You are on page 1of 8

Geothermics 46 (2013) 1421

Contents lists available at SciVerse ScienceDirect

Geothermics
journal homepage: www.elsevier.com/locate/geothermics

Metal corrosion in geothermal brine environments of the Upper Rhine


graben Laboratory and on-site studies
N. Mundhenk a, , P. Huttenloch b , T. Kohl a , H. Steger a , R. Zorn b
a
b

Institute of Applied Geosciences/KIT, Germany


European Institute for Energy Research (EIFER), Karlsruhe, Germany

a r t i c l e

i n f o

Article history:
Received 3 July 2011
Accepted 31 October 2012
Available online 5 December 2012
Keywords:
Geothermal brine
Metal corrosion
Upper Rhine graben
Weight loss measurements
Electrochemical polarization (Tafel
method)

a b s t r a c t
Corrosion of construction materials in geothermal brine environments may play a major role in the
long-term operation and stability of geothermal power plants. Herein the results obtained from laboratory and on-site experiments are compared in order to evaluate candidate materials as to their
stability in hot brine environments. Weight-loss experiments in the Soultz-sous-Forts (Upper Rhine
graben, France) enhanced geothermal system, and electrochemical measurements in the laboratory
were conducted using 8 metals (unalloyed steels, stainless steels, and Ni-based alloy). Weight loss
and electrochemical measurements of the corrosion rate for unalloyed steels were found to be in good
agreement.
Both methods reveal unalloyed steels suffer surface recession with uniform corrosion rates <0.2 mm/y,
accompanied by surface deposits, providing minor protection. Stainless steel 1.4104 shows insufcient
stability during 4-week exposure, whereas 1.4539 and 1.4404 do not exhibit any visible corrosion. However, electrochemical measurements on stainless steels result in pitting corrosion with pitting potentials
dened for each steel quality. The very low corrosion rates obtained by electrochemical measurements
for the Ni-based alloy 2.4856 were conrmed by on-site exposure tests.
2012 Elsevier Ltd. All rights reserved.

1. Introduction
One of the major problems associated with the utilization
of deep geothermal energy is corrosion and degradation of the
construction materials (Corsi, 1986; Ellis, 1985). Assessment of
material behavior under these conditions is of crucial importance to
the utilization of geothermal energy. Material degradation may put
the long-term functionality of the materials used at risk (Shannon,
1975; Amend, 2009). For tubing steels, corrosion may be tolerated
as long as material thickness is sufcient. Functional components,
such as the heat exchanger, armatures and pumps, may lose their
functionality as a result of material degradation. For this reason, high-quality materials are used. The selection of construction
materials for a geothermal power plant is a trade-off between
investment costs vs. life time and replacement costs (Shannon,
1975).
The Upper Rhine graben as part of a large rift system in Central
Europe is characterized by a high heat ow of 100120 mW/m2 ,
compared to 80 mW/m2 in the surrounding area. The most prominent thermal anomaly is located at Soultz with temperatures
of 120 C at 800 m depth (Pribnow and Schellschmidt, 2000).

Corresponding author. Tel.: +49 721 608 47764; fax: +49 721 608 43374.
E-mail address: mundhenk@kit.edu (N. Mundhenk).
0375-6505/$ see front matter 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.geothermics.2012.10.006

Consequently, the Rhine graben offers favorable conditions for


geothermal power generation. Several power plants have been
built here over the last few decades. Some of these rely on water
from the granitic basement (e.g., the Soultz-sous-Forts EGS in
France) while others use water from overlying permotriassic sediments (e.g., the hydrothermal projects in Bruchsal and Landau,
Germany).
Regarding corrosion as a complex system behavior, many inuencing factors have to be taken into account. Apart from water
chemistry and pressure, temperature and hydrodynamic conditions have a major inuence. Furthermore, steel properties (e.g.,
composition, microstructure) must be considered (Carvalho et al.,
2005). The corrosive potential of the geothermal brines acting on
metals mainly depends on chemical key species, such as hydrogen ion, halogenides, and dissolved gases at high temperature and
pressure (Kaya and Hoshan, 2005). Deep reservoir waters from the
Upper Rhine graben are characterized by a high mineralization,
mainly by NaCl. In addition, high gas contents, mainly CO2 , are dissolved in the brine and increase the acidity. Typical pH values for
Upper Rhine graben brines are between 4.5 and 5.5, obtained by
direct measurements and theoretical considerations. Under operating pressure, a two-phase mixture of liquid and gas may form
during operation. In this environment, different modes of corrosion
(uniform and localized) are likely to occur. Uniform corrosion may
be tolerated, as long as material thickness is sufcient for longer

