You are on page 1of 16

Geochemical Journal, Vol. 37, pp.

351 to 366, 2003

Na-carbonate waters of extreme composition:


Possible origin and evolution

GIAMPIERO VENTURELLI,1* TIZIANO BOSCHETTI 1 and VITTORIO DUCHI 2


1
Department of Earth Sciences, Parco Area delle Scienze 157/A, 43100 Parma, Italy
2
Department of Earth Sciences, via A. Lamarmora 4, 50121 Firenze, Italy

(Received July 18, 2002; Accepted December 27, 2002)

Na-carbonate (NaCW) waters are not concentrated in well defined areas, but usually widespread in
areas where other water types (e.g., Ca-carbonate) are dominant. NaCW are the product of long-term
water-rock interaction with dissolution of Na-silicates in presence of phyllosilicates, silica phases, and
calcite. NaCW circulating in calcite-bearing sediments very probably have a Ca-carbonate parent with
moderate to low PCO2 , which changes its composition assuming increasing Na character as the water-rock
interaction proceeds. However, the moderate to low PCO2 values of the potential parent Ca-carbonate
waters do not account for the high Na content of many NaCW. The higher PCO2 required may be due to
oxidation of organic matter of the sediments and, perhaps, to further addition of CO2 coming from deeper
crustal levels, from the mantle, or from other sources. Na-Ca exchange involving a Na-exchanger could be
an alternative genetic hypothesis. At present, however, at least for some areas (e.g., Northern Apennines,
Italy), this hypothesis is not supported by mineralogical evidence.

water types (e.g., Ca-carbonate) are dominant.


I NTRODUCTION
Some analyses of different types of NaCW (mostly
Definitions from Italy) are reported in Table 1 together with
The Na-carbonate waters (NaCW) are here indications about the rocks involved in water-rock
defined as waters having [Na total] ≥ ([K total] + interaction.
2[Ca total] + 2[Mg total]), ([HCO3 total] + 2[CO 3 In the following pages we will focus on wa-
total]) ≥ ([Cl–] + 2[S total]) and [Na total] > [K ters of extreme composition, i.e., with na and calc
total], where the values in square bracket are mo- higher than 0.85. In the Northern Apennines
lar concentrations. The waters considered in this (Italy), waters with na and calc > 0.85 are wide-
paper have moderate values of pH (<9.7); thus spread and thus, frequently, we shall implicitly
alkalinity (Alk) is practically only due to carbon- refer to them.
ate alkalinity, i.e., Alk ≅ [HCO3 total] + 2[CO3
total]. The Na-carbonate character for each water Data processing
sample may be expressed by the minimum value For water speciation and for other calculations
of the indexes na = [Na total]/([Na total] + [K to- we have used a recent version of PHREEQC
tal] + 2[Ca total] + 2[Mg total]) and calc = ([HCO3 (Parkhurst and Appelo, 1999), a flexible ad user-
total] + 2[CO3 total])/([HCO3 total] + 2[CO 3 to- friendly software. The software is adequate for low
tal] + [Cl–] + 2[S total]). ionic strength solutions (Debye-Hückel expres-
The Na-carbonate waters are not very frequent; sion), and suitable at higher ionic strength for so-
they are commonly present in areas where other dium chloride dominated systems. The limits of

*Corresponding author (e-mail: gvt@unipr.it)

351
352 G. Venturelli et al.

Table 1. Analyses of Na-carbonate waters from different localities

the software depend on the following points: tified by the parameter S.I. = log(K/Q), where K
(i) the data are taken from different reference is the equilibrium constant of the considered dis-
sources and thus the data may lack internal con- solution reaction of the phase and Q is the activ-
sistency; (ii) for some reactions the temperature ity product of the aqueous species involved in the
dependence of the equilibrium constant is not reaction.
known. These limits, however, do not invalidate
our semi-quantitative approach and results. If ther-
MAIN G EOCHEMICAL F EATURES AND
modynamic data were not found in the PHREEQC
GENETIC HYPOTHESES
data base, for some compounds or reactions data
from EQ3/6 (Wolery, 1992) and SUPCTR 92 NaCW may have both high (e.g., Larderello
(Johnson et al., 1992) were used. The saturation geothermal field, central Italy: Duchi et al., 1992)
degree of pure phases in the water has been quan- and low temperature (e.g., Northern Apennines,
Na-carbonate waters 353

Table 1. (continued)

Italy: Toscani et al., 2001; Po plain, Italy: Gorgoni Tuscany as well as in several other Italian regions
et al., 1982) and salinity. Salinity may range from Permian-Triassic or Messinian evaporites may
very low (e.g., some waters from the Gotthard occur at depth and sometimes at the surface (for
Tunnels, Switzerland: Michard et al., 1966) up to geochemical data see, for instance, Cortecci et al.,
some hundred of milliequivalent/L (mEqle/L) of 1981, and Longinelli, 1979/1980).
total dissolved ions (Σions) (e.g., waters from Oxygen and hydrogen stable isotope data are
Gambassi and Casciana, Central Italy: Bencini et not available for all the waters reported in Table
al., 1977). Some waters may have high calcium 1. However, the available δ 18 O and δ D values
and sulphate contents: Ca up to 13–17 (mEq/L) mostly indicate a recent meteoric origin (Fig. 1),
and SO 4 up to 16–22 mEq/L at Casciana and the main variations being related to altitude and
Gambassi. These high values are most likely due continental effects (cf., Longinelli et al., 2000).
to calcium sulphate dissolution; actually, in This, however is not true for some NaCW from
354 G. Venturelli et al.

