You are on page 1of 13

Available online at www.sciencedirect.

com

Geochimica et Cosmochimica Acta 72 (2008) 5009–5021


www.elsevier.com/locate/gca

Calcite precipitation instability under laminar, open-channel flow


Ø. Hammer a,*, D.K. Dysthe a, B. Lelu a,b, H. Lund a,c, P. Meakin a,d, B. Jamtveit a
a
Physics of Geological Processes, University of Oslo, P.O. Box 1048, 0316 Blindern, Oslo, Norway
b
Ecole Normale Supérieure de Lyon, 46, allée d’Italie, 69364 Lyon Cedex 07, France
c
Norwegian University of Science and Technology, 7491 Trondheim, Norway
d
Idaho National Laboratory, P.O. Box 1625, Idaho Falls, ID 83415-2211, USA

Received 2 October 2007; accepted in revised form 22 July 2008; available online 22 August 2008

Abstract

We present a 2D numerical model for the growth of calcite from supersaturated aqueous solutions under laminar, open-
channel flow conditions. The model couples solution chemistry, precipitation at solution/calcite interfaces, hydrodynamics,
diffusion and degassing. The model output is compared with experimental results obtained using an oversaturated calcite solu-
tion produced by mixing CaCl2 and Na2CO3. The precipitation rate is observed to increase when the supersaturated solution
flows over an obstruction, leading to a growth instability that causes the formation of terraces. At relatively high flow rates,
the most important mechanism for this behaviour seems to be hydrodynamic advection of dissolved species either towards or
away from the calcite surface, depending on location relative to the obstruction, which deforms the concentration gradients.
At lower flow rates, steepening of diffusion gradients around protrusions becomes important. Enhanced degassing over the
obstruction due to shallowing and pressure drop is not important on small scales. Diffusion controlled transport close to
the calcite surface can lead to a fingering-type growth instability, which generates porous textures. Our results are consistent
with existing diffusive boundary layer theory, but for flow over non-smooth surfaces, simple calcite precipitation models that
include empirical correlations between fluid flow rate and calcite precipitation rate are inaccurate.
Ó 2008 Elsevier Ltd. All rights reserved.

1. INTRODUCTION interesting example which involves not only transport of


the precipitating solutes but also transport of CO2 and its
Precipitation of minerals from aqueous solution is often exchange with the atmosphere.
studied primarily in terms of chemical kinetics and diffusion The increase in the calcite precipitation rate on protru-
(e.g. Buhmann and Dreybrodt, 1985). However, advective, sions under open-channel flow is responsible for a number
hydrodynamically controlled transport will also influence of pattern formation phenomena in hot springs, karst
precipitation rates in systems with relatively high-velocity caves and similar systems. Travertine terracing is the
flow such as springs, streams, high-velocity hydrothermal prime example of this positive feedback instability, where
conduits and water pipes. For flow in the complicated precipitation causes faster growth of the step. Several
geometries found in nature, the coupling between hydrody- mechanisms have been proposed for this phenomenon.
namics and precipitation become complex, and the causal Buhmann and Dreybrodt (1985) developed a model in
relationships difficult to establish. A modeling approach is which higher flow rates under turbulent conditions, which
therefore useful. Calcite precipitation in open channels, may be found near an obstruction, lead to thinning of the
forming travertine, is a particularly ubiquitous and laminar boundary layer. This results in faster diffusion
across the boundary layer, which is rate limiting during
diffusion-limited precipitation. This phenomenon, con-
*
Corresponding author. firmed experimentally by Liu and Dreybrodt (1997), was
E-mail addresses: ohammer@nhm.uio.no (Ø. Hammer), also used by Wooding (1991) to explain travertine terrac-
d.k.dysthe@fys.uio.no (D.K. Dysthe). ing. The dynamics of travertine precipitation generally

0016-7037/$ - see front matter Ó 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.gca.2008.07.028
5010 Ø. Hammer et al. / Geochimica et Cosmochimica Acta 72 (2008) 5009–5021

leads to an expansion of the flow channel (rather than The OH used up by precipitation according to this equa-
localization as is the case for erosion), producing a wide tion is balanced by loss of H+ through the right-to-left reac-
and shallow sheet of water with low Reynolds numbers tion path of Eq. (1a), or supplied directly through Eq. (1b),
and little turbulence (Baker and Smart, 1995). The equa- producing CO2 in solution.
tions of Buhmann and Dreybrodt (1985) are still valid, The focus of this study is not calcite precipitation
but complete mixing in a turbulent core cannot be as- kinetics in isolation, but development of a better under-
sumed. Other authors (Chen et al., 2004) have suggested standing of the detailed processes responsible for the
that the reduced volume/surface ratio, caused by shallow- growth instability that leads to travertine terracing. We
ing, itself enhances CO2 degassing, and therefore leads to therefore need to carry out modeling and experiments in
increased precipitation. Agitation of the surface at and geometry and at a scale comparable to natural terracing
downstream from the obstruction can also facilitate degas- conditions, and to be able to map out the spatial and tem-
sing. Finally, pressure drop in areas of fast flow (the Ber- poral distribution of precipitation at this scale. Well devel-
noulli effect) could conceivably enhance degassing by oped travertine terraces in nature have minimum heights
reducing gas solubility (Chen et al., 2004). of about 5 mm and lateral dimensions of 1 to several
Some authors (Goldenfeld et al., 2006; Hammer et al., 100 cm. Travertine growth surfaces have RMS roughness
2007) have developed models that simply assume a linear of the order 10 lm to 1 mm measured over 4 mm2 areas.
relationship between flow velocity and precipitation rate, Open channel flow over well developed travertine terraces
as observed in nature, without addressing the mechanism has three main characteristics: (i) slow, laminar flow, (ii)
for this correlation. The pattern formation demonstrated shallow water depth over terrace edges (2–5 mm), and
in these models is primarily due to this relationship. (iii) the flow direction, water flux and water depth up-
In this paper, we develop a full 2D model of carbonate stream of, and over terrace edges vary slowly and contin-
precipitation under laminar, open-channel flow over an uously due to the growth of the travertine. A wealth of
obstruction, without any a priori assumptions with regard experimental studies of growth and dissolution of carbon-
to the relationship between flow rate and precipitation rate. ates and other minerals focus on controlling the hydrody-
The model couples fluid dynamics with a detailed simula- namic conditions at the mineral surface. Examples include
tion of the solution chemistry of the carbonate system, cal- rotating disk experiments (e.g. Compton and Daly, 1987)
cite precipitation kinetics, species transport (diffusion and and experiments performed with channel flow cell (CFC)
advection) and non-equilibrium degassing. setups (Compton and Unwin, 1990). The CFC is similar
The results indicate that precipitation is enhanced at an to that used in our work, but all studies that we are aware
obstruction, even without turbulence. The mechanism in- of attempt to achieve homogeneous flow conditions on the
volves compression and deformation of concentration gra- mineral surface. In this paper, we describe a study of the
dients due to advection. The fluid flow rate seems less growth instability caused by flow inhomogeneity. A wide
important in controlling the precipitation rate under lami- range of different techniques (with or without lateral spa-
nar flow conditions, and the field observations of a correla- tial resolution) have been used to quantify growth/dissolu-
tion between the two may be due in part to the negative tion kinetics in such cells. Since the early 90s the methods
correlation between depth and the flow rate imposed by of choice to study growth and dissolution with spatial res-
hydrodynamics. olution have been atomic force microscopy (AFM)
The calcium carbonate system in aqueous solution (e.g. (Hillner et al., 1992), X-ray reflectivity (Chiarello et al.,
Buhmann and Dreybrodt, 1985) involves the conversion be- 1993), and interferometry (van Enckevort, 1992). The geo-
tween CO2, H2CO3, H+ and HCO3  , where the latter can chemical studies with these techniques have focused on the
further be assumed to be in equilibrium with CO3 2 . This nanometer scale. Of these methods, only low coherence
conversion involves two reaction paths: interferometry can be used to characterize surface rough-
ness on the millimeter scale and growth inhomogeneities
CO2 þ H2 O $ H2 CO03 $ Hþ þ HCO3  ð1aÞ
on lateral scales of the order of 1 cm, as found on
and travertine surfaces.
CO2 þ OH $ HCO3  ð1bÞ
Other reactions taking place in solution are fast and can be 2. METHODS
assumed to be in equilibrium.
At the water–calcite surface, precipitation and dissolu- 2.1. Carbonate system model
tion involve the three reactions:
The solution chemistry of the carbonate system is simu-
CaCO3 þ Hþ $ Ca2þ þ HCO3  ð2aÞ lated with a coupled reaction–diffusion–advection model
CaCO3 þ H2 CO03 $ Ca2þ þ 2HCO3  ð2bÞ based on the three molar concentrations [CO2], [Ca2+]
and the sum of the carbonate and bicarbonate concentra-
and tions ½c ¼ ½HCO3   þ ½CO3 2  (that is, total dissolved car-
CaCO3 þ H2 O $ Ca2þ þ HCO  bon concentration excluding [CO2] and ½H2 CO3 0 ). The
3 þ OH ð2cÞ
rate of CO2 concentration change due to conversion of
Under typical conditions for calcite precipitation, with CO2 into H+ and HCO3  (Eq. (1)) and the transport of
pH > 6 and P CO2 < 0:01 atm, Reaction (2c) is dominant. CO2 in bulk solution are given by
Calcite precipitation instability under open-channel flow 5011

