Sie sind auf Seite 1von 7

Understanding localized

corrosion
The breakdown of a protective passive film leading to accelerated
dissolution at localized sites is an important practical issue and a vexing
scientific problem. The small dimensions, short timescale, and dynamic
interplay between a heterogeneous surface and changing potential
and solution concentration gradients complicate the development of a
complete understanding of the phenomena. This review touches on some
of the recent developments in the field, including scanning tunneling
microscopy imaging of the earliest stages of pitting which supports a
new model explaining the localization of attack, pitting in thin aqueous
layers relevant to atmospheric corrosion, the factors controlling crevice
corrosion, and predictive modeling of localized corrosion.
G. S. Frankela* and N. Sridharb
aFontana Corrosion Center, The Ohio State University, Columbus, OH 43210, USA
bDNV Research and Innovation, Dublin, OH 43017, USA

*E-mail: frankel.10@osu.edu

Localized corrosion is the accelerated attack of a passive metal that many pits cease to grow at this stage); stable growth of localized
in a corrosive environment at discrete sites where the otherwise corrosion sites that can grow quite large; and finally – if conditions
protective passive film has broken down1. Common forms of permit – repassivation or cessation of attack. Localized corrosion is
localized corrosion include pitting on a boldly exposed surface, known to initiate above a critical potential and repassivate below
corrosion in a creviced region shielded from the bulk environment, another, lower potential1. Once initiated, the attack is stabilized by
intergranular corrosion of an alloy with a susceptible grain the localized development of aggressive conditions as the result of
boundary region, and exfoliation corrosion, which is a form of metal cation hydrolysis and Cl– migration. A conservative criterion
intergranular corrosion involving the prying apart of elongated for localized corrosion is the prediction that it can occur at potentials
grains by a voluminous grain boundary corrosion product. above the repassivation potential and should not occur if the corrosion
The various localized corrosion phenomena have been studied potential remains below the repassivation potential by a margin of
for decades and considerable understanding has been achieved. The safety that should be at least 100 mV.
process of localized corrosion can be divided into a sequence of steps1: Despite the advances that have been made, fundamental
initiation by the breakdown of the passive film; metastable growth of understanding and prediction of localized corrosion is a difficult
small pits on the verge of stability (metastable in this context indicates problem for a number of reasons. First of all, the events take place on

