Anda di halaman 1dari 16

Journal

of

Hydrology
ELSEVIER

Journal of Hydrology 167 (1995) 263-278

[2]

Geochemistry of carbon dioxide in six travertine-depositing


waters of Italy
Allan Pentecost
Division of Life Sciences, King's College London, Campden Hill Road, London W8 7AH, UK

Received 30 January 1994; revision accepted 3 August 1994

Abstract
The chemical compositions of six travertine-depositing hot spring waters in Italy are
described with emphasis on the carbon dioxide system. All springs contained high concentrations of CO 2 (>20 mM 1-1) with equilibrium partial pressures well above those which
could have been formed in contact with a soil atmosphere. After surfacing, the CO2 is rapidly
lost to the atmosphere, with evasion rates close to the springs ranging from 0.45-4.41 mM m -2
s - l . Partial pressures of CO 2 showed an exponential decline with distance, which is consistent
with the static film model where temperature and turbulence are constant. Downstream CO2
transfer coefficients, which ranged from 66 to 360 cm h - l were consistent with moderately
turbulent flow, however, there was no correlation between turbulence, measured as the mean
shear stress, and CO2 evasion rate. The channels investigated had all been modified by man and
most possessed even widths and gradients.
All waters became increasingly supersaturated with aragonite and calcite downstream and
both of these minerals were present in fresh travertine deposits. The supersaturation was driven
almost exclusively by gas evasion. Comparison of daytime and nighttime evasion rates demonstrated that photosynthetic activity was an insignificant source of CO2 flux in the reaches
investigated. Carbon dioxide evasion is therefore primarily responsible for the supersaturation
and probably also the deposition of travertine at these sites.
The CaCO 3 content of the travertines ranged from 91.3 to 96.0 wt% with 1.7-4.1% CaSO4,
traces of organic matter and acid-insoluble minerals.

1. Introduction
S o m e o f the w o r l d ' s largest t r a v e r t i n e d e p o s i t s o c c u r in Italy, a n d m o s t are believed
to have been f o r m e d f r o m h o t springs highly c h a r g e d with c a l c i u m a n d c a r b o n
dioxide. L a r g e n u m b e r s o f h o t springs t h a t are h i g h l y c h a r g e d with CO2 o c c u r in
0022-1694/95/$09.50 1995 - Elsevier Science B.V. All rights reserved
SSDI 0022-1694(94)02596-7

264

A. Pentecost / Journal of Hydrology 167 (1995) 263-278

Italy (Waring, 1965; Barnes et al., 1978), but less than a quarter of these are known to
deposit travertine.
Most of the large travertine sites, which are extensively quarried, are now inactive,
but small travertine-depositing springs are widespread and a number are associated
with recent volcanic centres. Several sites are clustered around the Vican centre in the
Roman volcanic province, where fault-controlled springs deposit mounds and sheets
of travertine near Viterbo, Lazio.
There have been a number of geochemical studies on these deposits, but most
research has concentrated on inactive sites (Dall'Aglio and Tedesco, 1968;
Malesani and Vannuchi, 1975; Manfra et al., 1976). However, it is known
that levels of carbon dioxide are much higher than those in equilibrium
with soil atmospheres and probably linked to recent volcanic activity. Previous
investigations of water chemistry have tended to concentrate on trace, rather
than bulk constituents and little progress has been made in either identifying the
origins of the solutions or the chemical changes occurring on contact with the
atmosphere.
Travertine deposition is frequently associated with biological activity and many hot
springs possess a rich phototrophic bacterial flora. Phototrophic microbes can
remove dissolved carbon dioxide by photosynthetic uptake, resulting in the direct
precipitation of carbonates (Krumbein, 1979). Microbes can also provide a suitable
framework for crystal nucleation and accretion (Pentecost and Riding, 1986; Emeis
et al., 1987). Photosynthetic activity must be weighed against the direct transfer of
carbon dioxide from water to atmosphere, which is independent of biological activity.
Evasion of carbon dioxide to the atmosphere has been shown to increase with
turbulence (Dandurand et al., 1982; Herman and Lorah, 1987), leading to rise in
pH and carbonate ion activity. This in turn leads to calcite and aragonite
supersaturation, favouring the precipitation of these minerals. The rate of transfer
at these hot springs is unknown but assumed to be high because of the high CO2
partial pressures in these waters.
The aims of this investigation are threefold: first, to investigate the chemical
composition of travertine-depositing thermal waters, with emphasis on the CO2
system; second, to estimate CO2 evasion rates from in situ analyses and finally to
determine the travertine composition. The results are used to evaluate the significance
of gas evasion and photosynthetic activity in carbonate deposition.