N. Mundhenk et al. / Geothermics 46 (2013) 1421

periods, but local attack of the material may cause severe damage
within short time spans.
Many studies focused on corrosion in geothermal systems (Pye
et al., 1989; Frick et al., 2011; Corsi, 1986; Ellis, 1985; Shannon,
1975; Kaya and Hoshan, 2005; Zhao et al., 2008; Banas et al., 2007;
Cui et al., 2004; Davis and Munir, 1977; Sampedro et al., 1988;
Batis et al., 1997; Thorbjrnsson, 1995; Bilkova et al., 2003; Baticci
et al., 2010; McCright, 1980; Pound et al., 1985; Pohjanne et al.,
2008; Lopez et al., 2003; Harrar et al., 1977; Stahl et al., 2000;
Culivicchi et al., 1985; Carter and Cramer, 1992; Zhao and Zuo,
2007). Moreover, corrosion in Mexican geothermal elds has been
widely studied including scaling and microbiologically induced
corrosion (Valdez et al., 2000, 2009).
However, studies that use original geothermal waters involving potentiodynamic polarization (PP) measurements and on-site
weight loss measurements (WLMs) are rare. This work focuses on a
comparison of PP and WLM using geothermal brines sampled at the
surface installation in Soultz-sous-Forts (France). WLM were conducted by on-site exposure tests in a corrosion bypass downstream
the heat exchanger in Soultz-sous-Forts. Table 1 summarizes the
scope and limitations of the different methods.

Fig. 1. Schematic current density vs. potential curve with relevant parameters; Tafel
region with Tafel slopes, thermodynamic domains: passivitytranspassivity.

According to the SternGeary equation (Eq. (1)), the corrosion current density icorr can be calculated from the resulting Tafel
parameters a and c :
icorr =

2. Electrochemical measurements
Electrochemical techniques were developed some 50 years ago
and are widely applied in corrosion research, because they provide
a sensitive and instantaneous value of the corrosion rate (Pound
et al., 1985). Corrosion processes in nature are spontaneous. In
electrochemical polarization, these processes are induced by external DC voltage which causes anodic and cathodic reactions on the
metal surface. In this way, processes can be accelerated by several orders of magnitude (France, 1969). The corrosion rate may be
expressed in terms of current density. The response on polarization can be shown by current density potential curves. The current
density measured is the sum of the anodic and cathodic partial
current densities. Knowing the anodic and cathodic reactions, the
resulting corrosion current density icorr can be determined by Tafel
slopes (Siebert, 1985). From the Tafel slopes, the corrosion rate can
be determined by the linear polarization resistance according to
Stern and Geary (1957).

1
a c
B

=
Rp
2.303(a + c ) Rp

(1)

where icorr is the corrosion current density (A/cm2 ); a , c are the


anodic and cathodic Tafel slopes (mV/decade); B is the constant
(mV) and Rp is the polarization resistance ( cm2 ).
The corrosion rate (mm/y) dened as the mean rate of removal
of metal while assuming uniform corrosion is calculated from the
corrosion current
CR = 0.00327

icorr EW


(2)

where CR is the corrosion rate (mm/y); 0.00327 is a constant (for


mm/y); EW is an equivalent weight (g/equivalent) of the corroding
species and  is the density of the corroding metal (g/cm3 ).
Fig. 1 shows a schematic current density vs. potential curve
obtained by potentiodynamic polarization. Relevant parameters
used for the calculation of corrosion rates are indicated. Dashed
lines indicate the extrapolated tting lines of anodic and cathodic
branches (Tafel slopes).

Table 1
Intended use and limitations of the methods applied to both material groups.
Laboratory PP measurements

Exposure tests

Intention
Short-term material assessment
Determination of corrosion rate and mode of corrosion
within a few hours

Intention
Long-term tests
Determination of corrosion rate and mode of corrosion under in-situ conditions
Scale investigation

Limitations
Pressure-less laboratory method

15

Limitations
Corrosion rate of full-size components may deviate from corrosion rate of coupon
sample of the same alloy

No in-situ conditions (max. Temp. 80 C)


No scale formation
Unalloyed steels
Calculation of uniform corrosion rate by Tafel equation
Interpretation of current densitypotential curves

Unalloyed steels
Calculation of uniform corrosion rate by weight loss
Time-dependent corrosion behavior
Analysis of corrosion products and scaling
Effect of crevice congurations
Effect of mechanical stress and welds

Stainless steels & Ni-based alloy


Interpretation of current densitypotential curves
Determination of critical potentials (free corrosion
potential Ecorr and pitting potential Epit )

Stainless steels & Ni-based alloy


Pitting-resistance evaluation
Time-dependent corrosion behavior
Analysis of corrosion products and scaling
Effect of crevice congurations
Effect of mechanical stress and welds

16

N. Mundhenk et al. / Geothermics 46 (2013) 1421

Note that different thermodynamic domains can be identied


for a passivating metal. The sharp increase of current density at
Epit marks the end of the passivity range and initiation of pitting
corrosion.
Apart from the determination of corrosion rates, critical potentials were dened, allowing for an evaluation of pitting resistance.
For each steel quality, signicant potentials (corrosion potential
Ecorr and pitting potential Epit ) can be specied, whereas Epit Ecorr ,
the range of passivity, is an indicator of pitting susceptibility
(Sedriks, 1979). The intersection of Eredox with the individual polarization curves indicates the thermodynamic state of the steel in
the solution. This may be either the passive state, immunity, or the
state of increased metal dissolution. Passive state (or passivation)
means the formation of a hard non-reactive oxide lm acting as a
non-conductive barrier between the metal and attacking solution.
Current density in the passive potential range is strongly reduced
(Fig. 1).
3. Experimental procedures
Electrochemical polarization measurements in the laboratory
and on-site exposure experiments were performed using unalloyed
and stainless steel samples. Due to different corrosion mechanisms,
unalloyed and stainless steels have to be studied separately.