o½CO2  ½c ¼ ½HCO3   þ ½CO3 2  together with Eq. (7e) of


¼ ðk 1 þ k 2 ½OH Þ½CO2  þ ðk a ½Hþ  þ k 2 Þ
ot Buhmann and Dreybrodt (1985), we obtain the following
 ½HCO3   þ DCO2 r2 ½CO2   ur½CO2 ; ð3Þ equations:

where u is the fluid velocity [modified from Buhmann and aHþ cCO3 ½c
½HCO3   ¼ ; ð7aÞ
Dreybrodt (1985, Eq. (5a))]. The rate constants k1 and k2 K 2 cHCO3 þ aHþ cCO3
(forward rate of the first part of Eq. (1a), and forward rate K 2 aHCO3
of Eq. (1b)), k2 (reverse rate of Eq. (1b)) and the diffusion ½CO3 2  ¼ ; ð7bÞ
aHþ cCO3
constant DCO2 are given in Table 1. The rate constant ka ac-
counts for the fast dissociation of carbonic acid (second where K2 is the equilibrium constant for the dissociation of
part of Eq. (1a)) and is calculated according to Kaufmann bicarbonate (Table 1), and
and Dreybrodt (2007): ½OH  ¼ K W =½Hþ : ð7cÞ
k 1 cH cHCO3 In these equations, a are activities, calculated from concen-
ka ¼ ; ð4Þ
K5 trations using the corresponding activity coefficients. We
with the values of k1 (reverse rate of first part of Eq. (1a)) then compute the total charge concentration q:
and K5 (equilibrium constant for the fast second reaction of q ¼ 2½Ca2þ  þ ½Hþ   ½HCO3    2½CO3 2   ½OH : ð7dÞ
Eq. (1a)) given in Table 1. The activity coefficients c are cal-
culated from ionic strengths according to the extended For the simulations of the experimental model (see be-
Debye–Hückel theory (Dreybrodt, 1988). We also have low), the concentrations [Na+] and [Cl] are also in-
cluded in Eq. (7d). At near-neutral pH, Eq. (7d) can be
o½Ca2þ  approximated by
¼ Dr2 ½Ca2þ   ur½Ca2þ  ð5Þ
ot
q ¼ 2½Ca2þ  þ ½HCO3  : ð7eÞ
and
2+
Since [Ca ] and [c] are given by Eqs. (5) and (6), Eqs. (7a)–
o½c
¼ ðk 1 þ k 2 ½OH Þ½CO2   ðk a ½Hþ  þ k 2 Þ½HCO
3 (7d) can be combined to produce an explicit function
ot
q([H+]):
þ Dr2 ½c  ur½c ð6Þ
q ¼ 2½Ca2þ  þ ½Hþ 
 2 !
with ½c ¼ ½HCO3  þ ½CO3  as defined above. Eq. (6) is
the sum of Eqs. (5b) and (5c) in Buhmann and Dreybrodt cH ½Hþ cCO3 ½c K 2 cHCO3
 1þ2
(1985). K 2 cHCO3 þ cH ½Hþ cCO3 cH ½Hþ cCO3
The initial concentrations [CO2]0, [Ca2+]0 and [c]0 are set
 K W =½Hþ : ð7fÞ
to values known from the chemistry of the starting solution
+
at the upstream source. Subsequent compositions are then Charge balance requires that q([H ]) = 0, and we therefore
calculated from Eqs. (3)–(6). For each time step, the con- search for a root of this equation using a bisection algo-
centrations of the charged species H+, OH, rithm (Press et al., 1992). Inserting the resulting [H+] back
HCO3  and CO3 2 to be inserted into Eqs. (3)–(6) are into Eqs. (7a)–(7c), we then finally have the concentrations
found by assuming that they are in equilibrium with each [H+], ½HCO3  ; ½CO3 2  and [OH] to be used in Eqs. (3)–
other (due to fast reactions). This computation uses the fol- (6). After the bisection iteration has been used to determine
lowing method. First, ½HCO3  ; ½CO3 2  and [OH] are ex- [H+], the carbonic acid concentration can also be calculated
pressed as explicit functions of [H+] and [c]. Using using

