You are on page 1of 29

J.

of Supercritical Fluids 29 (2004) 129

Corrosion in high-temperature and supercritical water and aqueous solutions: a review


Peter Kritzer
Freudenberg Vliesstoffe KG, D-69 465 Weinheim, Germany Received 22 April 2002; received in revised form 6 January 2003; accepted 14 February 2003

Abstract The aim of the present article is to review some of the common corrosion phenomena and describe the predominant corrosion mechanisms in high-temperature and supercritical water. Corrosion in aqueous systems up to supercritical temperatures is determined by several solution-dependent and material-dependent factors. Solution-depending factors are the density, the temperature, the pH value, and the electrochemical potential of the solution, and the aggressiveness of the attacking anions. Material-dependent parameters include alloy composition, surface condition, material purity, and heat treatment. Corrosion phenomena that are observed include intergranular corrosion, pitting, general corrosion, and stress corrosion cracking. The solubility and dissociation of both attacking species and corrosion products play the most important role for corrosion in high-temperature water. Both solubility and dissociation processes are strongly inuenced by the density, or the ionic product, respectively, of the solvent. High values of both parameters favor ionic reactions, and thus, accelerate electrochemical forms of corrosion. At low densities, water behaves like a non-polar solvent, and thus, ions associate. In these cases, the concentation of e.g. aggressive H+ drops down and thus, solutions containing species such as HCl become neutral and thus less aggressive. Further, corrosion products plug the surface and material loss stops. Materials parameters have inuence especially on the initiation of corrosion. In the present article, these factors are linked with the physical and chemical properties of high-temperature and supercritical water. An outlook is also given for future research needs. 2003 Published by Elsevier B.V.
Keywords: Corrosion; High-temperature water; Supercritical water; Supercritical water oxidation; Density; Ionic product

1. Physical and chemical properties of high-temperature and supercritical water The physical and chemical properties of hightemperature and supercritical water (T > 374 C; p > 22.05 MPa) have been investigated and reviewed in detail by E.U. Franck and co-workers [15]. A

Tel.: +49-6201-80-4003; fax: +49-6201-88-4003. E-mail address: peter.kritzer@freudenberg.de (P. Kritzer).

computer program is available to calculate these properties [6]. Fig. 1 shows schematically the course of density when crossing the critical values of pressure or temperature. At a certain temperature, a drop can be observed, at which density and ionic product of high-temperature water fall down drastically. With increasing pressure, this drop shifts toward higher temperatures and its step height decreases (Fig. 2). It will be shown in the following sections that this drop is responsible for the uncommon

0896-8446/$ see front matter 2003 Published by Elsevier B.V. doi:10.1016/S0896-8446(03)00031-7

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129

Fig. 1. Phase diagram of water and schematic course of density versus pressure and temperature.

Fig. 3. Ionic product of high-temperature water and steam at different temperatures versus pressure. At low pressures, water behaves as a non-polar solvent with low self-dissociation. High pressures can increase the ionic product to values above those found for water at ambient conditions.

corrosion behavior of most materials in high-temperature water. Supercritical water has both liquid-like and gas-like characteristics like density between both states, high diffusivity and good heat-transporting properties. From this point of view, supercritical water can be seen as a dense gas. Thus, supercritical water is a medium with excellent transport properties and possesses a complete solvency for most gases [7,8], and organic compounds [911]. The ionic product of high-temperature water and steam versus temperature is illustrated in Fig. 3. Steam and low-density supercritical water behave like non-polar solvents

with low solvency for ionic compounds. Typical solubility values for sodium chloride and sodium sulfate in low-density supercritical water are 100 and 1 ppm, respectively. The solubility of NaCl versus pressure and temperature is illustrated in Fig. 4 (after Ref. [12,13]). These characteristics can be tuned by a change of pressure. In the vicinity of the drop of density and ionic product mentioned above, a slight modication of temperature and pressure has a huge effect on the physical and chemical parameters of the solvent. While chemical reactions occurring in water of high density are dominated by ionic pathways, low-density water favors radical reactions [1426]. High-density supercritical water is still a good solvent for organic compounds, but also for gases and salts. All these special features lead to interesting applications of supercritical water. In the present article, the following expressions are used as follows: Supercritical water: T > Tc (374 C); p > pc (22.05 MPa) Subcritical water: T < Tc ; p > pSaturation (p may also be above pc ) Steam: T < Tc (T may also be above Tc ); p < pSaturation 2. Applications of high temperature and supercritical water In the last two decades, high-temperature and supercritical water have become interesting mediums for

Fig. 2. Drop of physical properties of high-temperature water at different pressures. Note that the drop shifts towards higher temperatures at higher pressures. Data after Steamtables, see Kritzer et al. [171]; reprinted with kindly permission of the National Association of Corrosion Engineers (NACE).

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129

Fig. 4. Solubility of NaCl in supercritical water at different pressures versus temperature. Salt solubility follows the density drop mentioned above (e.g. Fig. 2). Data after Martynova [13]. Table 1 Applications of supercritical water Application Chemical reactions Hydrothermal syntheses Waste oxidation Radioactive waste reduction Biomass conversion Plastic degradation Synthesis of nano-particles Properties exploited High solvency for organics, tunable conc. of H+ and OH Solubilities High solvency for organics and oxygen High solvency for organics and oxygen; solubilities High solvency for organics High solubility of the monomers Low solubility of salts Refs. [1826] [27] [20,2833] [3436] [3741] [42] [4346]

chemistry [1846]. A number of reviews dealing with the general characteristics of supercritical water have been published recently [4,5,1417]. Here, only a brief overview is given of the enormously growing number of different applications so of which are listed in Table 1.

3. Corrosion in high-temperature and supercritical water 3.1. Short glossary of the typical forms of corrosion In this section, the principle corrosion mechanisms observed in high-temperature water will be described. An evaluation is given, which corrosion phenomena are most relevant for different applications.

3.1.1. Pitting corrosion Pitting corrosion is a localized form of corrosion occurring in the passive state of the metal. An example is shown in Fig. 5. In this case, the oxide lm in principle has a protective nature. However, aggressive anions such as chloride or bromide can penetrate into the oxide lm and destroy the lm locally. Typical initiation points are inclusions or grain boundaries. Small pits are formed in the rst step. The oxidation and dissolution of metal components such as nickel and/or chromium ions followed by their reaction as Lewis acids with the water leads to a strong acidication of the solution inside the pits. Due to migration processes from the bulk solution, the concentration of aggressive anions increases. Thus, the solution becomes more and more corrosive inside the pits and corrosion progresses. Pit growth generally occurs in

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129

Fig. 5. Typical form of pitting corrosion. The phenomenon is stochastic and due to complicate and changing local chemistry inside the pits, penetration rates are not linear. Note that the neighbored surface remains unattacked, underlining the principle stability of the oxide lm. Corrosion rates are in the range of 1000 m in 100 h (nickel-base alloy 625; conditions: [HCl] = 0.05 mol/kg; [O2 ] = 0.48 mol/kg; T = 160 C; p = 24 MPa; t = 124 h; see Ref. [173]).

Fig. 6. Typical wave-like form of general corrosion (shallow pitting). Since this form of corrosion is diffusion-controlled, corrosion rates are linear. Note that due to the general instability of the oxide lm, the complete surface is attacked. Grain boundaries at the surface might be etched (not visible in this gure due to too low magnication). Corrosion rates are in the range of 500 m in 100 h (nickel-base alloy 625; conditions: [HCl] = 0.05 mol/kg; [O2 ] = 0.48 mol/kg; T = 350 C; p = 24 MPa; t = 124 h; see Ref. [173]).

high rates. Increasing temperature additionally weakens the oxide lm, and thus, pitting occurs much easier at higher temperatureswhich is indicated, e.g. by the decrease of the lower limit of the electochemical potential, at which pitting is observed with increasing temperature [4767]. The reason for this behavior is (a) the higher number of locally limited defects in the lm [47,48] and (b) the increased tendency of the oxide lms to incorporate anions at higher temperatures [5053]. Note that still large areas of the surface remain unattacked, while in neighbored areas, corrosion rates can exceed some 10 m/h. Due to its stochastic and non-predictive nature, pitting is a dangerous form of corrosion. Review articles dealing with pitting corrosion have been published by several researchers [6879]. 3.1.2. General corrosion In contrast to pitting corrosion, the reason for general corrosion is a general instability of the oxide lm, and thus, corrosion attacks the entire surface. The typical morphology is a shallow, wave-like pitting (Fig. 6). General corrosion happens in such cases, where none of the alloy components could form a protective layer. This is the case during the active and transpassive dissolution of materials. Both processes

are described below. The starting points of general corrosion are the same like these found for pitting, the weakest points of the surface. It has been shown for different alloys and stainless steels that pitting switches to general corrosion above a certain temperature called the inversion temperature [60,62]. At this temperature, which is typically in the range of 200250 C, also the pitting potential no longer decreases, but shifts into that of the transpassive dissolution. In case of chromium-containing alloys, chromate is released under these conditions. It was shown that concentrations of oxygen even below some 100 ppm are sufcient for chromate formation [60]. The absolute material loss caused by general corrosion might be high, but due to the diffusion-controlled character, the corrosion rates are linear and thus, can be predicted. General corrosion is a common phenomenon in oxidizing high-temperature water. In cases, where chromate is released, grain boundaries on the surface might be attacked since chromate act as a grain-boundary etching agent [60,62]. It must be mentioned that chromium(III) oxides are still found under these conditions, although the chromate formation is thermodynamically favored, which underlines the low conversion rates of solid Cr(III) towards Cr(VI).

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129

Fig. 7. Intergranular corrosion. Due to the difference between grain boundaries and grains, corrosion occurs preferentially in this regions. Corrosion at this picture occurred at supercritical temperature; signicant material loss could not be observed. Corrosion rates are in the range of 10 m in 100 h (nickel-base alloy 625; conditions: [HCl] = 0.05 mol/kg; [O2 ]=0.48 mol/kg; T = 450 C; p = 24 MPa; t = 150 h, see Ref. [173]).