2. Methods
2.1. Field sites

Four of the six sites were grouped around Viterbo in Lazio: Bagnaccio, BullicameDante, Bullicame-West and Le Zitelle. Two sites, Bagni san Fillipo and Bagni di
Vignone are situated further north (Figs. 1 and 2(a)-(c)). At all sites travertine is
being deposited in stream beds which have been altered by channelling to provide
bathing water. These modifications were in most cases an advantage for the

A. Pentecost / Journal of Hydrology 167 (1995) 263-278

265

?
/
TOSCANA

Radicofani ",,

;
(

UMBRIA

r""
,

J.... N Q
f - ' " Lago di
.J

Bolsena

=Canino

1,
Orte

Tuscania

I Viterbo

LAZIO
20 krn

t
~

o-~_ Lago di
~

Vico

Fig. 1. Location of sites investigated. 1, Bagnaccio;2, Bagni san Fillipo; 3, Bagni di Vignoni;4, BullicameDante; 5, Bullicame-West;6, Le Zitelle.
estimation of C02 transfer, as they provided fairly even stream gradients, depths and
widths.
2.2. Water sampling and analysis

Sampling was carried out on a series of field trips between February 1988 and April
1993. Most water samples were collected in 130 ml glass bottles (excluding Si, where
polythene was used). Water samples were collected at the springs, and at one or more
points downstream, at distances of 7.8-111 m from the springs (Table 1). Gas
analyses (CO2, 02, H2S) were performed immediately after the samples had cooled,
but most remaining analyses were performed in the laboratory after preserving with
either 0.01 M HC1 or 0.001% HgC12.
The pH was determined in situ using a Corning 120 pH meter and glass electrode
calibrated with NBS buffers of pH 4.0 and 7.4. A cooled water sample was then
diluted fivefold with CO2-free distilled water, and the pH raised immediately to

A. Pentecost / Journal of Hydrology 167 (1995) 263-278

266

a) Bullicarne-Dante

b) Bullicame-West
~}

pools

9
t

~25rn

50m 2J!
I

c) LeZitelle
.............................................................
li

i
.

I
St rada Valore . . . . . . . . . . . . . . . . .

.......

I 50rn

"~

Fig. 2. Details of watercourses at the Viterbo hot springs. (a) Bullicarne-Dante; (b) gullicame-West; (c) Le
Zitelle. Stars show positions of the springs.

between 8.2 and 8.8 with CO2-free NaOH to prevent gas transfer to the atmosphere.
The sample was then placed in a fixed-volume titration flask (see Edmond, 1970), and
the alkalinity determined potentiometrically using a microburrette filled with 1.0 M
HCI. The total CO 2 was obtained using a computer program which calculated the
correct endpoint of the sample by iteration (Pentecost, 1992). Program inputs were in
situ and postalkali pH, spring and titration temperature, normality and volume of
titrant, and solution specific conductivity. Alkalinity corrections for other bases
(borate, silicate and sulphide) were determined for all samples. At least fivefold
dilution of samples was necessary for these waters to prevent the CO2 released during
titration escaping from solution and forming bubbles in the flask. Total sulphide,
thiosulphate and oxygen were determined by the combined method of Ingvorsen and

A. Pentecost / Journal of Hydrology 167 (1995) 263-278

267

Table 1
Hydrological data relating to carbon dioxide evasion rates
Characteristic

Bagnaccio-Paula Springs (1)

Discharge (1 s-l)
System length (m)
Mean width (m)
Mean depth (era)
Water area (m2)
Mean water
temperature C
Mean gradient
degrees
Mean shear stress
(gcm -l s-l)
Bed charactera

A-B

B-C

3.7
7.6
0.2
3.4
1.5
62.0

3.7
12
0.2
3.4
2.4
61.3

1.01

C-D

D-E

3.5
11.4
0.17
3.0
1.9
60.3

3.1
32
0.38
3.0
12
55,8

3.2
48
0.42
5.8
20
40.7

5.4
78
0.74
3.4
58
50

3.7
23
0.13
11.2
3.0
55

4.4
68
1.02
2.5
69
55

3.75

1,54

7.80

1.43

0.60

1.85

1.11

60

65

190

80

780

80

115

80

S/R

L/H

a Stream bed characteristics: S, smooth; L, loose travertine crust with microbial mats; R, rough with small
(1 cm) protuberances greater than 1cm.
Site key: 1, Bagnaccio, A represents spring orifice, with B, C, etc. consecutively downstream; 3, Bagni di
Vignoni; 4, Bullicame-Dante; 5, BuUicame West; 6, Le Zitelle.
J o r g e n s e n (1979). A t s o m e sites, the presence o f o x y g e n was tested q u a l i t a t i v e l y using
the sensitive r e s a z u r i n m e t h o d (Caldwell et al., 1984). M o s t o f the r e m a i n i n g c a t i o n s
were d e t e r m i n e d using a t o m i c a b s o r p t i o n s p e c t r o p h o t o m e t r y . C h l o r i d e a n d s u l p h a t e
were d e t e r m i n e d b y a r g e n t o m e t r y a n d t u r b i d i m e t r y respectively ( G r e e n b e r g , 1985).

2.3. Mineral saturation quotients, pC02 and travertine composition


T h e s a t u r a t i o n q u o t i e n t s f~a, f~c a n d 9tg o f a r a g o n i t e , calcite a n d g y p s u m ,
respectively, were d e t e r m i n e d using W A T E Q (Truesdell a n d Jones, 1974).

(Ca)s(cO3)s
~a,g = (Ca)m(CO3) m
W h e r e s refers to the s a m p l e a n d m the i o n activity p r o d u c t for the m i n e r a l ( a r a g o n i t e ,
calcite) at the m e a s u r e d t e m p e r a t u r e a n d pressure. W h e n f~ = 1 the w a t e r s a m p l e is at
e q u i l i b r i u m (i.e. s a t u r a t e d with the mineral), values a b o v e u n i t y indicate supers a t u r a t i o n . T h e s a m e p r o g r a m o u t p u t s the e q u i l i b r i u m C O 2 p a r t i a l pressures o f
these solutions, plus the ' a q u e o u s ' (that is, unionised) CO2 c o n c e n t r a t i o n .
T r a v e r t i n e c a r b o n a n d o x y g e n stable i s o t o p e c o m p o s i t i o n was d e t e r m i n e d a n d
X - r a y diffraction used to c o n f i r m the m i n e r a l o g y .

