Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Journal of Cultural Heritage 4 (2003) 109–115

www.elsevier.com/locate/culher

Original article

Crystallization damage by sodium sulfate


Nicholas Tsui a, Robert J. Flatt b,*, George W. Scherer b
a
Materials Science and Engineering, 8-032, MIT, Massachusetts Avenue 77, Cambridge, MA 02139, USA
b
Department of Civil and Environmental Engineering, E-Quad, Princeton University, Princeton, NJ, USA
Received 20 May 2002; accepted 21 January 2003

Abstract

Experiments demonstrate that a stone containing thenardite suffers great damage when exposed to water below the temperature limit of
mirabilite stability. This is due to a transition between thenardite and mirabilite, and not to thenardite reprecipitation. Damage occurs whether
or not thenardite was produced previously by mirabilite decomposition. Together with recent results from the literature, these results indicate
that damage occurs because thenardite dissolution can produce solutions highly supersaturated with respect to mirabilite, so that precipitation
of this mineral can lead to large crystallization pressures. Finally, it appears that there is a salt content threshold beyond which damage
increases substantially.
© 2003 Éditions scientifiques et médicales Elsevier SAS. All rights reserved.

Keywords: Sodium sulfate; Mirabilite; Thenardite; Crystallization pressure; Salt; Deterioration; Durability

1. Research aims commonly encountered, sodium sulfate is known to be the


worst, which is why it has become widely used in accelerated
The general aim of the research we are conducting on durability testing [3]. There are two major ways in which
crystallization is to develop a detailed understanding of the such tests are run. In the first, samples undergo cycles of
mechanisms by which salts can damage building materials. It impregnation in a saline solution followed by drying. In the
is through the careful examination of the thermodynamics of second, samples are in permanent contact at their base with a
this problem that a novel solution to preventing salt damage saline solution, while low humidity is imposed above to drive
has emerged. continuous capillary rise compensating for water evapora-
In the present study, we focus on the case of sodium tion.
sulfate. A detailed examination of the magnitude of the
stresses involved has been presented elsewhere. In this paper, Two main mechanisms are proposed to explain the exten-
we present simple experimental results that support the idea sive damage caused by sodium sulfate: hydration pressure
that sodium sulfate is so damaging because of a process of and crystallization pressure. In fact, it has been shown that
dissolution and precipitation rather than volume change as is the expression for hydration pressure accurately predicts the
usually stated. As we explain in the conclusion section, this upper bound of crystallization pressures [4]; nevertheless,
finding has serious implication as to the relevance of crystal- even though the predicted pressures are the same, it is impor-
lization tests commonly used to judge the durability of build- tant to know which phenomenon is responsible for the ob-
ing stone. served damage. Hydration results in a large volume expan-
sion (about 314%) as the anhydrous phase of sodium sulfate
(thenardite) converts to its decahydrate phase (mirabilite).
2. Introduction However, recent use of the environmental scanning electron
Crystallization of salts in stones or concrete is an impor- microscope (ESEM) demonstrated that this transition occurs
tant mode of decay of these materials [1,2]. Of all the salts by dissolution and precipitation of mirabilite [5], ruling out
hydration pressure as a credible cause for the extensive dam-
* Corresponding author. Present address: Sika AG, Corporate Research
age caused by sodium sulfate. It is important to point out that
and Analytics, Tüffenwies 16, 8048 Zurich, Switzerland. this precipitation involves an overall expansion [6] that might
E-mail address: flatt.robert@ch.sika.com (R.J. Flatt). cause a destructive hydraulic pressure as in the case of freez-
© 2003 Éditions scientifiques et médicales Elsevier SAS. All rights reserved.
DOI: 10.1016/S1296-2074(03)00022-0
110 N. Tsui et al. / Journal of Cultural Heritage 4 (2003) 109–115