Fig. 8. Stress corrosion cracking starting at the bottom of a shallow pit and passing through the whole wall of a tubular reactor in the temperature transition region between passive and transpassive states. In this example, failure occurred in less than 50 h in a solution containing HCl and oxygen (nickel-base alloy 625; conditions: [HCl] = 0.10 mol/kg; [O2 ] = 0.48 mol/kg; T = 220 C; p = 38 MPa; t = 24.5;Von Mises stress values are rd. 100 N/mm, see Ref. [173]).

3.1.3. Intergranular corrosion (intercrystalline corrosion, IC) The phenomenon of IC has found much attention in literature [8085]. Grain boundaries and the neighbored areas of the grains normally are chemically different compared to the bulk grains themselves. Additionally, new phases can be formed at the grain boundaries such as metal carbides or nitrides. Further, an enrichment, or segregation, respectively of trace elements at the grain boundaries lead to detrimental conditions. During IC, either the grain boundaries or neighboring grain areas might be attacked (Fig. 7). Local electrochemical elements might be formed. Different corrosion mechanisms are observed, so IC can be observed under nearly all conditions. Both, penetration depth and amount of dissolved material regularly are low. Therefore, IC is not as critical as the other forms of corrosion. However, whole grains may be dissolved at longer times and by the inuence of mechanic stress, IC may lead to the dangerous stress corrosion cracking (SCC). 3.1.4. Stress corrosion cracking SCC is an extremely dangerous form of corrosion, since its nature and its occurrence are stochastic (Fig. 8). Thus, failures can be catastrophic. Therefore, publications dealing with this form of corrosion

are numerous [86103]. SCC is observed along the grain boundaries (inter-granular) or through the grains (trans-granular). SCC is commonly present in the transition ranges between the active and the passive, or the passive and the transpassive potential, respectively. Thus, SCC was observed in high-temperature water in the presence of either hydrogen (active region) or oxygen (transpassive region). Most detrimental anions are chloride, bromide, and sulde. SCC needs both a chemical and a mechanical component. At higher mechanical stress of the material, the chemical aggressiveness of the solution does not need to be so high. On the other hand, in highly aggressive environments, relatively low values of stress can cause SCC. SCC commonly leads to a failure of the entire reactor due to leakage. If further burnable gases like hydrogen are released from such a leakage in the heating section of a reactor, res and even explosions cannot be excluded. 3.2. Brief historywet air oxidation and supercritical water oxidation The most widespread use of high-temperature water is as a heat transporting and heat transmission medium in power station cycles. Water in this environment

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129

contains low concentrations of corrosive species such as salts and oxygen. Additionally, suitable corrosion inhibitors are added to the water in such systems as noted later. Thus, in practice, corrosion phenomena that are dominant, are those, in which the mechanical component plays the most important role. Corrosion occurring under these conditions has been extensively investigated [104113]. In nuclear power generation cycles, also the effect of radioactivity has to be taken into account. Some 30 years ago, concentrations of inorganic compounds and oxygen above the ppm level in high-temperature water were only present in the industrial application of wet air oxidation (WAO) [114118]. WAO is the oxidative treatment of organic wastewaters at temperatures up to 320 C and pressures up to 20 MPa. Reaction times are 3060 min. The reactor material mostly is titanium grade 2, which has been proven to be resistant under these conditions. The purpose of this kind of waste treatment is to crack stable organic molecules and form harmless compounds like acetic acid, which can be easily destroyed subsequently with biological treatment. Currently, over 100 WAO plants are run successfully world-wide [114,118]. The typical oxidizer in such systems is oxygen from air, but also nitric acid has been used. The process of supercritical water oxidation (SCWO)the most widely investigated supercritical water application so far, can be seen as an extension of WAO to higher temperatures and pressures. Typical process parameters of SCWO are reaction temperatures of 400700 C and pressures of 2450 MPa. During the process, organic compounds react completely with the oxidantmostly oxygenin a single-phase reaction to form CO2 and H2 O. Hetero-atoms like halogens, sulfur, or phosphorus present in the organic wastes are transformed into their mineral acids HF, HCl, HBr, H2 SO4 or H3 PO4 , respectively. HI is not formed since iodides are rapidly converted to elemental iodine in the presence of oxygen. Organically bound nitrogen forms N2 predominately and small amounts of N2 O. Thereby, N(III) nitrogen is oxidized by the oxidant, while N(+V), being a powerful oxidant itself, is reduced. Undesirable by-products common in incineration like dioxins, or higher NOx are normally not formed, since reaction temperatures not favorable for their formation. SO2 is transferred to SO3 and removed with the aqueous

efuent as sulfate. The variety of employed wastes include warfare agents, rocket propellants, radioactive wastes, and wastes from the paper and chemical industries [2836]. In contrast to WAO, SCWO is able to achieve complete oxidation of most organic materials at short reaction times that range from seconds to minutes. Common reactor materials like stainless steels at service times of some 1000 h used in SCWO of aqueous waste streams of organic compounds containing no hetero-atoms do not exhibit remarkable corrosion rates. Unfortunately, the degradation of such compounds is economically uninteresting, since they can be treated cheaper with alternative methods such as incineration. It has been found in different studies that SCWO could become economically interesting for the treatment of highly toxic material like chlorinated dioxines or S- and P-containing warfare agents. However, the highly corrosive efuents resulting from the treatment of these waste streams, together with the precipitation of salts and a subsequent reactor plugging, has hindered a wider application of the process so far [33]. The number of investigations of the corrosion resistance of different reactor materials under SCWO conditions is enormous [119201]. However, a material that can withstand every conceivable chemical conditions has not yet been found. Most reviews dealing with the SCWO process are overly optimistic, and critical publications dealing with the process problems are few [28,33]. To solve these existing problems, the following routes for future research have been suggested [33]: Determination the corrosion mechanisms occurring under high-temperature conditions. Understanding the inuence and interaction of corrosion-determining parameters of the solution such as temperature, pressure, solution components, and pH. Understanding the inuence of the reactor material parameters, the material itself, alloy composition, heat treatment and the level of impurities. Determination of conditions, where corrosion is low. It is probable that no single reactor material can withstand every conceivable reaction conditions. Evaluation, of whether a particular corrosive species can be excluded, compensated or avoided so that a

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129

corrosion-resistant material for the desired application can be chosen. Construction of special reactor designs to limit contact of the corrosive species with the reactor wall or temperaturepressure regimes, in which highly corrosive species are formed, are avoided [201]. Several SCWO pilot plants are under operation world-wide. A recent overview over these activities is given elsewhere [28]. In recent years, also the application of high-temperature and especially supercritical water as medium for chemical reactions, synthesis and crystallization processes (see Table 1) has become a eld of wide interest. In these applications, the presence of salts, acids, or gases might be necessary for the chemical reaction, since they might be educts and/or products. 3.3. Solution parameters inuencing corrosion 3.3.1. Ionic reactions and oxide lm stability As mentioned above, high-temperature water is a medium of tunable density, ionic product and dielectric constant (see Figs. 13). Hence, its solvent character can vary from highly polar at high densities to nearly non-polar at low densities. High polarity favors the solubility and/or the dissociation of ionic species like salts, acids, and bases, and thus favors ionic reactions. Howeverand this stands in contrast to water at ambient conditionsalso non-polar gases and organic compounds might be completely soluble at sufciently high temperatures and pressures [7,8]. Low-density water suppresses ionic reactions and favors radical reaction pathways, especially at high temperatures. Generally, the high temperatures drastically accelerate all chemical reactions, and thus, kinetic effects become less and less important. Aqueous corrosion processes commonly are seen as ionic reactions. The rate-determining step generally is the dissolution of a protective surface saltmostly an oxide. In high-temperature water, this oxide might be of different nature or composition than the surface oxides found under ambient conditions. In general, oxides are preferentially formed compared to hydroxides at high temperatures, and the content of crystal water is lower. At ambient conditions, thermodynamically unstable compounds might form a protective layer in such

Fig. 9. Stability island of a protective oxide and the principle mechanisms for its dissolution. Horizontal direction means chemical dissolution (variation of the pH without changing the potential), while vertical direction means electrochemical dissolution (variation of the electrochemical potential without changing the pH). The stability island correlates with the passivity region of the metal.

cases, when their kinetic stability is high, and thus, their dissolution rate is very low. On the other hand, in high-temperature water of high density, thermodynamic stability is a necessary condition for a protective lm, since all chemical dissolution rates are accelerated. The principle reactions to dissolve such an oxide lm are similar to those found at room temperature and are illustrated in Fig. 9. A detailed description of the different dissolution processes is given in the following sections. 3.3.2. Temperature Corrosion rates in water generally increase with temperature. Indeed, most corrosion processes have a minimum temperature, below which corrosion is limited or does not occur. For example, stainless steels are attacked by chloride-induced pitting corrosion above some 80100 C, while the materials are resistant at lower temperatures. However, these ndings are only true as long as the other solution properties remain unchanged over the investigated temperature range. The rst experimental corrosion experiments in supercritical aqueous solutions gave surprising results: Corrosion was lower by orders of magnitude at the highest supercritical experimental temperatures of 500 C compared

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129

with corrosion at subcritical 300 C [119121,130 132,135,136]. This behavior was traced to the change of physical and chemical properties of water. 3.3.3. pH-value: chemical dissolution The pH value normally is based on the equilibrium reaction of the self-dissociation of water H2 O = H+ + OH This reaction has an endothermic character and thus, the equilibrium shifts towards the right side with increasing temperature. For example, at a pressure of 25 MPa, it runs through a maximum at about 300 C (Fig. 2). At these temperatures, the concentrations of both H+ and OH are about three orders of magnitude above the values in ambient water, so that water can be considered to be both acidic and alkaline. A technical application for this phenomenon are acidor base-catalyzed reactions such as estherications or hydrolysis reactions in high-density subcritical water [26]. Under these conditions, strong mineral acids like H2 SO4 , which are necessary for catalyzing such reactions at room temperature, are not used, and thus, a separation of the acidic catalyst in a second step is unnecessary. Higher temperatures at constant pressure lead to a reduction of the number of hydrogen bonds and thus favor the non-polar character of water [15,16]. Consequently, self-dissociation drastically drops down above a certain temperature. The dissociation of mineral acids follows the course of density of the solvent [170,202,203]. Since the dissociation of strong acids is exothermic, the decrease is strongly monotonous. In Fig. 10, the dissociation of HCl versus temperature is shown (data after Ref. [203]). Other strong acids behave more or less similar. Again, the large pressure inuence at high temperatures is remarkable and, e.g. HCl dissociation correlates well with the solution density [203]. The pH value, or dissociation rate, respectively, is one of the factors, which has most inuence on corrosion in high-temperature water. High or low values of pH lead to a chemical dissolution, which is described as dissolution of the protecting oxide at constant electrochemical potential, and indicated as the horizontal direction in Fig. 9. This form of dissolution is caused by the amphoteric character of most oxides; they can be dissolved either in acidic or in alkaline solutions.

Fig. 10. Dissociation of HCl in water versus temperature after Frantz and Marshall [203]. Dissociation follows the density of the solvent. For comparison, the pH of 1 M acetic acid at room temperature is indicated. It can bee seen that HCl at high temperatures still acts as acid, if the pressure is high enough. The behavior of other acids is similar (see, e.g. Ref.[202]).

Typical examples are the oxides of iron, nickel, and chromium. The pH region of highest stability depends on the isoelectrical point of the oxide and can change with temperature. Theoretical work of Adschiri et al. determined the solubility of the protecting oxides of iron, nickel and chromium in supercritical water versus pH value at non-oxidizing conditions [188]. For all three metal oxides, the solubility at neutral or slightly alkaline conditions is lowest. Chromium oxide show the lowest solubility of the three oxides, nickel oxide the highest. In a later publication, the same authors investigated the effect of oxygen on the corrosion of pure iron, nickel, chromium and different alloys versus pressure at slightly supercritical temperatures [194,195]. In all cases, an increased pressure increases the corrosion rate for the pure metals [194]. Chromium shows the highest stability compared to iron and nickel, in coincidence with the solubility data. This behavior also explains the observations, why chromium-containing alloys show a better corrosion resistance than chromium-free ones [195]. In a extensive study, Schroer et al. investigated the inuence the chromium content in binary NiCr alloys and the effect of an addition of aluminum and tungsten as alloying elements [198]. The authors also found a strong pressure dependency of the corrosion. At comparable pressuretemperature conditions 410 C/40 MPa, the corrosion rates for all these alloys were comparably high, and lie within one order of magnitude. The

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129

different corrosion rates were claimed on the different morphologies of the Cr(III) oxide/hydroxide layer. It must be mentioned that in special cases, no corrosion might occur, even although the oxide layer is removed. This is the case for solutions, in which no oxidative species are present that could further oxidize the pure metal after initial oxide removal. This state normally is called immunity of the metal or the alloy. The value of pH also inuences the solubility of the primary corrosion products. These are formed intermediately at the rapidly corroding surface, when local super-saturation occurs leading to a precipitation of these salts. For example, it has been shown that corrosion rates of nickel-base alloys in strongly oxidizing high-temperature solutions of different acids are inuenced by the solubility of the nickel salts of the acid anion [167,172,173]. On the other hand, the substance with the lowest solubility, which for stainless steels and nickel-base alloys in acidic solution are Cr (III) oxides or hydroxides, are present as solid corrosion products in form of thick scales on the corroded surface, while nickel is dissolved selectively. Typical concentrations of metal ions released from a nickel-base alloy 625 reactor (composition: 62 wt.% Ni, 22 wt.% Cr, 9.0 wt.% Mo) under oxidizing alkaline, neutral, and acidic conditions are shown in Table 2 [166]. In the same table, enrichment factors (EF) for Cr and Mo, normalized to the nickel content in the alloy, and solution, respectively, are listed. Acidic solution leads to high concentrations of chemically dissolved Ni in the efuent. Cr and Mo show low EF values, which means that Ni is dissolved preferentially. Increased pH values reduce the Ni concentrations by

orders of magnitude according to the thermodynamic stability of NiO under these conditions (see Section 3.3.7), while Cr and Mo are not inuenced that much. Chromium and molybdenum behave similar to each other. Both metals are dissolved chemically, mainly in acidic solutions, and electrochemically as thermodynamically favored CrO2 or MoO2 , mainly in al4 4 kaline solutions. The low concentrations of dissolved Cr under acidic conditions underlines the low solubility of the chromium(III) oxides. Under neutral, and especially alkaline conditions, Cr and Mo show a high enrichment compared to Ni. This, together with the decrease in concentration of all metals at alkaline conditions can be attributed to the protecting NiO, which suppresses a further dissolution of the alloy [165,166,172]. 3.3.4. The electrochemical potential and the solubility of gases: electrochemical dissolution At room temperature, non-polar gases like oxygen, hydrogen, or nitrogen are water-soluble only in low ppm concentrations according to their Henrys constants. Increasing the pressure at room temperature leads to only slightly higher gas solubility. Increasing the temperature at these high pressures results in a slight decrease of the gas solubility up to around 100 C. Above that temperature, gas solubility increases drastically [204208]. At supercritical parameters of the solvent, gases are completely miscible with the supercritical solvent [7,8]. Since corrosion reactions are generally oxidation reactions, they are very sensitive to changes in the electrochemical potential. Under these conditions, the

Table 2 Concentrations of metals released from a tube reactor of nickel-base alloy 625 corroded by acidic, neutral and alkaline chloride solutions (Tmax = 350 C; p = 24 MPa; [O2 ] = 0.5 mol/kg; [HCl] = [NaCl] = [NaOH] = 0.05 mol/kg; further experimental details are described in Ref. [166]). [Nidissolved ] ppm HCl + O2 NaCl + O2 NaCl/NaOH + O2 1560 10 0.1a [Crdissolved ] ppm 17 24 1.8 EF (Cr) 0.03 6.9 51.6 [Modissolved ] ppm 12 6 0.7 EF (Mo) 0.05 4.2 48.8

The enrichment factor (EF) of an element X compared to nickel is calculated after the following equation by taking into account the concentrations of dissolved metals in solution and the alloy composition (Ni: 62.8 wt.%, Cr: 21.9 wt.% Mo 9.0 wt.%). EFNi (X) =
a

[X]/[Ni] (dissolved) [X]/[Ni] (alloy)

Value is indeed below 0.1 ppm (the detection limit of the analytical technique), so the EF values for Cr and Mo are even higher.

10

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129

so-called electrochemical dissolution, which is a dissolution at constant pH value as indicated by the vertical direction of Fig. 9, occurs. In general, when a metal is protected by an oxide layer, it can be considered to be in its passive state. For the formation of the oxide lm, a minimum electrochemical potential is necessary. Below that potential, the metal might undergo a so-called active dissolution. In most stainless steels and nickel-base alloys, chromium is the oxide-forming element. In acidic solutions, the pure metal and its alloys are passivated by thermodynamically stable, solid Cr(III) compounds (Cr(OH)3 ; CrOOH; Cr2 O3 ). The so-formed oxides perfectly cover the underlying metal and protect it from further degradation. With increasing temperature, the water content of the oxide decreases and the cristallinity increases. At some 100 C, amorphous CrOOH is the favored species, while at temperatures above some 400 C, crystalline CrOOH (Grimaldiite), and especially Cr2 O3 predominate [166,173,198]. Below the potential, at which Cr(III) is stable, Cr(II) compounds might be formed, but these do not form a protective layer. Consequently, the underlying metal is dissolved rapidly by active dissolution (Cr0 Cr(II)). Increasing the electrochemical potential above that of the passive range might lead again to a high dissolution of the metal or alloy. Chromium then forms the soluble hexavalent species chromate (CrO2 ), hydrogen chromate (HCrO4 ) or chromic 4 acid (H2 CrO4 ), which cannot protect the metal anymore [60,62,166,170,172]. The former insulating effect of the oxide lm is lost. This process of high dissolution is called transpassive dissolution and is observed for many metals. However, titanium, niobium, or tantalum, respectively, do not undergo a transpassive dissolution at room temperature up to anodic potentials of some 100 V. In this case, the oxides are perfect insulators, and also oxygen release caused by water oxidation is suppressed. This leads to their outstanding corrosion performance in highly oxidative environments. It should be noted that due to the changing properties of oxide lms with temperature, observations made for room-temperature corrosion may not be valid at higher temperature in most cases. For example, the number of surface effects in the oxide increases with temperature. An increased number of defects leads to reduced corrosion resistance. Thus, metals with high corrosion resistance at low

Fig. 11. Schematical course of the electrochemical potential for the formation of soluble chromate in acidic and alkaline solution (after Chen et al. [209]) and the solubility of oxygen (after Zoss et al. [208]). Figure after Kritzer et al. [170].

temperatures may dissolve rapidly at increased temperatures. With an increased solubility of oxygen in water, the oxidizing power of the solution increases, and electrochemical processes become more and more important (Fig. 11). As an example, the process of supercritical water oxidation is given and compared with an oxidative treatment in air at the same temperatures, but ambient pressure. Taking into account SCWO-typical oxygen concentrations of around 5 mol/kg and a reaction pressure of 25 MPa, this results in an oxygen partial pressure of 2.5 MPa. This value is approximately two orders of magnitude above the corresponding value for applications in air at ambient pressures. Nevertheless, nickel-base alloys have shown their corrosion resistance against these strong oxidizing conditions due to formation of protecting NiO [168]. However, if ionic chlorides are present under these harsh conditions, and thus, the oxide lm does not protect perfectly any more, corrosion initiation occurs rapidly and the following corrosion rates are high [166]. Additionally, the oxidation of the metal occurs easier at higher temperatures due to the decrease in the protective nature of the oxide layer. The electrochemical potential for the transpassive transformation of chromium and Cr(III) by forming soluble Cr(VI) compounds decreases with increasing temperature [209] (Fig. 12; [169,170,172,173]). This means that chromium can be much more easily oxidized in high-temperature water. Additionally, higher pH-values favor the Cr(III)Cr (VI) transformation. As mentioned above, already low ppm values of

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129

11

3.3.5. Inuence of anions Anions can promote or inhibit the rates of different corrosion processes, according to how they interact with the protective oxide lm layer of the metal. Several possible inuences are: Oxide-lm destruction: this locally restricted form of corrosion is commonly known for the halides chloride, bromide, and iodide, but not uoride, in high-temperature water [166,167]. Under harsher conditions, also other anions like sulde or sulte may lead to a localized destruction of the oxide layer. Such localized corrosion (e.g. pitting corrosion, SCC) is extremely dangerous for technical applications, since its occurrence is stochastic and corrosion rates are high and not linear. Corrosion product dissolution rate: as mentioned before, the rate of diffusion-controlled corrosion reactions is determined by the solubility of the primary corrosion products. For example, the general corrosion of nickel-base alloy 625 at high subcritical temperatures in oxidizing solutions of HCl and HNO3 occurs in the same temperature range, but the corrosion rates differ by an order of magnitude (12; [174]). This can be explained by the higher solubility of Ni(NO3 )2 compared with that of Ni(Cl)2 and the diffusion-controlled character of the general corrosion. Anions as oxidizing agents: especially nitrate is known to be a powerful oxidizing agent in high-temperature water and thus can corrode metals also in the absence of oxygen. Furthermore, ions like sulfate, which are not seen as oxidizing agent at room temperature, may act as strong oxidizers in high-temperature water [210216]. Under these conditions, the thermodynamically favored suldes, sultes, or elemental sulfur can be formed. Thereby, the metals might undergo a fast active dissolution [165]. This dissolution is even worsen by the formation of thin sulfur lms on the dissolving metal surface, observed by Marcus and Protopopoff during the corrosion of chromium and nickel in high-temperature sulfate solutions [217,218]. Oxide-lm supporting by incorporation: several anions are incorporated in the oxide lm of nickel-base alloys and lead to their increased stability. Such modied layers then might possess a drastically reduced solubility leading to secondary passivation.

Fig. 12. Corrosion rates of general corrosion caused by solutions of different acids under similar conditions (Tmax = 500 C; p = 24 MPa; [O2 ] = 0.5 mol/kg; [HCl] = [HNO3 ] = 0.1 mol/kg). The high corrosion rates caused by nitric acid can be explained with the different solubilities of the main corrosion products Ni(NO3 )2 and NiCl2 . The nitrate is by far more soluble. Data after Ref. [174]; reprinted with kindly permission of Kluwer Academic Publishers.

oxygen are sufcient for chromate formation in neutral high-temperature solutions [60,62]. Both, the increased oxidative aggressiveness of oxygen-containing solutions and the reduced resistance against oxidation of chromium result in a high dissolution of the pure metal. Chromium-containing alloys will corrode in similar high corrosion rates, if no other alloy component can form an alternative protective lm. This behavior is the reason, why nickel-base alloys and stainless steels are rapidly corroded in oxidizing acidic high-temperature water of high density. Under these conditions, chromium is dissolved oxidatively, while the oxides of the other two main components, iron and nickel, are dissolved chemically. Molybdenum and tungsten behave chemically similar to chromium and are also dissolved oxidatively by formation of hexavalent species [165,166,172,173]. Due to the lower solubility of chromium and iron oxides compared to that of nickel oxides [188,194], iron, and especially chromium enriched solid corrosion products can be found under such acidic conditions, while Ni is missing [60,62,165,166]. Further, nickel-base alloys and stainless steels behave similar under these conditions their high dissolution rates vary by below one order of magnitude (see, e.g. Ref. [141,163,172,198,199]).

12

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129

Table 3 Inuence of inorganic ions on the corrosion of nickel-base alloys and stainless steels in high-temperature water Ion F Cl ; Br SO2 ; SO2 ; S2 O2 3 4 3 S2 NO 3 CO2 ; PO3 3 4 OH + H
a b

Mode of action Weak complex former Penetrate into & destroy protecting oxide-lm Oxidative in high-temperature water by forming S2 and S0b Reductive in high-temperature waterb Strongly oxidizing; main corrosion products well soluble Low-soluble salts Low-soluble salts Enhanced solubility of protecting oxides

Result Homogeneous corrosion possiblea ; passivating inuence? Strong localized corrosion: pitting and SCCa Strong homogeneous degradation possible Release of H2 possible; SCC possible Strong general corrosion possible Corrosion-inhibition possible Strongly passivating; corrosion-inhibition possible Strong general corrosion possiblea

In the presence of oxidizing compounds. See Refs. [166,210216].

This positive behavior was proven experimentally for carbonate, phosphate, uoride, and hydroxide in high-temperature solutions [56,60,167,175]. It must be the subject to future research, to what extent these anions could inhibit the detrimental character of, e.g. other halides. An overview over the inuence of different inorganic ions on the corrosion in high-temperature water is given in Table 3. 3.3.6. The connection of the corrosion-determining factors The interdependencies between the solution parameters and corrosion are illustrated in Fig. 13 (data after Ref. [170]). Generally, corrosion rates can be directly correlated with the density of the solution. Since density and ionic product show a similar drop in the vicinity of the critical point, a correlation between corrosion

rates and ionic product would lead to a similar result. This attempt rst was made by Kriksunov and Macdonald, who created a suitable model for predicting corrosion of stainless steels and nickel-base alloys in oxidizing HCl solutions as function of density [120]. It has been proven experimentally for many chemical solutions later [161175,186,194]. The dependency of the upper limit of general corrosion from the density drop of water is shown in Fig. 14. As mentioned above, high-temperature water of high density (above 200300 kg/m3 ) still has an ionic character and thus shows high solvency for ionic species. Thus, density inuences directly the dissociation of acids, salts, and bases and the solubility of salts.

Fig. 13. Connection between solution parameters and corrosion. Thickness of arrows indicate importance of pathways. Figure taken from Ref. [170].

Fig. 14. Experimentally found corrosion rates of general corrosion caused by HCl solutions at different pressures (Tmax = 500 C; p = 24/38 MPa; [O2 ] = 0.5 mol/kg; [HCl] 0.05 mol/kg). Note that the lower temperature limit of general corrosion is independent of pressure, while the upper temperature limit strongly increases with pressure (and correlates with the density drop). Data Ref. [172]; reprinted with kindly permission of the National Association of Corrosion Engineers (NACE).

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129

13

The dissociation of acids, salts, and bases correlate directly with the density (see, e.g. Figs. 2 and 10) or the ionic product of water. For example, HCl is completely dissociated by forming H+ and Cl at high densities, while at low densities, exclusively the non-dissociated HCl is present (and the solution becomes neutral). In this case, HCl is completely miscible with the supercritical water. As mentioned above, both H+ and Cl are corrosive species, while HCl itself is relatively harmless at temperatures up to about 600 C. High densities thus create a highly corrosive environment in the presence of acids. The solubility of salts is directly inuenced by solution density. Low-density water can only slowly remove salts in their ionic form. Under these conditions, the energies necessary to break the salt crystal lattice cannot be compensated by hydration energies. Consequently, salts, which are highly soluble at ambient conditions, are nearly insoluble in low-density high-temperature water. Typical values for the solubility of NaCl at different temperatures and pressures (densities) are shown above in Fig. 4. The solubility of Na2 SO4 is even lower by around two orders of magnitude. Both, the protecting oxides on metal and alloys and the primary corrosion products are salts. Enhancing their solubility automatically means enhancing corrosion. In low-density water, a removal of formed ionic compounds by the bulk solution in an open equilibrium process does not occur in remarkable rates. Thus, the formed primary corrosion products can plug the surface and cause a new kind of passivity. In principle, the complete miscibility of oxygen with low-density supercritical water and thus high partial pressure of oxygen should accelerate the cathodic corrosion reaction. On the other hand, the low solubility of ions hinders corrosion processes in the following way: The cathodic reaction at the metal | solution interface follows the equation O2 + 2H2 O + 4e 4OH

Fig. 15. Density range of high corrosion at different temperatures. General corrosion is low at densities below about 200300 kg/m3 . At 300 C, a pressure of already 10 MPa leads to high corrosion rates, while at 500 C, pressures above around 50 MPa are necessary for high corrosion. Increasing the pressure at constant temperature increases the rate of electrochemical corrosion. Note that the shift from no corrosion towards strong corrosion is only sharp at lower temperatures, while there is no clear dividing line at higher temperatures.

However, this reaction leads to an enrichment of negative charged hydroxyl ions at the metal surface, which cannot be dissolved by the non-polar solvent. Thus, a further cathodic reaction cannot occur. As conclusion, corrosion in high-temperature and supercritical water strongly correlates with density or ionic product. Corrosion is high at high densities and

negligibly low at lower densities (density below about 200300 kg/m3 ). Consequently, the generalization that supercritical water is not corrosive at all is true for supercritical water at low, but not for such at high densities. Highly pressurized supercritical water causes comparably high corrosion rates like dense subcritical water (Fig. 14). Fig. 15 illustrates the pressuretemperature regime where high ionic corrosion is expected. For nickel-base alloys, at a temperature of 300 C and pressures above only around 10 MPa (i.e. above the saturation pressure), high rates of corrosion are expected. On the other hand, temperatures of 500 C need minimum pressures of approximately 50 MPa for high rates of ionic corrosion. The density of steam is much lower compared with the density of water above the saturation pressure (see Fig. 15). Consequently, electrochemical corrosion rates caused by steam are by orders of magnitude below those found for high-density water at the same temperature. The mechanism of high-temperature corrosion is explained in Section 3.5. It must be noted that the correlation between the dielectric constant and corrosion does not lead to a similar exact result. This could be explained by the less inuence of the dielectric constant on the solubility of salts in high-temperature water.

14

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129

Stress corrosion cracking (SCC) has been observed in chloride-containing media at high densities [146,166]. It has been shown that the effect of bromide, which is also known for its SCC-inducing behavior, is at least not much worse than that of chloride [167]. Iodide causes pitting at room temperature, but is oxidized rapidly to non-corrosive iodine in the presence of oxygen at higher temperatures [167]. As mentioned above, the most dangerous areas for SCC are the potential transitions ranges between active and passive, or passive and transpassive region, respectively. SCC in the activepassive transition is well known from high-temperature water in power applications, where traces of hydrogen gas might be formed and described in detail elsewhere [86,91,101]. In special cases, SCC is also observed at other potentials. For example, SCC can occur, if pitting corrosion has thinned the reactor walls leading to an increased mechanical stress. However, no observation of SCC in reactors at low-density supercritical solutions has been reported in literature thus far. Recently, Fournier et al. investigated the SCC behavior of high nickel alloys in de-aerated and aerated low-density supercritical water [178]. The authors found a dependency of SCC and oxygen content, but this behavior was also observed in a pure oxygen atmosphere. The main reason for SCC susceptibility are niobium carbides. Nevertheless, the investigations do not cover the high-density region, where SCC occurred predominantly as showed in several screening tests. There are several reasons for the current lack of SCC in low-density water. First, the susceptibility of the materials against corrosion is much higher at other temperatures and higher densities, so if the material fails, it will not fail under low-density supercritical conditions. Second, for SCC to begin, the protective oxide layer has to be destroyed locally, which can occur through the penetration of chloride. In low-density solutions, the concentration of free ionic chlorides and other potentially detrimental anions is extremely low. Consequently, the oxide lm is either attacked generally by chemical dissolution processes or is not attacked at all. Finally, like other ionic corrosion phenomena, SCC requires the removal of dissolved metal ions out of the small crevices. However, the capacity of low-density water to dissolve salts and thus to remove the corrosion products is too low for crack propagation.

Nevertheless, the results of Founier et al. are important for the long-time stability of reactor materials under supercritical water conditions. It must be also the aim of future work to evaluate, whether SCC is a problem in the hydrogen-generating process of supercritical biomass conversion. While hydrogen in trace concentrations extremely accelerates SCC, there are indications that high nickel alloys are much less susceptible for SCC at high hydrogen concentrations [219,220]. Nevertheless, most critical for initiation of hydrogen-induced SCC are the presence of impurities, which are enriched at the grain boundaries of the alloy [86,101,221]. It must be again noted that such an enrichment might occur in-situ during long-time operations at temperatures above 600 C, so that the initial SCC resistance might be much better than the resistance during operation. 3.3.7. Low corrosion pHpotential regions This section should give an overview how to reach pHpotential-regions, in which low corrosion is expected. As mentioned above, at least one of the alloying componentschromium or nickelmust be able to form a protecting oxide. In Fig. 16, the stability islands of chromium and nickel are shown schematically. At an acidic pH and moderate electrochemical potential (state A), chromium protects the alloy. Note that pure nickel would undergo a chemical dissolution at these conditions. An increase of potential at constant pH leads to an instability of both chromium and nickel (state B). Chromium thereby is dissolved oxidatively as chromate, while nickel is dissolved chemically. Experimental investigations performed with pure nickel and

Fig. 16. Stability islands of chromium and nickel. Chromium is more stable against acidic solution, while nickel better tolerates oxidizing conditions. For the reactions leading to instability/stability, see text. E = electrochemical potential.

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129

15

chromium under such acidic, oxidizing conditions showed a fast dissolution of nickel [171]. Chromium was also completely unstable, but dissolution rates were much slower [171]. The mechanism leading to this behavior is described in Section 3.3.4 and illustrated in Fig. 11. By increasing the pH to neutral values at these high electrochemical potentials, a state C can be reached, where nickel is able to protect the alloy. Under these conditions, protecting NiO is formed, which has been found experimentally protecting nickel-base alloys in low-density supercritical water [166,167,172,197]. Note that pure chromium would undergo an electrochemical dissolution at these conditions by forming the well-soluble chromic acid H2 CrO4 . The pH increase can be obtained as follows: an addition of small amounts of a base like NaOH leads to a pH shift into the alkaline direction. Experimental results have shown that under these conditions, nickel-base alloys are not corroded also by oxygen- and chloride-containing high-density solutions [166,177].

A similar pH shift is reached when the subcritical supercritical transition is crossed, e.g. by increasing the temperature, and the density of the water and the dissociation of acids become low. This explains the low corrosion of low-density supercritical water also if acids are present [120,170,172]. Alternatively, the reduction of the electrochemical potential can bring the process back to state A, in which the material is protected by chromium oxides. This could be obtained by the exclusion of oxidizing compounds, or possibly by an electrochemical protection of the reactor wall. Whether such an protection works in practice would have to be the subject of further investigations. 3.3.8. The corrosion mechanism of nickel-base alloys and stainless steels in oxidizing acidic solutions The corrosion mechanism of nickel-base alloys in oxidizing high-temperature HCl solution has been published earlier (see Ref. [166]). This mechanism, which is a summary of the above sections, is shown in Fig. 17. Corrosion in oxidizing sulfuric acid solutions

Fig. 17. Mechanism of the corrosion of nickel-base alloys in oxidizing HCl solution (slightly modied after Ref. [166]). Corrosion in the passive state is pitting corrosion (see Fig. 5), while in the transpassive region, general corrosion (see Fig. 6) occurs. Note that the rates of pitting can be much higher than these of general corrosion, while the whole material loss can be less. Stress corrosion cracking (SCC; see Fig. 8) is observed primarily in the transition region between passive and transpassive region, but can also start at the bottom of pits in the passive region (increased mechanical stress!). In the low-density region, NiO passivates the alloy. Salts or corrosion products dissolved at lower temperatures might precipitate here (esp. NiCl2 ).

16

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129

Table 4 Typical corrosion temperatures and corrosion rates of intergranular corrosion (IC), pitting corrosion, general dissolution and stress corrosion cracking (SCC) observed for alloy 625 in oxidizing aqueous solutions of various acids ([acid] = 0.050.2 mol/kg; [O2 ] = 0.5 mol/kg; p = 24 MPa) Temperature of occurrence IC Pitting > 100 > 150200 Corrosion rate ( m/100 h) 1020 5001000a Occurs in All solutions HCl/O2 NaCl/O2 HBr/O2 HCl/O2 NaCl/O2 HBr/O2 H2 SO4 /O2 HNO3 /O2 HCl/O2 HBr/O2 Inhibitors No OHc CO2 c 3 PO3 c 4 Fd OHc CO2 c 3 PO3 c 4 Fd OHc CO2 d 3 PO3 c 4 Fd

General dissolution

> 250300

500b

SCC

250300

1000a

Data for HCl, H2 SO4 , H3 PO4 or HNO3 , respectively, were taken from Kritzer et al. [see text]. a Corrosion rates not linear. b Rates in sulfuric acid and hydrochloric acid. Values in nitric acid are signicantly higher ( > 1500 c Effect proven. d Effect has to be proven.

m in 100 h).

follows a similar behavior, but in this case, no pitting corrosion in the passive state is observed [165]. Mechanisms explaining the corrosion in oxidizing aqueous solutions of HBr [167], HF [167], HNO3 [174], and H3 PO4 [175] solutions are published elsewhere. The temperatures and corrosion rates of the various forms of corrosion observed for nickel-base alloy 625 is shown in Table 4. 3.3.9. Chemical water treatment strategies in power stations Since heat-transport applications in power station cycles are the most important application of high-temperature and supercritical water to date, in this section, a short overview will be given on the various types of

conventional water treatment to minimize corrosion. These different treatments are summarized in Table 5. Until the early seventies, phosphate treatment was commonly used for subcritical circulation boilers. In this kind of boiler, steam is continuously removed from the boiler system. Since salts are more soluble in the aqueous phase, an enrichment can occur in this phase. This has led to high corrosion rates and sudden breakdown of some reactors in the past. It was found that the concentrated phosphate solutions were able to dissolve the protecting iron oxides of the stainless steels, followed by a rapid oxidation and dissolution of the underlying alloy. Due to the low solubilities of phosphate in low-density water, this strategy cannot be applied for supercritical reactors.

Table 5 Typical concentrations of additions in different water-treatment strategies in power stations (after Ref. [228]) Phosphate treatment Feed water Boiler water pH value [O2 ] [N2 H4 ] N 2 H4 Na2 HPO4 9.5 < 7 ppb 1030 ppb Volatile treatment N2 H4 /NH3 N/A 9.5 < 7 ppb 1030 ppb Oxygen treatment (combined water treatment) O2 /NH3 N/A 89 100 ppb 0

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129

17

To avoid precipitation, the so-called all-volatile treatment was introduced. The key issue of this kind of treatment are the reducing, alkaline conditions caused by an ammonia/hydrazine containing solutions. Unfortunately, it was shown that such a treatment could cause elevated corrosion of stainless steels and nickel-base alloys. The oxide lm on stainless steels in alkaline high-temperature solutions contains of a double-layer with porous Fe3 O4 in the outer and protecting Fe2 O3 in the inner layer [222]. However, under reducing conditions, the formation of the protecting inner lm might be incomplete and the alloy shows elevated corrosion due to active dissolution. In case of nuclear applications, the materials of construction are weakened by high doses of radioactive rays. Further, any metal ions released from these materials carry off the radioactivity and cause radioactive waste. Problems in supercritical boilers arise from the possibility of salt plugging if corrosion products, which are released at lower, subcritical temperatures, deposit at higher temperatures. This again could lead to a plugging of the lines. In the combined water treatment method, the solution is oxidizing through an addition of around 100 ppb oxygen. By addition of ammonia, the pH value is shifted to the weakly alkaline region. Under these conditions, the formation of the protecting iron oxide lm is supported. Consequently, corrosion rates observed in power stations using this kind of treatment, are signicantly lower. However, chromium and molybdenum, both are frequent alloying elements of stainless steels, are not stable under these conditions and transformed oxidatively to soluble chromate and molybdate, respectively. Both metals are transformed to their corresponding soluble acid H2 MeO4 (Me = Cr, Mo) in low-density supercritical water and thus can pass through the reactor without causing plugging problems. Consequently, a small amount of both compounds is released during the operationespecially during warm-up procedures. In subcritical once-through reactors, the presence of both NH3 /O2 does not cause problems, since oxidation of ammonia is slow. In contrast, the oxidation rate of ammonia is strongly accelerated under supercritical conditions. The main oxidation product is nitrogen gas, but up to several percent of N2 O might be formed. Both reactions lead to ammonia and oxygen consumption.

Do to the consumption of ammonia, the pH is lowered to less alkaline values. Since iron oxides have their highest stability at a pH around 10, this could lead to an enhanced dissolution of the oxide layer. However, as ammonia is present in clearly higher concentrations than oxygen, this effect should be of minor relevance. Nevertheless, the under-stoichiometric reaction could lead to favored formation of N2 O. Since oxygen is present clearly below the stoechiometric amount for ammonia oxidation, it can be assumed that the consumption of oxygen is close to 100%. It must be kept in mind that oxygen was introduced to support the protective character of the iron oxide lm, so after its consumption, corrosion is enhanced. It should be clear that none of the present chemical treatment strategies can eliminate all corrosion problems. An alternative treatment could be the addition of small amounts of NaOH, or KOH instead of NH3 as PH stabilizer, respectively, which cannot be oxidized by the solution and thus might lead to acceptable corrosion rates inside the reactor. 3.4. Material parameters 3.4.1. Stainless steels and nickel-base alloys: alloy composition It is obvious that the composition of an alloy has a large inuence on its corrosion resistance. Chromium as alloying element improves the resistance against acidic and oxidizing media and reduces pitting corrosion, nickel improves the corrosion behavior in alkaline environments. Molybdenum causes a passivating effect at low, reducing potentials, at which other alloys show active dissolution. Consequently, the rst choice of an alloy used in highly oxidizing, acidic solutions at moderate high temperatures is one with high chromium content. However, as mentioned above, the tendency towards chromate formation increases with temperature and thus, chromium loses its protective effect. As shown in different studies, the conventional high-chromium nickel- and iron-based alloys show a similarly high corrosion rate under these conditions. To found a stable alloy, the alloying elements have to form an oxide layer which completely covers the alloy. The oxides of niobium, tantalum, aluminum, zirconium and yttrium are known to be chemically

18

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129

stable in oxidizing high-temperature water. It could be an item of future research to evaluate, whether alloys containing these elements in concentrations, which are high enough for oxide lm formation, show the desired corrosion resistance. Additionally, solubility screening tests with further oxides, e.g. these of the rare earth elements, or with the metals themselves, and nally with model alloys containing these metals, could also lead to an improved further strategy for the development of a stable alloy. In neutral or weakly alkaline oxidizing high temperature solution, as well as in low-density supercritical solution, nickel forms the protective layer, while chromium is unstable. It must be the item of future research to prove, if chromium is need as alloying component for reactor materials under these conditions at all. The presence of chromium always bears the risk of a release of chromate, which is crucible in the efuent solution. In reducing high-temperature water, e.g. during the reductive conversion of biomass, a further increase of molybdenum in the alloy composition might improve corrosion resistance, if chromium fails. 3.4.2. Stainless steels and nickel-base alloys: heat treatment and surface condition Most commonly used reactor materials are stainless steels and nickel base alloys. For these alloys, material parameters have a high inuence on general corrosion, and especially pitting corrosion and SCC. The presence of non-metallic inclusions in the metal matrix were found to be starting points of pitting and general corrosion [68,171]. Most detrimental inclusions are manganese sulde and titanium nitride. These inclusions lead to weak points of the oxide layer, and thus to preferential starting points of corrosion (see Fig. 18). Also, the composition of grain boundaries of the alloy has a remarkable effect on corrosion (see sections above). Depending on the kind heat-treatment of the alloy, the grain boundaries may be depleted of one or more of the alloy components. Most detrimental from corrosion side is a chromium depletion in areas of some hundreds of nanometers around the grain boundaries, at which chromium carbides are formed. In the neighbored regions of these carbides, chromium concentrations can decrease to values of 50% of the concentration in the bulk alloy [81,82,84,86]. This

Fig. 18. Micro-pit with a diameter of approximately 2 m starting at an inclusion under transpassive conditions (nickel-base alloy 625; polished surface, [HCl] = 0.05 mol/kg; [O2 ] = 0.48 mol/kg; T = 350 C; p = 24 MPa, t = 0.75 h; see Ref. [171]. Corrosion spreads out over the surface rapidly leading to the shallow pits shown in Fig. 6). Reprinted with kindly permission of the National Association of Corrosion Engineers (NACE).

leads to a less noble alloy in small restricted areas. Additionally, impurities like phosphorus or sulfur may become concentrated in the neighboring areas [85]. Non-metallic inclusions, grain boundary depletion, and strong enrichment of non-metallic contamination can form and disappear in certain temperature zones. At temperatures above some 900 C, depending on the alloy, the solubility of inclusions and non-metallic forms of contamination in the bulk metal matrix is high enough, so the formation of separated phases does not occur. At temperatures below some 600 C, the migration velocity of the impurities towards the grain boundaries and the formation of these phases is too slow, and thus, the time frame, during they will occur, exceed some 10,000 h. As a result, the most detrimental temperature region is between about 600 and 900 C, where the solubility of the impurities is relatively low, but their migration to the grain boundaries and precipitation is fast enough to happen within certain times. These impurities are known to be responsible for intergranular attack and stress corrosion cracking (see section above). With the shifting of the process temperatures of either SCWO and supercritical biomass conversion into these temperature ranges, also the risk for impurity-induced SCC increases. It should be an object of future research to evaluate the effect

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129

19

The surface condition also has an inuence on corrosion behavior. Polished surfaces are known to be more resistant to corrosion than non-polished ones, since the number of starting points of corrosion is lower [171]. However, after corrosion once has started, the corrosion rate of polished surfaces might be signicantly higher, since the total anodic current available for metal dissolution is distributed among a lower number of parallel growing pits resulting in an accelerated growth rate of the individual pit [171]. 3.4.3. Other materials Different other reactor materials other than stainless steels or nickel-base alloys have been suggested for supercritical water applications. For an literature overview over the mechanisms of these materials, see (Table 6). Such materials generally have lower mechanical strength and thus, need special reactor constructions for their use in supercritical water applications. Among them were noble metals like gold and platinum, the valve metals titanium, niobium and tantalum and different oxide ceramic materials. Gold, platinum, and titanium have been successfully used as liners inside a stainless steel pressure tube [141,185]. Ceramics can be applied in form of a swimming tube inside an outer high-grade alloy pressure tube [155]. In such reactors, the outer tube is only in contact with pressure-transmitting de-ionized water, while corrosive reactions can be performed at the same pressure inside the ceramic tube.

Fig. 19. A series of Micro-pits starting at grain boundaries also under transpassive conditions (nickel-base alloy 625; polished surface, [HCl] = 0.05 mol/kg; [O2 ] = 0.48 mol/kg; T = 350 C; p = 24 MPa, t = 0.75 h; see Ref. [171]. Reprinted with kindly permission of the National Association of Corrosion Engineers (NACE).

of the purity of the alloy on long-term stability of a SCWO/supercritical biomass conversion reactor. Also, the very rst step of general corrosion starts at inclusions, or grain boundaries, respectively, and leads to micro-pits, which is shown in Fig. 19. It must be the subject of future work, whether such precipitations can be formed or worsen under high-temperature supercritical water applications leading to the occurrence IC or SCC at longer operation times, or if alloys with higher purity are required under these conditions.

Table 6 Literature data for the corrosion and corrosion mechanisms of different reactor materials in oxidizing supercirtical solutions of different species Material Stainless Steels Nickel base alloys Corroding agent and Refs. HCl [119,126,127,133,139,151,163,196]; H2 SO4 [185,194]; NaCl [195] HCl [127,129,131,139,141,150,151,163,166,173,179,180,183,184,192,196,197]; H2 SO4 [165,173,185,194]; H3 PO4 [173,175]; HNO3 [174]; HF [167]; HBr [167]; NaCl [166,173,183,184]; Na2 SO4 [165,173]; NaOH [169,173] HCl [182,187,193] HCl [171,173,183,187,193] HCl [171,173,183,187,193] HCl [171,173,183] HCl [131,142,157,159,183,190,200]; H2 SO4 [159,186]; H3 PO4 [159,186]; NaCl [159] HCl [176]; H2 SO4 [176] HCl [156,190]; H2 SO4 [156]; H3 PO4 [156] HCl [140,141,190] HCl [153] HCl [153,155]; H2 SO4 [155] H3 PO4 [155]; HF [167]; HBr [167]

Iron Chromium Nickel Molybdenum Titanium Niobium Tantalum Gold and platinum non-Oxide ceramics Oxide ceramics

20

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129

3.5. Exceptions and cautions In the following section, exceptions and cautions are described, which cannot be classied by the previous sections. These show that corrosion in high-temperature aqueous solutions can still be unpredictable. The presence of high concentrations of NaOH in low-density supercritical solutions leads to the formation of a liquid NaOH phase. This phase heavily attacks, e.g. alumina ceramics [155], which are chemically dissolved. In presence of oxygen, most metallic materials are attacked [169,170,172]. Under these circumstances, the formed metal oxides like NiO are dissolved rapidly in the liquid melt and thus, passivation cannot occur [172]. However, in the absence of oxygen, nickel-base alloys are immune [173]. Niobium and tantalum, which show an excellent corrosion resistance in high-density subcritical solutions, surprisingly show a fast corrosion at higher temperature, although their oxides are chemically stable [156,176]. The reason for this unexpected behavior was found in a phase transformation of the protective oxide layer: At lower temperatures, an amorphous, protective oxide layer is present, which is converted to a non-protective crystalline one at higher temperatures. The metals are oxidized completely in short times and only the stable oxides remain. In special cases, the unexpected occurrence of new phases creates a highly corrosive environment. While, e.g. an oxidizing solution ([O2 ] = 0.5 mol/kg; T = 430470 C; p = 24 MPa) containing 0.1 mol/kg phosphoric acid causes almost no corrosion of nickel-base alloys, an increase of the acid concentration to 0.2 mol/kg leads to deadly high corrosion rates of half a millimeter per hour at the upper side of the reactor. The corrosion mechanism describes this behavior with the immediate presence of nickel(III) phosphates, which possess a melting point lower than the process temperature and thus are removed rapidly [175]. At temperatures above about 600700 C, another mechanism of corrosion, the so-called high-temperature corrosion may occur. At these temperatures, the common reactor materials iron, nickel or chromium begin to form volatile corrosion products in the presence of acids and salts such as NiCl2 or NiBr2 . The corrosion products can be removed quickly or even melt on

the surface leading to high general corrosion rates. Although these reactions contain oxidation processes of the metals, they are not seen as electrochemical reactions. Reviews dealing with the process of high-temperature corrosion are published [223228].

4. When and where is the corrosion low? The text above might lead to the impression that corrosion is always high in high-density, hightemperature water. However, it has been shown that corrosion of nickel-base alloys and stainless steels is low under the following conditions: At low concentrations of salts and especially certain acids, corrosion rates are low (also in the presence of high concentrations of oxygen). All strong acids (e.g. HCl, H2 SO4 , HNO3 , HBr)also in tracesshould be avoided. The presence of bases increases the stability of protective iron- and nickel oxides and thus has a corrosion-inhibitive effect. In the presence of phosphate, carbonate or uoride, these ions are incorporated in the oxide lm leading to a protective layer. This mechanism might also work in the presence of aggressive ions like chloride, but this assumption has to be proven in future experimental work. Tolerance of the metal surface against chloride or bromide in the presence of a certain passivating species has to be evaluated. It has been shown that some materials might be corrosion resistant in cases, where high corrosion is observed for other materials. For example, titanium is resistant against oxidizing high-temperature solutions of HCl also at high pressures [159,187], where nickel-base alloys fail. On the other hand, sulfuric acid attacks titanium at conditions, where nickel-base alloys are immune [187]. Therefore, it is absolutely necessary to know the composition of the attacking solution. For each solution, an optimized reactor material can be chosen. It must be critically mentioned that all corrosion tests that have been performed for SCWO applications have lasted not more than several hundreds hours. This is mainly because of the highly corrosive environment that lead to early failure. Long-time tests are necessary to obtain information about the corrosion

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129

21

resistance for power generating systems at typical operation times of months to years. It must also be mentioned that salt-doping (esp. sodium phosphates) of high-temperature water, which is common for subcritical power systems to reduce its aggressiveness, will lead to severe salt precipitation problems. The use of a combination of oxygen, to passivate the material, and ammonia, to adjust the pH, in the so-called oxygen treatment might cause problems since both components are able to react with each other. In every case, most problematic in corrosion considerations is the high-temperature, high-density regime.

5. Outlook With the increased number of applications of high-temperature and supercritical water, the need for a corrosion-resistant reactor material increases. Unfortunately, there is still a lack in basic physical and chemical data. Without these data, supercritical water may be a dangerous medium in some cases. Especially lacking are: Research on the solubility of oxides and salts: the solubility of inorganic compounds (e.g. oxides) in high-temperature water under various pH and electrochemical potential conditions. A lot of work in this eld has been done by geochemical researchers, but these investigations mainly are focused on pressures above some 100 MPa, which is out of scope for most supercritical water applications. Nevertheless, a collaboration between geologists, corrosion scientists and supercritical water users should be very fruitful for a better understanding of dissolution behavior and thus corrosion occurring in high-temperature, high-pressure water. Research on the phase behavior of supercritical aqueous systems. Few investigations exist for systems containing more than two compounds. But such multi-phase systems are present if supercritical water acts as a reaction medium. Since not all physical phenomena are understood in detail, extrapolations and predictions obtained from experiments at lower temperatures are sometimes inaccurate. Research on the possibility of inhibit corrosion by an addition of certain compounds (see Section 4). Such corrosion inhibitors could offer the

possibility to tolerate higher concentrations of corrosive anions inside the reactors as a preferable technical alternative over a dilution of the solutions. Research on the inuence of metal purity on stress corrosion cracking. This item is important for applications, where the temperature of the reactor materials are in a range above the sensibilization temperature of some 600 C, and especially, if hydrogen is present. So far, the corrosion item in the process of biomass conversion, where hydrogen is released, has not reached much attention. Corrosion tests under realistic conditions. Such tests are necessary and cannot be replaced by theoretical work. It has been shown in the past that calculations dealing with the transition from subcritical to supercritical water do not always account for the drastic change of physical and chemical properties. Tests should also consider the mechanical inuence. Lots of corrosion tests done in the past have been performed with coupon samples and thus do not take into account stress corrosion phenomenaalthough SCC is one of the most dangerous forms of corrosion for technical applications. Embrittlement can occur suddenly after long operation times. Until extensive long-time studies have not been performed, a too optimistic attitude rather hinders supercritical water applications.

Acknowledgements The experimental work was carried out during my PhD at the Institut fr Technische Chemie at the Forschungszentrum Karlsruhe, between 1995 and 1998. I would like to thank especially Gnter Franz, Claus Friedrich, Wilhelm Habicht, Michael Schacht and Basil Kanellakopoulos for the help and fruitful discussions. Further, I would like to thank Nikolaos Boukis and Eckhard Dinjus for the opportunity of performing this work in their group/institute and the support they gave. Finally, I want to acknowledge the National Association of Corrosion Engineers (NACE), Houston, TX, USA, for honouring my work with the A.B. Campbell Award 2000. Parts of this work were presented at the First International Symposium on Supercritical-cooled Reactors (SCR-2000), Tokyo, 2000.

22

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129 E. Kiran, P.G. Debenedetti, C.J. Peters (Eds.), Supercritical uidsfundamentals and applications, NATO Science Series E, Vol. 366, Kluwer Publishers, Dodrecht, 2000, p. 345. T. Mizuno, Properties and application of supercritical water, Zairyo-to-Kankyo 47 (1998) 298. B. Kuhlmann, E.M. Arnett, M. Siskin, Classical organic reactions in pure superheated water, J. Org. Chem. 59 (1994) 3098. P.E. Savage, S. Gopalan, T.I. Mizan, C.J. Martino, E.E. Brock, Reactions at supercritical conditions: applications and fundamentals, AICE J. 41 (1995) 1723. H. Schmieder, N. Dahmen, J. Schn, G. Wiegand, Industrial and environmental applications of supercritical uids, in: R. van Eldik, C.D. Hubbard (Eds.), Chemistry under Extreme or Non-Classical Conditions, Spektrum, Heidelberg, 1996, p. 273. T. Adschiri, Material synthesis in supercritical water, Zairyo-to-Kankyo 49 (2000) 126. P.E. Savage, Organic chemical reactions in supercritical water, Chem. Rev. 99 (1999) 603. P.G. Jessop, T. Ikariya, R. Noyori, Homogeneous catalysis in supercritical uids, Chem. Rev. 99 (1999) 475. A. Baiker, Supercritical uids in heterogeneous catalysis, Chem. Rev. 99 (1999) 453. J.A. Darr, M. Poliakoff, New directions in inorganic and metalorganic coordination chemistry in supercritical uids, Chem. Rev. 99 (1999) 495. D. Broll, C. Kaul, A. Kramer, P. Krammer, T. Richter, M. Jung, H. Vogel, P. Zehner, Chemistry in supercritical water, Angew. Chem. Int. Ed. Engl. 38 (1999) 2999. A. Rabenau, Die Rolle der Hydrothermalsynthese in der prperativen Chemie, Angew. Chem. 97 (1985) 1017. R.W. Shaw, N. Dahmen, Destruction of toxic organic materials using supercritical water oxidation: current state of the technology, in: E. Kiran, P.G. Debenedetti, C.J. Peters (Eds.), Supercritical FluidsFundamentals and Applications, NATO Science Series E, Vol. 366, Kluwer Publishers, Dodrecht, 2000, p. 425. H. Schmieder, E. Dinjus, Supercritical water oxidation: state of the art, Chem. Eng. Technol. 22 (1999) 903. J.W. Tester, J.A. Cline, Hydrolysis and oxidation in subcritical and supercritical water: connecting process engineering science to molecular interactions, Corrosion 55 (1999) 1088. H.J. Bleyl, J. Abeln, N. Boukis, H. Goldacker, M. Kluth, A. Kruse, G. Petrich, H. Schmieder, G. Wiegand, Hazardous waste disposal by supercritical uids, Sep. Sci. Technol. 32 (1997) 459. S.F. Rice, R.R. Steeper, Oxidation rates of common organic compounds in supercritical water, J. Hazardous Mater. 59 (1998) 261. P. Kritzer, E. Dinjus, An assessment of supercritical water oxidation (SCWO)existing problems, possible solutions and new reactor concepts, Chem. Eng. J. 83 (2000) 207. G. Limousin, C. Joussot-Dubien, Y. Garrabos, S. Sarrade, Modelling of dodecane hydrothermal oxidation in a tubular