2.4. Carbon dioxide flux and transfer coefficients


T h e c a r b o n d i o x i d e flux was e s t i m a t e d f r o m the difference between the a n a l y t i c a l
c o n c e n t r a t i o n s o f c a r b o n d i o x i d e between two fixed p o i n t s , a l l o w a n c e being m a d e for

268

A. Pentecost/Journal of Hydrology 167 (1995) 263-278

any CaCO3 precipitated in between. Rates could not be corrected for photosynthesis
directly, but samples were taken at midday and midnight to provide an estimate of
biological activity, i.e. between photosynthesis which results in net removal of CO2
from the water in the day, and respiration which provides a small net input of CO2
into the water at night. At most sites, samples were taken only in the upper reaches of
the streams, but at Bagnaccio, a sequence of downstream measurements was made.
The surface area of water exposed between the two points was determined by
measuring the average stream width, using ten equally spaced points. Discharge
was estimated along a short length by calculating the cross-sectional area of flow
and the flow rate using five timed transits of surface-floating paper discs. A small
correction was made for drag (Morisawa, 1968). The flux is expressed as mMole CO2
m -2 s -1. The CO2 transfer coefficients were calculated using the static film model
(Gislason, 1989) where the flux, F is expressible as
F = k ( C s - Cw)
= k[(Pco2K

) - Cw]

where k is the transfer coefficient, Cs is the concentration of gas at the film top, Cw is
the concentration of gas in water below the film, Pc% is the partial pressure of CO2 in
air above the film and Ks is Henry's Law constant. The film thickness, z is given by
z = D / F where D is the coefficient of diffusion of CO 2 in water, obtained from
Broecker and Peng (1974). For the calculations, Pc% has been taken as its
atmospheric value, 10 -3.44 bar and Cw as the bulk concentration of CO2 (aq).
2.5. M e a n s h e a r s t r e s s

Mean shear stress can be estimated from


r = pgRS

where p is the density of water, g is the acceleration owing to gravity, R is the


hydraulic radius of the stream bed and S is the stream gradient, obtained by levelling
(Richards, 1982). This relationship strictly applies to stream sections where there is no
acceleration, i.e. without cascades. These were absent in all of the stream sections
investigated.

3. Results
3.1. S i t e h y d r o l o g y

Some pertinent hydrological characteristics are shown in Table 1. Discharge at all


springs is less than 5 1 s -l , and the channels leading from the springs rarely exceeded I
m in width and 10 cm in depth. Most streams seeped into the ground after a short
distance and only at Le Zitelle did they flow into a second-order stream, at which
point travertine deposition ceased. Springwater temperatures varied from site to site,

A. Pentecost / Journal of Hydrology 167 (1995) 263-278

269

but the Viterbo group (Bagnaccio, Bullicame and Le Zitelle) all had temperatures close
to 62C. Gradients generally were slight along the man-made channels. At Bagno
Vignoni, however, the stream flowed down a steep gradient before finally cascading
over the main travertine mound. Flow rates rarely exceeded 1 m s-I at any of the sites.
The average shear stresses (Table 1) reflect differences in the gradient and shows that
Bagnaccio A-B, B-C, D - E , Bullicame-Dante and Le Zitelle all have about the same
mean shear stress (40-80 g cm -1 s-l). Channel bed characteristics were variable and at all
sites consisted of travertine. The deposits were often smooth and hard, as at Bagnaccio
and Bagni di Vignoni, but at Le Zitelle, the main channel possessed a few deep pools with
jagged, upstream-directed concretions but these did not appear to give rise to high shear
stresses (see Table 1). The highest shear stresses were obtained at Bagni di Vignoni where
the waters descended the steep, upper section of the travertine mound.
3.2. Water chemistry

All spring waters contained high concentrations of carbon dioxide, calcium,


magnesium and sulphate, with lower levels of sodium and chloride (Table 2).
Springwater pH was close to 6.4 and all waters contained less than 0.2 mM oxygen
and total sulphide.
The pCO2 of the emerging waters was high and often approached atmospheric
pressure. All waters arose apparently supersaturated with respect to both aragonite
and calcite and both minerals occurred in the associated travertines. Supersaturation
was particularly high at the Bagni san Fillipo spring, which also had the lowest pCO2.
This was probably because the waters were conducted on to the travertine from a
large pipe, which connected with the source about 200 m distant, permitting some
degassing within the pipe. F o r this reason, CO2 flux measurements were not made at
this site. Three of the springwaters were also close to gypsum saturation (Table 2).
3.3. C02 f l u x and travertine deposition

At Bagnaccio, the fall in T D I C (total dissolved inorganic carbon) and CO2 (aq) is
clearly demonstrated (Fig. 3). The rate of fall for both is non-linear with distance, but
for CO2 (aq) a logarithmic fall is evident (Fig. 4). As CO2 evasion proceeded, pH rose
from around 6.4 to 7.4. Oxygen invaded the water with an apparently linear rate of
uptake with distance and at the lowest site, 65 m distant from the source, oxygen
saturation had reached 76%. The mean oxygen import rate was estimated as 30 # M
m -2 s -I . The carbon dioxide flux differed very little between day and night (Table 3).
Travertine deposition in the channels was monitored by measuring dissolved
calcium. Rates of CO2 deposition (into CaCO3) were low when compared with
degassing, the largest amount CaCO3 precipitated between the highest and lowest
sites at Bagnaccio being 0.3 m M l -l. At the other Viterbo springs and at Bagni di
Vignoni, the differences between midday and midnight measurements were again
small and calcium carbonate deposition amounted to a maximum of 0.3 mM 1-1.
At Bullicame-Dante and Le Zitelle the maximum CO2 'loss' to CaCO3 deposition was
0.76 m M 1-1 and 1.5 m M 1-l, respectively.