38 OCTOBER 2008 | VOLUME 11 | NUMBER 10 ISSN:1369 7021 © Elsevier Ltd 2008


Understanding localized corrosion REVIEW

that of the passive film outside of the pits, suggesting that the pits
were never depassivated but rather that there was accelerated attack
through the passive film at these sites. Such a view is consistent with
the adsorption mechanism of passivity breakdown4.
A recent publication has shown that the nanometer-sized pits,
which can form on Ni single crystals even in Cl–-free solutions such as
H2SO4, are found primarily at the boundaries between grains in the
passive oxide film5. A model for pit initiation based on the reactivity of
the oxide grain boundaries was also presented5. This model considers
that these oxide grain boundary regions have a lower resistance to
ionic transport and are thus the sites of oxide breakdown. Cl– competes
with OH– adsorption and accelerates ion transport at the local defect
sites. The experimental evidence cannot distinguish between local
thinning and suboxide void formation as the initiation mechanism.
However, this model presents the first explanation for the localization
of attack on the surface of a homogeneous metal sample.
Localized corrosion has been studied almost exclusively in bulk
solutions relevant to complete immersion conditions. However, there
Fig. 1 STM image of pits formed in single-crystal Ni(111). The surface was
passivated first at 0.96 V SHE in 0.05 M H2SO4 + 0.095 M NaOH and then NaCl
are a number of important applications where localized corrosion of
was added to 0.05 M. z range = 12.3 nm 3. passive metals under atmospheric corrosion conditions might play a
role. The proposed site for disposal of high-level radioactive waste
a very small scale, with a passive film nanometers in thickness and in the United States is Yucca Mountain in Nevada. This location
initiation sites of a similar size. Immediately after initiation, the rate is well above the water table but the metal waste canisters in the
of pit growth can be extremely high, even tens of A/cm2 2. Therefore, emplacement tunnels will be exposed to atmospheric corrosion
the situation is extremely dynamic with rapidly moving boundaries and conditions associated either with dripping water or deliquescence of
rapidly changing chemistries. Moreover, initiation events are relatively surface salts or dust in a humid environment6,7. Atmospheric localized
scarce, with maybe a few sites per cm2, and stochastic in nature. As a corrosion can also occur in the vapor space above a closed storage
result, it is not possible to predict exactly when and where breakdown tank8. No standard electrochemical methods exist for the study
will occur, which makes high-resolution observation of initiation of atmospheric corrosion9, but Stratmann and coworkers10–13 and
events extremely difficult. Finally, the exact susceptibility of sites for Nishikata and coworkers14–19 have developed different approaches
breakdown and the ability of other sites to sustain cathodic reactions for electrochemical measurements under thin electrolyte layers.
depend on details of the surface film properties, such as catalytic Stratmann and coworkers11–13 developed Kelvin probe methods, but
activity, that are not well understood, difficult to measure on the scale did not focus on localized corrosion until a recent study that showed
of the sites, and change with time as the process evolves. no effect of solution layer thickness down to tens of micrometers
The following will describe some recent advances in the on the breakdown potential of 304 stainless steel10. Nishikata and
understanding of localized corrosion. The topics discussed include coworkers17–19 also studied pitting corrosion of stainless steel under
scanning tunneling microscopy (STM) imaging of the earliest stages of thin electrolyte layers. They used samples composed of a working
pitting, pitting in thin aqueous layers (which is relevant to atmospheric electrode (WE), a counter electrode (CE), and a chloridized Ag wire
corrosion), controlling factors in crevice corrosion, and modeling of reference electrode (RE) embedded in epoxy under thin electrolyte
localized corrosion. layers and exposed them to cyclic wet–dry conditions. The corrosion
As mentioned above, the earliest stages of localized corrosion rate associated with pitting was assessed by electrochemical
breakdown are the least understood. The smallest of pits have in impedance spectroscopy measurements during cyclic wet–dry cycling
recent years been imaged with STM3. Fig. 1 shows a single-crystal on different stainless steels18. Corrosion potential measurements were
Ni surface passivated at a controlled potential in a Cl–-free solution made on an embedded WE using a different design in which an agar-
followed by addition of NaCl to the solution. The resulting pits are filled hole in the epoxy allowed ionic connection to a remote saturated
extremely small, with an average depth of only 3 nm. Such features are calomel RE19. A decrease in the corrosion potential by hundreds of
not present before the addition of Cl– and the charge associated with millivolts coincided with a sharp decrease in polarization resistance
the dissolution is much too small to discern via a current transient. and the onset of pitting corrosion as a critical Cl– concentration was
The bottom of the pits has a passive film with a structure identical to reached during drying.

OCTOBER 2008 | VOLUME 11 | NUMBER 10 39


REVIEW Understanding localized corrosion

(a) (b)

Fig. 2 Schematic diagram of (a) a crevice with the cathode covered by a thinfilm of electrolyte containing particulates, and (b) a decoupled cathode simulating the
effect of area blockage by particles 22.
Tada and Frankel20,21 took an approach similar to that of Nishikata available to sustain dissolution at the anode (Fig. 3). The cathode
et al. to investigate the effects of a particulate layer on the breakdown capacity increases with increasing electrode kinetics as reflected by the
of 304L stainless steel in thin layers of electrolyte. They used layers exchange current density, increasing bulk solution conductivity, and
of silica particles deposited by electrophoretic deposition to simulate decreasing repassivation potential. The particle loading strongly affects
particles that might deposit on a surface under atmospheric conditions. the cathode capacity for high solution resistance and fast kinetics.
Using open circuit potential and galvanic current transients, they A variety of models have been developed to describe localized
found that the silica layer had a greater effect on propagation than on corrosion. The most atomistically based model available for the
initiation of pitting corrosion21. The propagation of pitting corrosion for prediction of localized corrosion susceptibility is the point defect
silica-coated samples was slower than for uncoated 304 stainless steel. model (PDM) developed by Macdonald et al.23,24. The PDM considers
The effects of inert particulate in a thin electrolyte film on the the transport of cation vacancies in the passive film towards the film–
ability of cathodic sites to provide the necessary current for spatially metal interface, where they can condense to form voids or nascent pits
separated localized corrosion sites to continue growing was reported at the metal–passive film interface if the flux of vacancies arriving at
recently by Agarwal et al.22. Particles decrease the electrolyte the interface is greater than what can be consumed by annihilation via
conductivity by volume blockage and also alter the available area. a metal oxidation reaction. The effect of an aggressive species, such
Numerical methods were used to solve the Laplace equation and as Cl–, is hypothesized to alter the generation and transport of cation
generate three-dimensional simulations of the current and potential vacancies. The model correctly predicts the logarithmic dependency
distributions. The anode was considered to be in a shielded crevice of pit initiation potential on Cl– concentration, the effects of some
and the particles were located in a thin electrolyte region outside of alloying elements, and the effect of parameters such as scan rate on
the crevice (Fig. 2a). The particles blocked the cathode area, which pit initiation. In a recent development of the PDM, Macdonald and
was modeled by a segmented decoupled cathode as shown in Fig. Yang25 modeled the inhibitive effect of oxyanions, such as NO3– and
2b. Calculations were performed to predict the effects of a number BO3–, on pit breakdown potential of 316L stainless steel by considering
of different parameters on the cathode capacity, which is the current the competitive adsorption of aggressive and inhibitive species at