Table 1. (continued)

*Sulphide-bearing waters; waters from: s = spring, w = well, ? = unknown; Alk = total alkalinity; S.I.cc and S.I.chal, Saturation
Index for calcite and chalcedony respectively. Samples BIS 13 to (4) are from Northern Apennines.
References: 1, Marini et al., 2000a; 2, Marini et al., 2000b; 3, Toscani et al., 2001; 4, Cortecci et al., 1999; 5, Elliot et al.,
1999; 6, Duchi et al., 1987; 7, Bencini et al., 1977; 8, Caboi et al., 1993; 9, Pastorelli et al., 2001; 10, Michard et al., 1996.
Lithology and mineralogy of interacting rocks (when reported in the papers).
BIS13-K06: marly sediments and interbedded pelites (illite, smectite-illite mixed layers, albite, silica phases, minor K-feldspar,
micas, chlorite, very rare kaolinite) (Venturelli and Frey, 1977; Fontana et al., 1994). F8-F10: calcareous sandstones, marly
pelites and sandy pelites. F14-(4): quartz-feldspatic sandstones with interbedded siltites and pelites (silica phases, plagioclase
and K feldspar in decreasing amount, and illite, chlorite, some smectite). 12c, 14c: limestones with nodular and tabular flints
(minor quartz, montmorillonite, mica, apatite, paligorskite, feldspar, marcasite, cristobalite; kaolinite and mixed layers re-
stricted to the lower chalk) (reported in Edmunds et al., 1987). G5, Ga4: amphibole gneiss of the Tremola series. Ga5, Ga6:
orthogneiss. Ga14: Permocarboniferous schists. SIB2-SIB7: crystalline basement. 1–17: mostly trachydacites (plagioclase, K-
feldspar, orthopyroxene, clinopyroxene, quartz, and minor mica and olivine) (Ferrari et al., 1996). 31, 50: ultramafic rocks,
metasediments, Triassic-Jurassic rocks (?) (Marini et al., 2000a). 97,103: limestones and pelitic cherts. 249: carbonaceous
flysch, pelites. 283: limestones, quartz-feldspatic sandstones. 407, AB: calcalkaline ignimbrites (plagioclase, and in place K-
feldspar, orthopyroxene, clinopyroxene, rare biotite, glass, amphibole; kaolinite and smectite as alteration products). 489:
limestones. SM: andesites (plagioclase, orthopyroxene, clinopyroxene, amphibole, occasional olivine and smectite as altera-
tion product).
Na-carbonate waters 355

waters from waters from


carbonate-sandstone-pelitic carbonate-sandstone-pelitic
rocks rocks
0.0
-20 waters from crystalline waters from crystalline
and volcanic rocks and volcanic rocks

WL
GM
-30 waters with high PCO2 -0.5 waters with high PCO2

-40
-1.0

-50

log (Eq/L Na)


-1.5
δ H (‰)

-60
2

-2.0
-70

-80 -2.5

-90 -3.0

-100
-3.5
-18 -16 -14 -12 -10 -8 -6 -4 -2 0
-6.0 -5.0 -4.0 -3.0 -2.0 -1.0 0.0
δ O (‰)
18
log (Eq/L Cl)

Fig. 1. Oxygen and hydrogen isotopes in H 2O of some Fig. 2. log[Na] vs. log[Cl] (Eq/L) for the waters re-
Na-carbonate waters. References: waters from carbon- ported in Table 1.
ate-pelitic and sandstone-pelitic rocks, Italy and
England, Cortecci et al., 1999, Elliot et al., 1999,
Toscani et al., 2001; waters from crystalline rocks,
Switzerland, Pastorelli et al., 2001; high PCO2 waters residual salts in sediments (e.g., Calzetti, 1999)
from Sardinia, Italy, Caboi et al., 1993. GMWL, global and to (iii) NaCl release from fluid inclusions
meteoric water line, δ 2 H = (8.20 ± 0.07) δ 18 O +
(11.27 ± 0.65) (‰) (Rozansky et al., 1993).
during water-rock interaction (e.g., Michard et al.,
1966). The importance of the different sources
varies from place to place. For instance, in
Tuscany, NaCl is probably mostly leached from
the Po plain, Italy (Gorgoni et al., 1982) not re- evaporites and, perhaps, fluid inclusions of the
ported in Table 1. The few tritium data support a crystalline basement.
long-term circulation at depth (Gorgoni et al., (2) Waters with PCO 2 less than 10–1.5 bar, low
1982; Toscani et al., 2001). The waters reported salinity (Σions < 40 mEq/L), variable but com-
in Table 1 may be distinguished into two main monly high pH (up to 9.7), and moderate to very
groups: low sulphate and chloride concentrations; many
(1) Waters with PCO 2 higher than 10–0.5 bar, waters of this group have extreme composition (na
high salinity (Σions > 80 mEq/L), low pH (usu- and calc > 0.85) with very low calcium, magne-
ally < 7) and, sometimes, high sulphate and chlo- sium, potassium, chloride and sulphur contents
ride contents; most waters of this groups come (e.g., Toscani et al., 2001).
from Tuscany, where CO 2 emissions from the Field evidence indicates that NaCW are not
ground are common (Gianelli, 1985; Minissale, related to a well defined lithology; they circulate
1991), and from Sardinia. These waters also ex- in and emanate from different rock types and in
hibit some correlation between Na and Cl (Fig. 2) areas where other types of waters may be domi-
which indicates addition of NaCl component. The nant (e.g., Ca-carbonate or Ca-sulphate waters).
NaCl component may be referred to (i) contribu- It is noteworthy, however, that waters with the
tion of NaCl-rich formation waters, (ii) dissolu- lowest salinity circulate in crystalline rocks (see
tion of halite present in (a) evaporites or (b) as Gotthard Tunnels) whereas most waters coming
356 G. Venturelli et al.

3 (2) Significant dissolution of Na-silicate is


2 made possible by addition of deep-seated CO 2
1 during some stage of the water-rock interaction.
S.I.

