used in [34] to carry out monotonic uniaxial compression tests with an axial displacement control rate of 10 μm/s. Each LVDT was inserted on each side of the specimen between the plates of the compression testing machine, and the four LVDTs allowed the average axial displacement between platens to be determined. Stress–strain diagrams were obtained in order to determine the behaviour of 50 × 50 × 100 mm3 sandstone prismatic samples under uniaxial compression.
The Drilling Resistance Measurement Test (DRMT) was performed with a portable microdrilling device with a drilling head on a slide, which controls the stepper movement towards the stone building blocks on a façade or on a holder. Diaber tungsten drill bits with a 5 mm diameter and a triangular tip were used in this study to reach maximum drill hole depths of up to 25 mm through the measurement of a load cell. These drill holes were made on building sandstones of St. Leonard’s portal of the south façade (Figure 3) and on similar specimens collected close to this monument [27]. The drilling parameters were a speed of 200 rotations per minute (rpm) and an advancing rate of 20 mm/min, which were defined by preliminary tests. The power used to operate the DRMT device was electricity.
Tip wear due to abrasion was corrected on drilling strength profiles according to Singer et al. [35], based on linear dependence on the drilled depth. The correction factor was determined by linear regression on data recorded from a DRMTs performed on the same stone type at increasing drilled lengths with the same drill bit. The test results are given in graphs with drilling strength (σd) on the vertical axis and drilled depth (d) on the horizontal axis.
A balance was considered between respect for the integrity of the monument and the unknown time required for the preservation efforts for this heritage building. The maximum number of DRMTs carried out in the present study was only 2 to 3 per block of sandstone varieties in order to allow a future follow-up of weathering progression. Performing all of the DRMTs on the portal of the south façade, analysing their results and obtaining the drilled data took one working day.

3. Results and Discussion

3.1. Petrographic Study

Hand specimens of four varieties of sandstones were observed, and the classifications of their surface colour according to the Munsell colour system are the following: A HUE 5Y 5/3, olive; B HUE 5Y 6/4, olive; C HUE 2.5Y 7/2, light grey; M HUE 2.5Y 6/6, olive-yellow.
These samples of sandstones were collected from the Lourinhã Formation, which has been dated to the Late Jurassic, according to [36].
Thin sections of representative samples of the sandstone varieties were studied under a polarizing microscope (Figure 4), and a modal analysis was carried out based on thousand-point counting, allowing the composition to be obtained, as presented in Table 1.
Table 2 shows the general results of this micropetrographic study including the observations made by scanning electron microscopy (SEM) and X-ray diffraction (XRD) analyses (Figure 4, Figure 5 and Figure 6). Quartz and feldspars are the major minerals, making up about 30–50% and 13–18% of the composition. The carbonates range between approximately 20% and 40% of the total amount and act as the cementing material, secondarily generated. The A, B, C and M varieties of sandstones were classified as lithic arkose according to [37], with the carbonate content playing an important role. These sandstone varieties were then classified as lithic arkose with carbonate cement, and two lithotypes were defined: A + B and C + M.
The properties studied and indicated in Table 2 show considerable similarities between the two lithotypes in terms of the following:
  • The constitutive minerals are essentially the same in the grains, matrix and grain-bonding cement, although there are differences in their respective contents, as per the previous reference in Table 1.
  • Micas present preferential orientations, except for the B variety, where there is a random distribution, and the M variety, which exhibits orientations in both directions. The macroscopic arrangement of the orientation of lineations in typology B is apparent, as microscopic observation revealed they were randomly distributed. In typology M, bedding was not observed macroscopically due to mica orientations in the two directions, according to microscopic observation.
  • The cement of carbonate and silicious materials, which predominate, is of secondary origin.
  • The rock fragments have the same origin, i.e., debris due to weathering, carried from Berlengas Islands in this specific subsector of the Lusitanian Basin, close to its western margin.
  • Chemical weathering is frequently shown in feldspars, e.g., sericitization and the partial chloritization of biotite, which also occurs in association with iron oxides and hydroxides, with their dissemination in the matrix, and of clinochlore, which constitutes oxidized chlorite.
Concerning the fabric, i.e., the geometric arrangement and the spatial distribution of the constituent elements, some differences are highlighted regarding the following aspects:
  • Grain size: In the A + B lithotype, the average grain sizes of quartz and feldspars range between 0.1 and 0.13 mm, with maximum values between 0.3 and about 0.5 mm, while in the lithotype C + M, they vary between 0.15 and 0.24 mm, with maximum values between around 0.5 and 0.95 mm. The designation of very fine sands of the classification [38] is not adequate for the C + M lithotype, only fine sands. Variety M has grains with average dimensions slightly larger than those of variety C and maximum dimensions similar to those of coarse sands.
  • Bonds and contacts between grains: In both lithotypes, the grains are supported by a predominantly calcite matrix, with direct contacts between them being rare in lithotype A + B, while in the C + M lithotype, there is a higher frequency of contacts.
The A + B lithotype is more compact than C + M.
Detailed micropetrographic observations carried out by SEM and XRD analyses are shown in Figure 4, Figure 5, Figure 6 and Figure 7. The minimum observable size of micropores obtained by SEM is less than about 0.5 μm (Figure 5e).

