Sunteți pe pagina 1din 15

Engineering Geology 74 (2004) 113 127

www.elsevier.com/locate/enggeo

Durability estimation of porous building stones from pore


structure and strength
D. Benavente a,b,*, M.A. Garca del Cura b,c, R. Fort b,c, S. Ordonez a,b
a

Departamento de Ciencias de la Tierra y del Medio Ambiente, Universidad de Alicante, Apartado 99, 03080 Alicante, Spain
b
Laboratorio de Petrologa Aplicada, Unidad Asociada CSIC-UA, Spain
c
Instituto de Geologa Economica, CSIC-UCM, Madrid, Spain
Received 24 September 2003; accepted 12 March 2004
Available online

Abstract
The influence of both pore structure and strength on the estimation of stone durability is evaluated. Salt crystallisation may
limit the durability of porous building stones because it can be considered as one of the most powerful weathering agents. Salt
crystallisation produces stress over the pore surface. Consequently, stone durability is closely related to both pore structure and
strength, which is the material resistance to crystallisation pressure. We propose a novel petrophysical durability estimator
(PDE) as the ratio between parameters and estimators, which are based on pore structure and the strength of the material.
In this study, we have used stone parameters and estimators that have an influence on durability and depend only upon pore
structure, i.e., connected porosity, water absorption, the water absorption coefficient, specific surface area (SSA), the saturation
coefficient, the durability factor, the durability dimensional estimator (DDE) and others derived from porosity, such as microporosity
and adjusted microporosity. We also used stone parameters and estimators with an influence on strength: flexural strength, uniaxial
compressive strength, Youngs dynamic modulus and compressional wave velocity. These parameters and estimators, and the
proposed petrophysical estimator are compared with a salt weathering test. Our study shows that there is a very strong correlation
between salt weathering and the proposed petrophysical estimator, whereas only moderate correlation exists with the estimators that
depend on pore structure and strength. We conclude that the proposed estimator contains the information necessary to understand and
estimate the durability of porous materials which have an impact on buildings, civil constructions and historical monuments.
D 2004 Elsevier B.V. All rights reserved.
Keywords: Building stone; Durability; Limestone; Petrophysics; Pore structure; Rock strength; Salt crystallisation tests; Weathering

1. Introduction
Weathering agents may cause a rapid change in
the initial petrophysical properties of rocks, and thus
* Corresponding author. Departamento de Ciencias de la Tierra
y del Medio Ambiente, Universidad de Alicante, Apartado 99,
03080 Alicante, Spain. Tel./fax: +34-965-903727.
E-mail address: david.benavente@ua.es (D. Benavente).
0013-7952/$ - see front matter D 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.enggeo.2004.03.005

limit their durability. Durability is an important


characteristic for controlling not only the stability
of slopes, surface and underground excavations but
also that of a great number of stone structures,
buildings and monuments. In fact, the durability of
a building stone can be defined as the measure of the
ability of natural building stone to endure and
maintain its essential and distinctive characteristics
of strength, resistance to decay and appearance in

114

D. Benavente et al. / Engineering Geology 74 (2004) 113127

relation to a specific manner, purpose and environment of use (Anon., 1983).


Salt crystallisation is one of the most powerful
weathering agents, especially when combined with
frost activity (Jefferson, 1993). In porous materials,
crystallisation pressurewhich depends on pore structure, the saturation degree of salt and the energy
difference between the crystal and the pore wallis
the most important decay mechanism that occurs during
salt weathering. It is well known that crystallisation
pressure is lower in larger pores (Wellman and Wilson,
1965; Fitzner and Snethlage, 1982; Gauri et al., 1988;
Benavente et al., 1999; Scherer, 1999, 2000; Flatt,
2002), and high saturation levels lead to high crystallisation pressure (Correns, 1949; Benavente et al.,
1999; Scherer, 1999, 2000; Flatt, 2002). The energy
difference between the crystal and the pore wall originates from surface forces. As a result, a substantial
amount of energy is normally required to produce direct
contact between salt and the material, setting the upper
limit of crystallisation pressure and, therefore, producing high levels of damage (Scherer, 1999, 2000).
Many studies have been carried out using experimental laboratory simulations to assess stone durability
produced by salt weathering. The differences between
them are the experimental conditions, which can be
distinguished by the size of the sample, type of saline
solution, salt concentration, temperature, relative humidity and number of cycles (Sperling and Cooke,
1985; Richardson, 1991; Jefferson, 1993; Goudie,
1993, 1999; Goudie and Viles, 1995; Ordonez et al.,
1997; Benavente et al., 2001). Moreover, these tests
have been widely used to evaluate stone durability
produced by other decay mechanisms, such as frost
action and wetting and drying cycles (Anon., 1983;
Jefferson, 1993, Ordonez et al., 1997; Winkler, 1997;
Scherer et al., 2001).
1.1. Durability estimation
Stone durability depends heavily on both strength
and pore structure properties, although it has mainly
been estimated either from strength or from pore
structure properties. The durability estimators are
defined by the petrophysical parameters of the stone,
which are usually obtained for its characterisation. By
applying the estimators, the time needed to carry out
the experiment is reduced from 3 4 weeks to a matter

of hours. Considerable practical advantages are offered for rapid assessment of the suitability of porous
stones for use under salt weathering conditions. However, the estimators may produce errors in predicting
stone durability because they are commonly defined
by the correlation of salt crystallisation tests and
related petrophysical parameters. This fact may cause
an inaccuracy to occur when the estimators are used in
other types of rocks.
Stone durability can be estimated from both
strength and pore structure properties. For example,
stones with high compressive strength, Youngs modulus or ultrasonic wave propagation velocities tend to
have a longer durability (Zezza, 1990; Valdeon et al.,
1996; Goudie, 1999; Nicholson, 2001).
Estimators are more frequently used and developed
from pore structure than from strength. The simplest
estimator is porosity, P, which is defined as the ratio of
the volume of voids to total volume of stone, expressed
as a percentage. More specifically, connected porosity
is related to the flow of weathering agents, decay
processes, such as salt crystallisation and frost activity
and, therefore, to stone durability. Connected porosity
can be obtained by different procedures. In this study,
mercury intrusion porosimetry (MIP) and capillary
imbibition are used for obtaining connected porosity.
An important parameter that describes the porous
media of materials is the specific surface area (SSA),
which is defined as the interstitial surface area of voids
and pores per unit mass of the porous material (Gregg
and Sing, 1982; Dullien, 1992; Tiab and Donaldson,
1996). SSA plays an important role in many applications of porous media such as ion exchange columns,
fluid conductivity and permeability, as well as in the
durability of porous materials. SSA may be used as a
durability estimator because high SSA values mean that
a greater surface area of the material will be decayed. It
is also inversely related to pore size and, therefore,
directly related to salt crystallisation.
Hirschawald (1908) defined the saturation coefficient, CS, as the ratio of the volume of water absorbed
(or accessible porosity to water Pabs) and total void
volume, Vtot (or total porosity, P),
CS