0 This interpretation is suggested, for instance, for


-1 some thermal waters in Tuscany, Italy (Bencini et
-2 1
al., 1977), in Mt. Amiata geothermal field (Duchi
-3 et al., 1987), at Bagno di Romagna (Emilia-
0 5 10 15 20 25 30 35 40 45 50 Romagna Region, Italy; Cortecci et al., 1999), and
T(°C) for some CO 2-rich waters of Sardinia ( PCO 2 up to
about 1 bar: Caboi et al., 1993).
Fig. 3. S.I. chalcedony (open symbols) and S.I.calcite (filled
symbols) vs. temperature ( °C) for the waters reported (3) The Na enrichment is due to interactions
in Table 1. Circle = waters from carbonate-pelitic and of water with alkali-rich volcanic rocks (e.g.,
sandstone-pelitic rocks; triangle = waters from crys- Duchi et al., 1987, for the Mt. Amiata volcanic
talline and volcanic rocks; diamond = high- PCO2 wa- area, and Duchi et al., 1991, for the volcanic area
ters. The continuous line indicates the difference of southern Latium, Central Italy) or, in general,
S.I. qtz - S.I.chalcedony vs. temperature (S.I.qtz - S.I.chalcedony
with magmatic and metamorphic rocks (e.g.,
≅ –3.13 × 10 –3 T( ° C) + 0.51). Only sample 1 from
Marmoreto, Tuscany, is strongly undersaturated in cal- Michard et al., 1996; Grasby et al., 2000;
cite. Pastorelli et al., 2001).
(4) NaCW are generated by prograde dissolu-
tion of silicate minerals—including albite—under
from carbonaceous-pelitic formations are of ex- conditions of saturation with respect to other sili-
treme Na-carbonate composition (na and calc > cate phases. Recently, this process has been pro-
0.85) and contain reduced sulphur. posed to explain the origin of sulphide-bearing
On average, the Saturation Index for calcite NaCW occurring in the Bisagno valley, north of
(S.I.calcite) is about 0.3, but many waters deviate Genoa, Italy (Marini et al., 2000b) and of some
considerably from this value. Most NaCW com- strongly reduced waters in the Northern Apennines
ing from carbonate-sandstone-pelitic rocks are (Toscani et al., 2001).
saturated to oversaturated in calcite, whereas low- (5) The sodic character is due to ionic exchange
PCO 2 waters from crystalline and volcanic rocks involving Na-rich clay which releases Na and ac-
and high- PCO 2 waters from several litologies may quires Ca (and Mg) (e.g., London and Berkshire
be undersaturated in this phase. The waters are in basins, UK: Elliot et al., 1999; northern Apennines
general undersaturated in quartz and saturated in and Po plain, Italy: Gorgoni et al., 1982; Venturelli
chalcedony. Both S.I.calcite and S.I.chalcedony are not et al., 2000).
apparently related to the temperature (Fig. 3) of Any of the processes listed above may con-
the waters suggesting that, as observed in many tribute to the production of NaCW in the different
areas worldwide, dissolution/deposition of these areas; frequently, however, they are not suffi-
phases does not need high temperature. ciently constrained. Thus, in the following pages
The hypotheses on the origin and evolution of we consider different hypotheses and put con-
NaCW commonly have been only considered in straints on their reliability taking into account
relation to local situations. The most common speciation in solution and water-rock interaction
hypotheses are summarised below. processes compatible with the geological situa-
(1) The excess of dissolved inorganic carbon tions. More specifically, we will try to answer the
in respect to Ca (and Mg) is related to addition of following questions. (i) Can Na-carbonate waters
CO2 which is generated by oxidation of organic be produced by interaction with carbonate-rich
matter during bacterially-mediated reduction of rocks? (ii) Which is the role of the silicate phases?
S(VI) (e.g., Cortecci et al., 1999). (iii) Which is the role of carbon dioxide?
Na-carbonate waters 357

neous and, if reducing bacteria are present, sul-


D ISCUSSION
phur reduction and CO2 production may be fast.
Carbon dioxide In the underground waters, which have ac-
An increase in the carbon dioxide partial pres- quired oxygen from the atmosphere and which are
sure during water-rock interaction influence the then subtracted to any interaction with an oxygen-
process of mineral dissolution. Carbon dioxide bearing gas phase, the content of dissolved O 2
may be generated by several processes. For in- cannot exceed about 0.3 mmole/L (actually
stance, in central Italy emission of large amounts [O2] ≅ 1.5 × 10–3 PO 2 , at 0°C and 1 bar total pres-
of CO 2 from soil, fractures and springs may be sure, and in the atmosphere PO 2 ≅ 0.2); thus, as-
due to mantle degassing (Minissale, 1991), to ther- suming that in reaction (1) all oxygen is consumed
mal breakdown of carbonate at depth—as a con- for carbon oxidation, the increase in dissolved
sequence of a high geothermal gradient (Gianelli, C(IV) will be no more than about 0.3 mmole/L.
1985) generated by intrusions of magmatic bod- On the contrary, CO2 generation via reactions (2)
ies at depth—or to both processes (Marini and and (3) may vary largely depending on the con-
Chiodini, 1994). centration of S(VI) and N(V) as well as, of course,
Carbon dioxide may also be produced by oxi- on the amount of organic matter present. Actu-
dation of organic compounds generated from or- ally, whereas the content of oxidised nitrogen is
ganic matter entrapped into the sediments at the usually low to moderate (<<1 mmole/L), sulphur
time of their deposition. The compound CH2O is may reach high values in some sulphate-bearing
used as simplified representation of organic mat- waters (cf., Toscani et al., 2001). In the case sul-
ter. Neglecting the speciation (aqueous, gas, solid) phate is the oxidant, high CO2 production would
of the different substances, the following reactions be coupled with high S(-II) generation (reactions
may be written: (2b) and (2c)). The role of these processes for the
origin of NaCW will be discussed later.
oxidation by dissolved O 2
“CH2O” + O2 = CO2 + H 2O; (1) Dissolution of albite, analcime, nepheline
Albite (or albitic plagioclase) is very common
oxidation by dissolved S(VI) in the pelitic portions of the sedimentary rocks
3“CH2O” + 2SO42– + 4H+ = 3CO2 + 2S + 5H2O involved in water-rock interaction. Analcime and
(2a) nepheline may occur in some volcanic rocks (e.g.,
2“CH2O” + SO 42– + 2H+ = 2CO2 + H2S + 2H2O in the potassic rocks of the Roman Province, Italy;
(2b) see Table 1); analcime is typically an alteration
2“CH2O” + SO 42– + H + = 2CO2 + HS– + 2H 2O; product of leucite (e.g., Cundari and Graziani,
(2c) 1964), nepheline is a primary phase. These min-
erals may constitute main Na sources in the wa-
oxidation by dissolved N(V), ters, whereas the carbon source may be carbon
5“CH2O” + 4NO 3– + 4H+ = 5CO2 + 2N2 + 7H2O dioxide of variable origin as discussed before, and
(3a) carbonate minerals. Simulation of the interaction
2“CH2O” + NO3– + 2H+ = NH4+ + 2CO2 + H2O between albite, analcime, nepheline, ±calcite and
(3b) pure water is summarized in Table 2(A). At
2“CH2O” + NO3– + H+ = NH3 + 2CO2 + H2O. log PCO 2 = 0, the maximum obtained value of Na
(3c) (S.I. Na-mineral = 0), is about 4.3, 1.6 and 0.6
mequivalent/kg water (mEq/kgW) for nepheline,
If oxygen fugacity and the activity of the dis- analcime and albite dissolution respectively. It is
solved species are appropriate, oxidation of or- noteworthy that, at least for albite and nepheline,
ganic matter may be thermodynamically sponta- these values, are hypothetical upper limits of Na
358 G. Venturelli et al.