(a) (b) (c)

Fig. 3 The effects of various properties on the cathode capacity assuming 25 µm3 particles at an area and volume fraction of 0.306: (a) exchange current density for
the cathodic reaction; (b) bulk solution conductivity; (c) repassivation potential22.

40 OCTOBER 2008 | VOLUME 11 | NUMBER 10


Understanding localized corrosion REVIEW

O vacancies at the film–solution interface. The model predicts that the a pit or crevice associated with hydrolysis and electrolytic migration
breakdown potential should vary with of Cl–, which combine to form an aggressive local environment that
stabilizes attack27–29.
⎛ [X − ] ⎞
log ⎜⎜ ⎟, Recent work has utilized coupled multielectrode arrays to
⎝ [Y z −] ⎟⎠
investigate the spatial and temporal crevice corrosion of various
where [X-] is the concentration of aggressive species such as Cl- and corrosion-resistant alloys30. Using scaling laws, these experiments have
[Yz-] is the concentration of inhibiting oxyanions25. Experimental interpreted data from multiple-crevice assembly electrodes, which are
results found this relationship to be true for 316L stainless steel in commonly used for studying the susceptibility of an alloy to crevice
Cl– + NO3– solutions (Fig. 4)25. It should be noted that criticisms of corrosion. This scaling model predicted a variation of the critical
the PDM have focused on the linearization of the transport equations crevice depth with the square root of crevice gap distance30. The
and how the potential drop at the oxide/electrolyte interface is multielectrode arrays consisted of 100 close-packed circular wire cross-
handled4. Furthermore, the PDM predicts the pitting potential, whereas sections covered by a polymer crevice former. The less resistant 316
the breakdown occurs at much lower potentials as evidenced by stainless steel dissolved in the crevice at the mass-transport-limited
metastable pitting transients in the data25. A bilogarithmic dependence condition, resulting in crevice corrosion that started near the mouth of
of pitting potential on Cl– concentration is often observed, but is not the crevice and then spread inward and outward over time. In contrast,
predicted by the PDM at present. Finally, the PDM does not explain the more resistant alloy 625, a Ni-base alloy with 20–23% Cr and 8–
why breakdown is localized at discrete sites. 10% Mo, exhibited an active–passive transition in the crevice solution
Most models for describing the stability of growing localized so that the attack started at a location deeper in the crevice where the
corrosion have focused on crevice corrosion. A growing pit is essentially ohmic drop was sufficiently high. The location could be predicted using
identical to a growing crevice, though the transport distance varies a model based on polarization data in simulated crevice solutions and
much less with time for a crevice as it is set primarily by the crevice the potential distribution down the crevice30.
geometry. There are two primary models for crevice stability, which In recent years, Anderko et al.31–34 have developed a different
focus on local potential or local chemistry, respectively. The critical approach to modeling localized corrosion by adapting Okada’s35,36
potential drop theory promoted by Cho and Pickering26 describes irreversible thermodynamic model and combining it with a model
the ohmic potential drop associated with ionic current flow that of thermodynamic speciation37,38. This approach focuses on the
must exist if the anodic current in the crevice is connected to an repassivation potential, which is lower than the breakdown potential
external cathode. According to the potential drop theory, the potential and therefore a more conservative value to use as a design parameter.
decreases from the passive region at the outside of the crevice to the As mentioned above, localized corrosion might be possible when
active region within the crevice. These models depend on the presence the corrosion potential is higher than the repassivation potential,
of an active–passive transition and therefore do not fully describe and is very unlikely if the corrosion potential remains lower than
the situation for alloys that do not exhibit such a transition except the repassivation potential by a margin of 100 mV or more. The
under low pH conditions. In fact, for alloys that do not exhibit an repassivation potential is predicted as described below and compared
active–passive transition, the decrease in potential associated with the to a value of the corrosion potential calculated using mixed potential
ohmic potential drop consumes overpotential in the pit or crevice and theory37. Fig. 5 shows schematically how these quantities might
destabilizes attack. Other models focus on the change in chemistry in vary with various experimental parameters and where the regions of
corrosion susceptibility might lie34.
The repassivation potential is the threshold condition at which
pitting or crevice corrosion is stable and the effects of alloy
composition and solution chemistry on this potential must be
understood to predict the likelihood of localized corrosion. According to
the model33, the metal (M) undergoes dissolution underneath a layer of
concentrated metal halide solution MX as shown schematically in Fig.
6. This conceptualization of repassivation of a stable pit is consistent
with Raman spectroscopic studies of artificial pits on Ni39. In the
process of repassivation, a thin layer of oxide is assumed to form at the
interface between the metal and the hydrous metal halide. The model
assumes that, at any given instant, the oxide layer covers a certain
Fig. 4 Breakdown potential of 316L stainless steel measured as a function of fraction of the metal surface. This fraction increases as repassivation is
Cl– and NO3– concentration25. approached. The dissolution rate of the metal under the oxide is lower