crystallization stress that is determined by the disjoining


forces of the salt/mineral interface. In the case of mirabilite,
it is conceivable that this maximum is higher than for other
salts [13], though we expect that stone and concrete are too
weak to resist crystal growth up to the maximum crystalliza-
tion pressures of most salts [11]. Therefore, as a general rule,
crystallization damage in these materials will depend on the
degree of supersaturation that can be achieved, the pore size
distribution and the tensile strength of these materials, rather
than on chemistry or mineralogy of the host [2,8,11].
Accepting that crystallization pressure is the source of
damage, Rodriguez-Navarro et al. [14] recently argued that
precipitation of thenardite rather than mirabilite may be the
culprit in damage from sodium sulfate, particularly in situa-
tions of constant capillary rise. These authors found that
Fig. 1. Phase diagram for sodium sulfate. The continuous lines indicate the
boundaries of the stable phases. Triangles and squares are experimental data under conditions of fast evaporation driven by low ambient
for mirabilite and thenardite, respectively [10]. The discontinuous line relative humidity, both mirabilite and thenardite could be
corresponds to a solution in metastable equilibrium with respect to thenar- precipitated, instead of the precipitation of mirabilite and
dite and supersaturated with respect to mirabilite. subsequent dehydration expected from equilibrium thermo-
dynamics. With even lower relative humidities they were also
ing of water [7]. The magnitude of this pressure is likely to be
able to precipitate only thenardite. Since these authors also
relatively small compared to ice because of the limited ex-
observed that in experiments of constant capillary rise, dam-
pansion involved and the probably low crystallization rates
age increased substantially with low relative humidities and
[8]2.
that the spallings of their samples contained thenardite of
On the other hand, sodium sulfate can develop particularly
similar morphology to the one produced in their experiments
large crystallization pressures. The first convincing argument
described above, they inferred that thenardite crystallization
for this idea was presented by Chatterji and Jensen [9]. They
was probably responsible for the enhanced damage at low
noted that the solubility of thenardite becomes increasingly
relative humidity. On the other hand, the fact that in some
larger than that of mirabilite below 32 °C. Therefore, any
instances, both minerals form indicates that mirabilite is
porous material containing thenardite that is exposed to wa-
precipitating under conditions of high supersaturation and
ter or rising humidity will host both dissolution of thenardite
that the concentration in the solution must be at least satisfy-
and precipitation of mirabilite. The key aspect is that the
ing thenardite saturation. Using thermodynamic data to ana-
thenardite dissolves until it reaches its equilibrium concen-
lyze the conditions of this experiment, it was found that
tration, which at ambient temperature is highly supersatu-
mirabilite precipitation alone could generate stresses suffi-
rated with respect to mirabilite as indicated in Fig. 1 (data are
cient to damage most stones and concrete [8]. This demon-
from Kracek [10], construction of equilibrium lines is de-
strates that contribution from thenardite crystallization pres-
tailed in [8]). Such situations are systematically repeated in
sure is not a necessary condition, though it does not exclude
the durability test that involves cycles of impregnation and
the possibility of its contribution.
drying. The expected tensile stresses are substantially larger
Another observation of Rodriguez-Navarro et al. [14] is
than the tensile strength of most stone and concrete, which
that when thenardite that was formed by mirabilite dehydra-
means that mirabilite precipitation alone is sufficient to dam-
tion is exposed to increasing humidity at ambient tempera-
age these materials during the rewetting [8].
ture, then precipitation of both phases can be observed (mira-
These stresses originate from repulsive surface forces
bilite appears and thenardite recrystallizes). In this case, the
between the pore walls of most inorganic building materials
formation of thenardite can be seen as an attempt to minimize
and salt crystals [11,12] and each system has a maximum
the surface energy of the sub-micron crystallites produced
during mirabilite decomposition, a process that leads to
2
Hydraulic pressure is created during freezing, because water is pushed pseudo-morphic substitution for mirabilite crystals. This ob-
away from the site of growing ice crystals, owing to the expansion in volume servation again raises the question about which sodium sul-
as ice forms [7]. As water is driven through the pores ahead of the advancing
fate phase causes most damage by precipitation. This time
crystals, frictional resistance between water and the pore walls produces a
back-pressure. The magnitude of this pressure increases with the rate of however, the question applies to situations of cyclic impreg-
growth of the crystals, so lower temperatures and faster cooling rates nation and drying rather than constant capillary rise.
increase the damage from hydraulic pressure during freezing. In the case of In this paper, we present an approach similar to that of
mirabilite crystallization there is a volume increase, so water is displaced Jensen and Chatterji [9], who subjected bricks to impregna-
during growth. However, the magnitude of the volume change is much
smaller (nine times at 20 °C) and the rate of growth of salt crystals is far
tion cycles above 32 °C and then split these into two series for
slower than that of ice, so the hydraulic pressure generated by mirabilite further cycles. One series impregnated at 30 °C did not
crystallization is expected to be very small. undergo any damage, while the other series impregnated at
N. Tsui et al. / Journal of Cultural Heritage 4 (2003) 109–115 111