PP scans were made for all steel qualities available. Except for
repassivation scans carried out at room temperature (22 C), all
PP measurements were conducted at 80 C. This temperature was
chosen for comparability reasons, since it represents the average
temperature in the bypass in Soultz (Baticci et al., 2010). To stabilize
the pH, CO2 entered the cell during the experiments.
After determining the free corrosion potential Ecorr , the scans
were run with a slew rate of 0.16 mV s1 . Each scan begins at
200 mV from Ecorr in anodic direction.
Reverse polarization scans were used to study the repassivation
behavior, since repassivation depends on chloride concentration.
The direction of polarization was switched by reaching 100 A in
these experiments.
Electrochemical measurements are particularly sensitive. Apart
from the aging of electrodes, laboratory procedures, such as sample preparation, have an inuence on the results. To assess the
accuracy of the results, an error analysis has been performed by a
series of measurements for 1.4404 in untreated 80 C Soultz brine.
As can be seen in Fig. 2, the shape of the curves is almost identical and the positions of Ecorr and Epit only vary in a narrow range.
Signicant deviation appears after switching to cathodic direction.
However, this implies that the reproducibility of measurements is
fairly good. Moreover, reference experiments were performed at
regular intervals according to ASTM G5-82.
3.2. Exposure tests: on-site corrosion bypass

3.1. Potentiodynamic polarization measurements


Potentiodynamic polarization (PP) measurements were carried
out with a KMZ 5 AEH by Sensortechnik Meinsberg (Germany).
This glass cell has a volumetric capacity of 500 ml. The main part
of the device is a three-electrode system with a steel sample
as a working electrode (exposed surface: 0.76 cm2 ), a platinum
counter electrode, and an Ag/AgCl reference electrode. The LMremote software provided by Sensortechnik Meinsberg controls the
potentiostat/galvanostat (PS 2000) and allows the simultaneous
measurement of additional parameters (electric conductivity,
redox potential, pH). To ensure homogeneous mixing, a mechanical stirrer (50 rpm) was used. Nevertheless, PP measurements were
conducted under quasi-stagnant conditions.

Fig. 2. Reproducibility measurements for 1.4404 in untreated 80 C Soultz brine.

Exposure tests at elevated temperatures and pressures are of


great importance to material evaluation, since they represent the
original corrosion conditions. The on-site bypass at Soultz has
3 chambers and is located on the re-injection side downstream
of the heat exchanger (Fig. 3). Previous studies using the on-site
corrosion bypass at Soultz were carried out by Baticci (2009).

Fig. 3. Corrosion bypass at Soultz-sous-Forts (Baticci, 2009).

N. Mundhenk et al. / Geothermics 46 (2013) 1421

17

Table 2
Material compositions (%) as obtained by WDX.
Sample Matl no.

AISI

Unalloyed steels
N80
P110
P235GH
Stainless steels
1.4104
1.4404
1.4539
1.4571
Ni-alloy
2.4856a (alloy 625)
a

UNS

Si

Mn

Cr

0.321
0.306
0.075

0.369
0.708
0.015

1.05
0.901
0.404

0.018
0.016
0.011

0.014
0.005
0.013

0.477
0.503
0.620
0.596

1.490
1.470
1.120
1.860

0.049
0.044
0.027
0.035

0.227
0.023
0.001
0.013

430F
316L
904L
316Ti

S43020
S31603
N08904
S31635

0.170
0.260
0.014
0.030

N06625

0.030.10

0.5

0.5

0.02

0.015

0.076
0.055
0.032

Mo

Ni

Cu

Ti

Fe

Nb

0.007
0.001
0.004

0.065
0.078
0.029

0.106
0.153
0.093

0.001
0.003
0.000

97.9
97.6
99.2

16.46
16.52
19.37
16.70

0.232
2.100
4.200
2.140

0.234
10.09
24.03
10.72

0.119
0.438
1.220
0.494

0.005
0.011
0.017
0.208

80.3
68.3
48.9
66.7

2023

810

3.154.15

58

0.5

0.4

Nominal composition.