Vabs Pabs
:

Vtot
P

Absorbed water can be obtained by several methods: by capillarity imbibition (Richardson, 1991;

D. Benavente et al. / Engineering Geology 74 (2004) 113127

Jefferson, 1993) or by complete immersion in water


for 24 h (Modd et al., 1996). Differences between
these methods can be obtained when macroporous and
microporous stones are compared (Richardson, 1991;
Benavente et al., 2001; Benavente, 2003). Different
behaviour is recorded in the water absorption due to
the fact that in the immersion test, air is trapped within
the microporous sample, giving low water absorption
results. In the present study, the volume of water
absorbed (or accessible porosity to water, Pabs) is
obtained by capillary imbibition, P(Cap). A porous
material that absorbs a high amount of water is more
susceptible to decay because it facilitates different
decay mechanisms. Thus, high saturation coefficient
values lead to lower levels of stone durability.
The water absorption coefficient, C, obtained in the
capillary imbibition test, is closely related to pore
structure by the effective radius and the porosity.
Thus, rocks with narrow pores produce a low water
absorption coefficient value (Dullien et al., 1977;
Hammecker et al., 1993; Hammecker and Jeannette,
1994; Leventis et al., 2000; Benavente et al., 2002).
Due to the strong influence of pore structure on the
water absorption coefficient, it is reasonable to suggest that this coefficient may be inversely related to
stone durability, and therefore, it can be used as a
durability estimator.
Water absorption, Cabs, is also used as a durability
estimator and is very similar to the saturation coefficient. Water absorption is defined as the ratio of
absorbed water (difference between the weight of
the saturated and the dried sample) and the weight
of the dried sample.
Richardson (1991) defined the durability factor, D,
as the product of the saturation coefficient squared
and the connected porosity (D = CS2P). Durability
factor is related to salt weathering by the following
empirical correlations (Richardson, 1991; Bell, 1993):
DWL % 3:19D  9:60 Richardson; 1991;
DWL % 2:61D  7:60 Bell; 1993;
where DWL [%] is the percentage of dry weight loss
by salt crystallisation.
Modd et al. (1996) defined several durability
estimators by the correlation of DWL [%] to pore

115

structure parameters in 221 limestones from the UK.


These estimators are defined as follows:


Microporosity, Pmicro [%]: the percentage of pores


which are less than 5 Am in diameter, although in this
study the microporosity is chosen for the percentage
of pores which are less than 1 Am in radius.
 Adjusted microporosity, Padj: the product of total
porosity and measured microporosity ( Padj =
PPmicro).
 Other durability estimators are: PPmicro, PPmicroCS,
( PPmicro)CS, [( PPmicro)CS]0.5.
The best correlation between DWL [%] and the
previous estimators was found with [( PPmicro)CS]0.5,
where the linear regression coefficient was 0.840
(Modd et al., 1996).
Ordonez et al. (1997) defined the durability dimensional estimator (DDE) by the relationship between crystallisation pressure and pore structure as
follows:
X Dv ri
DDE Am1
Pconc ;
ri
where Dv is the pore size distribution; ri is the pore
size; and Pconc is the connected porosity. This
durability estimator contains full information about
the pore structure and can estimate stone durability
given that several important decay mechanisms are
inversely related to pore size, such as frost and salt
action as well as capillary pressure during the
wetting and drying cycles (Winkler, 1997; Scherer
et al., 2001).
Because salt crystallisation produces stress over
pore surfaces, it is reasonable to assume that stone
resistance to salt crystallisation is related to both pore
structure and stone strength, which is the material
resistance to crystallisation pressure. Thus, we propose
a new petrophysical durability estimator (PDE), which
represents the relationship between crystallisation pressure and material resistance. Because crystallisation
pressure is closely related to pore structure, it can be
described by parameters and estimators depending on
pore structure and having an influence on durability
(porosity, specific surface area, durability dimensional
estimator, etc.). The strength of the stone is an essential
parameter in determining its durability and can be
characterised macroscopically from static measure-

116

D. Benavente et al. / Engineering Geology 74 (2004) 113127

ments (uniaxial compressive strength, Brazilian test,


flexural strength, etc.) and dynamic measurements
(ultrasonic wave propagation technique; Jaeger and
Cook, 1976; Goodman, 1989). In this study, rock
strength is characterised by flexural strength, uniaxial
compressive strength and Youngs dynamic modulus.
Therefore, we define the petrophysical durability estimator in relation to these parameters.
In conclusion, the proposed petrophysical durability estimator (PDE) is defined as the ratio of parameters or estimators that are based on pore structure, X,
to the strength of material, r, as follows:
PDE

X
:
r

The aim of this paper is to evaluate the influence of


both the pore structure and strength on the durability
estimation and also to provide the most appropriate
durability estimator. This has been achieved by comparing previously existing estimators and the proposed petrophysical durability estimator with a salt
crystallisation test.

2. Experimental procedure
2.1. Materials
In this study, 18 porous stones have been chosen
for their different petrophysical and petrographic
characteristics. These stones are marketed and used
as building materials and are highly homogeneous not
only in the hand specimen but also in the quarry
(Benavente, 2003). These stones are divided into three
types of limestones, according to the Folk (1962)
classification: biocalcirudite (BR), biocalcarenite
(BC) and biomicrite (BM); and a siliceous stone
named quartz arenite (QA).
Biocalcirudites (and/or sandy biocalcirudites; BR)
are unsorted detrital stones with fragments of bryozoans, red algae, molluscs and echinoderm from 0.2 to
10 mm. The terrigenous fraction is comprised of
quartz, limeclasts and feldspars. These biocalcirudites
have abundant interparticle porosity, whereas intraparticle porosity is variable. Cement is scarce and
consists mainly of microcrystalline drusy calcite.
Biocalcarenites (BC) or sandy fossiliferous limestones are well-sorted arenites. These stones contain

foraminifers (mainly Globigerinae) ranging in size


from 0.2 to 0.5 mm. Foraminifera shells are generally
filled by glauconite and/or siliceous cement. The
terrigenous fraction is comprised of quartz, feldspars,
micas, dolostones and other rock fragments. Both
interparticle and intraparticle porosity are variable.
The most abundant type of cement present on these
stones is equant equicrystalline mosaics of calcite
spar.
The biomicrite (BM) or biomicrite limestone contains foraminifers (mainly Globigerinae), ranging in
size from 0.1 to 0.2 mm. The terrigenous fraction is
mainly comprised of quartz, feldspars and micas. BM
shows intraparticle and vuggy porosity. The micrite is
around 2 Am.
Quartzarenite (QA) is a well-sorted sandstone that
consists of monocrystalline quartz grains with rare
feldspars, metamorphic clasts, chert and muscovite
grains. Primary interparticle porosity has been partially filled by silica crystallisation, which forms frequent
overgrowths on quartz grains.
The amount of cracks and fractures can be considered negligible in the stones studied (Benavente,
2003). A detailed summary of the nature, approximate
quantity and mode of the clasts, as well as the type of
porosity and cement and matrix of the studied stones
are shown in Table 1.
2.2. Porous media characterisation
In order to characterise the porous media of stones,
mercury intrusion porosimetry (MIP), capillary imbibition and the nitrogen absorption technique were used.
Mercury porosimetry is extensively used for the
characterisation of porous media. The connected porosity, pore size distribution and mean pore size were
obtained by using the Autoscan-33 mercury porosimeter (Table 2).
The capillary imbibition test is widely used for the
characterisation of building materials and provides
information about their pore structure and durability.
Three samples were used in each test. The samples
were each in the form of 2.5 2.5 4 cm prisms and
distilled water was used. The results were plotted as
absorbed water per area of the sample throughout
imbibition versus the square root of time. Through
this kind of representation, the capillary imbibition
kinetic shows two parts. The first part defines water

