On The Mechanism of High Temperature Corrosion: December 2018

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/329488067

On the Mechanism of High Temperature Corrosion

Preprint · December 2018

CITATIONS READS
0 3,081

1 author:

Omar J. Yepez
Voxel innovations
52 PUBLICATIONS   180 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Methanol Ethanol Fuel Cells View project

Interfacial electrochemical reaction mechanisms View project

All content following this page was uploaded by Omar J. Yepez on 07 December 2018.

The user has requested enhancement of the downloaded file.


On the Mechanism of High Temperature Corrosion
Omar Yépez
Clariant Oil Services
2750 Technology Forest Blvd.
The Woodlands, Texas 77381

ABSTRACT

High temperature corrosion is defined as the degradation of the metallic material at


temperatures higher than 400 °C (750 °F) and at atmospheric pressures. In these conditions there
is no presence of water/electrolyte and corrosion occurs through direct chemical reaction
between the metallic material and different chemicals. Normally, these chemicals are part of the
environment in which the metallic material is being used. Examples of these corrosion processes
are: carburization, chlorination, nitration, oxidation and sulfidation. Occluded hydrogen attack
could be a consequence of any of these processes. Also, coke may build up and produce or help
to produce under deposit corrosion. This kind of corrosion is normally found in boilers, furnaces,
gas turbines and diesel engines. Therefore, it is of a wide interest for industrial applications in
general. In this chapter our understanding of these different corrosion mechanism will be
reviewed and updated.

Key words: high temperature, corrosion mechanism, material degradation.


1 Introduction

A metal in contact with a hot gas, typically at temperatures above 400 °C, in absence of liquid
water phase, can suffer corrosion, also called hot corrosion. While aqueous (wet) corrosion
processes are of electrochemical nature, hot corrosion is a chemical process, i.e., governed by
chemical kinetics at the interface metal/gas phase. Nevertheless, the oxide layer that forms at
the metal surface is influenced by ionic diffusion and electronic conductivity within the oxide, as
typical of an electrochemical mechanism. At elevated temperatures, the main chemical
interactions that can destroy an already formed protective scale are: 1) carburization, 2)
chlorination, 3) nitridation, 4) oxidation and sulfurization, 5) ashes, deposits and molten salts and
6) hydrogen embrittlement and decarburization.

These processes destroy the protective scale by different mechanisms. These mechanisms can
result in: thinning the scale, reacting and producing eutectics, inducing its metal segregation,
dissolving it in molten salts or inducing its evaporation. In the special case of hydrogen, destroying
the internal metal strength through embrittlement or decarburization. In this chapter, the
degradation mechanisms of these main hot corrosion phenomena will be presented.

2 Carburization

Corrosion of metals and alloys in industrial environments with high carbon activity, the so-called
metal dusting, which is carburization of the metal, has been known for 60 years. This
phenomenon occurs mainly in thermo-chemical factories, metallurgical, petrochemical and
refining industries. This means anywhere the metal is exposed to a carbon rich atmosphere and
at high temperatures. This is why a common feature of this metal degradation is a temperature
in the range of 450 - 800 °C. The first data on damage related to metal dusting was published by
Camp in 1945 [1]. A superheater feedstock unit was damaged in a refinery naphtha reforming
installation. In 1950, Burns made a series of case studies of corrosion of metal alloys under
different environments in refinery plants. He found that in the case of processing low sulfur crude
oils, significant corrosive losses on oil distillation units may occur at temperatures about 400°C
[2].

For example, carburization is a common problem in the production of synthesis gas by means of
reforming process, where natural gas is replaced by a mixture of CO, H2, CO2 and H2O. To
increase the production efficiency, carbon monoxide is increased in the mixture. This
modification increases the carbon activity in the environment to which the metal is exposed. In
turns, this lead to metal dusting. The mechanism of carburization goes in several stages. First,
elemental carbon deposit on the iron surface and supersaturates the ferrite phase (α-Fe). In the
next step, cementite (Fe3C) forms at the surface. Cementite formation occurs with an increase
in volume, which results in the generation of crystalline defects. More carbon diffuses through
the cementite phase, increasing the number of crystal defects. This eventually cause cementite
particles to separate into small particles. This ware out the cementite phase producing
metallurgical damage. This model explains metal dusting without decomposition of Fe3C [3].