Table 1
Chemical and physical parameters used in the model
Parameter Unit Value Reference
1
log10 k1 s 329.850  110.54 log10 T  17265.4/T Usdowski (1982)
log10 k M1 s1 329.850  110.54 log10 T  17265.4/T Usdowski (1982)
log10 k2 M1 s1 13.635  2895/T Usdowski (1982)
log10 k2 s1 14.09  5308/T Usdowski (1982)
log10 KW M2 22.801  4787.3/T  0.010365T  7.1321 log10 T Harned and Hamer (1933)
log10 K2 M 107.8871 + 5151.79/T  0.03252849T + 38.92561 log10 T  563713.9T2 Plummer and Busenberg (1982)
K5 M 1.707  104 Plummer and Busenberg (1982)
P atm
CO2 atm 3.5  104
log10 KH M atm1 108.3865  6919.53/T + 0.01985076T  40.45154 log10 T + 669,365T2 Plummer and Busenberg (1982)
k cm h1 10 Liss and Merlivat (1986)
log10 j1 cm s1 0.198  444/T Plummer et al. (1978)
log10 j2 cm s1 2.84  2177/T Plummer et al. (1978)
log10 j3 mM cm2 s1 5.86  317/T (valid for TC 6 25) Plummer et al. (1978)
DCO2 cm2/s 105(0.56 + 0.058TC) Buhmann and Dreybrodt (1985)
D cm2/s 0:7DCO2 Buhmann and Dreybrodt (1985)
Values are converted to SI units by the program. T is temperature in K, TC in centigrades.
5012 Ø. Hammer et al. / Geochimica et Cosmochimica Acta 72 (2008) 5009–5021
 
cH ½Hþ  F CO2 ¼ k K H P atm
½H2 CO3 0  ¼ : ð8Þ CO2  ½CO2  ; ð17Þ
K 5 cH2 CO3 0
where KH is Henry’s constant. An estimated gas transfer
At the calcite surface, the precipitation (negative) or disso- velocity of k = 10 cm/h = 2.8  103 cm/s (Liss and
lution (positive) flux into the fluid, in mmol/cm2/s, is mod- Merlivat, 1986) was used. This gives the following bound-
eled using the PWP equation (Plummer et al., 1978; ary conditions at the air–water interface Caw:
Buhmann and Dreybrodt, 1985) 
o½CO2 
F ¼ j1 aHþ þ j2 aH2 CO3 þ j3  j4 aCa2þ aHCO3  ; ð9Þ ¼ F CO2 ; ð18Þ
ot Caw

where o½Ca2þ 
¼ 0; ð19Þ
aH2 CO3 ¼ aH2 CO3 0 þ aCO2 : ð10Þ ot Caw

The constant j4 depends on T and P CO2 . Kaufmann and and


Dreybrodt (2007) used the following expression, derived 
o½c
by fitting to the empirical data of Plummer et al. (1978): ¼ 0: ð20Þ
ot Caw

2:375 þ 0:025T C P CO2 P 0:05 atm;
log10 j4 ¼ These boundary conditions are different from those given
2:375 þ 0:025T C þ 0:56ð log P CO2  1:3Þ P CO2 < 0:05 atm;
ð11Þ
by e.g. Buhmann and Dreybrodt (1985). These authors used
F CO2 ¼ F , assuming that the amount of CO2 released by
where TC is the temperature in degrees centigrade and P CO2 precipitation through the reactions of Eqs. (1) and (2c) is
is in atmospheres. balanced by the amount of CO2 lost to the atmosphere by
This approximation is valid for temperatures up to degassing. This balance is a reasonable assumption in the
35 °C. The equation for j4 given by Buhmann and case of plug flow over a large, flat surface, but it cannot
Dreybrodt (1985) contains a printing error (Dreybrodt, pri- be used in the general case in which advection and concen-
vate communication). We also tested a simpler equation tration fields vary in space. No physical mechanism can
due to Compton and Pritchard (1990, Eq. (53)): maintain equal fluxes between any two given points on
F CP ¼ 9:5  1011  7:14  103 ½Ca2þ ½CO3 2 ; ð12Þ the water–calcite and the water–air interfaces in the pres-
ence of complex flow and evolving water chemistry. We
with FCP in mol/cm2/s. This gave similar precipitation pat- must therefore resort to basic physical principles such as
terns, but with lower overall precipitation rates than the the solution chemical kinetics and fluid dynamics (Eqs.
PWP model. Finally, we implemented the equation of (3)–(7)), the kinetics of precipitation/dissolution (Eq. (9))
Brown et al. (1993, Eq. (18)): and the transport of CO2 across the aqueous/air interface
(Eq. (17)).
K sp  ½Ca2þ ½CO3 2 
F B ¼ k b K Ca K CO3 ; ð13Þ The constants used in these equations are summarized in
ð1 þ K Ca ½Ca2þ Þð1 þ K CO3 ½CO3 2 Þ Table 1.
again with FB in mol/cm2/s, and with To calculate the velocity vector field u, the Navier–
k b K Ca K CO3 ¼ 990 cm4 =mol=s, KCa = 106 cm3/mol and Stokes equations for open-channel (free-surface) flow in a
K CO3 ¼ 3  107 cm3 =mol (Brown et al., 1993). KCa and vertical 2D slice are solved using the code NaSt2D (Griebel
K CO3 are Langmuirian adsorption constants. This equation et al., 1997). The reactive transport equations (3)–(6) are
also gave precipitation patterns similar to those obtained approximated using explicit finite difference methods, and
from the PWP model, but with precipitation rates an order the reaction and transport terms are calculated in the same
of magnitude lower. The results reported below were ob- time step (one-step method). The advection term is imple-
tained using the PWP equation. mented with a second-order Lax–Wendroff difference
The precipitation of one mole of CaCO3 removes of one scheme (Lax and Wendroff, 1960).
mole of Ca2+ and one mole of c from the solution at the
water–calcite interface Cwc (and conversely for dissolution). 2.2. Experiments
This gives the following boundary conditions at the water–
calcite interface: The main barrier to reproducing natural carbonate
 growth patterns in the laboratory is the difficulty in main-
o½CO2 
¼ 0; ð14Þ taining control over the concentrations of Ca2+ and
ot Cwc CO3 2 in aqueous solutions over long periods of time. An

o½Ca2þ  attempt was made to reproduce the conditions typical of
¼ F; ð15Þ
ot Cwc hydrothermal systems, by saturating water with CO2 at
high pressure, passing the fluid through crushed limestone,
and releasing it onto a surface and allowing CO2 to escape to
 the atmosphere. Although a significant amount of calcium
o½c
¼ F: ð16Þ carbonate was dissolved, the precipitation rate was never
ot Cwc
sufficient to observe growth patterns of interest. We there-
The CO2 flux across the air–water interface, in mmol/cm2/s fore turned to a different method for producing calcite
(negative for degassing) is given by supersaturation: mixing Na2CO3 and CaCl2 solutions at
Calcite precipitation instability under open-channel flow 5013