D. Benavente et al. / Engineering Geology 74 (2004) 113127

117

Table 1
Nature, approximate quantity and mode of the clasts, and porosity type and cement and matrix type
Clasts

BR-1

BR-2
BR-3

BR-4

BR-5

BR-6

BR-7

BR-8

BC-1

BC-2

BC-3

BC-4

Porosity type

Cement and matrix type

Ref.

28

Interparticle (++++)

Drusy calcite spar (+++)

4, 5

0.03 0.1

Intraparticle (++++)

Syntaxial calcite spar (++)

100
<1
90

14
0.1 0.2
2 10

Interparticle (++++)
Intraparticle (+++)
Interparticle (++++)

Drusy calcite spar (++)

4, 5

Drusy calcite spar (+++)

4, 5

10
80

0.2 0.5
0.3 2

Intraparticle (++++)
Interparticle (++++)
Intraparticle (++++)

Micrite (+)
Drusy calcite spar (+++)

4, 5

20

0.2 0.5

95

0.2 10

Interparticle (++++)
Intraparticle (++)

Equant equicrystalline
mosaics of calcite spar (++)
Syntaxial calcite spar (++)

4, 5, 6

0.1 0.4

90

24

Interparticle (++++)
Intraparticle (+)

4, 5

10

0.1 0.2

95

5 10

Drusy calcite spar (++)


Equant equicrystalline
mosaics of calcite spar (++)
Syntaxial calcite spar (+)
Micrite (+)
Drusy calcite spar (++)

5
90

0.1 0.5
24

10
90

12
0.12 0.24

10

0.08 0.16

85

0.2 0.3

15

0.1 0.3

85

0.3 0.5

Interparticle (+++)

15

0.1 0.3

75

0.4 1

Intraparticle (+++)
Channel (+)
Interparticle (++++)

Nature

Percentage
(%)

Red algae, molluscs,


echinoderms
Quartz, feldspars,
(rock fragment)
Briozoans, red algae
Quartz, feldspars
Briozoans, red algae,
(echinoderms, foraminifers)
Quartz, feldspars
Briozoans, red algae,
echinoderms, (foraminifers)
Quartz, feldspars,
(rock fragments)
Foraminifers, briozoans,
molluscs, echinoderms,
red algae
Quartz, feldspars, rock
fragments, (muscovite)
Briozoans, red algae,
molluscs, echinoderms,
(foraminifers)
Quartz, feldspars, rock
fragments (chert)
Briozoans, molluscs,
equinoderms, red algae,
(foraminifers)
Quartz, (feldspars, clays)
Briozoans, molluscs, red algae,
foraminifers, echinoderms
Quartz, limeclasts
(feldspars, dolostone)
Foraminifers, calcispheres,
(echinoderm spines)
Quartz (feldspars, biotite,
muscovite, dolostones
and other rock fragments)
Foraminifers, briozoans,
echinoderms, molluscs
(red algae)
Quartz, (rock fragments,
feldspars)
Foraminifers, briozoans,
molluscs, echinoderms,
(red algae)
Quartz, rock fragments,
(feldspars)
Foraminifers, briozoans,
molluscs, echinoderms,
(red algae)

98
2

Mode (mm)

Interparticle (+++)
Intraparticle (+++)

Interparticle (++++)
Intraparticle (+)

Interparticle (+++)
Intraparticle (++)

Interparticle (+++)
Intraparticle (+++)
Channel (++)

4, 5

Equant equicrystalline
mosaics of calcite spar (+)
Drusy calcite spar (+)
Micrite (+)
Equant equicrystalline
mosaics of calcite spar (+++)
Silica (+)
Micrite (++)

Equant equicrystalline
mosaics of calcite spar (+++)
Syntaxial calcite spar (+)
Silica (++)
Micrite (+++)
Equant equicrystalline mosaics
of calcite spar (+++)

1, 2, 5, 7

Syntaxial calcite spar (+)


Micrite (+++)
Equant equicrystalline
mosaics of calcite spar (+++)

4, 5

1, 2, 5, 7

1, 2, 5, 7

(continued on next page)

118

D. Benavente et al. / Engineering Geology 74 (2004) 113127

Table 1 (continued)
Clasts

BC-4

BC-5

BC-6

BC-7

BC-8

BM

QA

Porosity type

Cement and matrix type

0.2 0.6

Intraparticle (++++)

Syntaxial calcite spar (++)

85

0.2 0.5

Interparticle (+++)

Equant equicrystalline
mosaics of calcite spar (+++)

15

0.1 0.3

Intraparticle (+++)
Channel (++)

70

1.5 2

Inrterparticle (+++)
Intraparticle (+)

30

0.5 1

Syntaxial calcite spar (+)


Silica (++)
Micrite (+++)
Drusy calcite spar (++)
Equant equicrystalline
mosaics of calcite spar (+)
Silica (+)
Micrite (+)

45

0.25 0.5

Interparticle (+)
Intercrystalline (+)
Intraparticle (+)

Equant equicrystalline
mosaics of calcite spar (++)
Micrite (++)

5, 6

55

0.12 0.25

90

0.12 0.25

Interparticle (+)

4, 5, 6

10

0.06 0.12

80
20

0.1 0.2
0.06 0.12

Intraparticle (+)
Vug (+)

95
5

0.15 0.30
0.15 0.30

Interparticle (+++)

Equant equicrystalline
mosaics of calcite spar (+++)
Drusy calcite spar (+)
Silica (+)
Micrite (+)
Equant equicrystalline
mosaics of calcite spar (+)
Micrite (++++)
Silica: microcrystalline quartz and
overgrowths on quartz grains (+++)

Nature

Percentage
(%)

Quartz, rock fragments,


feldspars, (muscovite,
tourmaline, oxides)
Foraminifers, briozoans,
molluscs, echinoderms,
(red algae)
Quartz, rock fragments,
feldspars, (mouscovite,
oxides)
Briozoans, red algae,
echinoderms, foraminifers,
molluscs
Quartz, rocks fragments
(chert, metamorphic rocks
and dolostone), feldspars
(muscovite, oxides)
Foraminifers, red algae,
echinoderms (briozoans,
echinoderm spines)
Quartz, feldespars
(muscovite, biotite,
rock fragments, tourmaline)
Foraminifers, (echinoderms,
red algae, molluscs, spicules,
radiolarians)
Quartz, feldspars (muscovite,
rock fragments)
Foraminifers
Quartz, feldspars, muscovite
(biotite, oxides, tourmaline)
Quartz
Feldspars, rock fragments

25

Ref.