3.2. The Physical Study of Both Lithotypes of Sandstones of the Lourinhã Formation

Table 3 shows the values of porosity accessible to water and the bulk and real densities obtained on 240 specimens of sandstone A, B, C and M varieties and also on small samples of each variety from the monument, aside from variety A, which was not found. Water-accessible porosity values vary between about 3 and 5% for specimens of type A, 6 and 8% for the B variety, and 11 and 14% for the C variety, and for specimens of M-variety sandstone, the values range between approximately 16.5 and 19.5%. The average values of bulk density range from 2179 kg/m3 (variety M) to 2510 kg/m3 (variety B) and 2594 kg/m3 (variety A), with the values of the coefficient of variation generally lower than 1%. The average value of real densities of all specimens is close to 2680 kg/m3, i.e., the mean value of variety C, and the values of the coefficient of variation are lower than 0.5%. The average values of the properties measured regarding the tiny samples collected from the monument are generally within the narrow range of variation of 10% of each average value.
Figure 8 and Figure 9 show the PSDs of sandstones of varieties B, C and M obtained by MIP.
The PSD curves of sample M87 and monument samples Ig1 and Ig6 are similar.
The value of the distribution mode of samples B1, B7 and C4 is a radius of 54 μm (Figure 8) due to the measurement of only three values of radius (213, 54 and 36 μm). For higher values of injection pressure, corresponding to radius values lower than 10 μm, the variation in radius between measurements is about 1 μm or even lower, being about 0.5 nm for a radius value of 0.01 μm. So, the dimension of 54 μm does not represent more than 15% of the total volume of the pore space determined by MIP (Figure 9).
The percentage of micropore radii smaller than 7.5 μm [40] ranges between approximately 60% and 96%, and the median pore radius obtained at 50% of the total volume of mercury injected varies between about 0.1 μm in sample B7 and 5 μm in sample Ig1 (Figure 8 and Figure 9 and Table 4). The value of the latter parameter is 96% in sample C10, reaching approximately 85% (samples B1, C4 and Ig6), more than 80% (samples B7 and Ig1) and around 80% (sample M87).
Table 4 shows the permeability values inferred by the Katz and Thompson method [32], ranging from around 2 to 100 mD. The higher values of permeability obtained on the B samples are due to the value of the distribution mode of the radius being 54 μm. The software used by the equipment to calculate permeability assumed the maximum hydraulic conductance for this pore radius value, overvaluing these permeabilities.
The results of the tests of absorption of water under low pressure (pipe method) obtained on prismatic specimens and on monument façades are shown in Table 5, i.e., the values of the coefficient of absorption of water, k.
The curves of capillarity water absorption are presented in Figure 10, and the corresponding values of the coefficient of water absorption are shown in Table 6.
The potential correlation between surface wettability measurements and water absorption rates in historical building materials is also an important area of research contributing to the understanding of the physical behaviour of sandstone materials in monuments [41].
The drying behaviour of both sandstone lithotypes is shown in Figure 11 and Figure 12.
In Figure 11, the occurrence of a high retention of water content at the end of the tests, compared to the initial water content, is approximately 20% in the specimens of variety B and more than 1/3 in type A samples.
In Figure 12, the complete drying of sandstone variety M was achieved at around 550 h. In variety C, this time around, 10% of the initial water content remained, and drying was not completed after about 1100 h.
The curves of sandstone varieties A, B and C stabilized and would not benefit from longer tests (Figure 11 and Figure 12).
Although the results for the two lithotypes are not directly comparable, given the different conditions of the experimental procedures, it is evident that the values of the final retained water content show a decreasing trend from the A + B to the C + M lithotype.
The drying index (DI) is less affected by the experimental conditions and was measured following [33], which states that the characterization of drying is carried out by this index. In Table 7 and Table 8, the results of the drying index are presented. This parameter was calculated for all parts of the evaporation curves, considering two evaporation regimes. Thus, the total drying index (DIt) was decomposed into the index obtained in the constant flow regime (DI1) and in the non-linear flow regime (DI2).
The values of the drying index obtained on the specimens of the two lithotypes are also not comparable, given the differences in the final drying times, around 500 h in the A + B and 1100 h in the C + M lithotype samples. Longer drying times in the first lithotype were not allowed for the reasons explained above.
The values of the drying index shown in Table 7 and Table 8 generally exhibit acceptable deviations.
The mean values of the DIt and DI2 indices provided in Table 7 decrease by around half from the A to the B variety, while DI1 doubles. In Table 8, a similar situation is shown in terms of DIt from variety C to M. The mean value of DI2 of variety M is one-third of the average value of DI2 of type C, and the mean DI1 value of variety M is more than twice the average DI1 value of variety C. An important change in variety M is shown: the average values of DI1 and DI2 are equal, which means a significant increase in the contribution of the constant flow regime to the drying process.
The analysis of the results leads to the conclusion that the increment in the contribution of the constant flow regime in drying occurs with the increase in porosity of each lithotype, meaning easy drying.
While the evaporation kinetics of the constant flow regime are essentially based on external parameters, in the non-linear flow regime, they depend on the characteristics of the pore space.
The increment in values of retained water content in the vapour phase of drying (non-linear flow regime) is related to the lower dimensions of microporosity, with a wide distribution of more than 80% of the pore space of the samples investigated by MIP. As the pore space dimensions decrease, the adhesion forces between the water molecules and the pore walls increase, along with the tendency for small air bubbles to trap water droplets.
The coefficient of absorption by capillarity and drying index data are only shown because they allow a better characterization of monuments through similar samples that were tested.

3.3. The Mechanical Study of Both Lithotypes of Sandstones of the Lourinhã Formation

The mechanical behaviour was assessed by stress–strain curves obtained from compression tests on more than forty prismatic specimens of varieties A, B, C and M [34], and the physical behaviour was evaluated from porosity, PSD, water absorption and drying tests carried out on the samples investigated in this experimental work.
In the present study, the mechanical characterization of sandstones on the monument’s portal on the south façade was performed by DRMTs on the same blocks of varieties B, C and M on which water absorption under low pressure tests were done previously (Figure 2 and Figure 13 and Table 5). The use of the results of the latter physical tests guided the authors to perform the DRMT on adequate locations. The link between compression and DRMT results is due to the similar values of water absorption under low pressure shown by the monument blocks and sandstone specimens. In one block of the B variety, this link was achieved through a porosity value because it was the available result of a physical parameter. The methodology described limited the number of drilling tests to what was required, allowing the monument’s integrity to be respected. From the comparison between the DRMT results obtained on the monument blocks (Figure 14) and the stress–strain diagrams of the corresponding sandstone specimens collected in the vicinity of the monument, direct correlations between compression parameters and drilling strength are given in Section 4.
Drilling depths of up to 20 mm were deep enough to obtain strength values for an outer envelope with a weathering thickness of up to 9 mm on blocks of the south portal. Values of peak strength recorded at approximately 1 and 8 mm drilled lengths are due to higher-resistance grains.