the inlet of a long channel flow cell. As mentioned in the and it was inclined at 0.8–1.7°. Various obstacles were
introduction, open channel flow over well developed traver- placed in the flow, 0.7 m down the ramp from the inlet
tine terraces is slow and laminar with a water depth of 2– point. The initial velocity down the ramp was 3–5 mm/s.
5 mm over terrace edges. The fact that the water flux and Small amounts of grease or other impurities caused the
water depth upstream of terrace edges vary slowly and con- water to wet the surface incompletely and made the flow
tinuously (due to the growth of the travertine) indicates that pattern inhomogeneous. This did not perturb the instability
the exact flow patterns upstream of the terrace edges are not under investigation, but in order to get quantitative results
important for the instability mechanism. A significant it is simpler to use a flow that is homogeneous along the
observation that has bearing on the construction of our step edge. To obtain uniform wetting properties everywhere
flow channel is that travertine terraces do not form imme- along the ramp we removed the deposited carbonate with
diately adjacent to the fluid source. The flow channel was hydrochloric acid, and sand blasted the ramp before each
therefore chosen to be almost one meter long. experiment.
Na2CO3 and CaCl2 solutions were brought together and The flow obstacle had a rectangular cross-section (in the
mixed at the top of a glass ramp with side walls, and the x–z-plane in Fig. 1) with a length of 6–9 mm and a height of
supersaturated solution flowed down the ramp and over a 2.9 mm. Both corners facing up towards the water flow
glass obstacle near its lower end (Fig. 1a). The solutions, were ground to a 45° bevel of length 0.5 mm. Different sur-
each approximately 4 mM, were prepared by injecting con- face treatments to obtain homogeneous flow and good opti-
centrated solutions into tubes with pure water flow rates of cal contrast between a dark background and the light
0.5–4 ml/s and the two tubes containing diluted solutions of precipitate were tried. These included polished glass, sand
Na2CO3 and CaCl2 were joined at the end of a common blasted glass and black matte lacquered surfaces.
tube where the solutions mixed. It was therefore possible We used two methods to measure the thickness of the
to maintain a steady, continuous flow on the ramp while calcite precipitate on the solid surfaces on and around the
the concentrations of Na2CO3 and CaCl2 were controlled. flow obstacle: white light interferometry (WLI) and imaging
The ramp had a width of 59 mm with 1 cm high side walls, of scattered light intensity. WLI is a standard technique for

a
camera

z y
10 cm
x

b z image c
x

lamp lens
d

air
e
water

glass
glass

Fig. 1. Sketch of the experimental setup for calcium carbonate growth. (a) 0.04 M solutions of Na2CO3 and CaCl2 were injected at the top of
a ramp, which was inclined at 0.8–1.7°. Flow obstacles were placed 0.6–0.7 m downslope from the top of the ramp, and a camera (C) was
mounted to follow the carbonate growth near and on the obstacle. (b) Geometry of illumination and imaging such that specularly reflected
light does not enter the imaging lens. (c)–(e) Different contributions to imaged light intensities. (c) The three planes reflecting light before
precipitation. (d) Forward dominated scattering (reflected and refracted) of first and second generation. (e) Varying contribution of specular
reflection and scattered intensity in different directions as function of thickness of precipitate.
5014 Ø. Hammer et al. / Geochimica et Cosmochimica Acta 72 (2008) 5009–5021

measuring surface topography with nanometer resolution. glass interface, which was dyed black, scatters forward with
We used a Wyco NT1100 white light interferometer with a very low intensity. When the obstacle is sand blasted (see
5 and 50 Wyco objectives. Most mineral growth/disso- Fig. 1d) the glass–water interface scatters light with a strong
lution studies depend on the high resolution capability of forward intensity distribution. The lacquered surface also
this technique (e.g. Arvidson et al., 2003). There are three scatters forward, but with a much lower intensity. When
potential problems when this technique is applied to the small, randomly oriented crystals grow on the glass–water
measurement of porous, microcrystalline precipitates with interface they become centers of light scattering (see
height differences reaching the millimeter scale. The tech- Fig. 1e). A single crystal scatters light in a few, discrete
nique measures only the topography of the outer surface directions depending on orientation, but the refracted light
and gives no information about the porosity of the mate- is in turn reflected from the original surface of the obstacle.
rial. Secondly, in order to measure the thickness the posi- Averaged over random orientations and including the con-
tion of the bottom or substrate must be known. This is tribution from the original surface the scattering intensity
simply achieved by using a flat substrate which facilitates has a strong forward distribution in the specular direction.
determination of the height of the substrate at all lateral As an increasing number of small crystals grow ‘randomly’
positions. Finally, small precipitated crystals are randomly on top of each other, the scattering distribution broadens
oriented and many parts of the surfaces are therefore at and the intensity in the direction of the imaging lens in-
high angles to the substrate plane (the plane normal to creases. When the local volume of precipitate exceeds the
the objective axis). High inclination angle surfaces may only volume of scatterers contributing to the imaged intensity
be measured by high numerical aperture (NA) objectives, will no longer increase, because the recording system is sat-
and even with a 50, 0.55NA objective many parts of the urated. Different flow obstacle surface treatments were tried
surfaces will not be registered. After the measurement, the out to test the robustness of the technique.
topographic map is interpolated to fill the entire surface A Nikon D100 with a 60 mm Nikkormat lens was used
(to estimate the surface heights in the high inclination angle and the images were recorded in the Tiff 8 bit format. The
parts of the surface) and then the average height relative to 2000  3000 pixel red–green–blue (RGB) images were aver-
the substrate is calculated. In most cases we believe the aged to 2000  3000 grayscale images and treated as fol-
interpolation underestimates the average precipitate thick- lows: The images were rotated, typically by a few degrees,
ness, and we estimate the accuracy of the calculated precip- to exactly match the y-axis of the image with the down-
itate thickness to be ±30%. Since each single topography stream step edge. The recorded intensities, I (measured in
measurement covers only 186  248 lm several measure- 8 bits: 0 = no light recorded, 255 = maximum intensity)
ments are needed to obtain a representative average of were averaged along the y direction to obtain a vector of
the precipitate thickness. For large and relatively homoge- 3000 intensity values corresponding to about 20 mm in
nous samples the standard deviation of repeated measure- the x-direction. These vectors were smoothed using a 10
ments is 10–20%. In addition to these sources of error, pixel moving average yielding a 300 pixel intensity profile
rough substrates (lacquered, ground or sand blasted sur- in the x-direction. Then the intensity profile at time 0 was
faces) make the determination of the reference surface less subtracted from the intensity profiles at all later times to
accurate. obtain the diffuse scatter intensity profile in the direction
Outside of the flat top surface of the flow obstacles the of the lens.
WLI technique is not practical. Although the WLI can be In order to obtain precipitate thickness from the inten-
used in a fluid cell it is not practical to use it in situ because sity profiles a series of calibration experiments was per-
any perturbation of the water level would change the mea- formed. The calibration experiments confirmed our prior
sured topography. In order to estimate the precipitate observation that precipitation occurs as small crystallites
thickness all along the flow direction on and around the that scatter light and not as more-or-less uniform layer
flow obstacle a non-intrusive measurement technique that growth on a substrate. The very first crystallites have a pre-
is sensitive in the 1–100 lm vertical range with 10–100 lm ferred orientation dictated by the substrate, but long before
lateral resolution is needed. The natural choice was an opti- the substrate surface is covered the crystallites have a high
cal technique, and we chose to analyze the scattered light degree of randomness in their orientations. Fig. 2 shows the
intensity image obtained from the precipitate. This in situ results of the calibration experiments that simply compare
measurement of precipitate thickness in the x–y-plane is measurement using the two techniques (WLI and scattered
based on scattering of light by small, randomly oriented light intensity) for the same precipitates. There is a region
calcite crystals. Fig. 1b shows how the lamps, flow obstacle over which the scattered light intensity depends linearly
and imaging system were configured. The dashed lines indi- on the precipitate thickness. At low thickness (<1 lm) the
cate the field of view (FOV) and the arrows indicate that reason for the deviation is not clear. At high thickness
light rays from the lamps do not contribute to the recorded (>50 lm) the scattered intensity becomes saturated as ex-
intensity in the image when reflected specularly from the plained above. In the intermediate region the relation be-
surfaces inside the FOV. Fig. 1c and d indicates typical dis- tween thickness and intensity is characterized by a single
tributions of reflected light for two of the surface treatments constant. Each set of experiments with equal imaging con-
used. The lengths of the arrows indicate the reflected (re- ditions (surface quality, dry/wet, illumination, shutter
fracted) intensity from different interfaces and in different speed, camera lens, camera distance, etc.) has a distinct cal-
directions. When the flow obstacle is polished (Fig. 1c) ibration constant a that must be determined by a few ex situ
the glass–water interface reflects specularly and the glass– topography measurements with the WLI.
Calcite precipitation instability under open-channel flow 5015