Mode (mm)

1, 2, 5, 7

5, 6

3, 4, 5, 6

Nature of clasts: minor components are in brackets.


Approximate amount: ++++ high, +++ medium, ++ low, + scarce.
(1) Ordonez et al. (1994); (2) Ordonez et al. (1997); (3) Benavente et al. (1999); (4) Benavente et al. (2001); (5) Benavente (2003); (6)
Benavente et al. (2002); (7) Fort et al. (2002).

absorption and the second part defines saturation. The


slope of the curve during water absorption is the water
absorption coefficient, C. The connected porosity is
calculated because the sample is completely saturated
in the saturation part (Table 2).
In this study, a total connected porosity, P(CapHg), is defined by the combination of MIP and
capillarity. The pore size interval characterisation by
MIP ranges from 0.002 to 100 Am. Water motion is
mainly driven by capillary forces when the pore size is
in the 0.1- and 2500-Am interval (Winkler, 1997).
Therefore, the total connected porosity, P(Cap-Hg), is

calculated by linking the porosity in the 0.001 0.1


Am range, from MIP, and the 0.1 2500 Am range,
from capillary imbibition (Table 2).
The determination of the specific surface area (SSA)
of the samples was accomplished via the nitrogen
absorption technique. Calculation of the SSA was
carried out through the BET method (Rouquerol et
al., 1994) and the points in the relative pressure interval
P/P0 = 0.1 0.35 were analysed by using the ASAP
2010 Micromeritics apparatus (Table 2).
The dry bulk density, qbulk, of a rock is defined as
the ratio of its mass to its volume, including the

BR-1

BR-2

BR-3

BR-4

BR-5

BR-6

BR-7

BR-8

BC-1

BC-2

BC-3

BC-4

BC-5

BC-6

BC-7

BC-8

BM

QA

< 0.01
0.01 0.1
0.1 1
1 10
10 100

0.00
6.26
17.08
20.55
56.11

0.14
8.15
12.26
14.23
65.22

0.93
3.92
14.54
39.34
41.27

2.45
15.96
18.93
26.45
36.21

2.72
10.87
18.64
50.32
17.45

1.48
21.97
16.47
19.88
40.20

0.83
9.84
18.91
14.71
55.71

0.88
11.38
71.35
15.66
0.73

13.69
34.86
48
1.97
1.48

10.12
68.27
20.28
0.36
0.76

2.32
20.01
57.65
18.65
1.37

1.58
16.42
36.33
41.87
3.71

6.46
38.10
42.82
10.16
2.45

2.45
15.96
18.93
26.45
36.21

0.88
11.38
71.35
15.66
0.73

11.68
24.70
37.18
22.59
3.85

2.20
17.88
66.24
11.86
1.82

1.93
12.27
16.71
49.78
19.31

P(Hg) [%]
P(Cap) [%]
P(Cap-Hg) [%]
C [kg/(m2 h0.5)]
SSA [m2/g]
qbulk [g/cm3]
rF [MPa]
rC [MPa]
vP [km/s]
vS [km/s]
E [GPa]
DWL [%]

11.86
11.71
12.45
3.81
1.06
2.14
6.03
25.73
4.60
2.17
27.35
0.80

9.23
13.52
14.29
9.75
0.55
1.92
5.00
15.70
3.83
1.95
19.35
1.20

14.01
12.69
13.37
7.29
0.98
2.03
4.58
19.89
4.45
2.18
25.90
1.00

4.51
14.16
15.03
4.28
1.22
2.32
5.07
36.98
4.96
2.02
26.52
4.20

27.55
17.32
21.06
6.11
9.57
1.94
4.08
14.31
3.07
1.83
15.91
11.00

6.46
14.24
15.75
11.12
0.67
1.95
3.10
8.09
4.51
2.14
24.20
2.10

11.81
12.88
14.14
3.55
0.58
2.11
4.90
24.20
4.17
2.11
24.95
1.40

18.58
17.49
18.37
8.20
0.52
2.08
3.00
20.90
4.36
2.47
32.07
1.50

15.54
15.03
22.57
0.91
9.64
2.00
7.91

3.80

12.70
11.59
22.36
1.00
9.88
2.25
5.90
35.50
3.75
1.70
17.83
13.80

14.40
13.60
18.33
0.97
6.34
2.15
5.10
27.90
3.83
1.80
18.92
3.20

16.00
15.69
19.73
3.17
7.80
2.12
6.30
23.50
3.34
1.58
14.35
7.20

15.35
15.08
28.74
1.12
7.82
2.22
8.70
34.30
3.77
1.79
19.27
8.10

17.78
12.511
15.78
3.66
1.37
2.12
3.70
21.30

7.60

17.75
16.32
18.50
2.61
3.35
2.06
1.20
13.70

16.30

15.31
13.74
19.31
0.73
6.76
2.25
8.70
33.50
3.43
1.97
21.90
0.50

22.79
20.56
25.14
2.44
12.11
2.00
0.95
4.27
2.52
0.94
5.02
33.20

6.40
7.66
8.57
1.18
1.75
2.42
11.30
67.80
4.04
2.21
30.29
0.50

D. Benavente et al. / Engineering Geology 74 (2004) 113127

Table 2
Pore size interval; connected porosity obtained by mercury intrusion porosimetry, P(Hg), and capillarity imbibition, P(Cap); total connected porosity, P(Cap-Hg); water absorption
coefficient, C; specific surface area (SSA); bulk density, qbulk; flexural strength; rF, uniaxial compressive strength, rC; Youngs dynamic modulus, E; compressional wave velocity, vP;
shear wave velocity, vS; percentage of dry weight loss by salt crystallisation test (DWL) of the stones