270

A. Pentecost / Journal of Hydrology 167 (1995) 263-278

Table 2
Chemical composition of spring waters
Determinand

t(c)
pH
TDICa(mM)
HCO3 (mM)
pCO2(atm)
log pCO2
CO2 (aq) (mM)
Ca (mM)
Mg (mM)
Sr (#M)
Na (mM)
K (mM)
CI (mM)
SO4 (raM)
total S (#M)
02 (mM)b
9tcc
f~a
f~g

Site
1

64.9
6.32
31.2
17.4
0.881
-0.055
13.8
14.3
6.91
92
2.00
0.21
2.00
12.9
32
0.01
5.70
4.02
0.86

47.0
6.72
36.4
27.8
0.395
-0.403
8.25
18.0
8.13
5.00
0.05
0.40
15.6
39
+
14.8
10.0
1.23

33.2
6.75
21.9
16.6
0.177
-0.752
4.96
18.0
8.55
2.85
0.60
1.57
11.3
34
0.26
7.22
4.82
-

55.5
6.30
28.5
15.8
0.705
-0.152
12.7
14.3
5.65

56.3
6.54
21.6
14.6
0.384
-0.416
6.80
14.0
5.75
2.80
0.85
0.40
11.4
82
0.22
6.46
4.51
-

62.9
6.32
23.0
12.8
0.632
-0.199
10.2
14.2
4.9
98
1.9
0.08
0.43
9.7
112
+
3.98
2.82
-

3.30
0.85
0.25
11.4
76
0.02
3.94
2.74
0.87

a TDIC Total dissolved inorganic carbon.


b + indicates waters resazurin-positive.
c ~c,a,g saturation ratios for calcite, aragonite and gypsum.
Sites: 1, Bagnaccio,Paula Springs; 2, Bagni san Fillipo, top spring; 3, Bagni di Vignoni;4, Bullicame-Dante;
5, Bullicame-West;6, Le Zitelle.
C a r b o n dioxide flux between sites was variable (Table 3), with a range of 0.45-4.41
m M CO2 m -2 s -1 The highest flux was o b t a i n e d at Bagnaccio, j u s t below the springs,
a n d the lowest were o b t a i n e d at Le Zitelle, in the upper, slow-flowing channel. O n
average, d a y t i m e rates of degassing (1.8 m M CO2 m -2 s - l ) were the same as nighttime
rates, A significant positive c o r r e l a t i o n was o b t a i n e d between the flux a n d b o t h the
m e a n pCO2 ( P < 0.01) a n d the t e m p e r a t u r e ( P < 0.1), b u t n o correlation was f o u n d
between the flux a n d stream gradient. A l t h o u g h there was some t e n d e n c y for sites
with higher shear stresses to have larger CO2 fluxes, there was n o significant
c o r r e l a t i o n between these variates.
The c a r b o n dioxide transfer coefficients ranged from 66 to 360 cm h -1 (Table 3)
with estimated film thicknesses in the range 0 . 1 - 0 . 4 cm.

3.4. Travertine composition


The m a j o r travertine c o n s t i t u e n t s are listed in T a b l e 4. I n all samples, the deposits
c o n t a i n e d m o r e t h a n 9 0 % calcium c a r b o n a t e b u t with significant a m o u n t s of gypsum,
which a m o u n t e d to 4.1% at Bullicame West. T h e Viterbo travertines n o r m a l l y consisted o f mixtures o f a r a g o n i t e a n d calcite a n d only at Le Zitelle was a r a g o n i t e the

A. Pentecost/Journal of Hydrology 167 (1995) 263-278

271
150

100

7.5

50
7.0

o/

6.5

20

15

0~0

" ~ 8

10

TDIC m M / l

10

20

m.

30
downst ream

40

50

Fig. 3. Downstream chemical changes at Paula spring, Bagnaccio on 8 April 1993. (a) Oxygen; (b) pH; (c)
total dissolved inorganic carbon; (d) CO2 (aq). Full lines denote midday measurements; broken lines,
midnight.
p r e d o m i n a n t mineral. Traces o f aragonite were f o u n d at Bullicame-West, and at
Bullicame D a n t e the travertine consisted entirely o f calcite.
The stable isotopic compositions show that all o f the travertines were enriched with
13C and 180 (Table 1) in relation to the P D B and S M O W standards.