Table 2. Composition of water (mEq/kg water) interacting with albite, analcime, nepheline and calcite at defined
saturation index (S.I.), at different PCO2 and T = 25 °C

The values S.I. = 0.3 for calcite and S.I. = 0 for chalcedony agree with the average values for the investigated waters. Illite =
K 0.6Mg 0.25Al 2.3Si 3.5(OH) 10 and Ca-montmorillonite = Ca 0.165Al2.33Si 3.67 O10(OH) 2.
(A) Note that, on the basis of mineral stoichiometry, we have: for albite dissolution Na = Al = 1/3Si, for analcime Na = Al = 1/
2Si, for nepheline Na = Al = Si.
Na-carbonate waters 359

concentration; in fact, at temperature lower than interaction involving albite, calcite, chalcedony,
about 100°C the silicate dissolution is a slow proc- Ca-montmorillonite, illite, kaolinite, and water at
ess which could lead to mineral saturation only variable PCO 2 is reported in Table 2(B). The main
after long time (see, for instance, Chou and results are the following: (i) a silica phase or illite
Wollast, 1984; Chou et al., 1989). In spite of it, saturation (and thus precipitation) is necessary to
comparison between the computed and the actual maintain low silica in solution; (ii) under condi-
NaCW data leads to the following remarks: (i) also tions of albite, chalcedony, ±Ca-montmorillonite,
assuming, as a boundary condition, that only the ±kaolinite saturation and slight calcite
Na that is in excess to Cl derives from silicates, oversaturation (S.I.calcite = 0.3, average values for
the actual difference [Na]-[Cl] in some NaCW is the investigated waters), the Na contents, alkalin-
higher than [Na] obtained by our simulation; ity and the Na/Ca ratio reach high values also at
(ii) moreover, the computed Si and Al (not re- moderate PCO 2 . The dissolving Na-silicate in-
ported in Table 2) are too high in comparison with volved in the examples of Table 2(B) is only albite,
the concentrations observed in the investigated but similar results may be obtained considering
low-temperature waters with pH comparable to analcime and nepheline.
those obtained by simulation. These results ex-
clude the possibility that only dissolution of sili- Prograde water-rock interaction and the origin of
cate phases and addition of carbon dioxide to the NaCW
aquatic systems is responsible for the high Na The data of Table 1 are plotted in Fig. 4 to-
content in many of the investigated waters. gether with trends obtained by simulations of
In the sediments, other sodic minerals could water-rock interaction. On average, the waters
occur, e.g., as Na-bearing smectites; however, the have S.I. calcite ≈ 0.3 and S.I.chalcedony ≈ 0; thus in
Na contribution due to their dissolution would be all the simulations these values of saturation in-
generally low (Na < 0.1 mEq/kgW) also under dex have been assumed. The trends of the elements
condition of water saturation in these phases and in solution have been obtained at 5 and 25°C at
at a high pressure of CO 2. defined PCO 2 (10–2, 10–1 and 1 bar) and assuming
that (i) calcite, Ca-montmorillonite and chal-
Albite, analcime, nepheline dissolution under con- cedony quickly react generating Ca-carbonate
ditions of stability of other phases water (S.I. calcite = 0.3, S.I. Ca-montmorillonite =
As discussed above, dissolution of albite, S.I.chalcedony = 0), (ii) saturation in these phases and
analcime and nepheline at those conditions can- oversaturation in calcite are maintained during
not alone explain the high Na content of most interaction with albite up to variable S.I. albite val-
waters. On the other hand, under conditions of ues (from –4.5 to saturation). Also simulation with
saturation in other minerals, such as kaolinite, kaolinite in place of Ca-montmorillonite has been
smectite and silica phases, water with high Na and performed; however, since the obtained results
alkalinity may be produced since dissolution of were very similar, only the trends calculated on
albite is enhanced by Si and Al removal. It is note- the basis of Ca-montmorillonite are reported in
worthy that crystallization of these phases is fa- Fig. 4.
cilitated if they are already present in some From Figs. 4A and B, we draw the following
amounts in the rocks interacting with the waters. conclusions. Among the waters which surely in-
From Table 1 it is evident that phyllosilicates teract with calcite-bearing rocks, the low- PCO 2
(mostly smectite, illite, kaolinite and mixed lay- group could be generated starting from Ca-carbon-
ers), silica phases and calcite are widespread, of- ate waters characterised by PCO 2 of about 10–2–
ten abundant or dominant, in the rocks interact- 10 –1 bar, whereas for the most saline waters the
ing with the investigated waters. Thus, as an ex- initial PCO 2 would be equal or higher than 1 bar.
ample, the results of a simulation of water rock- The simulation described above imposes the lim-
360 G. Venturelli et al.