Table 3
Typical composition of Soultz brine (Aquilina et al., 1997).
pH

Eredox (mV) (Ag/AgCl)

EC (mS/cm)

Na+ (mg/l)

Ca2+ (mg/l)

K+ (mg/l)

Mg2+ (mg/l)

Cl (mg/l)

5.33

13

108.7

27,900

6880

3180

142

48,507

Br (mg/l)

SO4 2 (mg/l)

Fe (tot) (mg/l)

Mn2+ (mg/l)

Rb+ (mg/l)

Sr2+ (mg/l)

Li+ (mg/l)

Cs+ (mg/l)

231.8

167

25.82

18.3

25.7

444.4

165.0

14.6

Operating and physico-chemical parameters are: ow rate 4 m3 /h,


temperature 80 C, pressure 1518 bar, pH 4.8. The chambers are
equipped separately with two coupons each, one for calculating
the corrosion rate, and one for the analysis of surface deposits.
Long-term tests of up to 5 months were performed in addition to
short-term tests by Baticci (2009).
Before exposure, the samples were prepared as described in Section 3.3. To calculate the corrosion rate by weight loss, the surface
area and the weight of each sample are necessary. After extraction, the samples were treated with inhibited HCl according to
ASTM G1-03 and weighed again. The uniform corrosion rate (surface recession) can be calculated using the following equation:
CR =

(K W )
(A T D)

(3)

where K is the constant (for mm/y): (K = 8.76 104 ); T is the time


of exposure (h); A is an area (cm2 ); W is the weight loss (g) and D
is the density (g/cm3 ).

typical composition is given in Table 3. The amount of dissolved


gases (>90% CO2 ) is considerably high, making the water slightly
acidic (pH 5). Under the reducing reservoir conditions, the content of dissolved O2 is negligibly low. The redox potential Eredox of
the geothermal brine is an important and a very sensitive parameter indicating the oxidation/reduction capacity of the solution.
On-site measurements under O2 -free sampling conditions yielded
13 mV vs. Ag/AgCl. Synthetic solutions for reverse polarization
experiments were prepared from pure water by mixing with NaCl.
Before the laboratory experiments, the water was de-aerated
by a vacuum pump and pressurized with high-purity CO2 (quality
4.8) to 6 bar. The acidity of the solution was adjusted to the pH
expected under high CO2 partial pressure by adding diluted HCl
before the experiment. To stabilize the pH, CO2 was added during
the experiments.
4. Results and discussion
4.1. Unalloyed steels

3.3. Metals
A total of 8 metals were tested in geothermal brine environment: commonly used casing and pipe steels P110, N80, and
P235GH, acid-resistant stainless steels 1.4104, 1.4404, 1.4571, and
1.4539, and the Ni-based alloy 2.4856. Chemical compositions
determined by wavelength dispersive X-ray (WDX) are given in
Table 2. These steels can be subdivided by their texture: ferritic
unalloyed steels, martensitic and austenitic alloyed steels. The steel
samples were either cut as rectangular samples or slices depending
on the received geometry. According to the ASTM G5-82 standard,
the metal samples were polished with SiC abrasive paper (120,
320, and 1000 grit), washed with pure water, degreased in acetone,
dried, and weighed.

Unalloyed steels are susceptible to corrosion under the test


conditions and suffer from uniform corrosion and shallow pitting.

3.4. Test solutions


The laboratory experiments were carried out with geothermal
brine and synthetic solutions prepared in the laboratory. The Soultz
geothermal brine was sampled from the surface installations by
ashing in a cooling coil and brought to the laboratory. For analysis, the samples were ltrated and acidied. The brine can be
classied as Na(Ca)Cl water and has a TDS of approx. 90 g/L. A

Fig. 4. Polarization curves of unalloyed steels in 80 C Soultz brine showing only


little difference in electrochemical behavior.

18

N. Mundhenk et al. / Geothermics 46 (2013) 1421

Table 4
Electrochemical data of metals exposed to geothermal water from Soultz-sous-Forts.
Steel
Unalloyed steels
P235GH
P110
N80
Stainless steels
1.4104
1.4404
1.4539
Ni-based alloy
2.4856
a

ba (mV)

bc (mV)

icorr (A/cm2 )

716
717
717

72.9
66.0
76.7

93.3
63.0
106.0

15.5
15.2
12.1

450
415
368

75.3
127
187.1

77.3
91.2
134.8

3.1
1.1
0.5

316
50
58

134
365
426

0.037
0.012
0.0055

257

67.4

57.8

0.2

0.0022

Ecorr (mV)

Epit (mV)

Epit Ecorr passivity (mV)

CR (mm/y)
0.180
0.177
0.140

2.4856 shows no pitting.

Table 5
Corrosion rates (CR) of unalloyed steels determined by weight loss after on-site
exposure at Soultz-sous-Forts.
Steel

Exposure time (h)

CR (mm/y)

P235GH
P110a
P110
N80a
N80

886.33
645.5
3693.5
645.5
3693.5

0.14
0.15
0.23
0.15
0.13

Baticci (2009).

During both time spans, scales formed on the surface were


observed to become successively adherent. Exemplarily SEM
images are shown in Fig. 6 for steel N80. The scale formed after
5 months could only be removed by acid treatment. The composition of the scale was inhomogeneous. Nevertheless, exemplarily
compositions from spot analysis of different samples are given in
Table 6. Scales on N80 and P235GH samples mainly consist of Fe
oxides and/or hydroxides. P110 shows a Pb-rich carbonate scale
with minor amounts of Sb. These results are consistent with the
observations made by Baticci (2009).
Fig. 5. SEMs of N80 after electrochemical polarization in Soultz brine; roughening,
shallow pitting, and undercutting are predominant corrosion modes.