150 flow. The computed velocity field (absolute values) is shown


in Fig. 3. Clearly, the protrusion has a number of effects on
the flow, including shallowing and consequently an increase
in velocity above and especially immediately downstream
100 from the protrusion. Both of these parameters have been
proposed as possible causes of increased calcite precipita-
tion – shallowing by the mechanism of enhanced degassing
(Chen et al., 2004), and velocity through the thinning of a
50
boundary layer through which solutes must diffuse in the
case of turbulent flow (Buhmann and Dreybrodt, 1985).
A standing wave (hydraulic jump) downstream from the
0 obstruction was also evident.
0 20 40 60 80 The carbon dioxide concentration field and the calcite
precipitation rate along the water–calcite interface are
shown in Figs. 4a and 5a. The most important feature is
Fig. 2. Calibration of precipitation thickness from imaged scatter-
the compression of the CO2 concentration profile and
ing intensity. The thickness in lm was measured by WLI. The
intensity was measured with 8 bits (0–255) resolution, I0 is the resulting steepening of the CO2 concentration gradient over
intensity before precipitation and a is a proportionality constant the obstruction. This enhances CO2 removal from the cal-
for each experiment series with equal imaging conditions. Data cite surface, leading to higher precipitation rates. Interest-
points with squares, asterisks and circles are results of different ingly, high precipitation rates are found not only in the
calibration experiments and the solid line is the fitted linear area of highest flow velocity around the downstream corner
calibration curve. Squares: Ex situ simultaneous imaging of four of the step, but also at the upstream corner and in the area
distinct experiments, each with homogeneous distribution of some distance downstream from the obstacle. The eddy be-
precipitate and averages of 35 topography measurements with hind the step produces a stagnant zone and a reduction in
a = 1. Asterisks: the same experiments using in situ imaging with
precipitation rate. The highest rates are at the corners of
a = .5. Circles: in situ imaging of one experiment with inhomoge-
the step, where the concentration gradients near to the sur-
neous precipitation and 16 topography measurements at different
sites on the sample with a = 1. face are steep. Also, the precipitation rate is elevated down-
stream because of the mixing. This precipitation
distribution would lead to a downstream migration of the
3. RESULTS step, together with outwards and upwards growth of the
step edge.
3.1. Model results Jamtveit et al. (2006) used stable isotopes to infer a cal-
cite precipitation rate of 14 mm/year at terrace edges in the
The model was run with water chemistry close to that Troll system. Using a density of 2.5 g/cm3 for travertine,
determined near the source of the Troll hydrothermal this corresponds to a precipitation rate of about 9.6 mg/
spring, Svalbard (Hammer et al., 2005). The model was ini- cm2/day, or about four times the modeled rate. On the
tially run with a low flow velocity of 2 mm/s in a smooth other hand, direct field measurement taken in a microter-
channel of depth 0.5 cm and length 100 cm, until a steady raced location with a pH similar to that used in the model,
state was reached after 500 s. In this initialization phase, indicated a rate of 0.72 mg/cm2/day, or about one fourth of
the diffusion constants were set to 50 times their nominal the modeled rate (Hammer et al., 2005, sample 5). The
values, to simulate the effects of vertical mixing due to a modeled precipitation rate is therefore within the range of
rough substrate (in the following runs the diffusion con- field measurements. Of course, the natural system has high
stants are not increased in this way). At the source, the spatial variability due to its complicated geometry, biolog-
log of the calcite saturation index SI = 0.04, pH 6.50 and ical activity and large-scale trends in water chemistry.
the calcite precipitation rate was F0 = 0.05 mg/cm2/day, In Fig. 4b, the flow rate has been reduced to zero while
while at the lower end of the channel, degassing led to val- keeping the water surface topography the same as that in
ues of SI = 0.57, pH 7.15 and F0 = 2.07 mg/cm2/day. We Fig. 4a, to facilitate comparison. In this simulation, the
use concentration fields were initially homogenous, and diffu-
sive gradients evolved over time. To the left of the obstruc-
ðCa2þ ÞðCO3 2 Þ
SI ¼ log10 ; ð21Þ tion, CO2 concentrations are elevated because of a greater
K sp water depth. In this case of pure diffusion, the elevation
with the calcite solubility product Ksp = 3.35  109 (molar of the precipitation rate on the flat top of the obstruction
units). is slight. This result shows that although reduction in water
The profiles for [Ca2+], [CO2] and [c] were extracted depth may cause some increase in precipitation rate due to
from the lower end of the modeled channel and used as in- enhanced degassing, this effect is relatively small compared
put for a second run. The depth-averaged concentrations at with the increase in precipitation caused by advection, at
this point are [Ca2+] = 3.9 mM (130 mg/l), [CO2] = least under the modeled conditions.
0.77 mM and [c] = 7.8 mM. In the second run, the modeled Figs. 4c and 5c show the results using a rounder obstruc-
domain was 3.0 cm long and 0.6 cm deep, and a 150  30 tion. Here the connection between flow rate and precipita-
grid was used. A rectangular protrusion obstructed the tion rate is even less marked, with high precipitation rates
5016 Ø. Hammer et al. / Geochimica et Cosmochimica Acta 72 (2008) 5009–5021

Fig. 3. Absolute values of flow velocity (mm/s) computed using the hydrodynamic model. Flow from left to right.