1.0

-3.5
0.0
-3.0 -2.5
-2.0
-1.5
-1.0
-1.0
log PCO2 = 0
-0.5
-2.0
log PCO2

0.0

-3.0

-4.0 log PCO2 = -1

-5.0
log PCO2 = -2
-6.0
A
-7.0
0 20 40 60 80 100 120 140
Na (or Na-Cl) (mEq/kgW)
20

B C
0.0 0.0
16
-0.5 -0.5
-1.0
-1.0
Na (or Na-Cl) (mEq/kgW)

-1.5
-1.5
-2.0
12 -2.0

-2.5 log PCO = -1


8 2

-3.0
4
log PCO2 = -1

log PCO2 = -2 -3.5


log PCO
-4.0 2 = -2

0
0 4 8 12 16 0 1 2 3 4
Alk (mEq/kgW) Ca (mEq/kgW)

Fig. 4. log PCO2 vs. [Na], [Na] vs. [Alk] alkalinity and [Na] vs. [Ca] diagrams for the investigated waters. Shape
of symbols as in Fig. 1. Open circle = actual [Na] data; filled symbols = [Na]-[Cl] for the same waters; dotted
line = line connecting [Na] and [Na]-[Cl] data for the same sample. The continuous lines indicate the evolution
of water interacting with calcite, Ca-montmorillonite, chalcedony and albite at different PCO2 (log PCO2 0, –1, –2)
and temperature (5 °C, ×; 25 °C, +). The trends start from waters oversaturated in calcite (S.I.calcite = 0.3) and
saturated in Ca-montmorillonite and chalcedony and indicate the evolution of waters maintaining oversaturation
in calcite (S.I. calcite = 0.3), saturation in Ca-montmorillonite and chalcedony (S.I.chalcedony = 0) and interacting
with albite (S.I. albite ranging from –4.5 to 0). Numbers close to the lines indicate the value assumed by S.I.albite;
numbers in bold italics refer to the initial log PCO2 . As a first approximation, mEq/L ≡ mEq/kg water is assumed.
Na-carbonate waters 361

iting condition [Na]/alkalinity ratio → 1 for [Ca] As discussed above, oxidation of organic mat-
→ 0. Some waters, however, have [Na]/alkalinity ter—which is present in variable amounts in the
ratio significantly higher than unity and ([Na]- sediments—may be the source of the added CO 2.
[Cl])/alkalinity less than or around 1. This is rea- The effect of a PCO 2 increase is modelled in Fig.
sonably due to the contribution of halite occur- 5, which reports the result of progressive water-
ring in evaporites and as residual salt in sediments rock interaction at 5°C after organic matter oxi-
(Calzetti, 1999), or to mixing with NaCl-rich for- dation by dissolved oxygen and sulphate at a wa-
mation waters. In the diagrams, the difference ter-mineral interaction stage characterised by
[Na]-[Cl] represents an approximate lower limit S.I. albite = –3.5. Details about the conditions of the
for the Na derived from silicate dissolution. simulation are reported in the captions of Fig. 5.
NaCW from carbonate-pelitic and sandstone- At the beginning of the oxidation process, the as-
pelitic rocks Most NaCW from carbonate-pelitic sumed dissolved O 2 and S(VI) are ≅0.3 and ≅0.9
and sandstone-pelitic rocks are of extreme com- mmole/L, respectively, the last value being an
position (na and calc both > 0.85) with very low approximate upper limit for total sulphur in the
concentration of components other than Na and investigated NaCW from carbonate-sandstone-
alkalinity. Because of the very simple water chem- pelitic rocks.
istry, the Na and Ca contents, the alkalinity and Oxidation of organic matter at expense of sul-
the PCO 2 of the investigated waters may be easily phate in presence of sulphate-reducing bacteria
compared with the those obtained by the simu- (cf., Toscani et al., 2001) may also produce el-
lated water-rock interaction. emental or strongly reduced sulphur (mostly as
The simulation indicates that, to obtain the H2S and HS–). In this case, the actual low, S(-II)
actual Na content found in NaCW from carbon- in the investigated waters (see foot note of Table
ate-pelitic and sandstone-pelitic rocks, the poten- 1) may be due to prevalent S° production or to
tial Ca-carbonate parent waters of NaCW would S(-II) oxidation during the rise of the water to-
have a PCO 2 in the approximate range 10–2–10–1 ward the surface.
bar (Fig. 4A). These values, however, are too high Although oxidation of organic matter by oxy-
if compared with those of the most common Ca- gen and sulphate may produce CO2 and H2S com-
carbonate waters in the studied areas ( PCO 2 in the ponents, the PCO 2 increase obtained according to
range 10 –3.2–10 –1.8 bar; see for instance Venturelli the simulation is not sufficient to shift the [Na]-
et al., 2000, and Marini et al., 2000b). Assuming log PCO 2 trend up to the most sodic waters (see
that the discrepancy is not mostly due to unreli- Fig. 5). Thus, the wide distribution of the waters
able thermodynamic data, we conclude that NaCW in the diagram of Fig. 5 needs additional CO 2
have a Na content higher than expected if they sources. As an example, we now consider only the
are generated from Ca-carbonate waters with ini- NaCW from the northern Apennine chain.
tial PCO 2 ≈ 10–3.2–10–1.8 bar and without any fur- As discussed in the literature, the Tuscany and
ther addition of CO2. Theoretically, the Na excess Emilia-Romagna regions are characterised by the
could be produced by dissolution of silicates with occurrence of dry emission of CO2 and CH4, ther-
Na/Si ratio higher than that in albite, e.g., analcime mal springs very rich in CO2, thermal springs rich
and nepheline; these phases, however, are not ap- in N2, and mud volcanoes with emission of domi-
parently present in the carbonate-sandstone-pelitic nant CH4 (Bencini et al., 1977; Bencini and Duchi,
rocks interacting with the waters of Table 1 and 1981; Mattavelli et al., 1983; Mattavelli and
thus their contribution may be excluded. Actually, Novelli, 1987; Minissale and Duchi 1988; Duchi
addition to the aquifer of CO2 coming from the and Minissale, 1995; Minissale et al., 1997;
rocks interacting with the waters or from other Minissale et al., 2000). South of the Arno river,
sources is the most simple explanation of the sus- the dominant gas is CO 2 (usually more than about
pected Na excess. 91.5 mole %); north of the Arno river, CO2 emis-
362 G. Venturelli et al.