A more comprehensive model of metal dusting, which was proven by thermogravimetric studies,
optical and electron microscopy on metallographic cross sections and high-resolution
transmission electron microscopy, was presented by Grabke [4]. In the temperature range 400-
650 °C in H2-CO-H2O mixture, the main reaction for carbon transfer from the atmosphere is CO
reduction:

CO  H 2  H 2 O  C (1)

The following reaction sequence occurs on iron and low alloy steels:

a. Carbon is transferred by reaction (1) into solid solution, up to oversaturation, concerning


equilibrium with cementite.
b. Cementite nucleates and grows, mainly at the surface but also at grain boundaries. A
rather irregular layer of cementite crystal grows with characteristic protrusions into the
bulk metal phase, Fig. 1 a. Due to the low carbon diffusivity in cementite, it is a barrier
against further carbon ingress, and therefore the carbon activity at the surface rises.
c. Locally, graphite nucleates. This means a decrease of carbon activity and thus cementite
becomes instable, Fig. 1 b.
d. Cementite decomposes to graphite and iron, according to Fe3C  3Fe  C , the graphite
growth into the cementite. Carbon atoms from Fe3C attach to graphite planes, growing
more or less vertically into the cementite. It is observed graphite roots invading the
cementite. The iron diffuses outward through the graphite, concentrations of 3-4% Fe
have been detected in the graphite. The iron atoms agglomerate under formation of fine
particles, of about 20 nm average diameter. This is the metal dust. Into these particles
carbon is transferred, again by reaction (1) from the atmosphere Fig. 1 c.

The carbon diffuses through the particle to some site where graphite nucleation is easy. There
graphite growth starts, often growth of graphitic carbon filaments is observed, Fig. 1 b. The rate
of carbon transfer into the particles determines the growth of the filaments and thus of coke
growth.
This mechanism can be retarded and altered by the presence of sulfur. Adsorption of sulfur on
the metal surface hinder the transfer of carbon into solid solution, i.e. step (a), and also the
carbon deposition is retarded. But more important, the nucleation of graphite is impeded and
reaction step (c) may not take place at all or is delayed for a long time. Therefore, suppression of
metal dusting depends on carbon and sulfur activity in the atmosphere, carburizing gas CO or
CH4, temperature and time.

At higher temperatures > 700 °C the morphology of the reaction products changes. The iron from
the Fe3C decomposition does not form fine particles but agglomerates to an iron layer through
which the carbon must diffuse, to attach to the outer graphite layer, Fig. 1 d. Thereby the
processes are slowed down and metal dusting becomes controlled by carbon diffusion in ferrite,
or at higher temperature in austenite. In the system: outer graphite layer, metal layer, cementite
layer, metal phase, the phase boundary graphite/metal is instable, since the diffusion control
causes that graphite intrusions into the metal layer grow faster, until that layer is disrupted and
perturbed.

Figure 1. Schematic illustration of the processes in metal dusting of iron. a) After carbon transfer
from the gas phase and oversaturation of the metal phase, nucleation and growth of cementite
Fe3C occurs b) Nucleation and growth of graphite into the cementite phase. Carbon filaments
grown behind particles detached from the metal phase by the graphite growth c) At
temperatures < 600 °C, inward growth of Fe3C which disintegrate outward under coke formation
d) At temperatures > 700 °C, formation of an iron layer between cementite phase and coke,
carbon diffusion through this layer and final disappearance of Fe3C.
At even higher temperatures, 900 and 1000 °C in CH4-H2 mixtures, no cementite formation
occurred, but from the oversaturated metal phase, carbon diffuses through an austenite layer to
the graphite growing on the surface. Again the strong growth of graphite protrusions indicates
carbon diffusion control and instable phase boundaries. After longer duration also metal
disintegration is to be expected which means that metal dusting occurs also at 1000 °C, but not
via cementite formation.

3 Chlorination

Chloride-induced high-temperature corrosion results in the degradation of materials used in


number of applications, such as municipal waste incinerators, power plants firing chloride
containing coal or biomass, and in some minor but important application such as automotive
exhaust systems. In energy producing combustion process, corrosion of the water walls and
superheater tubes by chloride-rich compounds limits the metal surface temperature and thus
the efficiency in electricity production. Table 1 shows the melting points of metal chlorides for
the most common alloying elements,

Table 1 Melting points of solid metal chlorides

Chloride Tm, °C
CrCl2 820
CrCl3 1150
FeCl2 676
FeCl3 303
NiCl2 1030

The relatively low temperature presented by iron chlorides is what makes chromium to be added
in most applications involving chloride atmospheres [5].