119

120

D. Benavente et al. / Engineering Geology 74 (2004) 113127

volume of voids and grains. In the characterisation


test, six dried samples were used in the form of
2.5 2.5 4 cm prisms (Table 2).
2.3. Strength characterisation
In this study, the static stone strength measurement
was performed by flexural strength and uniaxial
compressive testing, which are often used in the
geomechanical characterisation of materials.
Flexural strength is the ability of a rock to withstand
bending or flexure (Winkler, 1997). This test is a
combination of compression and tension and can be
considered as an indirect measurement of tension stress
(Jaeger and Cook, 1976). In the characterisation test,
six samples were used, each measuring 300 25 40
mm. The sample was supported on two knife edges,
which were 200 mm apart. Two other knife edges,
which were 100 mm apart, were loaded onto the
sample at a rate of 0.010 kN/s (Table 2). The samples
were tested in accordance with the UNE-EN 13161
(Anon., 2002).
The uniaxial compressive strength test is used to
determine the maximum value of stress attained
before failure. In the characterisation test, six samples were used in the form of 70-mm cubes and the
load rate was 0.36 kN/s (Table 2). The samples were
tested in accordance with the UNE-EN 1926 (Anon.,
1999).
The dynamic elastic constants of a material can be
calculated from the known ultrasonic wave velocities
and stone density, using the theory of elasticity (e.g.,
Christensen, 1990; Gueguen and Palciauskas, 1994;
Tiab and Donaldson, 1996). Compressional (P) and
shear (S) waves were measured with a Sonic Viewer170 using polarised Panametric transducers (500 kHz).
In the characterisation test, six dried samples were used
in the form of 70-mm cubes (Table 2).
2.4. Salt crystallisation test
Salt crystallisation test was performed using the
continuous partial immersion test described by Benavente et al. (2001), which adequately represents the
decay process of building stones. Three samples were
used in the form of 2.5 2.5 4 cm prisms and a
14% w/w Na2SO4 solution was used. The cycle
duration was 24 h, and 15 cycles were realised. In

each cycle, two stages can be distinguished: the


capillary stage (40 jC and 80% RH) and the cooling
stage (10 jC and 70% RH). In the capillary stage, the
solution was subsaturated in mirabilite and thenardite,
and this solution rose to the top of the sample. In the
cooling stage, the solution was supersaturated in
mirabilite but subsaturated in thenardite, thus mirabilite precipitation occurred. Each stage lasted 12 h and
the time required to change temperature between
stages was 30 min. After this, the tested samples were
cleaned with pure water until salt was eliminated,
which is understood as the point where the conductivity of the solution from the cleaned stones reaches
pure water conductivity. Samples were dried (60 jC)
until they reached a constant weight. Dry weight loss
(DWL) was calculated at this stage (Table 2).

3. Results and discussion


Microporosity, Pmicro; adjusted microporosity, Padj;
PPmicro, PPmicroCS, ( PPmicro)CS, [(PPmicro)CS]0.5; water absorption, Cabs; saturation coefficient, CS; durability factor, D; and durability dimensional estimator
(DDE) are shown in Table 3.
Durability estimators have been evaluated by comparing them to the percentage of weight loss after salt
crystallisation. Table 4 shows the Pearsons correlation coefficients between weight loss after salt crystallisation and durability estimators that depend only
on pore structure, X: the connected porosity obtained
by MIP, P(Hg), and capillarity imbibition, P(Cap); the
total connected porosity, P(Cap-Hg); microporosity,
Pmicro; adjusted microporosity, Padj; PPmicro, PPmicro
CS, ( PPmicro)CS, [(PPmicro)CS]0.5; water absorption,
Cads; water absorption coefficient, C, and its reciprocal, C 1; specific surface area (SSA); saturation
coefficient, CS; durability factor, D; and durability
dimensional estimator (DDE); and on material
strength, r: flexural strength, rF; uniaxial compressive
strength, rC; Youngs dynamic modulus, E; and compressional wave velocity, vP.
Moreover, the proposed petrophysical durability
estimator (PDE) is also evaluated by comparing it to
the percentage of weight loss after salt crystallisation.
Table 5 shows the Pearsons correlation coefficients
between weight loss after salt crystallisation and the
petrophysical durability estimator with respect to

D. Benavente et al. / Engineering Geology 74 (2004) 113127

121

Table 3
Microporosity, Pmicro; adjusted microporosity, Padj; PPmicro, PPmicroCS, ( PPmicro)CS, [( PPmicro)CS]0.5; water absorption, Cabs; saturation
coefficient, CS; durability factor, D; and durability dimensional estimator (DDE) of the stones
Stone

Pmicro
[%]

Padj

PPmicro

PPmicroCS

( PPmicro)CS

[(PPmicro)CS]0.5

Cabs
[%]

CS

DDE
[Am 1]

BR-1
BR-2
BR-3
BR-4
BR-5
BR-6
BR-7
BR-8
BC-1
BC-2
BC-3
BC-4
BC-5
BC-6
BC-7
BC-8
BM
QA

2.77
1.90
2.72
1.67
8.88
2.58
3.49
2.12
15.00
12.03
13.47
10.03
15.12
6.64
14.84
11.26
19.67
1.98

0.003
0.003
0.004
0.003
0.019
0.004
0.005
0.004
0.034
0.027
0.025
0.020
0.043
0.010
0.027
0.022
0.049
0.002

10.96
13.09
11.80
13.92
13.99
13.99
12.04
16.66
11.31
12.85
9.86
12.44
14.32
11.63
9.34
11.50
10.16
7.82

0.003
0.003
0.003
0.002
0.015
0.004
0.004
0.004
0.023
0.014
0.017
0.016
0.023
0.008
0.024
0.013
0.040
0.002

0.005
0.004
0.005
0.004
0.038
0.007
0.008
0.006
0.105
0.154
0.084
0.044
0.193
0.027
0.045
0.094
0.086
0.003

0.070
0.061
0.070
0.060
0.195
0.083
0.089
0.076
0.324
0.392
0.289
0.210
0.439
0.164
0.213
0.306
0.292
0.058

3.57
9.35
6.36
2.56
9.80
8.07
5.54
6.82
0.14
5.18
6.40
6.70
5.70
5.61
7.24
5.06
10.27
3.65

0.94
0.95
0.95
0.94
0.82
0.90
0.91
0.93
0.67
0.52
0.67
0.80
0.52
0.79
0.86
0.62
0.82
0.89

11.01
12.79
12.05
13.34
14.24
12.87
11.73
15.83
10.01
6.01
8.23
12.47
7.91
9.92
13.67
7.38
16.81
6.85

0.29
0.34
0.45
0.66
2.58
0.7
0.62
0.43
6.7
4.58
3.34
2.69
7.57
1.89
1.23
5.42
3.4
0.55

We claim that the petrophysical durability estimator, with respect to the specific surface area, provides
an indication of durability given that it is inversely
related to pore size and, therefore, to salt crystallisation. The same can be said when the petrophysical
durability estimator is expressed in terms of microporosity, showing the role of smaller pore diameters on
the breaking down of the stone.
The strong correlation of the petrophysical durability estimator with respect to the durability dimensional estimator (DDE) can be attributed to the fact
that DDE provides full information about pore structure. DDE estimates stone durability because crystallisation pressure is inversely related to pore size and
extrapolates salt stress in one pore to the whole porous
stone (Ordonez et al., 1997).

flexural strength, uniaxial compressive strength and


Youngs dynamic modulus.
In general terms, there is moderate correlation
between salt weathering and the estimators that are
based on pore structure and strength, although the
high correlation of dynamic parameters (Youngs
modulus and compressional waves velocity) and
PPmicroCS should also be mentioned. However,
there is a very strong correlation between the percentage of weight loss after salt crystallisation and
the petrophysical durability estimator. These results
show the role of both pore structure and strength on
the durability of the material. In other words, they
show the relationship between the effectiveness of
salt crystallisation and the materials resistance to salt
stress.