20 °C decayed rapidly. Results were used to infer the impor-


tance of mirabilite precipitation as the cause of damage.
However, in view of the above results they are insufficient to
discount the possibility that recrystallization of thenardite
could be the ultimate cause of damage during rewetting. The
slightly more complete experimental approach presented
here demonstrates clearly that it is indeed mirabilite that
causes damage in such situations. Furthermore, it is found
that the magnitude of damage greatly increases once enough
salt is present for mirabilite to fill all pores. This implies that
crystallization stresses will develop throughout the sample,
propagating strength limiting flaws. Durability classifica-
tions resulting from this test will, therefore, not necessarily
correlate with real exposure conditions in which stresses are
developed more locally by supersaturated solutions produced
by evaporation.

3. Experimental

3.1. Materials

Indiana limestone was cut into 2.5 cm cubes. The apparent


density of this stone was determined to be 2.31 g/cm3. The
skeletal density measured by helium pycnometry was found
to be 2.69 g/cm3. This gives a total porosity of 14%.
The sodium sulfate used was analytical grade obtained
from Merck and all water was demineralized.

3.2. Methods

Twenty-nine samples were subjected to cycles of impreg-


nation and drying. They were separated into three groups A, Fig. 2. Flow chart of experimental protocol.
B and C. All samples were initially dried to constant mass at The weighed samples were then placed on individual glass
105 °C. Petri dishes previously weighed. They were placed into a
The 10 samples in group A were impregnated at ambient drying oven at 105 °C until constant mass was reached. At the
temperature (20 °C) with mirabilite saturated solutions (16% end of the drying step, the Petri dish and its contents were
thenardite by mass). The 10 samples in group B were impreg- weighed again to determine the mass of water lost. Decoher-
nated at 50 °C with the same solutions. The nine samples of ing pieces were then scraped off the remaining sample. The
group C were also impregnated at 50 °C but with solutions of remaining sample and its residue were weighed separately.
thenardite saturated at that temperature (31% thenardite by This procedure was repeated six times for each group as
mass). indicated in Fig. 2.
Each sample was impregnated in its individual closed At the end of the sixth cycle, each group was split into two.
screw-capped polypropylene flask (ID: 45 mm, h: 43 mm). The first half was impregnated at 20 °C and the second at
The mass of the polypropylene flask was measured. The 50 °C.
solution volume was chosen to be slightly greater than the
pore volume in the sample to limit loss of salt by diffusion out
of the stone. Prior to impregnation, samples were always left 4. Results
to equilibrate thermally either in the dry oven at 50 °C or in a
closed container at room temperature. If the stones impreg- In Fig. 3, the average mass fraction lost per cycle during
nated at 50 °C were not preheated, precipitation at pore both the impregnation and the drying steps is reported. In
entries would occur upon contact with the solution and the each graph, the two sub-groups that received a different
amount of solution absorbed into the stone was much lower. treatment in the seventh cycle are plotted separately, along
Impregnations lasted 6 h, a time found to be sufficient to with the error bar for the 95% confidence interval on the
fully saturate these samples. The masses of the wet samples average. It can be seen that in each case, both sub-groups are
and the mix of residue and solution remaining in the flask indistinguishable before the seventh cycle. Any difference at
were then measured. Samples impregnated at 50 °C were that point cannot, therefore, be attributed to a difference in
taken individually out of the oven directly to be measured. average properties of those sub-groups.
112 N. Tsui et al. / Journal of Cultural Heritage 4 (2003) 109–115

Fig. 3. Mass loss evolution during impregnation and drying cycles. Filled symbols indicate impregnation at 20 °C, open symbols 50 °C. Diamonds indicate
sub-groups with same impregnation temperature for all cycles. Squares indicate sub-groups that had a different impregnation temperature in cycle 7. Groups A
and B had solutions saturated at 20 °C, while group C had a solution saturated at 50 °C. Group A was run at 20 °C for cycles 1–6, while groups B and C were
run at 50 °C for those same cycles.