Fig. 4. CO2 concentrations in the model. (a) Flow past a rectangular obstruction with parameters approximating natural hydrothermal spring
conditions. (b) Flow rate forced to zero (pure diffusion). (c) Flow past a step with a rough surface. (d) Parameters approximating experimental
conditions.

on the upstream flank of the obstruction. In addition, pre- as NaCO3  . Modeling using the Phreeqc ver. 2.15 software
cipitation rates are highly variable on both flanks, where (Parkhurst, 1995) shows that ½NaCO3   will be on the order
the surface is rough (here the spatial discretization grid is of 10% of ½CO3 2 , and this was ignored in the simulation.
used to advantage by using it to represent a real, stepped Other species involving Na+ and Cl are negligible. The
morphology). Each peak corresponds to a small, protrud- precipitation rate distribution is comparable to that of
ing corner of the surface, where the diffusive concentration Fig. 5a, but with a more marked peak in the well-mixed
gradients are steep. area downstream from the obstruction.
Figs. 4d and 5d show flow over a step with geometry and A modeling experiment was also carried out to study ef-
chemical concentrations closer to those used in the experi- fects of local degassing and possibly high gas transfer rates
ment. The depth-averaged concentrations at the inflow (Eq. (17)) in areas of high flow velocity, due to e.g. agitation
boundary were [Ca2+] = 2.0 mM (79 mg/l), (Liss and Merlivat, 1986). In this experiment, transfer
[CO2] = 3.1  105 mM and [c] = 7.1 mM. In this alkaline velocities were made dependent on surface flow rate with
(pH 10–11) regime, CO2 concentrations were very low a large factor of proportionality, causing transfer velocities
and there was, in fact, ingassing from the atmosphere. to increase by a factor 100 in regions of fastest flow. The
The CO2 distribution is likely to be a less important control resulting precipitation pattern was indistinguishable from
on calcite precipitation rate under such conditions. In this that for a constant transfer velocity, and therefore the re-
solute, there will be some binding of ions into species such sults are not shown in the figures. Increasing the flow rate
Calcite precipitation instability under open-channel flow 5017

4
a
3

1
4
b
Precipitation (mg/cm2/day)

1
4
c
3

1
100
d
50

0
0 10 20 30
mm
Fig. 5. Calcite precipitation rates, in mg/cm2/day, corresponding to the model runs shown in Fig. 4.

to 20 mm/s gave patterns that were similar to those ob- travertine (Guo and Riding, 1998). It should be emphasized
tained at low flow rates, except that the precipitation max- that no matter the surface treatment (polished glass, sand
ima at the corners of the protrusion were relatively reduced blasted glass, organic spray paint) there is no measurable
(not shown). barrier to precipitation of the first calcite crystal seeds.
Crystals form readily anywhere on any available interface.
3.2. Experimental results Fig. 6 shows rough spheroidal calcite growth (cf. Jettest-
uen et al., 2006) in the stagnant water upstream from the
3.2.1. Qualitative features of calcite growth in experiments injection point of the fluid. The porous crystal agglomerates
While the fluid concentrations were optimized for fast suggest a high rate of crystal nucleation. On the upstream
calcite growth we observed that at high concentrations side of the step edge, the precipitate is similarly porous
(>4 mM) there was homogeneous nucleation with a varying and with rounded growth shapes. This is consistent with
induction time and no heterogeneous crystal growth on the natural travertine steps that have a hard, smooth down-
substrate. When the concentration was reduced below stream ‘‘lip” and a more porous upstream character.
4 mM heterogeneous calcite growth on the ramp would
commence as soon as the homogeneous nucleation ceased.
This suggests that travertine growth by agglomeration of
homogeneously nucleated calcite particles in the fluid flow
is not likely. This result deserves further study to check un-
der which conditions this may be true in nature. On the
other hand, calcite would grow on any water interface
(water–solid, water–air). Under some conditions ‘‘calcite
ice” was formed on the water–air interface (as can be ob-
served for example in the Mammoth Hot Springs in Yel-
lowstone). Calcite was also observed to grow on the gas–
water interface of gas bubbles produced from degassing
of N2 or CO2. Such bubbles increase the area of the gas–
water interface and thereby the total precipitation rate in
that region, creating an additional feedback mechanism
for localization of precipitation. Spherical holes, due to car- Fig. 6. Rough spheroidal growth of calcite in stagnant water, after
bonate growth on gas bubbles, are very common in natural 2 days.
5018 Ø. Hammer et al. / Geochimica et Cosmochimica Acta 72 (2008) 5009–5021

Fig. 7 illustrates the feedback between fluid flow and cal- from Fig. 8a and f, showing a porous agglomerate of small
cite growth, as fluid pathways change continuously due to particles, that it is difficult to measure the thickness of the
growth on the rim. Upstream from the obstruction (left precipitate locally. Fig. 8a shows a WLI topography image,
side) calcite grows as small clusters. This may be attributed obtained with a 5 objective, of a region of the obstacle
to a high nucleation barrier (the fast nucleation on the indicated in Fig. 8e. Inset Fig. 8ai is a higher resolution im-
obstruction on the other hand makes this an unlikely expla- age (using a 50, 0.55NA objective to capture light re-
nation) or a fingering instability upstream into the concen- flected from surfaces at larger angles) of the part of the
tration gradient. large image. Inset 8aii was obtained from the dataset shown
in inset Fig. 8ai, but with two-dimensional interpolation to
3.2.2. Quantitative experimental measurements of growth fill in the unresolved pixels (black regions in the images). In-
rate distribution set Fig. 8aii was used to determine the calibration factor a
In the experiments reported here, the ramp was inclined of the experiment, which was used to calculate the precipi-
at 1.7° and the flow rate was 1.8 ml/s. A pure water flow tate thickness profiles shown in Fig. 8b and c. Fig. 8b shows
was turned on to wet the ramp, and the lighting and camera precipitate thickness profiles at different times for a black
were adjusted before Na2CO3 and CaCl2 were injected into lacquered surface and Fig. 8c shows the same for the sand
the flow. Images focused on the rectangular flow obstacle blasted surface.
(e.g. Fig. 8d) were taken every 30 s during experiments that After 80 min, the average thickness at the top of the
lasted for 1–5 h. obstacle was about 5 lm (a growth rate of 90 lm per
Fig. 8e indicates the regions shown in Fig. 8a and f, and day). Considering the porous texture of the experimental
shows that the depth of focus is sufficient to simultaneously precipitate, this compares fairly well with the computer
map localized precipitation at the top of the obstacle and model result of about 40 mg/cm2/day of dense precipitate
immediately to its right. The images in Fig. 8a (ex situ white (Fig. 5d), or about 16 lm/day. The history of precipitation
light interferometer) and Fig. 8f (ex situ optical image) studies demonstrates (e.g. Compton and Daly, 1987;
show that the precipitate consists of small crystals with a Compton and Unwin, 1990; Wilkins et al., 2001, 2002) that
narrow size distribution (typically 10–30 lm) and fairly ran- exact match of precipitation rates from experiment and
dom orientations. Thickness profiles at 20, 40, 60 and modeling would require a much higher degree of control
80 min for two experiments with different surface treatment of the following features: fluid and substrate purity, effect
are displayed in Fig. 8b and c. We concentrated on the first of surface tension at the air–fluid interface, detailed com-
hours of carbonate growth for three reasons: (i) The nonlin- parison of the diffusion boundary layer (in our model the
ear relationship between precipitate thickness and intensity solid is atomically flat, in the experiments there are crystal-
at later times; (ii) extreme difficulty in measuring the distri- lites of 10–30 lm with randomly oriented surfaces).
bution of precipitate at later times; and (iii) the growth of a More important for this study, we observe that the
rim protruding into the flow and perturbing it. shapes of the precipitation profiles are very similar in both
The vertical axis in Fig. 8b and c shows the precipitate the model and experiment. The peaks on the corners of the
thickness in micrometers. The linear mapping from light obstruction are stronger in the experiment, perhaps due to
intensity to precipitate thickness was obtained by removing the effect of surface tension and the limitations of our mea-
the rectangular obstacle from the ramp after the experiment surement technique. We consider these results to be a rea-
and measuring the topography by WLI ex situ. It is clear sonable experimental validation of the model.