4. Analytical Modelling of Stress–Strain Curves from DRMT Data

In the present study, a non-linear regression equation, Equation (2), was determined from the correlation between the compressive strength (σC) obtained on representative prismatic specimens of sandstones extracted in the vicinity of the monument and the drilling strength (σd) recorded on corresponding sandstone blocks of the south portal (Figure 15).
σ c = 10.751 e 1.161 σ d
The results of deformation at failure (εR) determined by compression tests carried out on sandstone specimens and σd results obtained on similar sandstone blocks of the south façade are shown in Figure 16.
Equation (3) correlates the deformation at failure (εR) and the drilling resistance (σd):
ε R = 8.186 0.571 σ d 1000
The coefficient of determination R2 = 1.
However, no specimens with low-pressure water absorption coefficients of around 21.7 kg/m2/√h were obtained in order to compare compression and DRMT results, and it was not possible to correlate deformation at failure with the drilling strength value from the IgII test.
Equation (4) describes the compression behaviour of sandstones of the south portal through the compressive strength (σC), given by Equation (2), and the deformation at failure (εR), given by Equation (3).
The prediction of stress–strain curves for monument sandstones based on Equation (4) is dependent on the drilling strength (σd) obtained on the monument blocks of varieties B, C and M.
σ = 1000 ε 8.186 0.571 σ d 3 + 1.47 1000 ε 8.186 0.571 σ d 2 + 0.5 1000 ε 8.186 0.571 σ d × 10.751 e 1.161 σ d
Figure 17 presents experimental curves and predicted stress–strain curves generated through the use of Equation (4). The comparison between the former and latter diagrams of varieties of sandstones tested revealed that they are similar to each other.
Table 9 shows a comparison between the predicted values of compressive strength (σCpred. DRMT) and the strain at failure (εRpred. DRMT) of sandstone blocks of the portal of the south façade and the corresponding experimental values for samples obtained near the monument (σCsample i and εRsample i). The absolute differences (∆) between them are lower than 5%, aside from variety-A values, which are lower than 10%.
This analytical modelling allowed simulated stress–strain diagrams to be generated based on the drilling strength values obtained on sandstone samples. It was also possible to predict the drilling strength (σd) from the compressive strength and the strain at failure of the variety-A sample and generate the simulated stress–strain diagram based on drilling strength, even without the previous experimental results of the DRMT on this sandstone variety.
Figure 18 highlights the simulation of stress–strain behaviour of two sound blocks of sandstones of lithotype C + M and of the corresponding salt-weathered envelopes with a depth of up to 9 mm from the DRMT results. These weathered envelopes developed over a period of almost 800 years.
Over a period of around 800 years, the compressive strength decreased by approximately 20% for sandstone blocks of variety C and around 30% for sandstone blocks of variety M on the south façade of St. Leonard’s Church. These general rates of the compressive strength decrease are about 0.025% for the former and approx. 0.0375% per year for the latter blocks in the outer weathered envelope. Its thickness of up to 9 mm translates to a size increment of more than 10 µm per year. These data provide insight regarding the determination of a time range for the weathering of the outer envelopes of blocks of the south portal, reaching lower stable values of half the compressive strength that can ensure that they last for more than 130 to 200 years, assuming that similar environmental conditions occurred up to the current time.
However, future research will be carried out to contribute to a better understanding of the long-term effects of the simulated stress–strain diagrams on the durability of historical building materials.
Stress–strain diagrams allow the assessment of the stability of sandstone building blocks or rock mass blocks by following the evolution of salt-weathered envelopes. The methodology described in this work is essential to the study of the durability of blocks that support carvings, reliefs of cultural heritage and features of medium-scale geoheritage landforms, assuming that the time at which salt weathering began is known.
Adequate conservation works based on this innovative methodology for following the weathering of historical building materials and geoheritage rocks are essential to ensuring sustainable cultural tourism, which plays a key role in the sustainable development of such regions.
The methodology and findings of this study can be applied to other historical buildings or geoheritage sites facing similar durability issues, once a nearby non-heritage source similar to heritage stone materials is available to cut an adequate number of specimens with sizes that ensure the physical and mechanical validity of the test results. Laboratory facilities are required to carry out all tests used in this study, and authorized access to heritage sites should be given to perform the referenced NDTs and quasi NDTs.

5. Conclusions

The façades of Saint Leonard’s Church show important decorative reliefs and carvings. The building blocks that support these Middle Ages’ artistic testimonies are arkose sandstones with carbonate cement from the Lourinhã Formation. A broad characterization was carried out, comprising studies on rock composition, physics and mechanics, which allowed a better understanding of the physical and mechanical behaviour of these stone materials in order to obtain a match between the building blocks of this monument and the specimens extracted from masonry walls close to it.
A methodology was used to generate simulated stress–strain diagrams of the building blocks of this church based on compression tests carried out on similar building stones collected in its vicinity and DRMTs performed on the actual blocks of this monument. These predicted stress–strain curves have good agreement with the experimental diagrams obtained.
An assessment of the structural stability based on the stress–strain curves of the sound stone and weathered envelope of these building blocks allows researchers to monitor the mechanical behaviour of both and to gain knowledge about their durability to ensure the long-lasting existence of relevant art elements shown, ensuring the development of sustainable cultural tourism.
Further research avenues will come from data obtained on other historical buildings or geoheritage sites with the use of this innovative methodology in order to extend its application and to increase the knowledge about structural stability related to durability issues.