Considering first the cycles 1–6, it is seen that only group at 20 °C and of thenardite at 50 °C, which are respectively the
A is damaged and that this occurs during the impregnation most stable phases at those temperatures. Degradation is
and not the drying step. That group was impregnated at 20 °C much greater in the second case because more salt has been
for cycles 1–6, while the other two were impregnated at accumulated.
50 °C. Appendix A shows how to determine the mass of stone
In the seventh cycle, each group was split into two sub- and salt within the sample at each cycle. Fig. 4 indicates that
groups. The first sub-group was impregnated at 20 °C and the in cycle 6, for group A, there is a simultaneous increase in the
second one at 50 °C. Fig. 3 indicates that regardless of the mass loss accompanied by a decrease in the fraction of stone
past history, degradation systematically occurs when impreg- within the sample. The salt content at that point of time
nation is done at 20 °C, and that no degradation occurs at corresponds to about 4% of the stone mass. In this case, it is
50 °C. Once again, no significant degradation occurs during the average of all samples of group A that has been plotted.
the drying.
Groups B and C are distinguished by the fact that the 5. Discussion
concentrations of the impregnation solutions are respec-
tively, 16% and 31%. These concentrations, given as mass All results indicate that damage occurs when the impreg-
fraction of thenardite, correspond to saturation of mirabilite nation is done at 20 °C. Results for groups B and C in the
N. Tsui et al. / Journal of Cultural Heritage 4 (2003) 109–115 113

during the next impregnation. Furthermore, the degree of


damage between the sub-groups of groups B and C impreg-
nated at 20 °C in cycle 7 is very different (5% vs. 30% weight
loss). The salt contents of those samples at the end of cycle 6
are respectively 3.2% and 7% salt mass with respect to the
stone. These values happen to be on each side of the critical
salt content estimated above, supporting the idea that exten-
sive damage occurs once the crystals no longer have the
possibility of growing in unrestrained directions. Having
pointed this out, a word of caution as to the exactitude of the
estimated salt contents of the samples must be made. Indeed,
the calculation described in Appendix A assumes that the dry
loss contains proportions of salt and stone equal to those
Fig. 4. Evolution of the mass the sample and the stone it contains (the
sample is the total of the remaining stone and the salt it contains). The found in the bulk sample. In reality, particularly during the
relative salt content (given with respect to the mass of stone in the sample) is first cycles the salt content will be higher due to efflores-
also reported. All data in this figure are for group A. cence. As a result, salt contents are somewhat over-
seventh cycle are particularly instructive since they show that estimated. However, the consequence should not be too large
severe damage occurs at 20 °C, even though the thenardite since dry loss is small.
that is present in the samples would not have been produced The substantial degree of damage reached in the ultimate
by mirabilite decomposition, since all prior impregnations cycles of the testing procedure involving cycles of impregna-
were performed at a temperature above the limit of stability tion and drying, therefore, seems intimately linked to the
of this mineral (Fig. 1). On the other hand, the sub-group of inability of crystals to find possibilities for unrestrained
samples from group A that is impregnated at 50 °C in the growth. Consequently stress fields develop over zones large
seventh cycle would undergo recrystallization of thenardite, enough to propagate strength-limiting flaws, which explains
but no decay is observed. This demonstrates that, indepen- the large scale of damage observed. In real conditions of
dent of the way thenardite was obtained, damage occurs only exposure, it is unclear whether such a situation would occur.
when mirabilite is allowed to crystallize. In particular, for salts like sodium chloride, evaporation-
We can, therefore, state that the conversion from thenard- driven supersaturation only develops in a limited zone quite
ite to mirabilite is the cause of damage. Such results could be close to the surface. In such a situation, durability would be
equally used in favor of hydration or crystallization pressure dictated by the smaller flaws, which require higher stresses to
as the source of damage. To satisfy the proposed mechanism be propagated. Damage would be expected to be more granu-
of hydration pressure, the crystals of thenardite would have lar as is often observed in field exposure. In view of this, the
to undergo a continuous expansion by progressive integra- relevance of the durability test involving cycles of impregna-
tion of water into their lattice, eventually forming mirabilite. tion in a sodium sulfate solution followed by drying must be
Due to lattice mismatches, such a process is more than questioned.
unlikely. Furthermore, as mentioned earlier, the experimental
observation that this conversion takes place through dissolu- 6. Conclusions
tion and precipitation [5] establishes the validity of the crys-
tallization pressure view-point over the one of hydration The origin of the extensive damage that sodium sulfate
pressure. In addition, calculations show that if mirabilite can cause in durability testing that involves cycles of impreg-
crystallizes from a solution at the saturation concentration of nation and drying is linked to mirabilite precipitation that
thenardite, then mirabilite precipitation can generate tensile occurs during the wetting step. Damage is severe because
stresses in excess of the tensile strength of most stones and when water enters the thenardite-containing material, disso-
concrete [8]. lution of this mineral creates a solution highly supersaturated
An important factor in the crystallization inside porous with respect to mirabilite. The concentration is high enough
materials is that, whenever given the possibility, a crystal for mirabilite precipitation to generate stresses in excess of
would rather grow in directions where it does not encounter the tensile strength of most stones and concrete. This process
resistance. One, therefore, expects damage to increase sub- does not rely on thenardite having been obtained by mirabi-
stantially once enough salt is present to fill the porous net- lite dehydration, implying that thenardite recrystallization
work. For the stone we are dealing with, the porosity is 14%. does not contribute to damage in such situations.
The molar volume of mirabilite being 220.7 cm3/mol and the Damage appears to increase abruptly as soon as enough
molar mass of thenardite being 144.0 g/mol, we would ex- salt is present to prevent unrestrained precipitation of mira-
pect such an increase to occur once the salt contents reaches bilite during impregnation. This implies that stresses are
3.9 wt.% of thenardite. exerted over a large fraction of the sample’s volume, a situa-
For the samples of group A, this condition is reached at tion that would lead to a different pattern of damage than in
cycle 5 (Fig. 3) and there is indeed a clear increase in damage field conditions for which the presence and amount of
114 N. Tsui et al. / Journal of Cultural Heritage 4 (2003) 109–115