OCTOBER 2008 | VOLUME 11 | NUMBER 10 41


REVIEW Understanding localized corrosion

(a) than at the metal–halide interface and corresponds to the passive


dissolution rate. Thus, as the repassivation potential is approached,
the dissolution rate tends toward the passive dissolution rate. The
effects of multiple aggressive and nonaggressive or inhibitory species
are taken into account through a competitive adsorption scheme.
The aggressive species form metal complexes, which dissolve in the
active state. In contrast, the inhibitory species and water contribute
to the formation of oxides, which induce passivity. In general, the
equations that describe these processes are complex and can only
be solved numerically. However, a closed-form equation (Eq. (1)) has
been found in the limit of repassivation, i.e. when the current density
reaches a predetermined low value irp (assumed to be irp = 10–2 A/m2)
(b)
and the fluxes of metal ions become small and comparable to those
for passive dissolution. For a system containing NA aggressive ions
and NI inhibitory ions, this equation is given in its most general form
by33:

NI ⎡⎛ i ⎞ l '' ⎤ ⎛ ξ k FErp ⎞ NA k j '' n j ⎛ α j FErp ⎞


1 + ∑ ⎢⎜ −1⎟ k ⎥ θ knk exp⎜⎜ ⎟ =∑ θ j exp⎜⎜ ⎟
rp
⎜ ⎟ ⎟ ⎟ (1)
k ⎢⎣⎝ i p ⎠ i rp ⎥
⎦ ⎝ RT ⎠ j i rp ⎝ RT ⎠

where ip is the passive current density, irp is the experimental current


density that defines repassivation (irp = 10–2 A/m2), θj is the partial
surface coverage fraction by solution species j, T is the temperature,
R is the gas constant, F is the Faraday constant, and kj’’, lk’’, nj, nk, αj,
(c) and ζk are electrochemical kinetic parameters. The summation on
the right-hand side of Eq. (1) is performed over all aggressive species
(j = 1, ..., NA) and the summation on the left-hand side pertains to
inhibitory species (k = 1, ..., NI). It should be noted that H2O molecules
are treated as inhibitory species because they contribute to the
formation of the oxide layer.
Eq. (1) has a number of adjustable parameters that need to be
determined from experimental data. These parameters are derived from
systems that contain only one or two electrochemically active anions
(e.g. Cl– and NO3–) and they are used to predict the behavior for more

(d)

Fig. 5 Schematic diagram of the effect of Cl– (a), temperature (b), inhibitors
(c), and oxidizing redox species (d) on the relative values of the repassivation Fig. 6 Schematic summary of the phases and interfaces considered in the
potential (Erp) and corrosion potential (Ecorr). The shaded areas denote the derivation of the model (M, metal; MX, metal halide; MO, metal oxide; the
ranges in which localized corrosion can be expected34. numbers indicate the interfaces between the phases)34.