4. DISCUSSION

In general, advective solute transport seems to be the


overriding factor in controlling calcite precipitation rates
in open-channel flow, except when the surface relief is small
compared with the water depth, or when the flow rate is
very low. If the calcite substrate is not parallel to the flow
velocity, advection of dissolved species towards and away
from the surface is orders of magnitude faster than molec-
ular diffusion, and advection controls the steepness of the
concentration gradients. We propose that hydrodynami-
cally induced deformation of concentration gradients can
have a major impact on calcite precipitation patterns under
laminar flow conditions. Also, under turbulent flow, the
gradients in the laminar boundary layer will be similarly de-
formed by hydrodynamics.
In most natural systems, faster flow is generally found in
Fig. 7. Fluid flow (left to right) pattern over terrace rim showing
the vicinity of regions in which the water depth is small, and
how precipitation builds up to block the flow. The black traces are
due to ink poured into the flow. Upstream the ink is fairly evenly it can be difficult to spatially separate regions with high flow
distributed over the ramp. Across the precipitated travertine rim rates from those with small depths. In our model, the loca-
there are only two openings through which the water can flow. tion of maximum flow velocity is slightly displaced from the
Calcite precipitation instability under open-channel flow 5019

Fig. 8. (a) White light interferometer topography of a 0.85 mm  0.65 mm area at a step edge after 80 min (exp. 06.12.06). Black areas are not
resolved by the method. The two inset topography images are the same data set with (lower) and without (upper) data interpolation, taken
with a magnification objective that resolves larger reflection angles. The upper inset covers the corresponding region in the large image. The
interpolated data set (lower inset) is used to estimate the average thickness of precipitate. (b) and (c) Precipitate thickness from imaged
scattering intensity in situ after 20 (black circles), 40 (red diamonds), 60 (green squares) and 80 (blue triangles) minutes. (b) Exp. 11.12.06
(black lacquered surface, a = 0.5), (c) 08.12.06 (glass surface with black colour underneath as in exp. 06.12.06, a = 1). The proportionality
constants of calibration were based on measurement of average precipitate thickness at different positions. We estimate the peak growth rate
to be 0.8 ± 0.4 lm/min. (d) In situ image of step in flow ramp after 412 h of growth. (e) Closeup of image (d) on top of and right behind step. (f)
Ex situ image of a region on the top of the step from experiment 06.12.06 which was stopped after 80 min growth. A similar region of this
experiment is also displayed in (a).

location of maximum precipitation rate, indicating that the tion, and more compact travertine would result. This effect
progressive shallowing and resulting compression of con- may be responsible for the morphologies shown in Fig. 8e,
centration gradients established upstream from the position with porous, spheroidal growth on the low-velocity up-
of maximum flow velocity facilitates precipitation, rather stream side and a more compact, smooth ‘‘lip” on the high-
than the increased flow velocity. The model demonstrates er-velocity downstream side of the rim.
how shallowing over an obstruction enhances precipitation, Both the model and experiment also demonstrate the
causing an instability in the growth process under purely process that produces overhangs on travertine steps. The
laminar flow. The flow velocity in itself is not a causal fac- precipitation rate is high on the downstream upper corner
tor under these conditions. Instead, it is another result of of the obstruction, but very low at the base of the step.
shallowing. However, this result does not exclude the possi- Chen et al. (2004) invoked the Bernoulli effect as one of
bility that other mechanisms act to strengthen the instabil- their explanations for higher precipitation rates over
ity. Such mechanisms include thinning of a boundary layer obstructions. Regions of higher flow velocities have reduced
at high flow rates (Buhmann and Dreybrodt, 1985), en- pressures, which would also reduce the partial pressure of
hanced degassing under accelerating flow (Banerjee et al., CO2 on the water–air interface. This would increase degas-
2004) and agitation of the surface film, and sticking of par- sing according to Eq. (17). However, it can be easily shown
ticles to the obstruction. that this effect is minor, at least under the conditions mod-
The inhomogeneous precipitation pattern observed on eled here. Bernoulli’s law states that
the flanks of the obstruction in Fig. 5c, which we interpret 1
as a consequence of diffusion control, would lead to the P þ qv2 þ qgh ¼ c; ð22Þ
2
growth of calcite ‘fingers’ due to a Mullins–Sekerka insta-
bility (Mullins and Sekerka, 1963, 1964). This could be where P is the pressure in a fluid of density q and velocity v,
partly responsible for formation of the porous fabric that at a height habove a reference surface. Choosing the local
is often observed in natural travertine. At high flow rates, height of the air–water interface as our reference height
this diffusive effect is relatively small compared with advec- (h = 0), we find that c = Patm and
5020 Ø. Hammer et al. / Geochimica et Cosmochimica Acta 72 (2008) 5009–5021