4. Discussion
4.1. The dissolved carbon dioxide and its origin
C a r b o n dioxide - -

rich discharges are f o u n d t h r o u g h o u t the western limb

272

A. Pentecost / Journal of Hydrology 167 (1995) 263-278


pCO2(atm)

1.0

"'~'8~o

0.1

0.01

'
0

'
50

'
100
m. d o w n s t r e a m

'
150

Fig. 4. Log (CO 2 {aq}) as a function of distance from springs, Paula springs, Bagnaccio. Full line, midday;
broken line, midnighL

of the Italian peninsula and are associated with all of the recent Italian volcanic
centres.
Numerous analyses of thermal waters may be found in Waring (1965) but
measurements of dissolved carbon dioxide have rarely been undertaken on
travertine-depositing thermal springs. Some information is available from Japan
(Kitano, 1963), Italy (Malesani and Vannuchi, 1975), Wyoming (Friedman, 1970)
and Bolivia (Risacher and Eugster, 1979). These data are listed with some previously
unpublished information in Table 5(a). Carbon dioxide levels from tectonically active
regions fall in the range 12.7-67 mM 1-l with equilibrium partial pressures of
0.22-1.0 atm. These partial pressures greatly exceed those of the soil atmosphere
which lie in the range 0.01-0.1 atm (Atkinson and Smith, 1976).
Table 3
Carbon dioxide flux and transfer coefficients
Site

Flux mM C O 2 (m -2 s-1)
Day rate

Bagnaccio A-B
Bagnaccio B-C
Bagnaccio C - D
Bagnaccio D - E
Bagni di Vignone
Bullicame-Dante
Bullicame-West
Le Zitelle

4.4
3.5
3.0
0.9
1.3
0.5
2.4
0.5

Transfer coefficient k
(cmh -1)

Night rate

2.2
3.4
1.4
0.7
1.7
0.5
3.6
0.5

Day

Night

246
274
347
249
187
93
109
80

122
275
183
203
360
66
157
70

A. Pentecost / Journal of Hydrology 167 (1995) 263-278

273

Table 4
Composition of fresh travertines
Determinand

Site
1

CaCO 3 (%)
CaSO4 (%)
SrCO3 (%)
Organic mattera (%)
Acid-insoluble (%)
minerals
Mg (ppm)
Sr (ppm)
K (ppm)
Na (ppm)
Fe (ppm)
Mn (ppm)
P (ppm)
Mineralogyb
~13CPDB
6180 SMOW

94.1
3.4
1.6
0.48
0.15
985
9600
15
440
135
10
43
A+ C
6.29
18.71

96.0
2.3
1.4
0.14
0.08
1900
8100
57
136
74
A+ C
6.50
19.90

92.4
3.3
1.4
0.80
1.73
4460
4000
35
730
230
34
8
C
4.30
20.71

96.0
3.1
0.3
0.37
0.11
4690
1510
39
480
145
102
82
C
5.91
19.56

91.3
4.1
2.4
1.84
0.32
1380
14200
55
136
146
C + A(tr)
5.92
19.28

95.4
1.7
1.6
0.94
0.24
920
9340
24
57
14
A + C(tr)
4.93
17.62

a Includes some sulphur


b A, aragonite; C, calcite; tr, trace.
Sites: 1, Bagnaccio; 2, Bagni san Fillipo; 3, Bagni di Vignoni; 4, Bullicame-Dante; 5, Bullicame-West; 6, Le
Zitelle.
A c a r b o n d i o x i d e source a d d i t i o n a l to the soil is necessary to p r o d u c e such high
pressures at the e a r t h ' s surface. L i m e s t o n e a n d / o r o r g a n i c m a t t e r d e c a r b o n a t i o n in
a r e a s o f r e g i o n a l or c o n t a c t m e t a m o r p h i s m h a s often been suggested to a c c o u n t for
the a d d i t i o n a l c a r b o n d i o x i d e ( H u r l e y et al., 1966; B a r n e s et al., 1984; C a t h e l i n e a u et
al., 1989; Deines, 1992). N o d e c a r b o n a t i o n processes h o w e v e r have yet been u n e q u i v o c a l l y identified with a n y g e o t h e r m a l area. T h e I t a l i a n v o l c a n o e s a n d a s s o c i a t e d
g e o t h e r m a l fields o c c u r in a c o m p l e x tectonic region where crustal tension, deep
faulting a n d s u b d u c t i o n are all in p r o g r e s s (Chester, 1985) t h o u g h it is n o t possible
at p r e s e n t to identify w h i c h CO2-evolving process is a s s o c i a t e d with the h o t springs
investigated. T h e i s o t o p i c c o m p o s i t i o n o f the travertines a n d their a s s o c i a t e d waters
(Panichi a n d T o n g i o r g i , 1976; M a n f r a et al., 1976) t o g e t h e r with the results o b t a i n e d
here (Table 4) i n d i c a t e a heavier source o f c a r b o n which c o u l d be p r o v i d e d b y limestone d e c a r b o n a t i o n at high t e m p e r a t u r e s .
T h e r m a l springs o c c u r r i n g in tectonically quiet regions (Table 5b) have low equilib r i u m p C O 2 values t h a t fall within the r a n g e o f r e g i o n a l soil p C O 2 a n d r e p r e s e n t
d e e p - f l o w i n g m e t e o r i c waters unlikely to have c o n t a c t e d t h e r m a l l y g e n e r a t e d CO2
sources.
4.2. Gas f l u x
T h e c a r b o n d i o x i d e e v a s i o n rates ( T a b l e 3) were f o u n d to be high in all cases,

274

A. Pentecost / Journal of Hydrology 167 (1995) 263-278

Table 5
Carbon dioxide levels in travertine-depositing thermal springs
Site/country