Author Contributions

M.L.-M. developed the experimental program; M.L.-M. and C.C. analysed the data, discussed the results and contributed to the writing of the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. UNI 11182; Beni Culturali—Materiali Lapidei Naturali ed Artificiali—Descrizione Della Forma di Alterazione—Termini e Definizioni. UNI: Reggio di Calabria, Italy, 2006.
  2. Ordaz, J.; Esbert, R. Glosario de términos relacionados con el deterioro de las piedras de construcción. Mater. Constr. 1988, 38, 39–45. [Google Scholar] [CrossRef]
  3. Fitzner, B.; Heinrichs, K. Photo Atlas of Weathering Forms on Stone Monuments; Geological Institute, RWTH Aachen University—Working Group Natural Stones and Weathering: Aachen, Germany, 2004; 24p. [Google Scholar]
  4. Vergès-Belmin, V. Illustrated Glossary on Stone Deterioration Patterns. Glossaire Illustré sur les Formes d’altération de la Pierre; ICOMOS: Warsaw, Poland, 2008. [Google Scholar]
  5. Fitzner, B.; Heinrichs, K.; La Bouchardiere, D. Weathering damage on Pharaonic sandstone monuments in Luxor. Egypt. Build. Environ. 2003, 38, 1089–1103. [Google Scholar] [CrossRef]
  6. Heinrichs, K. Diagnosis of weathering damage on rock-cut monuments in Petra, Jordan. Environ. Geol. 2008, 56, 643–675. [Google Scholar] [CrossRef]
  7. Wadeson, L. Nabataean façade tombs: A new chronology. Studies in the History and Archaeology of Jordan. Jordan Press Found. 2013, 11, 507–528. [Google Scholar]
  8. Heinrichs, K.; Fitzner, B. Stone monuments of the Nemrud Dag sanctuary/Turkey—Petrographical investigation and diagnosis of weathering damage. [Steinmonumente des Nemrud Dag heiligtums/Türkei petrographische untersuchung und diagnose der verwitterungsschäden]. Z. Dtsch. Ges. Geowiss. 2007, 158, 519–548. [Google Scholar]
  9. Ma, J. Leshan Grand Buddah: Investigations of traditional and alternative materials for repairs. Methods of Evaluating Products for the Conservation of Porous Building Materials in Monuments. In Proceedings of the International Colloquium, ICCROM 1995, Rome, Italy, 19–21 June 1995; pp. 29–39. [Google Scholar]
  10. Qiao, Z.; Ding, Z.; Wang, J.; Sun, B.; Wang, F.; Xie, Z. Enhanced mechanical properties and environmental erosion resistance with metakaolin in a kind of Chinese traditional Lime-based mortar. Constr. Build. Mater. 2022, 317, 126110. [Google Scholar] [CrossRef]
  11. Stein, E. Constructing Kanchi: City of Infinite Temples; Amsterdam Univ. Press: Amsterdam, The Netherlands, 2021. [Google Scholar]
  12. Singh, M. Studies on weathering of Kailasanatha Temple, Kancheepuram. Curr. Sci. Assoc. 1993, 64, 559–565. [Google Scholar]
  13. Deshpande, S.; Savant, S.; Cardoz, D.; Mehrotra, R.; Piplani, N. Stabilisation & Conservation of Walls, Bastions & Slopes of Jaisalmer Fort; Report Prepared by Bombay Collaborative Urban Design & Conservation Pvt. Ltd. for the Archaeological Survey of India and World Monuments Fund; World Monuments Fund: New York, NY, USA, 2008; 253p. [Google Scholar]
  14. Uchida, E.; Ogawa, Y.; Maeda, N.; Nakagawa, T. Deterioration of stone materials in the Angkor monuments. Cambodia Eng. Geol. 1999, 55, 101–112. [Google Scholar] [CrossRef]
  15. Hosono, T.; Uchida, E.; Suda, C.; Ueno, A.; Nakagawa, T. Salt weathering of sandstone at the Angkor monuments, Cambodia: Identification of the origins of salts using sulfur and strontium isotopes. J. Arch. Sci. 2006, 33, 1541–1551. [Google Scholar] [CrossRef]
  16. Siedel, H.; Pfefferkorn, S.; Plehwe-Leisen, E.; Leisen, H. Sandstone weathering in tropical climate: Results of low-destructive investigations at the temple of Angkor Wat, Cambodia. Eng. Geol. 2010, 115, 182–192. [Google Scholar] [CrossRef]
  17. Ashmore, W. Quiriguá Reports, Settlement Archaeology at Quiriguá, Guatemala; Univ. Pennsylvania Press: Philadelphia, PA, USA, 2007; Volume IV, 362p. [Google Scholar]
  18. Migoń, P. Sandstone geomorphology—Recent advances. Geomorph 2021, 373, 107484. [Google Scholar] [CrossRef]
  19. Graham, J. Rainbow Bridge National Monument Geologic Resources Inventory Report; Natural resource report NPS/NRPC/GRD/NRR—2009/131; National Park Service: Denver, CO, USA, 2009.
  20. Ostanin, I.; Safonov, A.; Oseledets, I. Natural erosion of sandstone as shape optimization. Sci. Rep. 2017, 7, 17301. [Google Scholar] [CrossRef] [PubMed]
  21. Bruthans, J.; Soukup, J.; Vaculikova, J.; Filippi, M.; Schweigstillova, J.; Mayo, A.; Masin, D.; Kletetschka, G.; Rihosek, J. Sandstone landforms shaped by negative feedback between stress and erosion. Nat. Geosci. 2014, 7, 597–601. [Google Scholar] [CrossRef]
  22. Queensland’s Sandstone Wilderness Parks; State of Queensland, Queensland Parks and Wildlife Service, Department of Environment and Science: Brisbane, Australia, 2022; 28p.
  23. Soleilhavoup, F. Sahara. Visions d’un Explorateur de la Mémoire Rupestre; Transboréal: Paris, France, 1999; 128p, ISBN 2-913955-04-5. [Google Scholar]
  24. Tiano, P.; Rodrigues, J.D.; De Witte, E.; Vergès-Belmin, V.; Massey, S.; Snethlage, R.; Costa, D.; Cadot-Leroux, L.; Garrod, E.; Singer, B. The conservation of monuments: A new method to evaluate consolidating treatments. Int. J. Restor. Build. Monum. 2000, 6, 133–150. [Google Scholar]
  25. Ludovico-Marques, M.; Delgado-Rodrigues, J. Consolidation of volcanic tuffs. Some physical and mechanical properties. In Proceedings of the EUROCK 2002, International Workshop on Volcanic Rocks, Funchal, Portugal, 25–28 November 2002; pp. 37–44. [Google Scholar]