42 OCTOBER 2008 | VOLUME 11 | NUMBER 10


Understanding localized corrosion REVIEW

Fig. 7 Calculated and experimental repassivation potentials for type 316L stainless steel in solutions containing Cl–, NO3– and CH3COO– as a function of NO3–
activity for systems with Cl– concentrations fixed at 0.42 M and CH3COO– concentrations equal to either 0 or 1.344 M33.

complex solutions, such as the Cl– + NO3– + CH3COO– mixture. This without inhibitory species. For example, simulation results for solutions
is illustrated in Fig. 7, where it is shown that a presence of CH3COO– containing only Cl– are compared to experimental data for a number
results in a decrease in the concentration of NO3– required to inhibit of commercial alloys in Fig. 832,40. The model was then used to predict
localized corrosion in a Cl– solution33. The effect of CH3 COO– on the localized corrosion behavior of several alloys in a chemical plant
the inhibition efficiency of NO3– was predicted purely by the model that operated with a proprietary brine chemistry41. In this case, the
without utilizing experimental data from the Cl– + NO3– + CH3COO– corrosion and repassivation potentials were modeled for three alloys in
mixture. Similarly, the effects of alloying elements Cr, Mo, W, and N, the process chemistry. The predicted values agreed with measurements
on the parameters are derived from experimental data of different performed in the laboratory using solutions obtained from the chemical
alloys in Cl– only solutions. plant41. The model was further used to predict maximum pit growth
This approach has been successfully used in modeling commercial rate. The current density predicted by the model as a function of
alloys in different solutions containing aggressive species with or potential is given by:

Fig. 8 Comparison of Erp values obtained from the generalized correlation with alloy composition with experimental data for various alloys at 368.15 K32.

OCTOBER 2008 | VOLUME 11 | NUMBER 10 43


REVIEW Understanding localized corrosion

⎛ α j FE ⎞ ⎛ ξ FE ⎞
∑ k ''θ exp ⎜⎜ ⎟⎟ + ∑ l i ''θ ini exp ⎜ i ⎟
nj
j j
j ⎝ RT ⎠ i ⎝ RT ⎠
i= (2)
1 ⎛ ξ FE ⎞
1 + ∑ l i ''θ ini exp ⎜ i ⎟
ip i ⎝ RT ⎠

Eq. (2) reduces to Eq. (1) for E = Erp and i = irp. Since Eq. (2) is a limiting
law, its accuracy gradually deteriorates as the potential increasingly
deviates from Erp. Eq. (2) cannot be regarded as a complete description
of the propagation rate of an actively growing pit because it does not
take into account factors such as the ohmic potential drop, diffusion
limitations, pit shape, etc. However, the current density predicted
using Eq. (2) for E > Erp is useful because it provides an estimate of the
maximum propagation rate of an isolated pit as a function of potential.
Fig. 9 Comparison of experimental apparent pitting rates with calculated
Such an upper estimate is particularly convenient because it relies only
maximum rates of pit propagation for corrosion potentials obtained from
on parameters that have been calibrated using repassivation potential short-term experiments. The vertical bars show the range of predicted pitting
data. The predicted propagation rate using Eq. (2) was then compared rates when the corrosion potential varies from –0.20 V to –0.14 V (vs. SCE) and
the temperature varies from 95 to 105 °C 41.
to electrochemical measurements conducted in the process piping41.
The results are depicted in Fig. 9. There is general agreement between strides toward advancing the understanding of localized corrosion.
the range of predicted and measured growth rates. However, there are many remaining challenges associated with the
understanding and prediction of the various phenomena. The ultimate
Concluding comments goal of such studies is to be able to predict the time, location, and
The complexities of localized corrosion require sophisticated extent of localized corrosion, and to understand fully the effects of
experimental and computational tools. Recent work has made microstructure and environment.