Table 4
Pearsons correlation coefficients between weight loss after salt crystallisation test and parameters and durability estimators that are based on
pore structure and strength
P(Hg)

P(Cap)

0.556

0.626

P(Cap-Hg)

Pmicro

0.590

PPmicro

Padj

0.707

PPmicroCS

 0.187

0.723

0.830

( PPmicro)CS

[( PPmicro)CS]0.5

0.383

0.462

Cabs

C 1

SSA

CS

DDE

rF

rC

vP

0.456

 0.259

 0.017

0.641

 0.216

0.360

0.189

 0.396

 0.409

 0.718

 0.824

122

D. Benavente et al. / Engineering Geology 74 (2004) 113127

Table 5
Pearsons correlation coefficients between weight loss after salt
crystallisation test and the petrophysical durability estimator with
respect to flexural strength (PDEF), uniaxial compressive strength
(PDEC), and Youngs dynamic modulus (PDEE)
P(Hg)
P(Cap)
P(Cap-Hg)
Pmicro
Padj
PPmicro
PPmicroCS
( PPmicro)CS
[( PPmicro)CS]0.5
Cabs [%]
C
C1
SSA
CS
D
DDE

PDEF

PDEC

PDEE

0.880
0.854
0.894
0.911
0.921
0.770
0.912
0.944
0.938
0.854
0.262
0.824
0.919
0.762
0.822
0.938

0.858
0.798
0.834
0.898
0.889
0.630
0.884
0.899
0.906
0.764
0.069
0.794
0.878
0.675
0.761
0.863

0.927
0.924
0.956
0.936
0.939
0.950
0.929
0.883
0.937
0.905
0.344
0.637
0.964
0.870
0.889
0.810

In order to quantify durability from the petrophysical durability estimator, it is interesting to note that
the percentage of dry weight loss by salt crystallisation and the proposed durability estimator have similar numerical values. For this purpose, the parameter
units that form the petrophysical estimator are modified. Thus, if flexural strength and uniaxial compressive strength are commonly expressed in MPa,
Youngs dynamic modulus in GPa and DDE in
Am 1, then the durability estimator can be respectively expressed in terms of flexural strength (PDEF),
uniaxial compressive strength (PDEC) and Youngs
dynamic modulus (PDEE) as follows:
DDE
;
rF
DDE
PDEC m=kg 10
;
rC
DDE
:
PDEE mm=kg 10
E
PDEF m=kg 10

For BC-3 (Table 2), DDE = 3.34 Am 1 and rC =


27.90 MPa, so that PDEC = 1.20 m/kg. Thus, the
durability estimator has a similar numerical value to
the percentage of dry weight loss by salt crystallisation
(3.20%).

Thus, high values of the proposed durability estimator indicate lower stone durability as crystallisation
pressure is higher than the resistance to salt stress
(Table 6). Fig. 1 shows the correlation between the
percentage of dry weight loss by salt crystallisation
(DWL) and the durability dimensional estimator
(DDE), petrophysical durability estimator with respect
to flexural strength (PDEF), uniaxial compressive
strength (PDE C) and Youngs dynamic modulus
(PDEE). The poor correlation of DDE and the strong
correlation of PDE with salt damage reflect the need
to use both pore structure and strength in the estimation of stone durability. Furthermore, Fig. 1 shows
that the percentage of dry weight loss by salt crystallisation and the proposed durability estimator have
similar numerical values.
As would be expected, there is a strong correlation
between weight loss after salt crystallisation and the
petrophysical durability estimator with respect to
compressive strength. However, the best correlation
in the petrophysical durability estimator is produced
when expressed in terms of flexural strength. Thus,
the crystallisation pressure of salt creates tensile stress
over the pore surface (Scherer, 1999) and can produce
microcracking by the initiation and propagation of

Table 6
Petrophysical durability estimator with respect to flexural strength
(PDEF), uniaxial compressive strength (PDEC), and Youngs
dynamic modulus (PDEE) of the stones
Stone

PDEF
[m/kg]

PDEC
[m/kg]

PDEE
[mm/kg]

BR-1
BR-2
BR-3
BR-4
BR-5
BR-6
BR-7
BR-8
BC-1
BC-2
BC-3
BC-4
BC-5
BC-6
BC-7
BC-8
BM
QA

0.48
0.67
0.99
1.3
6.33
2.27
1.26
1.44
3.74
7.76
6.55
4.27
8.7
5.1
10.22
6.24
35.8
0.49

2.60
0.43
0.68
0.12
0.38
4.20
3.41
0.21

0.19
1.20
2.01
0.78
3.55
0.90
1.62
6.04
0.08

0.11
0.18
0.17
0.25
1.62
0.29
0.25
0.13

2.57
1.77
1.87
3.93

2.48
6.78
0.19

D. Benavente et al. / Engineering Geology 74 (2004) 113127

123

Fig. 1. Correlation between the percentage of dry weight loss by salt crystallisation (DWL) and (a) the durability dimensional estimator (DDE)
and the petrophysical durability estimator respectively expressed in terms of (b) flexural strength (PDEF), (c) uniaxial compressive strength
(PDEC) and (d) Youngs dynamic modulus (PDEE).

new cracks or the extension and widening of existing


microcracks and pores. This process can lead to
disintegration, detachment and fracturation of rock
and can therefore strongly decrease the tensile
strength of rock (Scherer et al., 2001; Nicholson,
2001). Consequently, tensile strength can be considered as the resistance of stone to salt crystallisation.
This fact explains the excellent correlation between
the salt crystallisation test and the proposed estimator
because flexural strength is an indirect measurement
of the tensile strength of the material.
The petrophysical durability estimator related to
Youngs dynamic modulus has important advantages
compared to static characterisation because it is an easy
and quick nondestructive method. However, depending
on the petrological properties of stones, the dynamic
method could result in a misleading characterisation.
Thus, for example, Youngs modulus of the QA stone,
which mainly comprises quartz, is similar to the other
samples studied (Table 2). This is due to the fact that the

vP and vS of the quartz and calcite are alike (Christensen, 1990). The excellent mechanical properties of QA
are not reflected in the dynamic characterisation as they
are in both the flexural and the compressive tests.
The evolution of pore structure and mechanical
properties produced by salt crystallisation is an important fact that is also considered by the proposed
estimator. Stress of salts in pores produces an increase
in pore size and porosity, and a decrease in rock
strength (Fitzner, 1988; Winkler, 1997; Benavente et
al., 1999; Nicholson, 2001). On the basis of pore
structure estimators, the durability of a stone with large
pores should be greater than a stone with small pores.
Therefore, an improvement may be expected in the
durability of the decayed stone. However, the durability of a decayed stone decreases because its strength is
significantly reduced after test cycles (Winkler, 1997;
Goudie, 1999).
The petrographic characteristics of stones, including grain size distribution, the type of cement and the