salt
smaller flaws requiring higher stresses to be propagated We assume that moven is split between the dry residue and
would dictate the durability of these materials. Due to this, the sample in proportions equal to their masses, so the
salt
the ability of this durability test to produce a classification of amount of salt in the residue mresidue, dry is:

冉 冊
resistances that is relevant for in situ exposure must be given
S
further consideration. mdry, final
= 1−
salt salt
mresidue, dry moven P (A.7)
mdry, rest
Acknowledgements S
where mdry, final is the remaining dried sample, mdry, rest is the
P

dried residue mass.


N. Tsui was supported during the summer of 2000, by the
We can now determine the material lost during drying, not
NSF program Research Experience for Undergraduates.
including salt:
Support for R. Flatt was provided by a postdoctoral grant
from the Swiss National Science Foundation, and financial mlost, dry = mdry, rest − mresidue, dry
S P salt
(A.8)
support from Lafarge Co. and the National Center for Pres-
ervation Teaching and Training (Award No. MT-2210-9-NC- To get the amount of salt and stone in the remaining
21). sample, we assume that they are during impregnation in
identical proportions as found in the sample prior to that step:
salt
Appendix A. Calculations minitial
=
S S
mlost, wet mlost, wet S (A.9)
minitial
We determine the mass loss during impregnation and
drying cycles, as well as the mass of remaining stone (with- The amount of salt in the stone at the end of each cycle is
out salt content). For impregnation: given by:
mempty + madded + mdry, initial = mrest + mwet + mlost
F sol S F S W
(A.1) mfinal = minitial − mlost, wet + moven − mresidue, dry (A.10)
salt salt salt salt salt
F sol
where: mempty is the empty polypropylene flask mass, madded is
S
the solution mass, mdry, initial is the dry sample mass at the The mass of the stone remaining in the sample at the end
F of each cycle is then:
cycle start, mrest is flask and residue mass once the sample has
S
been removed after impregnation, mwet is the wet sample mfinal = minitial − mfinal
stone S salt
(A.11)
W
mass after impregnation, mlost is water lost probably because
of a badly closed flask.
F sol
The term mrest includes the remaining solution mflask and
S
the residue mlost, wet: References
= + +
F F sol S
mrest mempty mflask mlost, wet (A.2)
[1] A. Arnold, K. Zehnder, Salt weathering on monuments, in: F. Zezza
A mass balance over water gives:
(Ed.), Proceedings of the First International Congress on the Conser-
madded = msample + mFlask + mlost
W W W W vation of Monuments in the Mediterranean Basin, Bari, 7–10 June,
(A.3)
Grafo, Brescia, Italy, 1989, pp. 31–58.
W
where is the mass lost by the sample during drying.
msample
[2] G.W. Scherer, R.J. Flatt, G. Wheeler, Materials science research for
If the solution and water are identically split between the conservation of sculpture and monuments, MRS Bulletin 26 (1)
stone and flask then: (2001) 44–50.
W
msample [3] C.A. Price, Testing porous building stone, The Architects Journal 2
= −
sol sol W
msample 共 madded mlost 兲 (A.4) (33) (1975) 337–339.