TDIC (mM 1-1)

pCO2
(atm)

t(C)

Reference

Pentecost
(unpublished)
Kitano
(1963)
Pentecost
(unpublished)
aFriedman
(1970)
aRisacher and
Eugster (1979)
Pentecost
(unpublished)
Pentecost
(unpublished)

(a) Tectonically/volcanically active regtons


Heber Hot Sprs
Utah
Japan

21.0

0.40

45

12.7-67

0.66-1.02

26-100

Karlovy-Vary
Czech Republic
Mammoth Hot Springs
Wyoming
Pastos Grandes
Bolivia
Rapolano Terme
Italy
St Nectaire,
France

54.5

0.906

57

19.6-20.1

0.31-0.33

73

c. 15.3

0.22

37

60.2

0.50

26

65.2

0.64

30

Bath Spa
England
Bormio, Italy

4.38

0.058

46

6.07

0.056

38

Laguna Grande
Mexico
Matlock Bath
England

4.04

0.012

28

4.30

0.038

20

(b) Tectonically quiet regions

aEdmunds and
Miles (1991)
aDe Capitani
et al. (1974)
Pentecost
(unpublished)
Pentecost
(unpublished)

Calculated from data in text.

reflecting the high partial pressures of the gas in solution. There is little information
on rates of carbon dioxide evasion from freshwaters or hot-springs. In a study of gas
transport in a G e r m a n travertine-depositing hill stream, a rate o f 8 # M m -2 s -1 was
estimated (Usdowski et al., 1979). A considerably higher rate of 120 # M m -2 s 1 was
reported from the M o n t e z u m a Well source (Cole and Bachelder, 1969). Neither of
these rates approaches those obtained here (Table 3) due to the much higher pCO2
and temperatures of the issuing thermal waters.
A detailed investigation of some cold and w a r m travertine-depositing streams of
Virginia indicated a positive relationship between mean shear stress and degassing
(Hoffer-French and Herman, 1989) with m a x i m u m CO2 evasion rates of around 1
/.tM kg -l s -1 on the cascades where shearing stresses approached 2 kg cm -1 s -1 . These
rates compare with a mean evasion rate of 6 # M kg -1 s 1 for Bagnaccio with shearing
stresses of about 0.1 kg cm 1 s-1 (Table 1). The lack of correlation between degassing
rate and the mean shear stress was unexpected at the Italian sites but might be
explained by the lack of extreme differences in turbulence (i.e. cascades) down the
watercourses investigated.
The flux o f carbon dioxide across the a i r - w a t e r interface has been shown, for most

A. Pentecost/Journal of Hydrology 167 (1995) 263-278

275

conditions of flow to be consistent with the static film model (Usdowski and Hoefs,
1990) and is given by F = k A C (see above). The transfer coefficient k is dependent
upon temperature and turbulence and where these factors are constant, the flux is
dependent solely upon the concentration difference of carbon dioxide across the film.
A water sample subject to gas evasion under these conditions will undergo an
exponential decline in dissolved gas with time, providing there are no limitations
imposed by chemical reactions involving the gas. Carbon dioxide evasion at
Bagnaccio showed such a decline with distance despite the fact that temperature
varied downstream. A similar decline is also apparent in the data of Usdowski et
al. (1979) but these examples must be regarded as the exception rather than the rule.
At Bagnaccio the exponential decline must be attributed to the comparatively simple
and even form of the artificial channel. The weak positive correlation between evasion
rate and temperature follows directly from the reduced solubility and increased
diffusivity of carbon dioxide with increasing temperature. Values for the transfer
coefficient and estimated film thickness are similar to those obtained for a meltwater
stream investigated by Gislason (1989) who found that they were indicative of
turbulent flow. Coefficients in excess of 100 cm h -l fall within the breaking-bubble
regime at the sea surface (Broecker and Siems, 1984), but bubble breaking was rarely
observed at the hot springs. This suggests that the hot-spring waters were somewhat
less turbulent than that predicted by the Broecker and Siems (1984) model, but this is
accounted by the increasing magnitude of k with water temperature. Another
contributory factor, which would be difficult to demonstrate, is that the concentration of carbon dioxide at the film surface is probably higher than the mean
atmospheric value owing to the large amounts of gas being discharged into the
overlying atmosphere.
While carbon dioxide was lost from the hot waters, oxygen was absorbed. At
Bagnaccio, the rate of uptake was consistent with the transfer model of Tsivoglou
and Neal (1976) where uptake is a function of stream gradient and transit time.

4.3. Chemical composition and travertine formation


All of the springwaters are of the C a - H C O 3 - S O 4 type with high levels of
magnesium and total dissolved solids ranging from 3.1 to 3.9 g 1 1 (Table 2). The
low salinities and chloride levels show that the waters are non-marine in origin.
Chemistries of the Viterbo springs are markedly similar to each other, suggesting a
common source. High levels of sulphate are probably the result of gypsum/anhydrite
dissolution as sediments containing these minerals are widespread in central Italy.
The dissolution of CaSO 4 and CaCO 3 may lead to subsurface precipitation of CaCO3
through the common ion effect (Freeze and Cherry, 1979). Some evidence of
subsurface deposition is apparent at Bagnaccio (R. L. Folk, personal communication, 1991), and the spring waters appear to rise slightly supersaturated with
respect to both calcite and aragonite, though close to gypsum saturation (Table 2).
The gross composition of these waters is similar to other travertine-depositing springs
in central Italy (Malesani and Vannuchi, 1975).
Experimental work has shown that aragonite is precipitated in preference to calcite