PP measurements were performed using P110, N80, and P235GH


in geothermal brine. As their compositions are similar, the currentdensity potential curves are almost congruent (Fig. 4) and electrochemical data (Table 4) lead to similar results. Eq. (2) yields uniform
corrosion rates <0.2 mm/y. As revealed, free corrosion potentials
Ecorr of unalloyed steels are nearly identical. After extraction,
N80 samples were examined by SEM (Fig. 5). No scale formation
was observed during the experiments. In general, roughening of
the surface was noticed. In addition, uniform attack and surface
shallow pitting was found. Maximum penetration depth was about
26 m.
Unalloyed steel coupons were exposed to geothermal brine in
the on-site corrosion bypass at Soultz. Maximum exposure time
was 5 months. For all unalloyed steels, corrosion rates were below
0.23 mm/y. As shown in Table 5, the corrosion rate did not change
signicantly over time. The surfaces were roughened and both uniform corrosion and shallow pitting were observed. After 5 months,
the maximum penetration depth was 101 m for N80.

Fig. 6. SEM-images of N80: cross section of the corroded metal surface exposed
in the Soultz corrosion loop for 50 days (top) and 153 days (bottom); roughening,
undercutting, shallow pitting, and scale formation are visible.

Table 6
Scale composition from spot analysis (wt%).

N80
P110
P235GH

Cl

Na

Ca

Mn

Fe

As

Sr

Mo

Sb

Ba

Pb

3.5
16.5
4.5

25.8
7.4
26.4

3.5

2.6

1.8

1.2

1.2

1.3

39.5
1.1
30.2

3.1
2.8
6.5

0.7
1.2
0.6

3.4

3.8

1.7
8.3
2.8

4.8

1.6

9.8
61
16.1

N. Mundhenk et al. / Geothermics 46 (2013) 1421

Fig. 7. Polarization curves of stainless steels and the Ni-based alloy in 80 C Soultz
brine; measured redox potential Eredox 50 mV.

19

Fig. 8. Repassivation of 1.4571 in brine of various chloride concentrations.

4.2. Stainless steels


Stainless steels are susceptible to localized corrosion, involving
pitting. This mode of corrosion differs from that of unalloyed steels,
whereas the corrosion behavior of these 2 groups has to be studied separately. Fig. 7 shows the current density-potential-curves
of stainless steels. The Eredox value of Soultz brine measured onsite is indicated by dotted lines. The electrochemical data are given
in Table 4. All stainless steels tested respond to polarization with
pitting corrosion and, hence, are more or less susceptible to pitting corrosion. The individual resistance of a stainless steel can be
expressed by the relative positions of Ecorr , Epit , and Eredox .
Steel grade 1.4104 shows a massive increase of current density
near Ecorr and no passivity range. The intersection with Eredox is at
a very high current density of 105 A/cm2 . According to PP measurements, 1.4104 is not stable in 80 C Soultz brine. Steel grade
1.4404 exhibits a passivity range and the intersection with Eredox
is close to Epit . A slight increase in oxidation capacity may initiate
pitting. Grade 1.4539 shows a passive range of more than 420 mV
and a reduced passive current density of 1 A/cm2 in comparison
to 1.4404. 1.4539 performs best under the testing conditions. This
observation can be related directly to the alloy content (Table 2).
The content of alloy elements and their individual nobility have
a major effect on the electrochemical behavior during polarization. The order of corrosion potentials toward more positive values
is consistent with the ranking of the alloys toward a more noble
behavior.
Reverse polarization scans were performed to study the repassivation behavior after pitting initiation. In these experiments, the
chloride concentration of the solution was varied (0.550 g/L). The
results for 1.4571 in chloride solutions of different concentrations
are shown in Fig. 8. Since the chloride ion has a negative effect on
the stability of passive lms, the pitting potential is shifted toward
more negative (less noble) potentials. Furthermore, increasing hysteresis (higher current densities) at higher chloride levels can be
observed by reverse polarization at a current density of 100 A.
Low concentrations of chloride do not initiate pitting corrosion. The
increase of current density indicates transpassive dissolution. For
concentrations around 50 g/L, the repassivation potential is near
the free corrosion potential, meaning no repassivation is possible in
case of pitting. These experiments were carried out at room temperature (22 C). Elevated temperatures will increase the susceptibility
to pitting corrosion (Tousek, 1977). Since the concentration of chloride in Soultz brine is approx. 50 g/L, the vulnerability to pitting is
very high.

Fig. 9. Pitting corrosion of stainless steel 1.4104 after a 4-week exposure in the
corrosion bypass.