HCl and chlorides have a big impact on the high temperature oxidation of iron and steels. The
most striking effect is the rapid response of the corrosion phenomena on the introduction of the
chlorine containing contamination into the environment. As soon as sodium chloride is
introduced, as a vapor into the oxidizing atmosphere or as a grain on to the oxide scale of a steel,
the oxidation is strongly accelerated. The main effect is considerable damage to the oxide scale.
After introduction of the chlorine containing contamination, the oxide scale is not adherent and
protective, but very loose, cracked and porous. Since in this state there is no passivation by a
protective oxide layer, the oxidation in the presence of chlorine containing contaminants has
been named active oxidation, i.e. the oxidation of steel accelerated by the presence of chlorine
[6]. In Figure 2 the corrosion of 2.25 Cr-1 Mo is shown compared to simple oxidation. At 500 °C,
the corrosion markedly increases after addition of 500 ppm of HCl compared to oxidation.

Figure 2 Corrosion of 2.25 Cr-1 Mo at 500 °C after addition of 500 ppm HCl compared to
oxidation a) He-0.05 bar O2; b) 500 ppm HCl [6].

The mechanism of chlorination involve several reactions. Given that high chromium alloys are
normally used in processes where chlorination may occur. The following reactions illustrate the
complex mechanism by which chlorine attack chromium oxide,

a) Formation of chloride gas or hydrochloric acid

5
Cr2O3 ( s)  4 KCl( s )  O2  2 K 2CrO4 ( s )  2Cl2 (2)
2

In the case of water vapor in the atmosphere, this reaction slightly varies because of the
formation of hydrochloric acid,

2Cr2O3 ( s)  8KCl( s)  4 H 2O  3O2  4 K 2CrO4 ( s)  8HCl (3)


b) Penetration of the gaseous chlorine and chlorides

This occurs through the oxide film. Reaction with the chromium or its carbides at the oxide/metal
interface will form volatile metal chlorides,

Cr ( s )  Cl 2  CrCl 2 ( s ) (4)

Cr3C2 ( s)  Cl2  3CrCl2 ( s)  2C (s) (5)

c) Diffusion of the metal chlorides towards the surface

CrCl 2 ( s )  CrCl 2 ( g ) (6)

d) Reaction with the oxygen in the atmosphere to form metal oxides,

3
2CrCl2 ( g )  O2  Cr2O3 ( s )  2Cl2 (7)
2

The released chlorine is dispersed in the atmosphere but a portion of it diffuses back to the
metal/oxide and fed reactions (4) and (5) re-activating them. This makes this process self-
sustained, autocatalytic [7].

4 Nitridation

Corrosion of chromium becomes more complex when more than one oxidant is present in the
reacting gas. The chemical reaction between chromium and gas mixture components is
determined by a series of factors, among which the most important include compactness of the
scale and its permeability to gas species, the rate of individual compound formation and their
properties. An often investigated example of a multi-component corroding gas system is air.
Scales formed during oxidation of chromium in air consist of an inner layer containing the nitride
Cr2N and an outer layer of Cr2O7. Porosity has occasionally been observed in the outer scale, a
factor sometimes considered to be the reason for nitrogen penetration towards the metal/scale
interface. Since nitrogen has a lower chemical affinity for chromium than oxygen, it reacts with
chromium only at the scale/metal interface, where the oxygen potential is lowest. Figure 3
depicts the effect of the presence of nitrogen on pure chromium at 950 °C,
Figure 3 Isothermal weight uptake kinetics for pure chromium in different gases at 950°C.

Scaling reactions of chromium and chromia forming alloys vary in a complicated way with gas
composition. Different oxide scale morphologies and oxidation kinetics are observed, in the
presence or absence of water vapor depending on oxygen partial pressure. The additional
presence of N2 in the reactive gas leads to the formation of Cr2N. Figure 4 presents the formation
of Cr2N underneath chromia scale [8].