1 Griebel, M., Dornseider, T. and Neunhoeffer, T. (1997) Numerical


P ¼ P atm  qv2 : ð23Þ
2 simulation in fluid dynamics. SIAM monographs on mathe-
matical modeling and computation 3. Society for Industrial and
It follows from Eq. (23) that even flow velocities as high as Applied Mathematics, Philadelphia.
1 m/s would only reduce the pressure by about 0.5%. In Guo L. and Riding R. (1998) Hot-spring travertine facies and
addition, precipitation patterns were not changed when sequences, Late Pleistocene, Rapolano Terme, Italy. Sedimen-
the degassing across the obstruction was increased by a fac- tology 45, 163–180.
tor of 100 in the simulations, showing that locally enhanced Hammer, Ø., Jamtveit, B., Benning, L. and Dysthe, D. K. (2005)
degassing near the water surface, whether due to agitation Evolution of fluid chemistry during travertine formation in the
or the Bernoulli effect, does not appreciably affect the car- Troll thermal springs, Svalbard, Norway. Geofluids 5, pp. 140–
bon dioxide distributions near the solid–liquid interface. 150.
This is partly because of loss of lateral contrasts in CO2 con- Hammer Ø., Dysthe D. K. and Jamtveit B. (2007) The dynamics of
travertine dams. Earth Planet. Sci. Lett. 256, 258–
centration from the air–water interface down to the water–
263.
calcite interfaces due to advection and diffusion, and in the Harned H. S. and Hamer W. J. (1933) The ionization constant of
case of low flow rates because the dissolved CO2 near the water. J. Am. Chem. Soc. 55, 693–695.
air–water interface is in near equilibrium with atmospheric Hillner P. E., Gratz A. J., Manne S. and Hansma P. K. (1992)
CO2 (cf. Fig. 4b, remembering that [CO2] at equilibrium is Atomic-scale imaging of calcite growth and dissolution in real
0.014 mM for the assumed atmospheric P CO2 ) so that degas- time. Geology 20, 359–362.
sing is not locally rate limiting. Degassing effects may be- Jamtveit B., Hammer Ø., Anderson C., Dysthe D. K., Heldmann J.
come locally important at larger scales however. and Vogel M. (2006) Travertines from the Troll thermal
springs, Svalbard. Norw. J. Geol. 86, 387–395.
Jettestuen E., Jamtveit B., Podladchikov Y. Y., deVillier S.,
ACKNOWLEDGMENTS Amundsen H. E. F. and Meakin P. (2006) Growth and
characterization of complex mineral surfaces. Earth Planet.
This study was supported by a Center of Excellence grant to Sci. Lett. 249, 108–118.
PGP from the Norwegian Research Council. Kaufmann G. and Dreybrodt W. (2007) Calcite dissolution kinetics
in the system CaCO3–H2O–CO2 at high undersaturation.
Geochim. Cosmochim. Acta 71, 1398–1410.
REFERENCES Lax P. D. and Wendroff B. (1960) Systems of conservation laws.
Commun. Pure Appl. Math. 13, 217–237.
Arvidson R. S., Ertan I. E., Amonette J. E. and Lüttge A. (2003) Liss P. S. and Merlivat L. (1986) Air–sea gas exchange rates:
Variation in calcite dissolution rates: a fundamental problem? introduction and synthesis. In The Role of Air–Sea Gas
Geochim. Cosmochim. Acta 67, 1623–1634. Exchange in Geochemical Cycling (ed. P. Buat-Menard). D.
Baker A. and Smart P. L. (1995) Recent flowstone growth rates: Reidel Publishing Company, Dordrecht, pp. 113–127.
field measurements in comparison to theoretical predictions. Liu Z. and Dreybrodt W. (1997) Dissolution kinetics of calcium
Chem. Geol. 122, 121–128. carbonate minerals in H2O-CO2 solutions in turbulent flow: the
Banerjee S., Lakehal D. and Fulgosi M. (2004) Surface divergence role of the diffusion boundary layer and the slow reaction
models for scalar exchange between turbulent streams. Int. J. H2 O þ CO2 ! Hþ þ HCO 3 . Geochim. Cosmochim. Acta 61,
Multiphase Flow 30, 963–977. 2879–2889.
Brown C. A., Compton R. G. and Narramore C. A. (1993) The Mullins W. W. and Sekerka R. F. (1963) Morphological stability of
kinetics of calcite dissolution/precipitation. J. Colloid Interf. a particle growing by diffusion or heat flow. J. Appl. Phys. 34,
Sci. 160, 372–379. 323–329.
Buhmann D. and Dreybrodt W. (1985) The kinetics of calcite Mullins W. W. and Sekerka R. F. (1964) Stability of a planar
dissolution and precipitation in geologically relevant situations interface during solidification of a dilute binary alloy. J. Appl.
of karst areas. 1. Open system. Chem. Geol. 48, 189–211. Phys. 35, 444–451.
Chen J., Zhang D. D., Wang S., Xiao T. and Huang R. (2004) Parkhurst, D. (1995) User’s guide to PHREEQC: a computer
Factors controlling tufa deposition in natural waters at program for speciation, reaction-path, advective transport, and
waterfall sites. Sediment. Geol. 166, 353–366. inverse geochemical calculations. USGS Water Resources
Chiarello R. P., Wogelius R. A. and Sturchio N. C. (1993) In situ Investigations Report 95-4227.
synchrotron X-ray reflectivity measurements at the calcite– Plummer L. N. and Busenberg E. (1982) The solubilities of calcite,
water interface. Geochim. Cosmochim. Acta 57, 4103–4110. aragonite and vaterite in CO2–H2O solutions between 0 and
Compton R. G. and Daly P. J. (1987) The dissolution/precipitation 90 °C, and an evaluation of the aqueous model of the system
kinetics of calcium carbonate: an assessment of various kinetic CaCO3–CO2–H2O. Geochim. Cosmochim. Acta 46, 1011–1040.
equations using a rotating disk method. J. Colloid Interf. Sci. Plummer L. N., Wigley T. M. L. and Parkhurst D. L. (1978) The
115, 493–498. kinetics of calcite dissolution in CO2–water systems at 5–60 °C
Compton R. G. and Pritchard K. L. (1990) Calcite dissolution and 0.0–1.0 atm CO2. Am. J. Sci. 278, 179–216.
kinetics at pH > 7: kinetics and mechanism. Phil. Trans. Roy. Press W. H., Flannery B. P., Teukolsky S. A. and Vetterling W. T.
Soc. Lond. A 330, 47–70. (1992) Numerical Recipes in C. Cambridge University Press,
Compton R. G. and Unwin P. R. (1990) The dissolution of calcite Cambridge.
in aqueous solution at pH < 4: kinetics and mechanism. Phil. Usdowski E. (1982) Reactions and equilibria in the systems CO2–
Trans. Roy. Soc. Lond. A 330, 1–45. H2O and CaCO3–CO2–H2O (0–50 °C). Neues Jahrb. Mineral.
Dreybrodt W. (1988) Processes in Karst Systems. Springer, Berlin. Abhand. 144, 148–171.
Goldenfeld N., Chan P. Y. and Veysey J. (2006) Dynamics of van Enckevort W. J. P. (1992) Phase shifting interferometry of
precipitation pattern formation at geothermal hot springs. growth-patterns on the octahedral faces of natural diamonds. J.
Phys. Rev. Lett. 96, 254501. Cryst. Growth 119, 177–194.
Calcite precipitation instability under open-channel flow 5021

Wilkins S. J., Compton R. G., Taylor M. A. and Viles H. A. (2001) kinetics of calcareous materials after sulphuric acid pre-treat-
Channel flow studies of the inhibiting action of gypsum on the ment and urban exposure. Stud. Conserv. 47, 88–94.
dissolution kinetics of calcite: a laboratory approach with Wooding R. A. (1991) Growth of natural dams by deposition from
implications for field monitoring. J. Colloid Interf. Sci. 236, steady supersaturated shallow flow. J. Geophys. Res. 96, 667–
354–361. 682.
Wilkins S. J., Compton R. G., Viles H. A. and Taylor M. P. (2002)
A new technique to evaluate and quantify modified solution
Associate editor: Roy A. Wogelius

View publication stats

You might also like