Stainless steel coupons were exposed to geothermal brine in


the on-site corrosion bypass for 1 month. No visible corrosion and
weight loss were observed for 1.4404 and 1.4539 coupons, whereas
1.4104 exhibited some texture-related pits (Fig. 9). All stainless
steels were blank after extraction from the bypass.
4.3. Ni-based alloy 2.4856
In PP measurements, grade 2.4856 exhibits uniform corrosion
with very low corrosion rates (0.0021 mm/y; Table 4). At higher
potentials, it does not respond with pitting and shows a wide range
of passivity, whereas the passive current density is very low (Fig. 7).
Compared to stainless steels, corrosion current density icorr is very
low. The sharp increase in current density at 900 mV indicates
breakdown of passivity and massive anodic dissolution, but pitting
is not initiated even when this potential is exceeded.
After on-site exposure, grade 2.4856 does not show any visible
corrosion attack and has a negligibly small uniform corrosion rate
(Table 7). After extraction from the bypass, the coupons were blank.
4.4. Method comparison
Laboratory PP results and weight loss results from on-site experiments were compared with regard to corrosion rates and modes
of corrosion (Table 7). It is obvious that corrosion rates obtained
by these methods are difcult to compare, since PP measurements were used for an electrochemically derived instantaneous
measurement of the corrosion rate under scaling-free and quasistagnant conditions, while weight loss exposure tests take place in
long-term tests under ow conditions including scaling effects.
Nevertheless, good agreement was found for unalloyed steels,
with the results obtained by PP measurement being slightly higher

20

N. Mundhenk et al. / Geothermics 46 (2013) 1421

Table 7
Comparison of corrosion rates calculated from electrochemical data and determined by the weight loss method in on-site exposure tests.
Metal

Unalloyed steels
N80
P110
P235GH
Stainless steels
1.4104
1.4404
1.4539
Ni-based alloy
2.4856
a
b

Alloy content (%)

PP measurements

Exposure test (weight loss)

CR (mm/y)

Mode of corrosiona

CR (mm/y)

Mode of corrosiona

0.18
0.14
0.18

U
U
U

0.15b
0.15b
0.14

U
U
U

19.7
31.7
51.1

0.037
0.012
0.006

P
P
P

0.0028
0
0

P
No visible attack
No visible attack

0.0021

0.0008

No visible attack

P, pitting corrosion; U, uniform corrosion.


Baticci (2009).

than those measured on exposed coupons by weight loss method.


This indicates, that scalings formed during exposure period only
offer minor protection. In fact, experiences from geothermal industry lead to the assumption that corrosion under stagnant or
low-turbulence ow regimes is more severe.
For the stainless steels, uniform corrosion rates of <0.04 mm/y
are obtained by PP measurements. These values by far exceed
those of the exposure tests. For stainless steels, the mode of
attack involves pitting. Consequently, uniform corrosion rates are
of minor interest. The stability against pitting qualies a stainless
steel under specic conditions. Information regarding the vulnerability to pitting can be obtained by PP measurements, whereas
exposure tests yield qualitative results (pitting or no pitting). The
results from PP measurements correlate well with observations
made on coupons after 4-week exposure tests.
1.4104 Insufcient performance in exposure tests and PP measurements.
1.4404 No visible attack after exposure tests, PP measurements
insufcient.
1.4539 No visible attack after exposure tests, good performance
in PP measurements.
The corrosion rate of the Ni-based alloy 2.4856 obtained by
PP measurements is considerable higher than for exposure tests,
whereas the mode of attack is the same (uniform attack).
5. Summary and conclusions
Two different methods of corrosion research have been applied
potentiodynamic polarization measurements in the laboratory
and on-site exposure tests in Soultz-sous-Forts. Although both
methods are widely applied in corrosion research, material assessment using both methods with a geothermal focus is still uncharted
territory. For the Upper Rhine graben with its huge geothermal
potential no major efforts have been taken.
Two aspects were of principal interest in this work (1) a stability evaluation for different metals and alloys and (2) a methods
comparison between PP measurements and on-site exposure tests.
For unalloyed steels, corrosion rates (<0.2 mm/y) obtained by PP
measurements and on-site exposure tests correlate well, whereas
scaling is likely to occur in long-term tests. Slightly increased corrosion rates obtained by PP measurements represent a worst case
scenario. Additionally, the mode of corrosion was identied for
each steel quality and conrmed on exposure coupons. The results
can be used for a pre-assessment of the suitability of materials in
laboratory short-term tests.
The stainless steels show pitting as the predominant mode
of corrosion. The vulnerability to pitting can be assessed by PP