Figure 4 Metallographic cross-section after 24 h isothermal oxidation of pure chromium in Ar-1%


O2 (left) and N2-1%O2 (right) at 950 °C [8].
5 Oxidation and sulfurization

5.1 Oxidation
Hypersonic air vehicles are typically exposed to harsh, high temperature oxidative environments,
which can lead to issues of structural integrity due to long-term oxidation and degradation. The
oxidation reaction is one of the most important factors on the high temperature performance of
metals. As well known, the oxide film coating on metal protects the metal substrate from
oxidization. However, it is also known that a significant mismatch in the thermal expansion
coefficient between the metal and the oxide film causes great oxidative stress in the oxide film
and the metal substrate/oxide film interface. The stress can induce a number of cracks in the
oxide film which allow oxygen diffusion toward the metal substrate. This will produce the
oxidation of the metal substrate, i.e. its degradation.

Generally speaking, the oxidation process of metal at high temperature has three steps: a) oxygen
atoms are adsorbed on the surface of the metal, and the reaction between metal and oxygen
produces an oxide film; b) oxygen diffuses through the oxide film to the surface of the metal
substrate; c) metal substrate is further oxidized and the oxide film thickness increases gradually.
Given that the speed of the oxidation reaction is greater than the diffusion of oxygen through the
oxide film [9], a parabolic increase of the weight gain and layer thickness in coupons subjected
to oxidation at high temperatures is observed. As the thickness of the oxide layer increases, it
takes more time for oxygen to go through the oxide film and touches the metal/metal oxide
interface. Thus, the increase of the weight gain in coupon is lower as time progresses. Figure 4
presents an example of such behavior,

Figure 5 C263 superalloy oxidation kinetics at 700, 800 and 900 °C. A parabolic behavior is clearly
observed in the weight gain of the coupons and in its oxide layer thickness, from [10].
5.2 Sulfurization
In many branches of modern technology metallic materials are exposed to sulfur-containing
atmospheres at high temperatures. The corrosion of common metals and conventional oxidation
resistant alloys (chromia formers) suffer a very rapid and often catastrophic degradation.

The reaction of metals and alloys with sulfur is basically identical to oxidation with oxygen.
However, the sulfurization rate of most metals is far higher than the oxidation rate in air or
oxygen because of the following:

• Most sulfides show far greater disorder than the corresponding oxides. Therefore, mass
transport in them is greater.
• Some sulfides form relatively low-melting eutectics with metals. The protective effect of
the film is thereby lost. Examples of this are:

Eutectic Melting point (°C)


Fe-FeS 965
FeO-FeS 940
Fe-FeO-FeS 925
Ni-Ni2S3 645

The low melting point of the Ni-Ni2S3 eutectic is the reason for the high sensitivity of high nickel
materials to sulfur. The resistance to attack by sulfur can be improved by adding chromium.
However, the protective effect of chromium is not as high for attack by S2 as it is for attack by O2
[5].

5.2.1 Physicochemical properties of metal sulfides

The sulfides of common metals show much greater non-stoichiometry, and thereby defect
concentration, than the corresponding oxides. This means that most metals will not produce a
protective scale after sulfidation. On the other hand, refractory –metal1 sulfides seen to show
very low deviations from stoichiometry. The smallest deviation from stoichiometry is presented
by molybdenum sulfide, MoS2 (see Figure 6),

1
Class of metals that are extraordinarily resistant to heat and wear: molybdenum, niobium, tantalum, tungsten
and rhenium.
Figure 6 Collective plot of maximum non-stoichiometry in several metal sulfides and oxides as a
function of temperature at constant sulfur and oxygen pressure from [11].

A comparison of a self-diffusion coefficients in oxides and sulfides clearly indicates that the rate
of cation self-diffusion in common-metal sulfide is generally much higher than in the
corresponding oxides (see Figure 7). This is because in the majority of transition-metal sulfides
there are much higher defect concentrations.

In the case of refractory-metal sulfides the situation is different. MoS2 presents a very low defect
concentration, whereas NbS2 has a rather high one and nonetheless, the rate of niobium
sulfidation is comparable with that of molybdenum. Some studies indicates that the very good
protective properties of the NbS2 scale on niobium result mainly from very low defect mobility
rather than defect concentration [12].
Figure 7 Comparison of self-diffusion coefficients in some metal sulfides and oxides.