  26. Dassow, J.; Lia, X.; Leea, M.; Youngb, M.; Harkness, P. Ultrasonic drilling for the characterisation of building stones and salt induced decay. Ultrasonics 2020, 101, 106018. [Google Scholar] [CrossRef] [PubMed]
  27. Ludovico-Marques, M. Contribution to the Knowledge of the Effect of Crystallization of Salts in the Weathering of Sandstones. Application to the Built Heritage of Atouguia da Baleia. Ph.D. Thesis, Universidade Nova de Lisboa, Lisbon, Portugal, 2008; p. 314. (In Portuguese). [Google Scholar]
  28. Ludovico-Marques, M.; Chastre, C. Effect of salt crystallization ageing on the compressive behavior of sandstone blocks in historical buildings. Eng. Fail. Anal. 2012, 26, 247–257. [Google Scholar] [CrossRef]
  29. Chastre, C.; Ludovico-Marques, M. Nondestructive testing methodology to assess the conservation of historic stone buildings and monuments. In Handbook of Materials Failure Analysis with Case Studies from the Construction Industries; Makhlouf, H., Aliofkhazraei, M., Eds.; Butterworth Heinemann (Elsevier): Oxford, UK, 2018; pp. 255–294. [Google Scholar]
  30. RILEM. Recommended tests to measure the deterioration of stone and to assess the effectiveness of treatment methods. Mater. Const. Bourdais-Dunoud 1980, 13, 175–253. [Google Scholar]
  31. EN1936; Natural Stone Test Method-Determination of Real Density and Apparent Density, and of Total and Open Porosity. European Committee for Standardization: Brussels, Belgium, 1999.
  32. Katz, A.; Thompson, A. Quantitative prediction of permeability in porous rock. Phys. Rev. Online Arch. B 1986, 34, 8179–8181. [Google Scholar] [CrossRef] [PubMed]
  33. NORMAL–29/88; Misuradell’Indice di Asciugamento (Drying Index). Alterazioni dei Material Lapidei e Trattamenti Conservativi Proposte per L’unificazione dei Metodi Sperimentali di Studio e di Controllo. CNR–ICR: Rome, Italy, 1991.
  34. Ludovico-Marques, M.; Chastre, C. Prediction of stress–strain curves based on hydric non-destructive tests on sandstones. Materials 2019, 12, 3366. [Google Scholar] [CrossRef] [PubMed]
  35. Singer, B.; Hornschild, J.; Snethlage, R. Strength profiles and correction functions for abrasive stones. In Proceedings of the Workshop DRILLMORE—Drilling Methodologies for Monuments Restoration, Munich, Firenze, 16–17 March 2000. [Google Scholar]
  36. Taylor, A.; Gowland, S.; Leary, S.; Keogh, K.; Martinius, A. Stratigraphical correlation of the Late Jurassic Lourinhã Formation in the Consolação Sub-basin (Lusitanian Basin), Portugal. Geol. J. 2013, 49, 143–162. [Google Scholar] [CrossRef]
  37. Folk, R. Petrology of Sedimentary Rocks; Hemphill Publishing: Austin, TX, USA, 1974; 184p. [Google Scholar]
  38. Folk, R. Petrology of Sedimentary Rocks; Hemphill Publishing: Austin, TX, USA, 1968; 170p. [Google Scholar]
  39. Powers, M.A. New roundness scale for sedimentary particles. J. Sediment. Res. 1953, 23, 117–119. [Google Scholar] [CrossRef]
  40. Pellerin, F. La porosimetrie au mercure apliquee a l’ estude geotechnique des sols et des roches. Bull Liaison Lab Ponts Chauss 1980, 106, 105–116. [Google Scholar]
  41. Lee, J.; Derome, D.; Carmeliet, J. Drop impact on natural porous stones. J. Colloid Interface Sci. 2016, 469, 147–156. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Templar Cross visible on structural block of west portal of St. Leonard’s Church showing loss of material and lack of definition of carving.
Figure 1. Templar Cross visible on structural block of west portal of St. Leonard’s Church showing loss of material and lack of definition of carving.
Sustainability 16 06277 g001
Figure 2. Absorption of water in low-pressure tests: Karsten pipes on sandstone blocks of portal of south façade.
Figure 2. Absorption of water in low-pressure tests: Karsten pipes on sandstone blocks of portal of south façade.
Sustainability 16 06277 g002
Figure 3. DRM device carrying out tests on sandstone blocks of portal of south façade of St. Leonard’s Church.
Figure 3. DRM device carrying out tests on sandstone blocks of portal of south façade of St. Leonard’s Church.
Sustainability 16 06277 g003
Figure 4. Observations of thin sections under polarizing microscope (crossed nicols). Sandstone with carbonate cement, with grains supported by calcite matrix; direct contacts between quartz and feldspar minerals are rare. (a) Muscovite and clinochlore with dominant orientation on A variety (magnification: 40×). (b) Feldspar, quartz, micas, calcite and opaque minerals on A variety (magnification: 200×). (c) Quartz, feldspar, micas and calcite matrix of B variety (magnification: 100×). (d) B variety showing mica details, muscovite (green) and biotite (brown) (magnification: 100×). (e) Quartz overgrowths around calcite matrix of C variety (magnification: 100×). (f) Porous space with opaque and calcite minerals of C variety (magnification: 100×). (g) Matrix-supported grains and direct grain contacts (point contacts and waved sutures) of M variety (magnification: 100×). (h) Porous space with opaque minerals, quartz overgrowths and feldspar of M variety (magnification: 100×). (i) Matrix of C + M variety of monument (sample Ig1), mainly calcite, and direct grain contacts between quartz and feldspars (magnification: 100×). (Qz = quartz; Cc = calcite; Fsp = feldspar; Rf = rock fragment; Mi = mica; Cl = clinochlore; Cm = clay mineral and hematite; Op = opaque mineral).