measurements. A useful indicator is Epit Ecorr . The signicant


potentials Ecorr and Epit can be related directly to Eredox of the brine.
Current density-potential curves therefore contain information
about the thermodynamic state of the metal during immersion,
i.e., on immunity, passivity, or increased metal dissolution. The
corrosion performance of these steels varies under testing conditions. Corrosion resistance of the steels decreases in the following
order: 1.4539, 1.4404, and 1.4104. This directly correlates with the
alloy content. The latter steel failed in exposure tests, whereas the
other two were stable in 4-week exposure tests. Since corrosion is
a spontaneously induced process, long-term stability of stainless
steels remains uncertain. Long-term exposure tests are therefore
necessary.
A special remark has to be made on the chloride content of
geothermal brines. Particularly stainless steels are increasingly
susceptible to localized corrosion, when the chloride content
exceeds a critical concentration. Reverse polarization scans of
the stainless steel 1.4571 were therefore used to evaluate the
repassivation tendency. With increasing chloride concentrations,
this tendency is successively hindered. If pitting occurs in 50 g/L
chloride brine (comparable to Soultz brine), no repassivation
is possible and existing pits will grow steadily. Identifying
critical concentrations for different steels and alloys at elevated temperatures and pressures could be the focus of further
research.
The Ni-based alloy 2.4856 shows the best performance in terms
of corrosion resistance. It does not respond with pitting in PP measurements and exposure tests. A wide range of passivity could be
observed by PP measurements.
This study can help constructing engineers to decide which
material is applicable to specic conditions and may promote the
development of geothermal activity in the Upper Rhine graben.
Nevertheless, since challenges are still remaining, further research
has to be done on this eld. Our future work will include on-site
exposure tests of different materials (including polymer coatings)
on the high-temperature side at Soultz, more profound electrochemical measurements, laboratory autoclave exposure tests, and
experiments with corrosion-relevant H2 S. Indications, that corrosion occurs more severely in stagnant regions rather than under
ow conditions is also of special interest and will be tackled in the
future.
Acknowledgments
This project has been gratefully nanced by EnBW Energie
Baden-Wrttemberg AG and the Federal Ministry for the Environment, Nature Conservation and Nuclear Safety (BMU; Fkz:
0325111C). We would like to thank the GEIE (Exploitation
Minire de la chaleur) for giving access to the Soultz geothermal

N. Mundhenk et al. / Geothermics 46 (2013) 1421

installations. For on-site support Albert Genter and Joachim Place


are gratefully acknowledged.
References
Amend, B., 2009. Considerations in the testing and selection of materials for corrosive geothermal environments. GRC Transactions 33, 633637.
Aquilina, L., Pauwels, H., Genter, A., Fouillac, C., 1997. Water-rock interaction process
in the Triassic sandstone and the granitic basement of the Rhine graben: geochemical investigation of a geothermal reservoir. Geochimica et Cosmochimica
Acta 61, 42814295.
ASTM G1-03, 1885. Standard practice for preparing, cleaning, and evaluating corrosion test specimens. In: Haynes, G.S., Baboian, R. (Eds.), Laboratory Corrosion
Tests and Standards. ASTM Committee G-1 on Corrosion of Metals. ASTM Special
Technical Publication 866, pp. 534544.
ASTM G5-82, 1985. Standard practice for standard reference method for making potentiostatic and potentiodynamic anodic polarization measurements. In:
Haynes, G.S., Baboian, R. (Eds.), Laboratory Corrosion Tests and Standards. ASTM
Committee G-1 on Corrosion of Metals. ASTM Special Technical Publication 866,
pp. 513521.
Banas, J., Lelek-Borkowska, U., Mazurkiewicz, B., Solarski, W., 2007. Effect of CO2 and
H2 S on the composition and stability of passive lm on iron alloys in geothermal
water. Electrochimica Acta 52, 57045714.
Baticci, F., 2009. Material study on geothermal EGS (Enhanced Geothermal
System) power plant: application to the Soultz-sous-Forts site. Diploma
thesis of the Politecnico di Milano, Italia, Facolta di Ingegneria Industriale,
p. 202.
Baticci, F., Genter, A., Huttenloch, P., Zorn, R., 2010. Corrosion and scaling detection in
the Soultz EGS power plant, Upper Rhine graben, France. In: Proceedings World
Geothermal Congress, Bali, Indonesia, 2529 April, p. 11.
Batis, G., Kouloumbi, N., Kotsakou, K., 1997. Corrosion and protection of carbon steel
in low enthalpy geothermal uid. The case of Sousaki in Greece. Geothermics
26, 6582.
Bilkova, K., Seiersten, M., Muller, J., Faucher, J-P., 2003. Corrosion testing for hot
dry rock geothermal wells in Soultz-sous-Forts. EC Contract SES6-CT-2003502706, p. 10.
Carter, J.P., Cramer, S.D., 1992. Materials of Construction for High-Salinity Geothermal Brines. Report of Investigations/1992, RI-9402, U.S. Department of the
Interior, Bureau of Mines, p. 8.
Carvalho, D.S., Joia, C.J.B., Mattos, O.R., 2005. Corrosion rate of iron and ironchromium alloys in CO2 medium. Corrosion Science 47, 29742986.
Corsi, R., 1986. Scaling and corrosion in geothermal equipment: problems and preventive measures. Geothermics 15, 839856.
Cui, Z.D., Wu, S.L., Li, C.F., Zhu, S.L., Yang, X.J., 2004. Corrosion behavior of oil tube
steels under conditions of multiphase ow saturated with super-critical carbon
dioxide. Materials Letters 58, 10351040.
Culivicchi, G., Palmerini, C.G., Scolari, V., 1985. Behaviour of materials in geothermal
environments. Geothermics 14, 7390.
Davis, R.B., Munir, Z.A., 1977. Corrosion susceptibilities of various metals and
alloys in synthetic geothermal brines. Journal of Materials Science 12,
19091913.
Ellis, P.F., 1985. Companion study guide to short course on geothermal corrosion and
mitigation in low temperature geothermal heating systems. Report for the GeoHeat Center Oregon Institute of Technology, Klamath Falls, DCN-85-212-040-01,
p. 34.