W W
madded mlost
sol
[4] R.J. Flatt, G.W. Scherer, Hydration and crystallization pressure of
where msample is the amount of solution in the sample. In the sodium sulfate: a critical review, in: P.B. Vandiver, M. Gordway,
flask it is: J.L. Mass (Eds.), Proceedings of Materials Research Society, 712,
Materials Issues in Art and Archeology VI, Materials Res. Soc.,
W
mflask Warrendale, PA (2002) 29–34.
mflask = 共 madded − mlost 兲
sol sol W
(A.5)
madded − mlost
W W
[5] C. Rodriguez-Navarro, E. Doehne, Salt weathering: influence of
S evaporation rate, supersaturation and crystallization pattern, Earth
Eqs. (1)–(5) allow to calculate wet, the material lost
mlost,
Surface Process Landforms 24 (1999) 191–209.
during impregnation. Determining the mass lost during dry-
S [6] D.J. McMahon, P. Sandberg, K. Folliard, P.K. Mehta, Deterioration
ing, mlost, dry is more complex because of the prior salt uptake.
salt mechanisms of sodium sulfate, in: J. Delgado Rodrigues, F. Hen-
The salt mass moven in the dried sample and residue is:
riques, F. Telmo Jeremias (Eds.), Proceeding of the Seventh Interna-
mflask − madded
W W tional Congress on Deterioration and Conservation of Stone, Lisbon,
moven = msample
salt sol 15–18 June, Laboratório Nacional de Engenharia Civil, Lisbon, Por-
(A.6)
madded − mlost
sol W
tugal, 1992, pp. 705–714.
N. Tsui et al. / Journal of Cultural Heritage 4 (2003) 109–115 115

[7] T.C. Powers, The air requirement of Frost resistant concrete, Proceed- [12] G.W. Scherer, Stress from crystallization of salt in pores, in: V. Fassina
ings of Highway Research Board 29 (1949) 184–211. (Ed.), Proceedings of the Ninth International Congress on Deteriora-
[8] R.J. Flatt, Salt damage in porous materials: how high supersaturations tion and Conservation of Stone, Venice, 19–25 June, Elsevier,,
are generated, Journal of Crystal Growth 242 (2003) 435–454. Amsterdam, 2000, pp. 187–194.
[9] S. Chatterji, A.D. Jensen, Efflorescence and breakdown of building [13] G.W. Scherer, Reply to the discussion by S. Chatterji of the paper,
materials, Nordic Concrete Research 8 (1989) 56–61. “Crystallization in pores”, Cement and Concrete Research 20 (2000)
[10] F.C. Kracek, International Critical Tables 3, 1928, pp. 351 (NaCl: 673–675.
p. 370; Na2SO4: p. 371). [14] C. Rodriguez-Navarro, E. Doehne, E. Sebastian, How does sodium
[11] G.W. Scherer, Crystallization in pores, Cement and Concrete sulfate crystallize? Implications for the decay and testing of building
Research 29 (1999) 1347–1358. materials, Cement and Concrete Research 30 (2000) 1527–1534.

You might also like