Figure 8 Collective plot of the temperature dependence of the sulfidation and oxidation rates of
some metals from [11].
5.2.2 Sulfidation of pure metals

The main difference, in turn, is observed between sulfidation and oxidation rates of these metals.
In spite of compact-scale formation and the parabolic course of the reaction, the rate of sulfide
corrosion of such important metals as nickel, cobalt and even chromium, is many times higher
than their oxidation rates. This is illustrated in Figure 8 depicting the temperature dependence
of the sulfidation and oxidation rates of some metals.

5.2.3 Sulfidation of alloys

Despite great similarities in the growth mechanism of oxide and sulfide scales, the sulfidation and
oxidation rates of Fe-Cr alloys, for example, differ dramatically. This is illustrated in Figure 9 by
the dependence of the oxidation and sulfidation rates of iron-chromium alloys on their
composition.

Figure 9 The dependence of the sulfidation and oxidation rates of Fe-Cr alloys on composition
at 900°C from [13].

As can be seen, the oxidation rate of chromia formers (about 40 % Cr) is more than four order of
magnitude lower than the sulfidation rate of alloys, on the surface of which a homogeneous Cr2Sx
scale is formed.
More precisely, when Fe-Cr alloys were exposed to S2 gas at 1 atm at high temperatures (700-
800 °C), the reaction with sulfur depended on the amount of chromium in the alloy. At 1.86 %
Cr, an FeS layer formed below the Chromia layer. Between 1.86-38.3 % Cr, an outer Fe1-xS layer
and an inner (FeS, FeCr2S4) mixed layer formed and at percentages of chromium higher than
38.3 %, a solid solution of FeS-Cr2S3 formed. Although Cr decreased the sulfidation rate, even
Fe-Cr alloys with high Cr contents displayed insufficient corrosion resistance. This is attributed
to the fact that sulfidation rates of common metals are 10-100 times faster than oxidation rates
because the sulfides have much larger defect concentrations and lower melting points than the
corresponding oxides [14].

From a practical point of view chromium alone do not seen very promising in designing novel
materials resistant to sulfide corrosion. In contrast to this, refractory metals seem to be very
promising as they show excellent resistant to sulfur. Douglas [15] has shown that combined
alloying of iron with molybdenum and aluminum dramatically decreases the sulfidation rate of
iron as illustrated in Figure 10,

Figure 10 Temperature dependence of the sulfidation rate of Fe – 30 Mo and Fe – 30 Mo – x Al


alloys compared to the sulfidation rates of pure iron and molybdenum.
Thus, refractory metals such as molybdenum are highly resistant to sulfide corrosion, their
sulfidation rates being comparable to the oxidation rate of chromium, representing one of the
most-resistant metals to oxide corrosion.

The main reason why common metals are rapidly attacked by sulfur results from the very high
concentration of point defects in their sulfides. On the other hand, refractory metals are highly
resistant to sulfide corrosion because of either low defect concentration or low defect mobility
in the sulfides of these metals. The mechanisms of sulfide and oxide corrosion of pure metals are
very similar. The difference is due to a predominant disorder in transition-metal sulfide. All
conventional oxidation-resistant alloys (chromia formers) undergo very rapid degradation in
highly sulfurizing environments because of heterogeneous-scale formation of poor protective
properties. However, new prospects for the development of coating materials resistant to high-
temperature sulfide corrosion have been created by alloying common metals with molybdenum
and aluminum.