Figure 4. Observations of thin sections under polarizing microscope (crossed nicols). Sandstone with carbonate cement, with grains supported by calcite matrix; direct contacts between quartz and feldspar minerals are rare. (a) Muscovite and clinochlore with dominant orientation on A variety (magnification: 40×). (b) Feldspar, quartz, micas, calcite and opaque minerals on A variety (magnification: 200×). (c) Quartz, feldspar, micas and calcite matrix of B variety (magnification: 100×). (d) B variety showing mica details, muscovite (green) and biotite (brown) (magnification: 100×). (e) Quartz overgrowths around calcite matrix of C variety (magnification: 100×). (f) Porous space with opaque and calcite minerals of C variety (magnification: 100×). (g) Matrix-supported grains and direct grain contacts (point contacts and waved sutures) of M variety (magnification: 100×). (h) Porous space with opaque minerals, quartz overgrowths and feldspar of M variety (magnification: 100×). (i) Matrix of C + M variety of monument (sample Ig1), mainly calcite, and direct grain contacts between quartz and feldspars (magnification: 100×). (Qz = quartz; Cc = calcite; Fsp = feldspar; Rf = rock fragment; Mi = mica; Cl = clinochlore; Cm = clay mineral and hematite; Op = opaque mineral).
Sustainability 16 06277 g004aSustainability 16 06277 g004b
Figure 5. SEM observations of representative varieties of sandstone samples covered with gold. Porous space observations of (a) A-variety sandstone (magnification: 2000×), (b) B-variety sandstone (magnification: 1000×), (c) C-variety sandstone (magnification: 500×), (d) M-variety sandstone (magnification: 500×), (e) Ig1 sample corresponding to C + M lithotype (magnification: 3000×). (Qz = quartz; Cc = calcite; Ab = albite (Feldspar); Mu = muscovite; Mx = matrix.)
Figure 5. SEM observations of representative varieties of sandstone samples covered with gold. Porous space observations of (a) A-variety sandstone (magnification: 2000×), (b) B-variety sandstone (magnification: 1000×), (c) C-variety sandstone (magnification: 500×), (d) M-variety sandstone (magnification: 500×), (e) Ig1 sample corresponding to C + M lithotype (magnification: 3000×). (Qz = quartz; Cc = calcite; Ab = albite (Feldspar); Mu = muscovite; Mx = matrix.)
Sustainability 16 06277 g005
Figure 6. An X-ray diffraction diagram obtained on representative samples of A and B varieties of sandstones through the use of KαCu radiation, with the 2θ angle varying between 3 and 63°. (a) A variety, (b) B variety.
Figure 6. An X-ray diffraction diagram obtained on representative samples of A and B varieties of sandstones through the use of KαCu radiation, with the 2θ angle varying between 3 and 63°. (a) A variety, (b) B variety.
Sustainability 16 06277 g006
Figure 7. An X-ray diffraction diagram obtained on representative samples of C and M varieties of sandstones through the use of KαCu radiation, with the 2θ angle varying between 3 and 63°. (a) C variety, (b) M variety, (c) representative Ig1 sample corresponding to C + M lithotype of sandstone monument.
Figure 7. An X-ray diffraction diagram obtained on representative samples of C and M varieties of sandstones through the use of KαCu radiation, with the 2θ angle varying between 3 and 63°. (a) C variety, (b) M variety, (c) representative Ig1 sample corresponding to C + M lithotype of sandstone monument.
Sustainability 16 06277 g007
Figure 8. Pore size distribution of representative samples of B, C and M varieties of sandstones obtained from injection of differential volume of mercury (Ig1 and Ig6 are samples collected on south portal of monument).
Figure 8. Pore size distribution of representative samples of B, C and M varieties of sandstones obtained from injection of differential volume of mercury (Ig1 and Ig6 are samples collected on south portal of monument).
Sustainability 16 06277 g008
Figure 9. Pore size distribution of representative samples of B, C and M sandstones obtained from injection of cumulative volume of mercury.
Figure 9. Pore size distribution of representative samples of B, C and M sandstones obtained from injection of cumulative volume of mercury.
Sustainability 16 06277 g009
Figure 10. Curves of capillarity water absorption obtained on sandstone specimens.
Figure 10. Curves of capillarity water absorption obtained on sandstone specimens.
Sustainability 16 06277 g010
Figure 11. Graph of drying observed in sandstone varieties A and B.
Figure 11. Graph of drying observed in sandstone varieties A and B.
Sustainability 16 06277 g011
Figure 12. Graph of drying observed in sandstone varieties C and M.
Figure 12. Graph of drying observed in sandstone varieties C and M.
Sustainability 16 06277 g012
Figure 13. A DRMT on sandstone blocks of the south portal below the vault, related to the recorded coefficient of absorption of water (IgIII data obtained on a block only with the porosity value (n) and without the direct coefficient of absorption data).
Figure 13. A DRMT on sandstone blocks of the south portal below the vault, related to the recorded coefficient of absorption of water (IgIII data obtained on a block only with the porosity value (n) and without the direct coefficient of absorption data).
Sustainability 16 06277 g013
Figure 14. DRMT results on sandstone blocks of portal of south façade.