-1.0

-1.5

-3.5
-2.0
-3.0
-2.5 -2.5

-2.0
log PCO2

-3.0

-1.5
-3.5

-1.0
-4.0

-0.5
-4.5

-5.0 0.0

-5.5
0 5 10 15 20
Na (or Na-Cl)(mEq/kgW)

Fig. 5. log PCO2 vs. [Na] trends for water interacting with calcite, Ca-montmorillonite, chalcedony, and albite at
5 °C (numbers close to the lines indicate S.I.albite). Open circle, actual [Na] data for waters from carbonate-pelitic
and sandstone-pelitic rocks; filled circle, [Na]-[Cl] data for the same waters. As a first approximation, mEq/L ≡
mEq/kg water is assumed. (i) Dotted line with ×. Water with pH ≅ 7.43, alkalinity ≅ 5.96 mEq/kgW, and [Ca] ≅
2.98 mmole/KgW (S.I.calcite = 0.3 and log PCO2 = –2) interacts with calcite, Ca-montmorillonite, chalcedony and
albite under conditions of S.I.calcite = 0.3, S.I.Ca-montmorillonite = S.I. chalcedony = 0 and S.I. albite ranging from –4.5 to 0.
(ii) Dotted line with 䉫. Water with pH ≅ 7.40; alkalinity ≅ 5.63 mEq/kgW, Ca ≅ 3.73 and S(VI) ≅ 0.9 mmole/kgW
(S.I. calcite = 0.3 and log PCO2 = –2, S.I.gypsum = –1.35) interacts as at point (i). Continuous line with 䉫. Water of the
point (ii) interacts as at point (i) up to S.I.albite = –3.5. At S.I.albite = –3.5, organic matter is oxidised by dissolved
oxygen (≅ 0.3 mmole/kgW) and sulphate (≅0.9 mmole/kgW); PCO2 increases up to 10 –1.8 bar (see arrow along the
continuous line), and PH2 S becomes ≅ 10–2.64 bar. During this process; ≅2.1 mmole/kgW of C(IV) are added to the
water and ≅0.9 mmole/kgW of S(VI) are reduced to S(-II). For simplicity of the diagram, the two redox processes
are considered to occur instantaneously.

sion are scarce or lacking. maintained at low values: in detail, for the trends
Along the northern Apennine chain and north of Figs. 6A and 6B, at S.I.albite = 0, the carbon
of the Apennine watershed the dominant gas is added by CO 2 dissolution is about 75% and 90%
CH4, whereas is low to very low, in the approxi- of total dissolved C, respectively. Taking into ac-
mate range 10–1.6–10–2.9 bar for 1 bar total gas count the great amount of carbon which would
pressure. Although the concentration of CO2 in the need to be added from the external environment,
gas mixture is low, interaction of diffused emis- investigation on carbon isotopes of the dissolved
sion with ground waters could be sufficient to in- carbonic species and gaseous emissions could be
crease the silicate dissolution. This is modelled useful to support or invalidate the hypothesis of
in Fig. 6, which shows in detail how high the con- an open system for carbon.
centration of Na become if the water is open to The simulations discussed above are approxi-
appropriate CO2 input also under moderate CO2 mate and, at best, may give semiquantitative in-
partial pressure. For the trends shown in Fig. 6, dications. A more sophisticated kinetic approach
the CO2 input is very high also in the case PCO 2 is is not possible in our case since kinetic param-
Na-carbonate waters 363

-0.5

-1.0 A B
-1.5
-3.5 -3.0
-2.5
-2.0 -2.0
-1.5
-3.0 -3.0 -1.0
-2.5 -2.5
log PCO2

-2.0 -2.5

-2.0 -1.5 -2.0


-3.0
-1.0
-1.5 -1.5
-3.5 -0.5

0.0
-1.0
-4.0 -1.0

-4.5 -0.5 -0.5

-5.0 0.0 0.0

-5.5
0 4 8 12 16 0 4 8 12 16 20
Na (or Na-Cl) (mEq/kgW) Na (or Na-Cl) (mEq/kgW)

Fig. 6. log PCO2 vs. [Na] trends for water interacting with calcite, Ca-montmorillonite, chalcedony, and albite at
5 °C (numbers close to the lines indicate S.I.albite). Shape of symbols as in Fig. 1. Open circle, actual [Na] data for
waters from carbonate-pelitic and sandstone-pelitic rocks; filled circle, [Na]-[Cl] data for the same waters. As a
first approximation, mEq/L ≡ mEq/kg water is assumed. (A) Dotted lines as in Fig. 5, points (i) and (ii). Continu-
ous line with 䉫 indicates the water evolution down to PCO2 = 10 –3.95 bar starting from the dotted line with dia-
mond at S.I. albite ≅ –2.5 and progressive addition of CO 2 (range of PCO2 : 10–2.45–10 –3.95 bar). (B) Dotted lines as
in Fig. 5, points (i) and (ii); continuous line with 䉫 as in Fig. 5. Continuous line with square indicates the water
evolution down to PCO2 = 10–2.8 starting from the continuous line with diamond at S.I.albite ≅ –3 and progressive
addition of CO 2 (range of PCO2 : 10–1.9–10–2.8 bar).

eters, such as the mineral surface area, are not several authors (see above). In particular, for wa-
available for the involved rocks. ters from the Northern Apennines, cation exchange
NaCW from crystalline and volcanic rocks in ab- with release of Na and uptake of Ca by
sence of calcite As far as the waters from the crys- phyllosilicates, according to the general reaction
talline and volcanic rocks are concerned, an ori-
gin from a Ca-carbonate parent could not be the 0.5Ca2+ + NaR → Ca0.5R + Na+ (4)
rule; this in spite of the common saturation or
oversaturation in calcite. Actually, also in absence where R indicate the clay exchanger, has been
of calcite in the mineral assemblage, saturation in suggested by Venturelli et al. (2000) and criticised
calcite and high pH may be easily reached during by Toscani et al. (2001).
dissolution of Ca-plagioclase in water with dis- Under moderately dry conditions (relative hu-
solved CO2 coming from atmosphere and soil, and midity less than about 50%), Na,K-
under condition, for instance, of kaolinite satura- montmorillonites are characterised by d001 around
tion. For instance, considering the initial PCO 2 = 12 Å, whereas Ca,Mg-montmorillonites have a d-
10 –3 bar, for S.I.anorthite = –7 and S.I.kaolinite = –6, spacing of 14–15 Å (e.g., Alietti and Brigatti,
the resulting value of S.I.calcite is about 0.4. 1979). Although no systematic research on
montmorillonites of the Northern Apennine for-
Na-Ca cation exchange mations has been carried out, it is noteworthy that
The role of cationic exchange in the genesis no 12 Å-montmorillonites have been found so far.
of Na-carbonate waters has been suggested by Accordingly, recent cationic exchange experi-
364 G. Venturelli et al.