276

A. Pentecost/Journal of Hydrology 167 (1995) 263-278

at temperatures exceeding about 40C (Lippmann, 1973). Because most of the


thermal springs have temperatures exceeding 45C, the occurrence of aragonite was
not surprising (Table 4), though its absence at Bullicame-Dante (t 40-55.5C) was.
The comparatively high levels of Sr in the travertine (Table 4) may be attributed to its
high level in the water and the low partition coefficient for Sr in aragonite (Cipriani et
al., 1977). Although relatively high Mg levels are present to high levels in the water,
high magnesian calcite was not detected by X-ray diffraction in the travertine. The
occurrence of gypsum (Table 4) probably resulted from evaporation of the
CaSO4-saturated water.
The deposition of travertine from thermal springs has been observed on numerous
occasions and is the subject of many investigations. Two views are commonly
expressed concerning its formation, though they are not mutually exclusive
(Pentecost, 1990). One considers that deposition is the result of CO2 evasion, which
raises pH and leads to readjustment of the dissolved carbonate equilibrium.
Subsequently, increasing carbonate supersaturation and precipitation occur. The
other view emphasizes the role of biological processes in controlling carbonate
precipitation, either through the removal of CO2 via photosynthesis and/or a
catalytic effect operating at the organism surface. Because organisms are abundant
at all of these springs their possible effects cannot be ignored.
If photosynthesis is significant in removing CO2 from the water, then a marked
difference in the bulk T D I C and the evasion rate should be apparent between
measurements carried out at midday and midnight. No evidence was found here to
support significant photosynthetic activity, and it must be concluded that gas evasion
is the major driving force leading to carbonate supersaturation at all of these sites.
Whether organisms proceed to catalyse carbonate deposition from a water that has
become highly supersaturated as a result of CO2 evasion is another matter and is
beyond the scope of this study.

Acknowledgements
I wish to thank Dr. T. P Jones of the Institute for Geology, University of Tubingen
for the stable isotope analyses and Professor R.L. Folk for assistance with some of the
in situ analyses.

References
Atkinson, T.C. and Smith, D.I., 1976.The erosion of limestones. In: T.D. Ford (Editor), The Scienceof
Speleology.Academicpress, London, pp. 151-178.
Barnes, I., Irwin, W.P. and White, D.E., 1978.Global distribution of carbon dioxidedischarges,and major
zones of seismicity.U.S., Geol. Surv., Water Res. Invest. No. 78-39 (Open-fileRep.).
Barnes, I., Irwin, W.P. and White, D.E., 1984.Map showingworld distribution of carbon dioxide springs
and major zones of seismicity. Department of the Interior. U.S. Geol. Surv., Misc. Invest. Ser. Map
1-1528.
Broecker,H.C. and Peng, T.H., 1974. Gas exchange rates between air and sea. Tellus, 26: 21-35.