Figure 14. DRMT results on sandstone blocks of portal of south façade.
Sustainability 16 06277 g014
Figure 15. Results of σd from sandstone blocks of St. Leonard’s Church and σc from corresponding specimens extracted close to this monument.
Figure 15. Results of σd from sandstone blocks of St. Leonard’s Church and σc from corresponding specimens extracted close to this monument.
Sustainability 16 06277 g015
Figure 16. Results of εR and σd determined by compression tests on sandstone specimens and DRMT on sandstone blocks of south façade.
Figure 16. Results of εR and σd determined by compression tests on sandstone specimens and DRMT on sandstone blocks of south façade.
Sustainability 16 06277 g016
Figure 17. Analytical modelling of curves from DRMT results of sandstone specimens of varieties A, B, C and M. The thinner black curves are the stress–strain diagrams of experimental results, and the thicker dashed black curves are the analytical stress–strain diagrams.
Figure 17. Analytical modelling of curves from DRMT results of sandstone specimens of varieties A, B, C and M. The thinner black curves are the stress–strain diagrams of experimental results, and the thicker dashed black curves are the analytical stress–strain diagrams.
Sustainability 16 06277 g017
Figure 18. Stress–strain diagrams of drilled sound building blocks are dashed thinner curves, and those of their weathered outer envelopes are dashed thicker curves. The thicker black curve is an experimental curve of a representative specimen.
Figure 18. Stress–strain diagrams of drilled sound building blocks are dashed thinner curves, and those of their weathered outer envelopes are dashed thicker curves. The thicker black curve is an experimental curve of a representative specimen.
Sustainability 16 06277 g018
Table 1. The composition of the A, B, C and M varieties of sandstones obtained from modal analysis based on a thousand-point counting method (thin sections under a polarizing microscope, observations under crossed nicols).
Table 1. The composition of the A, B, C and M varieties of sandstones obtained from modal analysis based on a thousand-point counting method (thin sections under a polarizing microscope, observations under crossed nicols).
Minerals (%)AA ┴BB ┴CC ┴MM ┴Ig1 #
Quartz29.630.230.831.637.838.651.444.841.3
Feldspars *14.613.215.615.117.917.015.217.018.0
Carbonates **37.439.633.634.824.626.820.221.418.3
Micas/clinochlore ***4.93.85.84.64.43.83.63.65.4
Matrix ****6.76.87.87.57.35.83.24.88.8
Rock fragments *****5.85.45.45.27.07.25.26.67.2
Opaque minerals1.01.01.01.21.00.81.21.81.0
Total100100100100100100100100100
┴ Perpendicular direction to lineations; # Ig1 is a sample collected on the south façade of St. Leonard’s Church and corresponds to C + M varieties regarding its composition; * feldspars: albite, microcline; ** carbonates: calcite prevails, and ankerite is rare; *** micas: muscovite and biotite; **** matrix: clay minerals, clinochlore, calcite, hematite/goethite; ***** rock fragments: sedimentary rock fragments, chert, carbonates, metamorphic rock fragments and polycrystalline quartz.
Table 2. Results of micropetrographic analysis of A, B, C and M varieties of sandstones.
Table 2. Results of micropetrographic analysis of A, B, C and M varieties of sandstones.
Sandstones VarietyABCMC + M (Monument)
ClassificationLithic arkose with carbonate cement (after [37]).
Major mineralsQuartz, calcite, feldspars, rock fragments.
Minor mineralsMicas, clinochlore, clays, opaque minerals, hematite, goethite, heavy minerals.
FabricGrain size (quartz and feldspars)0.1 mm (average)
0.4 mm (maximum)
0.12 mm (average)
0.51 mm (maximum)
0.15 mm (average)
0.50 mm (maximum)
0.22 mm (average)
0.95 mm (maximum)
0.16 mm (average)
0.43 mm (maximum)
Very fine sand (according to [38])Very fine sand (according to [38])Fine sand (according to [38])Fine sand (according to [38])Fine sand (according to [38])
Grains with several sizes
SortingModerate.
Roundness (quartz and feldspars)Sub-angular to sub-rounded (after [39]).
Grain contacts (quartz and feldspars)Matrix-supported grains and rare direct grain contacts.Matrix-supported grains and direct grain contacts (point contacts and waved sutures).
OrientationMicas and clinochlore present a dominant orientationNot identified.Micas and clinochlore present a dominant orientationMicas and clinochlore present a dominant orientationMicas and clinochlore present a dominant orientation
Grain bondingMatrix-supported fabric (carbonate as matrix mineral is predominant).Matrix-supported fabric (carbonate as matrix mineral is predominant) and direct grain contacts (point contacts and waved sutures).
Mineral componentsQuartzSometimes with overgrowths and inclusions.
CarbonatesCarbonate, mainly calcite; ankerite is rare.
FeldsparsAlbite, anorthite, microcline; often weathered and sericitized or looking similar to quartz.
Rock fragmentsSedimentary rock fragments, chert, carbonate rock fragments, metamorphic rock fragments, polycrystalline quartz.
Mineral componentsMicas/
Clinochlore
Partly in large flakes; micas are biotite and muscovite; biotite is partly chloritized or occurs in combination with Fe-oxide (hematite); clinochlore occurs as single mineral or as fine-grained matrix mineral.
Clay mineralsKaolinite, sericite; partly in combination with Fe-oxide (hematite) and Fe-hydroxide (goethite).
Opaque min.Mainly ferrous.