5 Effect of Ash, Deposit and Salts

Many corrosion problems arise through deposition of ash and dust on the metal surface. These
deposits can react at elevated temperatures, even in the solid state, with the protective films of
metals and alloys to form new compounds with different transport properties. However,
corrosion becomes particularly intense when fusible phase form because either the deposits
contain components with relatively low melting points, or low melting eutectics arise between
the deposits and protective films. A well-known phenomenon is oil ash containing vanadium,
which in the presence of surplus oxygen creates a critical oxidation rate (see Figure 11). Other
oxides such as PbO, MoO3 and B2O3 can also form low-melting eutectics. Alkali-metal sulfates
are also often responsible for increased corrosion (see Figure 12). The corrosion kinetics are
significantly enhanced by several orders of magnitude compared to the oxidation without any
salt. As shown in Figure 13 for nickel, reacted beneath a sodium sulfate (Na2SO4) deposit at 900
°C in an oxygen-sulfurous anhydride (SO2) atmosphere. The corrosion mechanisms have to be
distinguish between “Type I and Type II hot corrosion”. Type I usually occurs at temperatures
higher than the melting point of Na2SO4 (T > 884 °C) and the oxide scale is dissolved in the salt
melt due to basic fluxing. Type II hot corrosion occurs at temperatures below the melting point
of Na2SO4. One example of this is the corrosion of nickel beneath a solid sodium sulfate salt
deposit in SO3 –containing gas. In the earliest stages of corrosion a sodium sulfate-nickel sulfate
solid solution is formed. As this reaction proceed, the amount of nickel sulfate increases and the
salt mixture melt. This results in accelerated corrosion. The melting point of the salt mixture
reduces as the partial pressure of sulfuric anhydrite (SO3) increases.

Figure 11 Dependence of the mass increase of steel X20 CrNiSi 254 after annealing for 6 h at 800
°C in a) V2O5-containing oil ashes on the oxygen content of the gas and b) Natural oil ash with
no vanadium.

Figure 12 Corrosion rate of carious high temperature and heat resistant steels embedded in pure
potassium sulfate after reaction for 700 h at 650 °C in an air-steam environment containing SO2.
Figure 13 Comparison of the reaction kinetics of pure nickel with and without deposit of sodium
sulfate in 101.3 kPa O2-4% SO2 at 900 °C. a) with sodium sulfate, b) without sodium sulfate.

The corrosion mechanism distinguish between basic and acidic dissolution and fluxing. Basic
solubility of oxide scales occurs by oxide ions, being present in the molten salt from the
dissociation of the sulfate ion according to,

SO42  SO2  12 O2  O 2 (8)

By the oxidation of the bare metal surface by sulfate in a basic melt, metal oxide and additional
oxide ions are formed according to:

Ni  SO42  NiO  SO2  O 2 (9)

Hence, the [O-2] on top of the metal oxide is higher than in the entire melt and dissolution of the
oxide takes places according to,

NiO  O 2  NiO22 (10)

In the case of basic dissolution, the metal oxide is dissolved as a complex oxide ion.

Acidic dissolution occurs by SO3. In principle the following reaction takes place,
NiO  SO3  Ni 2  SO42 (11)

The oxides are dissolved as metal ions in the sulfate melt and metal sulfate are formed [5].

6 Hydrogen embrittlement and decarburization


The last mechanism to be discussed is not related to the fate of the protective scale but to the
mechanical damage occurring by the occlusion of hydrogen in the body of the metal. The current
understanding of hydrogen embrittlement is that hydrogen from the environment dissolves into
steel, migrates as atomic hydrogen towards internal stress centers such as crack tips, and
ultimately facilitates nucleation and propagation of cracks, leading to failure. While it is well
accepted that hydrogen embrittlement is caused by accumulation of atomic hydrogen at internal
centers of high triaxial stresses, the actual micromechanism of failure is not fully understood.
The entrance of hydrogen gas into the metal structure goes through its dissociation at the
surface,

H 2  2H ad
*
(12)

These adsorbed hydrogen atoms then get dissolved into the metal,
*
H ad  H ab
*
(13)

Hab* travel through the metal matrix by diffusion.

This Hab* accumulates in regions of high hydrostatic stresses (e.g. near a crack tip) and this
accumulation can induce dislocation plasticity. It is well known that dislocation plasticity
generates excess vacancies. And hydrogen bind strongly to these vacancies, stabilizing them. A
tipping point is reached, through attainment of a critical local excess vacancy concentration,
where nucleation and growth of nanovoids occurs. This produce the ultimate failure through
nanovoid coalescence [16] (see Figure 14).
Figure 14 Schematic of the Nano-void coalescence mechanism.

At temperatures above 300 °C, the partial pressure is the decisive factor for the impact of
hydrogen (see Figure 15).

Figure 15 Influence of hydrogen pressure and temperature on the onset of damage.


Over 300 °C the hydrogen solubility in steel follows the Sievert law:

cH  kT  PH 2 (14)

The Arrhenius equation kT  k0  exp  Q RT  can be used to determine kT, where Q is the heat
of hydrogen dissolution in the metal = 27.2 kJ/mol.