124

D. Benavente et al. / Engineering Geology 74 (2004) 113127

Fig. 2. Photomicrographs of thin sections of some studied stones showing different cement types: (a) drusy calcite spar (indicated by d) and
equant equicrystalline mosaics (indicated by e) of calcite spar cements in the biocalcirudite BR-1; (b) drusy calcite spar (indicated by d) and
rim or overgrowth syntaxial calcite cement (indicated by r) in the biocalcirudite BR-4; (c) equant equicrystalline mosaics of calcite spar cement
(indicated by e) in the biocalcarenite BC-4; and (d) overgrowth quartz (indicated by o) and microcrystalline silica (indicated by m) cements in
the quartz arenite.

amount of cement/matrix, considerably affect weatherability (Table 1). Thus, the durability of biocalcirudites (BR) is longer than that of biocalcarenites (BC)
because BR pores are larger than that of BCs, given
that their grain size is also greater.
The biocalcirudites (BR) studied contain large
pores which are not completely filled with cement.
The most common type of cement present on these
stones is drusy calcite spar, which is mainly comprised of prismatic calcite and does not considerably
improve the strength of the stone (Fig. 2a and b).
The most important type of cement in the biocalcarenites (BC) studied is the equant equicrystalline mosaics of calcite spar. This cement completely
fills the interparticle porosity with equant crystals of
calcite (Fig. 2a and c). The texture of this cement is
frequently a consequence of the neomorphism process (Tucker, 1990). As grains are well cemented,

disaggregation is therefore more difficult and rock


strength is much improved. This fact can account for
the excellent resistance to salt weathering of BC8 (Table 2).
Syntaxial cement or rim cement is formed by
crystals that grow in optical continuity with singlecrystal grains and which produce excellent grain
cementation. Consequently, an improvement in the
strength of the stone is achieved. However, in the
carbonate stone studied, the syntaxial calcite cement
is only a minor component because few grains are
available for calcite crystal growth and neither does it
contribute to the strength of these stones. This calcite
cement grows especially on echinoderm grains and
some clasts of altered red algae (Fig. 2b). In the
quartz arenite, the overgrowth silica cement is abundant and greatly improves its strength and, therefore,
its durability (Fig. 2d; Table 2).

D. Benavente et al. / Engineering Geology 74 (2004) 113127

The low resistance of the biomicrite (BM) to salt


crystallisation is due to the low cement/matrix ratio,
which substantially decreases its strength. Thus, if the
durability of stone is only measured using the pore
structure estimators, e.g., using DDE, then it would be
expected that the durability of BC-1, BC-2 and BC-3
is lower than BM (Table 3). However, the durability
of BM is lower than that of BC-1, BC-2 and BC-3 due
to low stone strength (Table 2). Previous examples
show the importance of the petrography characterisation in the understanding of the durability of porous
building stones.
We consider that the proposed petrophysical durability estimator is a novel expression for estimating
stone durability and a significant improvement in
what has been used in the past. The proposed estimator provides full information on those properties
which essentially have an influence on the stones
weatherability, pore structure and strength. It is also
accurate in comparison with the experimental data and
is very easy to apply.
The proposed durability estimator does not consider
several factors such as lithologial variability and the
pore structure, e.g., the presence of cracks and fractures, pore shape, topology of the pore network.
Moreover, this estimator is only compared with salt
weathering and does not take into account other important decay mechanisms such as frost action, wetting
and drying cycles, thermal stress and acid attack.
Further research into a wider range of rock types and
durability tests are needed in order to be able to fully
apply the proposed estimator to stone durability.

4. Conclusions
A novel petrophysical durability estimator is proposed as the ratio of parameters or estimators that are
based on pore structure to strength of the material.
Our study shows that there is a strong correlation
between salt weathering and the petrophysical durability estimator. However, there is only moderate
correlation with the estimators that depend on pore
structure and strength. These results demonstrate the
importance of both pore structure and strength when
estimating stone durability.
In particular, we consider that the proposed estimatordefined as the ratio of the durability dimen-

125

sional estimator (DDE) to the flexural strength of the


stonecontains the fullest information for understanding and estimating the durability of stones. On
one hand, DDE is related to connected porosity and
pore size distribution and adequately represents the
salt stress in the whole porous stone, while, on the
other hand, the flexural strength of the material is an
indirect measurement of tensile strength which can
be considered as the resistance of stone to salt
crystallisation.
The practical application of the proposed estimator
is that a single-mercury porosimetry measurement and
strength characterisation (principally flexural
strength) of the stone be used to predict its durability.
Consequently, this estimator saves much time and
provides significant practical advantages for a rapid
durability assessment given that the estimator is
defined from the stones petrophysical parameters
which are themselves usually obtained for its initial
characterisation.

Acknowledgements
This study was supported by the Research Project
MAT 2000-0744 (MCYT, Spain) and a predoctoral
research fellowship from Generalitat Valenciana
awarded to D. Benavente. The authors would like to
thank D.T. Nicholson and another (anonymous)
reviewer for their many incisive comments and
criticisms in revising this article. Special thanks to
G.W. Scherer (Princeton University) and R. Flatt
(Sika AG, Zurich) for their helpful comments and
suggestions.

References
Anon., 1983. The selection of natural building stone. Digest. 260.
Watford: Building Research Establishment. Her Majestys Stationary Office.
Anon., 1999. Metodos de ensayo para piedra natural. Determinacion de la resistencia a la compresion. UNE-EN 1926.
Anon., 2002. Metodos de ensayo para piedra natural. Determinacion de la resistencia a la flexion a momento constante. UNEEN 13161.
Bell, F.G., 1993. Durability of carbonate rock as building stone with
comments on its preservation. Environmental Geology 21,
187 200.