21

France Jr., W.D., 1969. Controlled potential corrosion tests, their applications and
limitations. Test Methods Forum, Materials Research and Standards, 2126.
Frick, S., Regenspurg, S., Kranz, S., Milsch, H., Saadat, A., Francke, H., Brandt, W.,
Huenges, E., 2011. Geochemical and process engineering challenges for geothermal power generation. Chemie Ingenieur Technik 83 (12), 20932104.
Harrar, J., McCright, R., Goldberg, A., 1977. Field electrochemical measurements of
corrosion characteristics of materials in hypersaline geothermal brine. Report
UCRL-52376, Lawrence Livermore Laboratory, Livermore, California, p. 21.
Kaya, T., Hoshan, P., 2005. Corrosion and material selection for geothermal systems.
In: Proceedings World Geothermal Congress, Antalya, Turkey, 2429 April.
Lopez, D.A., Perez, T., Simison, S.N., 2003. The inuence of microstructure and
chemical composition of carbon and low alloy steels in CO2 corrosion. A stateof-the-art appraisal. Materials and Design 24, 561575.
McCright, R.D., 1980. Corrosion behaviour of materials exposed to hypersaline
geothermal brine. Master thesis, 1980. Lawrence Livermore Laboratory, UCRL85174, CONF-810402-4.
Pohjanne, P., Carpen, L., Hakkarainen, T., Kinnunen, P., 2008. A method to predict
pitting corrosion of stainless steels in evaporite conditions. Journal of Constructional Steel Research 64, 13251331.
Pound, B.G., Abdurrahman, M.H., Glucina, M.P., Wright, G.A., Sharp, R.M., 1985. The
corrosion of carbon steel and stainless steel in simulated geothermal media.
Australian Journal of Chemistry 38, 11331140.
Pribnow, D., Schellschmidt, R., 2000. Thermal tracking of upper crustal uid ow in
the Rhine graben. Geophysical Research Letters 13 (27), 19571960.
Pye, D.S., Holligan, D., Cron, C.J., Love, W.W., 1989. The use of Beta-C titanium for
downhole production casing in geothermal wells. Geothermics 18, 259267.
Sampedro, J.A., Rosas, N., Diaz, R., Dominguez, B., 1988. Developments in geothermal
energy in Mexico part nineteen. Corrosion in Mexican geothermal wells. Heat
Recovery Systems and CHP 8, 355362.
Sedriks, A.J., 1979. Corrosion of Stainless Steels, 2nd ed. John Wiley & Sons, New
York, p. 282.
Shannon, D.W., 1975. Economic impact of corrosion and scaling problems in geothermal energy systems. Master thesis, Battelle Pacic Northwest Laboratories
Richland, Washington 99352, 1975, BNWL-1866 UC-4, p. 115.
Siebert, O.W., 1985. In: Haynes, G.S., Baboian, R. (Eds.), Laboratory Electrochemical Test Methods, Laboratory Corrosion Tests and Standards, ASTM STP 866.
American Society for Testing and Materials, Philadelphia, pp. 6590.
Stahl, G., Patzay, G., Weiser, L., Kalman, E., 2000. Study of calcite scaling and corrosion
processes in geothermal systems. Geothermics 29, 105119.
Stern, M., Geary, L.A., 1957. Electrochemical polarization: I, a theoretical analysis
of the shape of polarization curves. Journal of the Electrochemical Society 104,
5662.
Thorbjrnsson, I., 1995. Corrosion fatigue testing of eight different steels in an
Icelandic geothermal environment. Materials and Design 16, 97102.
Tousek, J., 1977. Temperature dependence of pitting corrosion in CrNi stainless
steels. Werkstoffe und Korrosion 28, 619622.
Valdez, B., Rioseco, L., Schorr, M., Navarrete, M., 2000. Deterioration of materials in
geothermal elds in Mexico. Materials and Corrosion 51, 698704.
Valdez, B., Schorr, M., Quintero, M., Carrillo, M., Zlatev, R., Stoytcheva, M., Ocampo,
J., 2009. Corrosion and scaling at Cerro Prieto Geothermal Field. Anti-Corrosion
Methods and Materials 56, 2834.
Zhao, J.M., Lu, Y., Liu, H.X., 2008. Corrosion and control of P110 oil tube steel in CO2 saturated solution. Corrosion Engineering, Science and Technology 43, 313319.
Zhao, J.M., Zuo, Y., 2007. Corrosion behaviour in simulated solutions within pits
and crevices on carbon steel. Corrosion Engineering, Science and Technology
42, 203206.

You might also like