In addition at this temperature there is no longer a problem of the high energy threshold for the
dissociation, adsorption and absorption of hydrogen. After diffusion of the hydrogen the damage
at increased temperatures, however, is not caused by the material embrittlement but by the
structure changes in the material. In carbon steels the combined carbon reacts as cementite with
the hydrogen and forms methane,

Fe3C  4H ab
*
 3Fe  CH4 (15)

The steel loses its strength due to decarburization. The resistance to the formation of methane
can be improved by suitable alloying additions or carbide components such as chromium,
molybdenum, vanadium and tungsten or by using austenitic steels [17].

Bibliography
1 R.F. Hochman, “Catastrophic deterioration of high temperature alloys in carbonaceous atmospheres”
Proceedings of the Symposium on properties of High Temperature Alloys (1976) 715.
2 L.D. Burns, “Corrosion on new distillation unit processing low sulfur crude”, Corrosion 6 (1950) 169.
3 M. Szkodo and G. Gajowiec, “Studies of the mechanism of metal dusting of 10CrMo9-10 steel after 10 years of
operation in the semi-regenerative catalytic reformer”, Corr. Sci. 102 (2016) 279.
4 H. J. Grabke, “Metal dusting”, Materials and Corrosion 54 (2003) 736.
5 H. Grafen, E. Manfred Horn, H. Schlecker, H. Schidler and M. Spiegel, “Corrosion, 2. High Temperature” Ullmann’s
Encyclopedia of Industrial Chemistry, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim (2015).
6 H. J. Grabke, E. Reese and M. Spiegel, “The Effect of Chlorides, Hydrogen Chloride, and Sulfur Dioxide in the
Oxidation of Steels below Deposits”, Corr. Sci. 37 (1995) 1023.
7 D. Fantozzi, V. Matikainen, M. Uusitalo, H. Koivuluoto and P. Vuoristo, “Chlorine-induced high temperature
corrosion of Inconel 625 sprayed coatings deposited with different thermal spray techniques”, Surface & Coatings
Tech. 318 (2017) 233.
8 M. Michalik, S.L. Tobing, M. Hansel, V. Shemet, W.J. Quadakkers and D. J. Young, “Effect of water vapour on the
high temperature nitridation of chromium” Materials and Corrosion 65 (2014) 260.
9 C. Wang, S. Ai and D. Fang, “Effect of Oxidation-Induced Material Parameter Variation on the High Temperature
Oxidation Behavior of Nickel”, Acta Mechanica Solida Sinica, 29 (2016) 337.
10 N. Sheng, K. Horke, A. Meyer, M. R. Gotterbarm, R. Rettig and R. F. Singer, “Surface recrystallization and its
effect on oxidation of superalloy C263”, Corr. Sci. 128 (2017) 186.
11 S. Mrowec and J. Janowski, in “Selected Topics in High Temperature Chemistry”, O. Johannesen and A. G.
Andersen, eds. (Elsevier, New York, 1989), pp. 55-99.
12 K. Przybylski and S. Mrowec, “High temperature sulfide corrosion and transport properties of transition metal
sulfides” Proceeding of the 3rd international symposium on material chemistry in nuclear environment, JAERI-
CONF-2003-001, Japan 2003.
13 S. Mrowec, “The problem of sulfur in high-temperature corrosion” Oxid. Met., 44 (1995) 177.
14 M. J. Kim, M. A. Abro and D. B. Lee, “Corrosion of Fe-(9~37) wt. %Cr Alloys at 700-800 °C in (N2, H2O, H2S)-
Mixed Gas” Metals 6 (2017) 291.
15 G. Wang, D. L. Douglass and F. Gesmundo, “High-temperature sulfidation of Fe-30Mo alloys containing ternary
additions of Al”, Oxid. Met., 35 (1991) 349.
16 T. Neeraj, R. Srinivasan and J. Li, “Hydrogen embrittlement of ferritic steels: Observations on deformation
microstructure, nanoscale dimples and failure by nanovoiding”, Acta Materialia 60 (2012) 5160.
17 J. Woodtli and R. Kieselbach, “Damage due to hydrogen embrittlement and stress corrosion cracking” Eng. Fail.
Anal., 7 (2000) 427.

View publication stats

You might also like