Oxides and hydrox of ironMainly hematite and goethite as opaque minerals and fine-grained as matrix minerals.
Heavy mineralsRare; mainly tourmaline, zircon, rutile; as single grains or inclusions.
CementCarbonates as cementing material, secondarily generated. Siliceous cementing material as contact cement. Secondarily generated and often mixed with fine-grained calcite are clay minerals, fine-grained clinochlore or hematite as contact cement on fine borders or as pore-filling materials.
Table 3. Physical parameters obtained by Archimedes’ principle.
Table 3. Physical parameters obtained by Archimedes’ principle.
SamplesNumber of SamplesPorosity n (%)
(Average ± SD (CV %))
Bulk Density (kg/m3)
(Average ± SD (CV %))
Real Density (kg/m3)
(Average ± SD (CV %))
A variety344.1 ± 0.4 (9.8%)2594 ± 13 (0.5%)2705 ± 6 (0.3%)
B variety356.9 ± 0.5 (7.2%)
6.3 (monument)
2510 ± 19 (0.8%)
2375 (monument)
2697 ± 8 (0.3%)
2533 (monument)
C variety3412.7 ± 0.4 (3.1%)
11.5 * (12.5 and 10.5)
2343 ± 25 (1.1%)
2350 * (2375 and 2324)
2684 ± 10 (0.4%)
2655 * (2714 and 2596)
M variety14018.5 ± 0.4 (2.2%)
17.1 * (18.3 and 15.9)
2179 ± 13 (0.6%)
2222 * (2202 and 2241)
2671 ± 6 (0.2%)
2682 * (2697 and 2666)
* Mean physical parameters obtained on monument samples Ig1 and Ig6 collected on south portal; SD is standard deviation; CV is coefficient of variation.
Table 4. Physical parameters obtained by mercury intrusion porosimetry.
Table 4. Physical parameters obtained by mercury intrusion porosimetry.
SamplesPorosity n (%)Bulk Density (kg/m3)Real Density (kg/m3)Median of Pore Radius (µm)Permeability (mD)
B15.8250926650.3156.8
B77.4247226700.1496.7
C412.4234026710.842.8
C1013.1233326840.952.3
M8717215825913.5730.1
Ig1 *16.1213925504.7819.7
Ig6 *16.2219426163.2024.6
* Ig1 and Ig6 are samples collected on south portal of monument.
Table 5. Values of coefficient of water absorption, k, obtained on all varieties of sandstones.
Table 5. Values of coefficient of water absorption, k, obtained on all varieties of sandstones.
VarietySamples
(Prisms)
Water Absorption, k (kg/m2/√h)Water Absorption, k (kg/m2/√h)
Average ± SD
(CV %)
AAP380.70.9 ± 0.2 (21)
AP390.7
AP530.8
AP960.7
AP11.3
AP51.1
AP61
AP90.9
AP110.8
BBP62.42.4 ± 0.3 (14.4)
BP272.4
BP323.2
BP452.5
BP722
BP32.4
BP132.4
BP2
CCP187.66.2 ± 1.1 (17.5)
CP247.4
CP505.3
CP405.3
CP875.3
MMP126.425.6 ± 3.4 (13.3)
MP223.7
MP322.8
MP523.2
MP631.8
C + MT15.6-
T36.0-
T428.5-
T621.7-
SD is standard deviation, and CV is coefficient of variation.
Table 6. Values of the coefficient Cc of capillarity water absorption obtained on sandstone varieties.
Table 6. Values of the coefficient Cc of capillarity water absorption obtained on sandstone varieties.
VarietySamplesCc (kg/m2/√h)Average ± SD
(CV %)
AP10.40.4 ± 0.1 (24.5)
P20.5
P50.3
P60.3
BP30.60.6 ± 0.1 (9.2)
P80.6
P110.5
P130.7
CP42.52.1 ± 0.5 (25.0)
P122.2
P141.5
MP15.75.9 ± 0.8 (13.6)
P26.2
P35.1
P45.3
P57.1
Table 7. Values of drying index of lithotype A + B.
Table 7. Values of drying index of lithotype A + B.
VarietySamples
(Prisms)
Drying IndexDrying Index
Average ± SD (CV %)
DI1DI2DItDI1DI2DIt
AAP10.010.370.380.01 ± 0.00 (0.00)0.36 ± 0.02
(5.56)
0.37 ± 0.02
(5.41)
AP20.010.330.34
AP50.010.370.38
AP60.010.380.39
BBP30.020.180.200.02 ± 0.00 (0.00)0.18 ± 0.01
(5.56)
0.20 ± 0.01
(5.00)
BP80.020.160.18
BP110.020.180.20
BP130.020.180.20
Table 8. Values of drying index of lithotype C + M.
Table 8. Values of drying index of lithotype C + M.
VarietySamples
(Cubes)
Drying IndexDrying Index
Average ± SD (CV %)
DI1DI2DItDI1DI2DIt
CC60.10.020.160.180.02 ± 0.00 (0.00)0.18 * ± 0.07 * (38.89) *
0.15 ± 0.01
(6.67)
0.20 * ± 0.07 * (35.00) *
0.17 ± 0.01
(5.88)
C85.20.020.130.15
C79.10.020.140.16
C78.10.020.160.18
C47.20.020.160.18
C87.20.020.33 *0.35 *
MM1030.060.060.120.05 ± 0.01 (20.00)0.05 ± 0.01 (20.00)0.10 ± 0.01 (10.00)
M1040.050.050.10
M1050.050.050.10
M1060.060.050.11
M1070.050.040.09
M1080.050.040.09
* Higher values.
Table 9. Comparison between predicted and experimental values of compressive strength and strain at failure obtained through drilling strength simulation.
Table 9. Comparison between predicted and experimental values of compressive strength and strain at failure obtained through drilling strength simulation.
Specimen/TestAP11BP3Bpt(n = 6.3%)CP24T3MP1T4
k (kg/m2/√h)0.822762628.5
σd (MPa)2.25 *-1.9-1.2-0.5
σc (MPa)136.295.0-45.7-18.7-
σc pred DRMT (MPa)146.5-97.6-43.3-19.2
εR6.3 × 10−37.2 × 10−3-7.4 × 10−3-7.9 × 10−3-
εR pred DRMT6.9 × 10−3-7.1 × 10−3-7.5 × 10−3-7.9 × 10−3
∆σc pred − σc specimen (%)7.62.75.32.7
∆εR pred − εR specimen (%)9.51.41.40
* Predicted value.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ludovico-Marques, M.; Chastre, C. Modelling the Mechanical Effect of Salt Weathering on Historical Sandstone Blocks through Microdrilling. Sustainability 2024, 16, 6277. https://doi.org/10.3390/su16156277

AMA Style

Ludovico-Marques M, Chastre C. Modelling the Mechanical Effect of Salt Weathering on Historical Sandstone Blocks through Microdrilling. Sustainability. 2024; 16(15):6277. https://doi.org/10.3390/su16156277

Chicago/Turabian Style

Ludovico-Marques, Marco, and Carlos Chastre. 2024. "Modelling the Mechanical Effect of Salt Weathering on Historical Sandstone Blocks through Microdrilling" Sustainability 16, no. 15: 6277. https://doi.org/10.3390/su16156277

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop