Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Available online at www.sciencedirect.

com

ScienceDirect
Journal of the European Ceramic Society 34 (2014) 229–235

Electrochemical corrosion of silicon carbide ceramics in H2SO4


Mathias Herrmann a,∗ , Kerstin Sempf a , Michael Schneider a , Uwe Sydow a , Kerstin Kremmer a ,
Alexander Michaelis a,b
a Fraunhofer Institute for Ceramic Technologies and Systems, Winterbergstraße 28, 01277 Dresden, Germany
b Dresden University of Technology, Institute of Materials Science, 01062 Dresden, Germany

Received 22 June 2013; received in revised form 16 August 2013; accepted 20 August 2013
Available online 20 September 2013

Abstract
Sintered silicon carbide materials have found widespread use due to their high corrosion stability. This corrosion stability can be affected by
electrochemical processes. Electrochemical corrosion experiments conducted on a SSiC material in H2 SO4 at different voltages and subsequent
detailed investigation of the formed surfaces was carried out. The first time a systematic local measurement of the thickness of the oxide layers
was carried out. The measurements revealed the formation of SiO2 surface layers with thickness up to 125 ␮m. The measured values also showed
a strong deviation from grain to grain. The thickness of the layers does not correlate with the crystallographic orientation of the grains or the
SiC-polytypes. The data indicate that the behaviour is caused by the variation of the resistivity of the grain boundaries. The measured thicknesses
as a function of the electrical charge transferred indicate that the electrochemical oxidation results in the SiO2 and carbon dioxide.
© 2013 Elsevier Ltd. All rights reserved.

Keywords: Silicon carbide; Corrosion; Microstructure

1. Introduction the SSiC materials is the reason why they are used in seals and
other components in the chemical industry.1
Silicon carbide materials, and solid state-sintered silicon car- Silicon-infiltrated SiC (SiSiC) and liquid phase-sintered SiC
bide materials (SSiC) in particular, have found widespread use as (LPSSiC) are also dense SiC materials. SiSiC contains an
seals, bearings and valves in a variety of media in industrial wear additional free silicon phase, LPSSiC additional oxide grain
applications 1 (Literature source 1). Solid state-sintered silicon boundaries. In comparison with SSiC, both materials show a
carbide ceramics (SSiC) are usually sintered with small amounts lower chemical resistance due to the lower stability of the grain
of B, Al compounds and C. The sintering additives Al and B can boundary phases.1
be incorporated into the SiC lattice. Therefore SSiC normally SiC itself is an intrinsic semiconductor and can have specific
only contains isolated B4 C grains and a small volume fraction electrical resistivities ranging from  m to M m, depending on
of isolated pockets of carbon precipitations. The SiC phase of the additives and preparation method used. 1 (literature source
the ceramic itself consists of hexagonal or rhombohedral poly- 1) Commonly used sintering additives such as boron and alu-
types, which have the same structural units, but differ in terms of minium enhance the conductivity of the material and hence
the sequence of layers along the c-axis of the lattice.1 The SSiC electrochemical corrosion can be a cause of severe material
materials also exhibit excellent corrosion resistance in a wide degradation. Damage patterns for SiC seals strongly suggest-
range of media and conditions.1–4 Unlike other Si-based ceram- ing electrochemical corrosion as the primary cause for material
ics, SSiC shows a high stability even in HF solutions.1–4 Besides degradation have been reported.5 Nevertheless, the corrosion
the good tribological properties, the high corrosion resistance of rates of ceramics are generally very low compared with those of
metals.6–10 Electrochemical corrosion in acid involves the fol-
lowing half-reactions9 (the standard potentials are given relative
∗ Corresponding author at: Fraunhofer-Institut für Keramische Technologien to the normal hydrogen electrode (NHE)):
und Systeme, IKTS Dresden, Sintern, Charakterisierung/Sintering, Characteri-
sation, Winterbergstraße 28, 01277 Dresden, Germany. SiO2 + C + 4H+ + 4e−  SiC + 2H2 O
Tel.: +49 0351 25537527; fax: +49 0351 2554122.
E-mail address: Mathias.Herrmann@ikts.fraunhofer.de (M. Herrmann). E◦ = −0.673 V (1)

0955-2219/$ – see front matter © 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.jeurceramsoc.2013.08.024
230 M. Herrmann et al. / Journal of the European Ceramic Society 34 (2014) 229–235

SiO2 + 6H+ + CO + 6e−  SiC + 3H2 O


E◦ = −0.276 V (2)

SiO2 + 8H+ + CO2 + 8e−  SiC + 4H2 O


E◦ = −0.222 V (3)

H+ + 2e−  H2 E◦ = 0.000 V (4)


O2 + 4H+ + 4e−  2H2 O E◦ = 1.229 V (5)
It is still not completely clear whether during corrosion CO2 ,
CO or carbon is produced according to reactions (1)–(3). The
data in the literature8,9 suggest that at least up to potentials of
1–1.5 V no substantial carbon oxidation takes place. From these
reactions it can be assumed that in acidic media the forma-
tion of a surface film of SiO2 passivates the system, whereas
in alkaline media the preferential formation of soluble silicates
prevents the passivation due to the instability of silicon diox-
ide at higher pH values.6,8–11 The electrochemical corrosion
strongly depends on the resistivity of the SiC material. Mate-
rials with high resistivity show negligible corrosion rates over a
wide range of voltages, whereas materials with low resistivity
undergo pronounced corrosion.9
Although some features of the electrochemical corrosion of
SiC in acids such as formation of a passivating SiO2 layer are
established, no detailed information about the mechanisms and
the correlation between microstructure and corrosion resistivity
is available.
In particular, no detailed data regarding the corrosion rate
of the grains as a function of crystallographic orientation and
SiC polytypes can be found in the literature. The aim of this
investigation was to study the electrochemical corrosion of SSiC
materials in H2 SO4 by electrochemical methods and detailed
microstructural investigations using SEM and EBSD.

2. Materials and methods

2.1. Material and microstructural analysis

A dense commercially available solid state-sintered silicon


carbide material (EKASIC D, ESK, Germany) consisting mainly
Fig. 1. SEM micrographs of the SSiC material in the initial state and after
of ␣-SiC with small carbon precipitates was used in this study.
corrosion at 2 V for 3600 s (a) and after removal of the oxide layer with HF (b).
The main polytypes of SiC were 4H, 6H and 15R. The material
also contained a small amount of aluminium as the main sin-
tering additive which was incorporated into the SiC grains. The determined using the Oxford Instruments thin film measurement
specific electrical conductivity was determined to be 2.5  m. tool11,12 allowing layers between 1 nm and 1 ␮m in thickness
The silicon carbide samples were embedded in epoxy resin, to be measured. For SiO2 layers with relative low electron
ground and polished with diamond suspensions then cleaned density the lowest thickness which could be detected is approx-
with ethanol and dried. The microstructure of the material is imately 3–5 nm. Details are described elsewhere.11 An electron
given in Fig. 1a. back-scattered diffraction (EBSD) system (Channel 5, Oxford
The microstructures and compositions of the samples were Instruments) was used to investigate the grain orientations.
analysed using an field emission scanning electron micro-
scope (NVISION, Zeiss, Germany). Energy-dispersive X-ray 2.2. Electrochemical measurements
spectroscopy (EDX) was used to determine the chemical com-
position of the microstructural constituents (INCA, Oxford The electrochemical experiments were carried out at room
Instruments). The thickness of the formed corrosion layers was temperature in a common electrochemical three-electrode cell
M. Herrmann et al. / Journal of the European Ceramic Society 34 (2014) 229–235 231

75

-2
i / µA cm
O2
50

25 SiO2-formation

0
O2- reduction
-25
H2
-50

-75
-1.0 -0.5 0.0 0.5 1.0 1.5
ESCE / V

Fig. 3. Cyclic voltammogram of aerated SSiC on 0.5 M H2 SO4


(dE/dt = 10 mV/s) without HF pre-treatment.

Fig. 2. Cross section of the top-frame cell and embedded sample. The cell was
pressed together in the region of the sample with threaded rods and nuts (not
shown). RE, reference electrode; CE, counter electrode (platinum); WE, working
electrode. The bottom side of the SiC sample was attached to a metallic holder
which was contacted sideways through the epoxy resin.

housed in a Faraday cage. Potentiodynamic and chronoam-


perometric measurements were performed with a computer-
controlled potentiostat Autolab PGSTAT 30 (Metrohm). A
saturated calomel electrode (Sensortechnik Meinsberg GmbH)
was used as the reference electrode (RE) and a platinum sheet
as the counter electrode (CE). A sketch of the electrode setup
is shown in Fig. 2. Details were already described by Sydow
et al.9 The electrolyte was aerated 0.5 M H2 SO4 prepared from
analytical grade chemicals. The sample surface was pre-treated
by HF cleaning to remove the surface layer.

3. Results

Fig. 3 shows a cyclic voltammogram of SSiC in sulphuric


acid. The graph was limited due to the hydrogen evolution on
the cathode side (Eq. (4)) and due to the oxygen evolution on
the anode side (Eq. (5)). The two reactions characterise the elec-
trochemical window of the aqueous electrolyte. On the anode
side formation of a silicon dioxide film could be observed, as
illustrated by the increase in the current density at approximately
E = 0.5 V according to Eqs. (1)–(3). Prior to the hydrogen evolu-
tion a plateau in the cathodic sweep occurred due to the reduction
of dissolved oxygen.8–10
Fig. 4a shows the change in open circuit potential (OCP)
over time in H2 SO4 . The OCP shifted from E ≈ −0.2 V to
E ≈ 0.1–0.2 V within the first few hours of exposure and then
stabilised at E ≈ 0.2–0.3 V. A number of fluctuations indicated Fig. 4. Open circuit potential of SSiC as a function of time in 0.5 M H2 SO4 (a),
SEM micrograph of the surface (b) and EDX spectra of the surface (c).
local breakdown and repassivation of the oxide film. SEM
inspection of the surface yielded no evidence of corrosive attack.
232 M. Herrmann et al. / Journal of the European Ceramic Society 34 (2014) 229–235

200
lg(i / A cm )
-2 -2 a)
900mV

-300mV 150
-3

Charge, mC/cm2
100
-4
d lgi / d lgt ~ -1

50
-5

0
0 1000 2000 3000 4000 5000
-6
Time , sec
-1 0 1 2 3
lg(t / s) b) 150 Mean Minimal Maximal

125
Fig. 5. Change in current with time in 0.5 M H2 SO4 after switching of the

Thickness of oxide layer, nm


voltage from −300 mV to 900 mV. 100

75

EDX measurement likewise revealed no significant oxide layer 50


thickness (Fig. 4b and c). The thickness could not be determined
25
using the thin film tool either. These results suggested that the
thickness of the layer was less than 3–5 nm. Therefore it can 0
0 1000 2000 3000 4000 5000
be concluded that even thin SiO2 layers < 3–5 nm thick can Time, sec
strongly change the electrochemical behaviour of SiC materials. c) 70
Fig. 5 shows a double logarithmic plot of the current
60
Calculated thickness of the oxide layer, nm

response after a potential step experiment with a step from


E = −0.3 V to E = 0.9 V, in the range of anodisation, but below 50
the potential where extensive decomposition of water takes
40
place. The data show a slope of the current density versus time
d(log(i))/d(log(t)) = −1 indicating passivation of the surface by 30
formation of the SiO2 layer according to the high field mecha-
nism (Curie-von Schweidler Law).13 20

For the purposes of measuring the oxide layer thickness as 10


a function of corrosion time electrochemical polarisation was
carried out at 1.6 V for different amounts of time. The thickness 0
0 10 20 30 40 50 60
of the corrosion layer was measured intermittently with the thin Mean thickness of the oxide layer, nm
film tool on at least 20 grains. Fig. 6a shows the cumulative
Fig. 6. Time dependence of the Charge transferred during the SiC interface at
charge during anodic polarisation. The cumulative charge could
1.6 V (a), dependence of the thickness of the oxide layer on the time (b) and
be converted to a SiO2 film thickness based on the following correlation of the mean thickness of the formed oxide layer with the calculated
assumptions: thickness based on the accumulated charge an Eq. (3) (c).

• formation of the oxide film takes place according to Eq. (3),


• the current efficiency κ is 100%. images showed that the corrosion was less pronounced in the
grain boundaries than in the grains (Fig. 1b and c).
The results given in Fig. 6 reveal that the overall thickness of The influence of the SiC polytypes and crystallographic
the layer was proportional to the corrosion time and the accumu- orientations of the grains on the corrosion behaviour was investi-
lated charge. SiO2 layer thicknesses of greater than 100 nm were gated through an additional experiment. An area on the polished
observed (Fig. 6b). This thickness was much higher than typical cross section was marked with hardness indentations and cor-
thicknesses on valve metals formed by anodisation at 1.6 V.13 roded for 2400 s at 1.6 V. The thickness of the oxide layer was
Fig. 6c shows the correlation between the measured and cal- determined using the thin film tool and then the SiO2 layer was
culated mean thickness of the oxide layer formed. The slop of removed by HF treatment. Afterwards the grain orientations and
the regression line is 1.08 ± 0.5 (R2 = 0.995). This is a strong polytypes of the SiC grains were determined by EBSD. The
evidence that the electrochemical oxidation takes place accord- results, given in Fig. 7, did not allow for a simple correlation
ing reaction (3) at these conditions, i.e. the reaction products are between the grain orientation, polytype and corrosion rate to be
CO2 and SiO2 . The data also reflect the strong variation in oxide identified. EDX results for Al content also yielded no correlation
layer thickness for different SiC grains. Moreover the FESEM between the Al content and the degree (or extent) of corrosion.
M. Herrmann et al. / Journal of the European Ceramic Society 34 (2014) 229–235 233

6H 4H 15R
70

60
Thickness of the oxide layer, nm

50

40

30

20

10

0
20 30 40 50 60 70 80 90 100
Angle to c direcon of the polytypes

Fig. 7. Thickness of the oxide layer after corrosion at 1.6 V (2400 s) as a function
of the angle between the direction normal to the grain surface and the [0 0 1]
direction.

4. Discussion

The overall manifestation of the corrosion process observed


in the experiments carried out in this study confirmed pre-
viously obtained results.6–10 A passivating SiO2 layer was
formed in acid. This study represented the first systematic
measurement of the oxide layer thickness under different cor-
rosive conditions. A general correlation between the SiO2
layer thickness and the charge transferred per unit area of
the interface was observed (Fig. 6b). The FESEM images
Fig. 8. Schematic diagram of the mechanism of electrochemical corrosion of
indicated that in the majority of the grain boundaries SiO2
SSiC and LPSSiC in acids (The thickness of the grain boundary lines represent
formation was less pronounced or even completely hindered the resistivity of the grain boundaries (Rgb1  Rgb2 ) The thickness of the current
(Fig. 1b). Reduced electrochemical attack of the grain bound- lines indicate the current density).
aries was also found for the SSiC material corroded in
NaOH.14
Strong differences in the thickness of the SiO2 layers were of impurities, etc.) the current through the grain and hence the
observed. For some grains layer thicknesses of 5 nm were found, extent of corrosion will be different for the different grains,
whereas other grains exhibited SiO2 layers up to 100 nm under irrespective of polytype and orientation. The grains for which
the same conditions. No correlation with the crystallographic the grain boundary resistivity is low will exhibit a high cur-
orientation, the type of polytype or the Al content of the grains rent density and therefore form thicker SiO2 layers. This is
was observed. schematically shown in Fig. 8. The formation of SiO2 layers
The lack of dependence of the corrosion on the polytype and of differing thickness is also an indication that the formed layer
the crystallographic orientation is at first glance surprising, but is not completely insulating as otherwise the redistribution of
can be explained taking into account the observed electrochem- the current would result in a much more homogeneous thick-
ical stability of the grain boundaries. The electrical resistivity ness.
is higher in the grain boundaries than in the highly Al-doped The lack of dependence on the Al content in the grains
grains. Different investigations of the conductivity of SiC mate- can be rationalised by the fact that the overall Al concen-
rials also indicate the higher electrical resistivity of the grain tration was about 0.59 wt%, far above the concentration at
boundaries in comparison to the grains15–18 or have even demon- which the Al doping will strongly change the conductivity of
strated lower carrier densities at the grain boundaries.19 Given the grains.20 Therefore the distribution of the grain bound-
this and the fact that the highest current density corresponds to ary resistivities masked this effect. This is different to the
the path with the lowest resistivity, it can be concluded that behaviour observed in LPSSiC materials where core–rim struc-
the charge transfer through a grain is determined or at least tures were observed.9,10,21–23 The cores were free of Al and
strongly influenced by the properties of the grain boundaries therefore had a high resistivity, whereas the rims showed Al
separating the grain in question from the neighbouring grains. concentrations in the range of 0.2–0.3 wt%, resulting in a high
The grain boundaries two or three grains beneath the surface can conductivity.10,11,22 Hence the rims are more strongly corroded
be neglected because they are averaged due to the different paths in comparison to the core areas. This allowed for visualisation
available. Assuming that there is a distribution of grain bound- of the core–rim structure by electrochemical etching as shown
ary resistivities (depending on grain orientation, segregation in Fig. 9.
234 M. Herrmann et al. / Journal of the European Ceramic Society 34 (2014) 229–235

were attacked to a lesser extent than the grains. The behaviour


can hence be explained if the grain boundary resistivity is
assumed to be the limiting factor for the corrosion rate. Sim-
ilar behaviour was observed for electrochemical corrosion in
NaOH.14

Acknowledgement

The authors gratefully acknowledge financial support from


the German Research Foundation (DFG Mi 509/6-2; He
2457/13-2, Kl 615/8-2).

References

1. Schmalzried C, Schwetz KH. Silicon carbide and boron carbide based hard
Fig. 9. FESEM micrograph of the core–rim structure of an LPSSiC material materials. In: Chen RC, editor. Ceramic Science and Technology. Wiley
after electrochemical etching in NaOH (for details see Sydow et al.10 and Sempf VCH; 2010. p. 131–225.
et al.11 ) (nonconductive oxide phases YAG-Y3 Al5 O12 , ␣-Al2 O3 ). 2. Schwetz KA, Hassler J. Zur Beständigkeit von Hochleistungskeramiken
gegen Flüssigkorrosion. cfi/Ber DKG 2002;79(10):D15–8.
3. Schwetz KA, Hassler J. Zur Beständigkeit von Hochleistungskeramiken
The proposed model does not mean that the corrosion is
gegen Flüssigkorrosion. cfi/Ber DKG 79 2002;11:D14–9.
independent of the conductivities (resistivities) of the grains 4. Herrmann M, Schilm J, Michael G. Corrosion behaviour of different tech-
themselves. The distribution of the resistivity grains is super- nical ceramics in acids, basic solutions and under hydrothermal condition.
imposed on the distribution of the grain boundary resistivities; cfi 2003;80(4):E27–34.
i.e. grains having the same grain boundary resistivity will cor- 5. Meschke F, Kailer A. Electrically driven cold water corrosion of SiC. cfi
2004;81:E19–22.
rode according to their conductivity (depending on dopant,
6. Cook SG, Little JA, King JE. Corrosion of silicon carbide ceramics
polytype and orientation). An indication for this is the cor- using conventional and electrochemical methods. Br Corros J 1994;29:
rosion behaviour observed in the LPSSiC materials (Fig. 9). 183.
The highly conductive rims underwent more extensive corrosion 7. Divakar R, Seshadri SG, Srinivasan M. Electrochemical techniques for
than did the non-conducting cores, as is shown schematically in corrosion rate determination in ceramics. J Am Ceram Soc 1989;72:
780–4.
Fig. 8b. Detailed investigation of the structure of the SiC/SiO2
8. Andrews A, Herrmann M, Sephton M, Machio C, Michaelis A. Elec-
interface also pointed to the formation of etching grooves for trochemical corrosion of solid and liquid phase sintered silicon carbide
certain crystallographic orientations,8,24 indicating that local in acidic and alkaline environments. J Eur Ceram Soc 2007;27:
electrochemical attack was dependent upon the crystallographic 2127–35.
orientation, but did not appear macroscopically due to the strong 9. Sydow U, Schneider M, Herrmann M, Kleebe H-J, Michaelis A.
Electrochemical corrosion of silicon carbide ceramics: part 1. Electro-
influence of the grain boundary resistivity on the corrosion
chemical investigation of sintered silicon carbide (SSiC). Mater Corros
rate. 2010;61:657–64.
10. Sydow U, Sempf K, Herrmann M, Schneider M, Kleebe H-J, Michaelis A.
Electrochemical corrosion of liquid-phase sintered silicon carbide. Mater
5. Conclusions
Corros 2011;62:9999.
11. Sempf S, Herrmann M, Sydow U. New ways of revealing the microstructures
Electrochemical corrosion experiments conducted on a SSiC of SiC materials. Prakt Metallographie 2012;49:64–74.
material in H2 SO4 at different voltages and subsequent detailed 12. Conrad T, Arstilla K, Hantschel T, Franquet A, Vanderforst W, Bauer F.
investigation of the formed surfaces revealed a strong influence Composition quantification of microelectronics multilayer thin films by
EDX: toward small scale analysis. Mater Res Soc 2013. IDS Number
of the formation of the SiO2 surface layer on the electrochemical
BND 16.
behaviour. 13. Lohrengel MM. Thin anodic oxide layers on aluminium and other valve
The shift of the open circuit potential (OCP) from −0.2 V metals. Mater Sci Eng R 1993;R11:243.
to 0.2–0.3 V versus time is caused by the formation of a native 14. Herrmann M, Sempf K, Wendrock H, Kremmer K, Schneider M, Sydow U,
oxide layer with thicknesses of less than 3–5 nm. The mean et al. Electrochemical corrosion of silicon carbide ceramics in NaOH. J Eur
Ceram Soc 2013 (submitted for publication).
thickness of the SiO2 surface layer formed by anodisation was
15. Kleebe H-J, Siegelin F. Schottky barrier formation in liquid-phase sintered
proportional to the charge transferred through the interface. silicon carbide. Z Metallkd 2003;94(3):211–7.
The measured and the calculated mean thickness of the oxide 16. Siegelin F, Kleebe H-J, Sigl S. Interface characteristics affecting
layer are in good agreement if CO2 and SiO2 are assumed electrical properties of Y-doped SiC. J Mater Res 2003;18(11):
as electrochemical oxidation products of SiC (Eq. (3)). How- 2608–17.
17. Schroeder A, Pelster R, Grunow V. Charge transport in silicon carbide:
ever, the formed layer was very inhomogeneous and did not
atomic and microscopic effects. J Appl Phys 1996;80:2260–8.
correlate with either the grain orientations or the SiC poly- 18. Sauti G, Can A, McLachlan DS, Herrmann M. The AC conductiv-
types. Thickness differed in places by a factor of 10. The ity of liquid-phase-sintered silicon carbide. J Am Ceram Soc 2007;90:
experiments also revealed that most of the grain boundaries 2446–53.
M. Herrmann et al. / Journal of the European Ceramic Society 34 (2014) 229–235 235

19. Kobayashi R. Evaluation of grain-boundary conduction of dense AlN–SiC 22. Ihle J, Martin HP, Herrmann M, Obenaus P, Adler J, Hermel W,
solid solution by scanning nonlinear dielectric microscopy. J Am Ceram Soc et al. The influence of porosity on the electrical properties of liq-
2010;93:4026–9. uid phase sintered silicon carbide. Int J Mater Res 2006;97(5):
20. Müller R, Künecke U, Weingärtner R, Maier M, Wellmann P. Anomalous 649–56.
charge carrier transport phenomena in highly aluminum doped SiC. Phys 23. Hu J, Gu H, Chen Z, Tan S, Jiang D, Rühle M. Core-shell structure from the
Stat Sol 2006;3:554–7. solution–reprecipitation process in hot-pressed AlN-doped SiC-ceramics.
21. Kleebe H-J, Sigl L-S. Core/rim structure of liquid-phase-sintered silicon Acta Mater 2007;55:5666–73.
carbide. J Am Ceram Soc 1993;76(3):773–6. 24. Kleebe, Kleebe H-J, in preparation.

You might also like