ments performed on Apennine clay sediments sug- Acknowledgments—The research has been supported
gest a calcic character for smectites (Calzetti, by the University of Parma (Local Grant). We are grate-
ful to M. F. Brigatti (University of Modena), L. Toscani
1999).
(University of Parma), R. Gieré (Purdue University),
and to an anonymous referee for their useful sugges-
tions.
CONCLUDING R EMARKS
(1) The waters considered in this paper are rep-
REFERENCES
resentative of different types of NaCW. In differ-
ent areas, NaCW coexist with the most common Alietti, A. and Brigatti, M. F. (1979) Behaviour of natu-
Ca-carbonate waters generated by fast interaction ral montmorillonites at suitable water vapour pres-
with calcite-bearing sediments. sure. Miner. Petrog. Acta (Bologna) 23, 145–150.
Bencini, A. and Duchi, V. (1981) Geochimical study of
(2) NaCW have widely variable composition some waters from Porretta Terme, Bologna. Rend.
as evidenced by their Na, Cl, and SO4 contents, Soc. It. Mineral. Petrol. 38, 1189–1195 (in Italian).
their alkalinities, Na/Ca ratio and PCO 2 . NaCW Bencini, A., Duchi, V. and Martini, M. (1977)
from Tuscany and Sardinia (Italy) may have high Geochemistry of thermal springs of Tuscany. Chem.
PCO 2 and salinity. Geol. 19, 229–252.
Caboi, R., Cidu, R., Fanfani, L. and Zuddas, P. (1993)
(3) NaCW represent the product of interaction
Geochemistry of the high-PCO2 waters in Logudoro,
between recent or old waters and different rock- Sardinia, Italy. Applied Geochemistry 8, 153–160.
types containing Na-silicates. Simple dissolution Calzetti, B. (1999) Interaction of Cr, Mn, Fe, Ni, Cu,
of Na-silicates also under high PCO 2 is not suffi- Zn, Cd, Pb, and fluids percolating from solid urban
waste with clays. Experimental data and environmen-
cient to produce the high Na contents and high tal implications. Dr. Sci. Thesis, Parma Univ., 169
alkalinity which characterise many waters. Dis- pp. (in Italian).
solution must occur under saturation/ Chou, L. and Wollast, R. (1984) Study of weathering
oversaturation of clay-minerals, silica phases and of albite at room temperature and pressure with a flu-
calcite; actually, these phases are usually present idised bed reactor. Geochim. Cosmochim. Acta 48,
2205–2218.
in the interacting rocks.
Chou, L., Garrels, R. M. and Wollast, R. (1989) Com-
(4) At least NaCW generated by interaction parative study of the kinetics and mechanism of dis-
with calcite-bearing rocks have Ca-carbonate par- solution of carbonate minerals. Chem. Geol. 78, 269–
ents, which, by interaction with albite, clay, and 282.
silica phases, acquire Na and loss Ca (mostly cal- Cortecci, G., Reyes, E., Berti, G. and Casati, P. (1981)
cite deposition). The Na vs. alkalinity and Na vs. Sulphur and oxygen isotopes in Italian marine sul-
phates of Permian and Triassic ages. Chem. Geol. 34,
Ca evolution is significantly dependent on the ini- 65–79.
tial water conditions (see Fig. 4). Cortecci, G., Calafato, A. and Boschetti, T. (1999) New
(5) Assuming that the simulation trends have chemical and isotopic data on the groundwater sys-
quantitative significance, the PCO 2 of the Ca-car- tem of Bagno di Romagna, Emilia-Romagna prov-
bonate parent is usually too low to obtain high Na ince, Italy. Miner. Petrogr. Acta (Bologna) 42, 89–
101.
values. Further addition of CO2 during the water-
Cundari, A. and Graziani, G. (1964) Products of leucite
rock interaction is necessary. For instance, addi- alteration in the Vico volcanics. Period. Mineral.
tional CO 2 may be produced by oxidation of or- (Roma) 33, 35–52 (in Italian).
ganic matter present in the sediments and/or by Duchi, V. and Minissale, A. (1995) Distribution of gas
mantle and crust degassing. emissions in the Tuscany-Latium peri-Tyrrhenian
(6) As far as the waters from northern sector and interaction with shallow groundwaters.
Boll. Soc. Geol. It. (Roma) 114, 337–351.
Apennines are concerned, at present, cationic ex- Duchi, V., Minissale, A. and Prati, F. (1987) Chemical
change involving a Na-exchanger, such as composition of thermal springs, cold springs,
smectite, is not supported by mineralogical evi- streams, and gas vents in the Mt. Amiata geothermal
dence. region (Tuscany, Italy). J. Volcanol. Geotherm. Res.
Na-carbonate waters 365