REFERENCES 23. Lin, L. F., et al., J. Electrochem. Soc. (1981) 128, 1194
1. Frankel, G. S., J. Electrochem. Soc. (1998) 145, 2186 24. Macdonald, D. D., J. Electrochem. Soc. (2006) 153, B213
2. Frankel, G. S., Corros. Sci. (1990) 30, 1203 25. Yang, S. F., and Macdonald, D. D., Electrochimica Acta (2007) 52, 1871
3. Maurice, V., et al., Electrochem. Solid-State Lett. (2001) 4, B1 26. Cho, K., and Pickering,H. W., J. Electrochem. Soc. (1990) 137, 3313
4. Strehblow, H. H., Mechanisms of pitting corrosion. In Corrosion Mechanisms in 27. Galvele, J. R., J. Electrochem. Soc. (1974) 123, 464
Theory and Practice, 2nd edn, Marcus, P. (ed.), Marcel Dekker: New York, (2002) 28. Oldfield, J. W., and Sutton, W.H. Br. Corros. J. (1978) 13, 104
5. Marcus, P., et al., Corros. Sci, (2008) in press 29. Oldfield, J. W., and Sutton, W. H., Br. Corros. J. (1980) 15, 31
6. Gordon, G. M., Corrosion (2002) 58, 811 30. Bocher, F., et al., J. Electrochem. Soc. (2008) 155, C256
7. Rebak, R. B., JOM (2008) 60, 40 31. Anderko, A., et al., Prediction of corrosion of nickel-base alloys and stainless
8. Subramanian, K. H., and Zapp, P. E., Pitting Corrosion in the Vapor Space and steels in oxidizing environments using thermodynamic and electrochemical
Liquid/Air Interface of High Level Radioactive Waste Tank, WSRC-MS-2003-00883, models, Paper 05053, Corrosion/2005. NACE International, 2005
Westinghouse Savannah River Company, 2003 32. Anderko, A., et al., Modeling localized corrosion in complex process environments
9. Frankel, G. S., J. ASTM Int. (2008) doi:10.1520/JAI101241 in the presence of inhibitors, Paper 06215, Corrosion/2006. NACE International,
2006
10. Frankel, G. S., et al., Corros. Sci. (2007) 49, 2021
33. Anderko, A., et al., Corros. Sci. (2004) 46, 1583
11. Stratmann, M., and Streckel, H., Corros. Sci. (1990) 30, 681
34. Anderko, A., et al., A computational approach to predicting the occurrence
12. Stratmann, M., and Streckel, H., Corros. Sci. (1990) 30, 697
of localized corrosion in multicomponent aqueous solutions, Paper 04061,
13. Stratmann, M., et al., Corros. Sci. (1990) 90, 715 Corrosion/2004. NACE International, 2004
14. Nishikata, A., et al., J. Electrochem. Soc., (1997) 144, 1244 35. Okada, T., J. Electrochem. Soc. (1984) 131, 241
15. Nishikata, A., et al., Corros. Sci. (1995) 37, 897 36. Okada, T., J. Electrochem. Soc. (1984) 131, 1026
16. Nishikata, A., et al., Corros. Sci. (1995) 37, 2059 37. Anderko, A., et al., Corrosion (2001) 57, 202
17. Tsutsumi, Y., et al., J. Electrochem. Soc. (2005) 152, B358 38. Anderko, A., et al., Fluid Phase Equlibria (2002) 123, 194
18. Vera Cruz, R. P., et al., Corros. Sci. (1996) 38, 1397 39. Sridhar, N., and Dunn, D. S., J. Electrochem. Soc. (1997) 144, 4243
19. Vera Cruz, R. P., et al., Corros. Sci. (1998) 40, 125 40. Anderko, A., et al., A general model for the repassivation potential as a function
20. Tada, E., and Frankel, G. S., J. Electrochem. Soc. (2007) 154, C312 of multiple aqueous species. 2. Effect of oxyanions on localized corrosion of Fe-
Ni-Cr-Mo-W-N alloys, Paper 08272, Corrosion/2008. NACE International, 2008
21. Tada, E., and G. S. Frankel, J. Electrochem. Soc. (2007) 154, C318
41. Anderko, A., et al., Corros. Eng., Sci. Technol. (2005) 40, 33
22. Agarwal, A. S., et al., J. Electrochem. Soc. (2008) submitted

44 OCTOBER 2008 | VOLUME 11 | NUMBER 10

Das könnte Ihnen auch gefallen