126

D. Benavente et al. / Engineering Geology 74 (2004) 113127

Benavente, D., 2003. Modelizacion y estimacion de la durabilidad de materiales petreos porosos frente a la cristalizacion
de sales. Biblioteca Virtual Miguel de Cervantes, http://www.
cervantesvirtual.com/FichaObra.html?Ref=12011, Accessed:
18/02/2004.
Benavente, D., Garca del Cura, M.A., Fort, R., Ordonez, S., 1999.
Thermodynamic modelling of changes induced by salt pressure
crystallisation in porous media of stone. Journal of Crystal
Growth 204, 168 178.
Benavente, D., Garca del Cura, M.A., Bernabeu, A., Ordonez, S.,
2001. Quantification of salt weathering in porous stones using
an experimental continuos partial immersion method. Engineering Geology 59, 313 325.
Benavente, D., Lock, P., Garca del Cura, M.A., Ordonez, S., 2002.
Predicting the capillary imbibition of porous rocks from microstructure. Transport in Porous Media 49, 59 76.
Christensen, N.I., 1990. Seismic velocities. In: Carmichael, R.S.
(Ed.), Practical Handbook of Physical Properties of Rocks and
Minerals. CRC Press, Boca Raton, FL, pp. 429 546.
Correns, C.W., 1949. Growth and dissolution of crystals under
linear pressure. Faraday Discussions 5, 267 271.
Dullien, F.A.L., 1992. Porous Media Fluid Transport and Pore
Structure. Academic Press, San Diego, CA.
Dullien, F.A.L., El-Sayed, M.S., Batra, V.K., 1977. Rate of capillary rise in porous media with nonuniform pores. Journal of
Colloid and Interface Science 60, 497 506.
Fitzner, B., 1988. Porosity properties of naturally or artificially
weathered sandstone. In: Ciabach, J. (Ed.), Proceedings of
the 6th International Congress on Deterioration and Conservation of Stone, Torun, Poland, pp. 236 245.
Fitzner, B., Snethlage, R., 1982. Ueber Zusammenhange zwischen
salzkristallisationsdruck und Porenradienverteilung. G.P. Newsletter 3, 13 24.
Flatt, R.J., 2002. Salt damage in porous materials: how high
supersaturations are generated. Journal of Crystal Growth
242, 435 454.
Folk, R., 1962. Spectral subdivision of limestone types. In:
Ham, W.E. (Ed.), Classification of Carbonate Rocks. Memoir-American Association of Petroleum Geologists, vol. 1,
pp. 62 84.
Fort, R., Bernabeu, A., Garca del Cura, M.A., Lopez de Azcona,
M.C., Ordonez, S., Mingarro, M., 2002. Novelda stone: a stone
widely used within the Spanish heritage. Materiales de Construccion 52, 19 32.
Gauri, K.L., Chowdhuy, A.N., Kulshreshtha, N.P., Punuru, A.R.,
1988. Geologic features and the durability of limestones at the
Sphinx. In: Marinos, K. (Ed.), Engineering Geology of Ancient
Works, Monuments and Historical Sites. A.A. Balkema, Rotterdam, pp. 723 729.
Goodman, R.E., 1989. Introduction to Rock Mechanics, 2nd edition.
Wiley, New York.
Goudie, A.S., 1993. Salt weathering simulation using a single-immersion technique. Earth Surface Processes and Landforms 18,
368 376.
Goudie, A.S., 1999. Experimental salt weathering of limestone in
relation to rock properties. Earth Surface Processes and Landforms 24, 715 724.

Goudie, A.S., Viles, H.A., 1995. The nature and pattern of debris
liberation by salt weathering: a laboratory study. Earth Surface
Processes and Landforms 20, 437 449.
Gregg, S.J., Sing, K.S.W., 1982. Adsorption, Surface Area, and
Porosity, 2nd edition. Academic Press, London.
Gueguen, Y., Palciauskas, V., 1994. Introduction to the Physics of
Rock. Princeton Univ. Press, Princeton, NJ.
Hammecker, C., Jeannette, D., 1994. Modelling the capillary
imbibition kinetics in sedimentary rocks: role of the petrographical features. Transport in Porous Media 17, 285 303.
Hammecker, C., Mertz, J.D., Fischer, C., Jeannette, D., 1993. A
geometrical model for numerical simulation of capillary imbibition in sedimentary rocks. Transport in Porous Media 12,
125 141.
Hirschawald, J., 1908. Die Prufung der naturlichen Bausteine auf
ihre Wetterbestandigkeit. Ernst & Sonh, Berlin.
Jaeger, J.C., Cook, N.G.W., 1976. Fundamentals of Rock Mechanics,
2nd edition. Chapman & Hall, New York.
Jefferson, D.P., 1993. Building stone: the geological dimension.
Quarterly Journal of Engineering Geology 26, 305 319.
Leventis, A., Verganelakis, D.A., Halse, M.R., Webber, J.B.,
Strange, J.H., 2000. Capillary imbibition and pore characterisation in cement pastes. Transport in Porous Media 39, 143 157.
Modd, B.K., Howarth, R.J., Bland, C.H., 1996. Rapid prediction of
building research establishment limestone durability class from
porosity and saturation. Quarterly Journal of Engineering Geology 29, 285 297.
Nicholson, D.T., 2001. Pore properties as indicators of breakdown
mechanisms in experimentally weathered limestones. Earth Surface Processes and Landforms 26, 819 838.
Ordonez, S., Garca del Cura, M.A., Fort, R., Louis, M., Lopez
De Azcona, M.C., Mingarro, F., 1994. Physical properties
and petrographic characteristics of some Bateig stone varieties. Proceedings 7th International IAEG Congress, Lisboa;
V. Balkema, Rotterdam, pp. 3595 3603.
Ordonez, S., Fort, R., Garca del Cura, M.A., 1997. Pore size distribution and the durability of a porous limestone. Quarterly
Journal of Engineering Geology 30, 221 230.
Richardson, B.A., 1991. The durability of porous stones. Stone
Industries 26 (10), 22 25.
Rouquerol, J., Avnir, D., Fairbridge, C.W., Everett, D.H., Haynes,
J.H., Pernicone, N., Ramsay, J.D.F., Sing, K.S.W., Unger, K.K.,
1994. Recommendations for the characterization of porous solids.
Pure and Applied Chemistry 66, 1739 1758.
Scherer, G.W., 1999. Crystallisation in pores. Cement and Concrete
Research 29, 1347 1358.
Scherer, G.W., 2000. Stress from crystallisation of salt in pores. In:
Fassina, V. (Ed.), Proceedings of the 9th International Congress
on Deterioration and Conservation of Stone, Venice. Elsevier,
pp. 187 194.
Scherer, G.W., Flatt, R., Wheeler, G., 2001. Materials science research for the conservation of sculpture and monuments. MRS
Bulletin 26, 44 50.
Sperling, C.H.B., Cooke, R.U., 1985. Laboratory simulation of
rock weathering by salt crystallisation and hydration processes in hot, arid environments. Earth Surface Processes 10,
541 555.

D. Benavente et al. / Engineering Geology 74 (2004) 113127


Tucker, M.E., 1990. Diagenetic processes, products and environments. In: Tucker, M.N., Wrigh, V.P. (Eds.), Carbonate Sedimentology. Blackwell, Oxford, pp. 314 336.
Tiab, D., Donaldson, E.C., 1996. Petrophysics: Theory and Practice
of Measuring Reservoir Rock and Fluid Transport Properties.
Gulf Pub. Co., Houston, TX.
Valdeon, L., de Freitas, M.H., King, M.S., 1996. Assessment of the
quality of building stones using signal processing procedures.
Quarterly Journal of Engineering Geology 29, 299 308.

127

Wellman, H.W., Wilson, A.T., 1965. Salt weathering, a neglected


geological erosive agent in coastal and arid environments. Nature 205, 1097 1098.
Winkler, E.M., 1997. Stone in Architecture: Properties, Durability,
3rd edition. Springer-Verlag, Berlin.
Zezza, U., 1990. Physical mechanical properties of quarry and
building stones. In: Veniale, F., Zezza, U. (Eds.), Analytical
Methodologies for Investigation of Damage Stones, Pavia,
pp. 1 20.

S-ar putea să vă placă și