31, 321–332. Marini, L. and Chiodini, G. (1994) The role of carbon


Duchi, V., Paolieri, M. and Pizzetti, A. (1991) dioxide in the carbonate-evaporite systems of
Geochemical study on natural gas and water dis- Tuscany and Latium (Italy). Acta Volcanologica 5,
charges in the southern Latium (Italy): circulation, 95–104.
evolution of fluids and geothermal potential in the Marini, L., Bonaria, V., Guidi, M., Hunziker, J. C.,
region. J. Volc. Geterm. Res. 47, 221–235. Ottonello, G. and Vetuschi Zuccolini, M. (2000a)
Duchi, V., Minissale, A. and Manganelli, M. (1992) Fluid geochemistry of the Acqui Terme-Visone
Chemical composition of natural deep and shallow geothermal area (Piemonte, Italy). Applied
hydrothermal fluids in the Larderello geothermal Geochemistry 15, 917–935.
field. J. Volcanol. Geotherm. Res. 49, 313–328. Marini, L., Ottonello, G., Canepa, M. and Cipolli, F.
Edmunds, W. M., Cook, W. G., Darling, D. G., (2000b) Water-rock interaction in the Bisogno Val-
Kinniburgh, D. G., Miles, D. L., Bath, A. H., ley (Genoa, Italy): application of an inverse approach
Morgan-Jones, M. and Andrews, J. N. (1987) Base- to model spring water chemistry. Geochim.
line geochemical conditions in the Chalk aquifer, Cosmochim. Acta 64, 2617–2635.
Berkshire, U.K.: a basis for groundwater quality Mattavelli, L. and Novelli, L. (1987) Geochemistry and
menagement. Applied Geochemistry 2, 251–274. habitat of natural gases in Italy. Org. Geochemistry
Elliot, T., Andrews, J. N. and Edmunds, W. M. (1999) 13, 1–13.
Hydrochemical trends, palaeorecharge and Mattavelli, L., Ricchiuto, T., Grignani, D. and Schoell,
groundwater ages in the fissured Chalk aquifer of the M. (1983) Geochemistry and habitat of natural gases
London and Berkshire basins, UK. Applied in Po basin, northern Italy. Am. Assoc. Pet. Geol. Bull.
Geochemistry 14, 333–363. 67, 2239–2254.
Ferrari, L., Ponticelli, S., Burlamacchi, L. and Manetti, Michard, G. D., Pearson, F. J. and Gautschi, A. (1996)
P. (1996) Volcanological evolution of the Monte Chemical evolution of waters during long term in-
Amiata, Southern Tuscany: new geological and pet- teraction with granitic rocks in Northern Switzerland.
rochemical data. Acta Volcanologica 8, 41–56. Applied Geochemistry 11, 757–774.
Fontana, D., Spadafora, E., Stefani, C., Stocchi, S., Minissale, A. (1991) Thermal springs in Italy: their re-
Tateo, F. and Villa, G. (1994) The upper Cretaceous lation to recent tectonics. Applied Geochemistry 6,
helminthoid flycsh of the Northern Apennines: prov- 201–212.
enance and sedimentation. Mem. Soc. Geol. It. Minissale, A. and Duchi, V. (1988) Gethermometry on
(Roma) 48, 237–250. fluids circulating in a carbonate reservoir in north-
Gianelli, G. (1985) On the origin of geothermal CO2 central Italy. J. Volcanol. Getherm. Res. 35, 237–252.
by metamorphic processes. Boll. Soc. Geol. It. Minissale, A. Evans, W., Magro, G. and Vaselli, O.
(Roma) 104, 575–584. (1997) Multiple source components in gas manifes-
Gorgoni, C., Martinelli, G. and Sighinolfi, G. P. (1982) tations from north-central Italy. Chem. Geol. 142,
Isotopic evidence of paleowaters in the Po sedimen- 175–192.
tary basin (Northern Italy). Geochem. J. 16, 51–61. Minissale, A., Magro, G., Martinelli, G., Vaselli, O. and
Grasby, S. E., Hutcheon, I. and Krouse, H. R. (2000) Tassi, G. F. (2000) Fluid geochemical transect in the
The influence of water-rock interaction on the chem- Northern Apennines (central-northern Italy); fluid
istry of thermal spring in western Canada. Applied genesis and migration and tectonic implications.
Geochemistry 15, 439–454. Tectonophysics 319, 199–222.
Johnson, J. W., Oelkers, E. H. and Helgeson, H. C. Parkhurst, D. L and Appelo, C. A. J. (1999) User’s guide
(1992) SUPCRT92: a software package for calculat- to PHREEQC—A computer program for speciation,
ing the standard molal thermodynamic properties of batch-reaction, one-dimensional transport, and in-
minerals, gases, aqueous species, and reactions from verse geochemical calculations. A.S. Department of
1 to 5000 bars and 0° to 1000°C. Laboratory of Theo- Interior, U.S. Geological Survey, Water-Resources
retical Geochemistry, University of California, Investigation Report 99-4259, Denver, Colorado.
Berkeley. Pastorelli, S., Marini, L. and Hunziker, J. (2001) Chem-
Longinelli, A. (1979/1980) Isotope geochemistry of istry, isotope values ( δD, δ18O, δ34S) and tempera-
some Messinian evaporites: paleoenvironmental im- tures of the water inflows in two Gotthard tunnels,
plications. Palaeo. Palaeo. Palaeo. 29, 95–123 Swiss Alps. Applied Geochemistry 16, 633–649.
Longinelli, A., Selmo, E. and Flora, O. (2000) Isotopic Rozansky, K., Araguàs-Araguàs, L. and Gonfiantini, R.
composition and tritium activity of atmospheric pre- (1993) Isotopic pattern in modern global precipita-
cipitation in Northern Italy. 5th Intern. Isotope Work- tion. Continental Isotopic Indicators of Climate, Am.
shop, Cracow, July 2000. Geophys. Union Monograph.
366 G. Venturelli et al.

Toscani, L., Venturelli, G. and Boschetti, T. (2001) Venturelli, G., Toscani, L., Mucchino, C. and Voltolini,
Sulphide-bearing waters in Northern Apennines, C. (2000) Study of the water-rock interactions of
Italy: general features and water rock interaction. spring waters in the Northern Apennines. Annali di
Aquatic Geochemistry 7, 195–216. Chimica 90, 359–368.
Venturelli, G. and Frey, M. (1977) Anchizone metamor- Wolery, Th. J. (1992). EQ3/6, a sotware package for
phism in sedimentary sequences of the Northen geochemical modelling of aqueous systems: package
Apennines. Rend. Soc. Mineral. It. (Milano) 33, 109– overview and installation guide (version 7). Lawrence
123. Livermore National Laboratory, Livermore, CA.

You might also like