A. Pentecost / Journal of Hydrology 167 (1995) 263-278

277

Broecker, H.C. and Siems, W., 1984. The role of bubbles for gas transfer from water to air at higher
windspeeds. Experiments in the wind wave facility in Hamburg. In: W. Brutsaert and G.H. Jirka
(Editors), Gas Transfer at Water Surfaces. Reidel, Dordrecht, pp. 229-236.
Caldwell, D.E., Kieft, T.L. and Brannan, D.K., 1984. Colonization of sulfide-oxygen interfaces on hot
spring tufa by Thermothrix thiopara. Geomicrobiol. J., 3: 181-200.
Cathelineau, M., Dubessy, J., Marignac, C., Valori, A., Gianelli, G. and Puxeddu, M., 1989. Pressuretemperature-fluid composition changes from magmatic to present day stages in the Larderello
geothermal field. In: D.L. Miles (Editor), Water-Rock Interaction. Balkema, Rotterdam, pp. 137-140.
Chester, D.K., 1985. Mount Etna, the anatomy of a volcano. Chapman and Hall, London.
Cipriani, N., Malesani, P. and Vannucci, S., 1977. I travertini dell'Italia centrale. Boll. Serv. Geol. Ital., 98:
85-115.
Cole, G.A. and Bachelder, G.L., 1969. Dynamics of an Arizona travertine-forming stream. J. Ariz. Acad.
Sci., 5: 271-283.
Dall'Aglio, M. and Tedesco, C., 1968. Rilievo idrogeochimico dell'area dei Monti Cimini. CNEN RT/GEO
(68), pp. 1-9.
Dandurand, J.L., Gout, R., Hoefs, J., Menschel, G., Schott, J. and Usdowski, E., 1982. Kinetically
controlled variations of major components and carbon and oxygen isotopes in a calcite-depositing
spring. Chem. Geol., 36: 299-315.
De Capitani, L., Fiorentine, P.M. and Terrani, M., 1974. Geochemistry of Ra in the supergene zone of a
radioactive spring. Atti Soc. Ital. Sci. Nat., 115:157-169.
Deines, P., 1992. Mantle carbon: concentration, mode of occurrence and isotopic composition. In: M.
Schidlowski, S. Golubic, M.M. Kimberley, D.M. McKirdy, and P.A. Trudinger, (Editors), Early
Organic Evolution. Springer, Berlin, pp. 133-146.
Edmond, J.M., 1970. High precision determination of titration alkalinity and total carbon dioxide content
of sea water by potentiometric titration. Deep Sea Res., 17: 737-750.
Edmunds, W.M. and Miles, D.L., 1991. The geochemistry of the Bath thermal waters. In: G.A. Kellaway
(Editor), Hot Springs of Bath. Bath City Council, UK, pp. 143-156.
Emeis, K.C., Richnow, H.H. and Kempe, S., 1987. Travertine formation in Plitvice National Park,
Yugoslavia: chemical versus biological control. Sedimentology, 34: 595-610.
Freeze, R.A. and Cherry, J.A., 1979. Groundwater. Prentice-Hall, New York.
Friedman, I., 1970. Some investigations of the deposition of travertine from hot springs - - 1. The isotopic
chemistry of a travertine-depositing spring. Geochim. Cosmochim. Acta, 34:1303-1315.
Gislason, S.R., 1989. Kinetics of water-air interactions in rivers: a field study in Iceland. In: D.L. Miles
(Editor), Water Rock Interaction. Balkema, Rotterdam, pp. 263-266.
Greenberg, A.E. (Editor), 1985. Standard Methods for the examination of Water and Wastewater, 16th
edn. American Public Health Association, Washington.
Herman, J.S. and Lorah, M.M., 1987. CO 2 outgassing and calcite precipitation in Falling Spring Creek,
Virginia, U.S.A. Chem. Geol., 62: 251-262.
Hoffer-French, K.J. and Herman, J.S., 1989. Evaluation of hydrological and biological influences on CO 2
fluxes from a karst stream. J. Hydrol., 108: 189-212.
Hurley, P.M., Fairburn, H.W. and Pinson, Jr., W.H., 1966. R b - S r isotopic evidence in origin of potashrich lavas of western Italy. Earth Planet. Sci. Lett., 5: 301-303.
Ingvorsen, K. and Jorgensen, B.B., 1979. Combined measurement of oxygen and sulfide in water samples.
Limnol. Oceanogr., 24: 390-393.
Kitano, Y., 1963. Geochemistry of calcareous deposits found in hot springs. J. Earth Sci., Nagoya Univ.,
11: 68-100.
Krumbein, W.E., 1979. Calcification by bacteria and algae. In: P.A. Trudinger and P.A. Swaine (Editors),
Biogeochemical Cycling of Mineral-forming Elements. Elsevier, New York, pp. 47-68.
Lippmann, F., 1973. Sedimentary Carbonate Minerals. Springer, Heidelberg.
Malesani, P. and Vannucci, S., 1975. Precipitazione di calcite o di aragonite dalle acque termominerali in
relazione alia genesi e all'evoluzione dei travertini. Atti R. Accad. Ital., 58: 761-776.
Manfra, L., Masi, U. and Turi, B., 1976. La composizione isotopica dei travertini del Lazio. Geol. Romana,
15: 127-174.

278

A. Pentecost / Journal of Hydrology 167 (1995) 263-278

Morisawa, M., 1968. Streams, their Dynamics and Morphology. McGraw-Hill, New York.
Panichi, C. and Tongiorgi, E., 1976. Carbon isotopic composition of CO 2 from springs, fumaroles, mofettes
and travertines of central and southern Italy. Preliminary prospection method of a geothermal area. In:
Proc. Second United Nations Symp. on the Development and Use of Geothermal Resources. San
Francisco, pp. 815-825.
Pentecost, A., 1990. The formation of travertine shrubs: Mammoth Hot Springs, Wyoming. Geol. Mag.,
127: 159-168.
Pentecost, A., 1992. Carbonate chemistry of surface waters in a temperate karst region: the southern
Yorkshire Dales, U.K.J. Hydrol., 139: 211-232.
Pentecost, A. and Riding. R., 1986. Calcification in cyanobacteria. In: B.S.C. Leadbeater and R. Riding
(Editors), Biomineralization of Lower Plants and Animals. Systematics Association Spec. Vol. 30.
Oxford University, Oxford, pp. 73-90.
Richards, K., 1982. Rivers: Form and Process in Alluvial Channels. Methuen, London.
Risacher, F. and Eugster, H.P., 1979. Holocene pisoliths and encrustations with spring-fed surface pools,
Pastos Grandes, Bolivia. Sedimentology, 26: 253-270.
Truesdell, A.H. and Jones, B.F., 1974. WATEQ, a computer programme for calculating chemical equilibria
of natural waters. J. Res. U.S. Geol. Surv., 2: 223-248.
Tsivoglou, E.C. and Neal, L.A., 1976. Tracer measurement of reaeration III., Predicting the reaeration
capacity of inland streams. Water Pollut. Control Fed. J., 48: 669-72.
Usdowski, E. and Hoefs, J., 1990. Kinetic 13C/12C and 180/160 effects upon dissolution and outgassing of
CO 2 in the system CO2-H20. Chem. Geol., 80: 109-118.
Usdowski, E., Hoefs, J. and Menschel, G., 1979. Relationship between 13C and 180 fractionation and
changes in major element composition in a recent calcite-depositing spring - - a model of chemical
variations with inorganic CaCO 3 precipitation. Earth Planet Sci. Lett., 42: 267-276.
Waring, G.A., 1965. Thermal springs of the United States and other countries of the world - - a summary.
U.S., Geol. Surv., U.S., Govt. Printing Office, Washington, Prof. Pap. 492.

Anda mungkin juga menyukai