References
[1] E.U. Franck, berkritisches Wasser als elektrolytisches Lsungsmittel, Angew. Chem. 73 (1961) 309. [2] E.U. Franck, Physicochemical properties of supercritical solvents, Ber. Bunsenges Phys. Chem. 88 (1984) 820. [3] E.U. Franck, Fluids at high pressures and temperatures, J. Chem. Thermodynamics 19 (1987) 225. [4] R.W. Shaw, T.B. Brill, A.A. Clifford, C.A. Eckert, E.U. Franck, Supercritical watera medium for chemistry, Chem. Eng. News 69 (51) (1991) 26. [5] E.U. Franck, Supercritical water and other uidsa historical perspective, in: E. Kiran, P.G. Debenedetti, C.J. Peters (Eds.), Supercritical FluidsFundamentals and Applications, NATO Science Series E, Vol. 366, Kluwer Publishers, Dodrecht, 2000, p. 307. [6] ASME Steam Tables, 6th ed., The American Society of Mechanical Engineers, New York, 1992 (computer program). [7] T.M. Seward, E.U. Franck, The system hydrogenwater up to 440 C and 2500 bar pressure, Ber. Bunsenges Phys. Chem. 85 (1981) 2. [8] M.L. Japas, E.U. Franck, High pressure phase equilibria and PVT-data of the wateroxygen system including waterair to 673 K and 250 MPa, Ber. Bunsenges Phys. Chem. 89 (1985) 1268. [9] K. Brllos, K. Peter, G.M. Schneider, Fluide Mischsysteme unter hohem Druck. Phasengleichgewichte und kritische Erscheinungen in den binren systemen cyclohexanH2 O, n-heptanH2 O, biphenylH2 O und benzol-D2 O bis 420 C und 3000 bar, Ber. Bunsenges Phys. Chem. 74 (1970) 682. [10] Z. Alwani, G. Schneider, Druckeinu auf die Entmischung ssiger Systeme. VI. Phasengleichgewichte und kritische Erscheinungen im System BenzolH2 O zwischen 250 und 368 C bis 3700 bar, Ber. Bunsenges Phys. Chem. 71 (1967) 633. [11] J.F. Connolly, Solubility of hydrocarbons in water near the critical temperature, J. Chem. Eng. Data 11 (1966) 13. [12] G.W. Morey, The solubility of solids in gases, Econ. Geol. 62 (1957) 225. [13] O.I. Martynova, Solubility of inorganic compounds in subcritical and supercritical water, in: D. Jones, G. de, J. Slater, R.W. Staehle (Eds.), NACE-4: High Temperature High Pressure Electrochemistry in Aqueous Solutions, NACE National Association of Corrosion Engineers, Houston, TX, 1976, p. 131. [14] K. Arai, T. Adschiri, Importance of phase equilibria for understanding supercritical uid environments, Fluid Phase Equilibria 158160 (1999) 673. [15] K.P. Johnston, P.J. Rossky, Solution chemistry in supercritical water: spectroscopy and simulation, in: E. Kiran, P.G. Debenedetti, C.J. Peters (Eds.), Supercritical FluidsFundamentals and Applications, NATO Science Series E, Vol. 366, Kluwer Publishers, Dodrecht, 2000, p. 323. [16] A.A. Chialvo, P.T. Cummings, Molecular simulation and modeling of supercritical water and aqueous solutions, in: [17] [18]

[19]

[20]

[21] [22] [23] [24] [25]

[26]

[27] [28]

[29] [30]

[31]

[32]

[33]

[34]

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129 coiled reactor, in: Proceedings of the 6th Meeting on Supercritical Fluids, Nottingham, UK, Institute National Polytechnique De Lorraine (INPL), Vandoeuvre Cedex, France, 1999, p. 471. A. Gidner, L.B. Stenmark, K.M. Carlsson, Treatment of different wastes by supercritical water oxidation, Proceedings of the 20th IT3 Conference, 2001, Philadelphia, PA. R.L. Smith Jr, P. Atmaji, Y. Hakuta, M. Kawaguchi, T. Adschiri, K. Arai, Recovery of metals from simulated high-level liquid waste with hydrothermal crystallization, J. Supercritical Fluids 11 (1997) 103. H. Schmieder, J. Abeln, N. Boukis, N. Dinjus, E. Kruse, A. Kluth, M. Petrich, G. Sadri, M. Schacht, Hydrothermal gasication of biomass and organic wastes, J. Supercritical Fluids 17 (2000) 145. H.R. Appell, Y.C. Fu, S. Friedman, P.M. Yavorsky, I. Wender, Converting Organic Wastes to Oil. A Replenishable Energy Source, Report of Investigations 7560, Bureau of Mines, Washington, DC, 1971, 20 pp. X. Xu, Y. Matsumura, J. Stenberg, M.J. Antal Jr, Carboncatalyzed gasication of organic feedstocks in supercritical water, Ind. Eng. Chem. Res. 35 (1996) 2522. T. Minowa, F. Zhen, T. Ogi, Cellulose decomposition in hot-compressed water with alkali or nickel catalyst, J. Supercritical Fluids 13 (1998) 253. M. Sasaki, B. Kabyemela, R. Malaluan, S. Hirose, N. Takeda, T. Adschiri, K. Arai, Cellulose hydrolysis in subcritical and supercritical water, J. Supercritical Fluids 13 (1998) 261. S. Hayashi, K. Nomaguchi, T. Okusawa, O. Yokomizo, Y. Ishagaki, H. Ishimaru, An energy balance study of a newly developed recycling system for waste plastics, Papers of Ship Res. Inst. 35 (1998) 133. M. Poliakoff, J.A. Darr, T. Ikehans, A. Cabanas, The synthesis of crystalline metal oxide particles in sub- and supercritical water mixtures, in: Proceedings of the 6th Meeting on Supercritical Fluids, Nottingham, UK, Institute National Polytechnique De Lorraine (INPL), Vandoeuvre Cedex, France, 1999, p. 483. M.M. Hoffmann, J.S. Young, J.L. Fulton, Unusual dysprosium ceramic nano-ber growth in a supercritical aqueous solution, J. Mater. Sci. 35 (2000) 4177. E. Tani, M. Yoshimura, S. Somiya, Formation of ultrane tetragonal ZrO2 powder under hydrothermal conditions, J. Am. Ceram. Soc. 65 (1983) 11. D.W. Matson, R.D. Smith, Supercritical uid technologies for ceramic-processing applications, J. Am. Ceram. Soc. 72 (1989) 871. G. Okamoto, Passive lm of 18-8 stainless steel structure and its function, Corrosion Sci. 13 (1973) 471. G. Bellanger, J.J. Rameau, Behaviour of hastelloy C22 steel in sulphate solutions at pH 3 and low temperatures, J. Mater. Sci. 31 (1996) 2097. P.E. Manning, D.J. Duquette, The effect of temperature (25289 C) on pit initiation in single phase and Duplex 304L stainless steels in 100 ppm Cl Solution, Corrosion Sci. 20 (1980) 597.

23

[35]

[36]

[37]

[38]

[39]

[40]

[41]

[42]

[43]

[44]

[45]

[46]

[47] [48]

[49]

[50] J.H. Wang, C.C. Su, Z. Szklarska-Smialowska, Effects of Cl concentration and temperature on pitting of AISI stainless steel, Corrosion 44 (1988) 732. [51] R.M. Carranza, M.G. Alvarez, The effect of temperature on the passive lm properties and pitting behaviour of a FeCrNi alloy, Corrosion Sci. 38 (1996) 909. [52] T. Fujii, Electrochemical study on the corrosion behavior of metals and alloys in aqueous solutions at high temperature and high pressure, Trans. Nat. Res. Inst. Metals 18 (1976) 101. [53] A.M. Olmedo, M. Villegas, M.G. Alvarez, Corrosion behaviour of alloy 800 in high-temperature aqueous solutions: electrochemical studies, J. Nucl. Mater. 229 (1996) 102. [54] J. Hickling, N. Wieling, Electrochemical investigations of the resistance of Inconel 600, Incoloy 800, and type 347 stainless steel to pitting corrosion in faulted PWR secondary water at 150250 C, Corrosion 37 (1981) 147. [55] L.F. Lin, G. Cragnolino, Z. Szklarska-Smialowska, D.D. Macdonald, Stress corrosion cracking of sensitized type 304 stainless steel in high temperature chloride solutions, Corrosion 37 (1981) 616. [56] W.F. Bogaerts, A.A. Van Haute, M.J. Brabers, Relative critical potentials for pitting corrosion of 304 stainless steel, Incoloy 800 and Inconel 600 in alkaline high-temperature aqueous solutions, J. Nucl. Mater. 115 (1983) 339. [57] G. Cragnolino, A review of pitting corrosion in hightemperature aqueous solutions, in: H. Isaacs (Ed.), Advances in Localized Corrosion, International Corrosion Conference Series, NACE National Association of Corrosion Engineers, Houston, TX, 1990, p. 413. [58] M. Karaminezhaad-Ranjbar, J. Mankowski, D.D. Mac donald, Pitting corrosion of Inconel 600 in high temperature chloride solution under controlled hydrodynamic conditions, Corrosion 41 (1985) 197. [59] J.R. Park, Z. Szklarska-Smialowska, Pitting corrosion of Inconel 600 in high-temperature water containing CuCl2 , Corrosion 41 (1985) 665. [60] W.F. Bogaerts, C. Bettendorf, Electrochemistry and Corrosion of Alloys in High-Temperature Water, EPRIReport NP-4705, Electric Power Research Institute, Palo Alto, CA, 1986. [61] Z. Szklarska-Smialowska, D. Grimes, The kinetics of pit growth on alloy 600 in chloride solutions at high temperatures, J. Park, Corrosion Sci. 27 (1987) 859. [62] W.F. Bogaerts, C. Bettendorf, High-Temperature Electrochemistry and Corrosion, EPRI-Report NP-5863, Electric Power Research Institute, Palo Alto, CA, 1988. [63] J. Postlethwaite, R.J. Scoular, M.H. Dobbin, Localized corrosion of molybdenum-bearing nickel alloys in chloride solutions, Corrosion 44 (1988) 199. [64] B. Stellwag, W. Beyer, N. Wieling, Electrode potential monitoring in hot water systemsa method to identify critical corrosion conditions, Dechema Monographs 101 (1986) 17.

24

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129 [82] G.S. Was, Grain boundary chemistry and intergranular fracture in austenitic nickel-base alloys, Mater. Sci. Forum 46 (1989) 335. [83] R.L. Cowan, C.S. Tedmon, Intergranular corrosion of iron nickelchromium alloys, in: M.G. Fontana, R.W. Staehle (Eds.), Advances in Corrosion Science and Technology, Vol. 3, Plenum Press, New York, 1973, p. 293. [84] C.L. Briant, S.K. Banerji, Intergranular failure in steel: the role of grain-boundary composition, Int. Met. Rev. 4 (1978) 164. [85] C.L. Briant, The effects of sulfur and phosphorus on the intergranular corrosion of 304 stainless steel, Corrosion 36 (1980) 497. [86] R.C. Newman, Stresscorrosion cracking mechanisms, in: P. Marcus, J. Oudar (Eds.), Corrosion Mechanisms in Theory and Practice, Marcel Dekker, New York, 1995, p. 311. [87] R.H. Jones, Stress Corrosion Cracking, ASM International, Materials Park, OH, 1992, 450 pp. [88] P.L. Andresen, Effects of temperature on crack growth rate in sensitized type 304 stainless steel and alloy 600, Corrosion 92, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 89, 1992, 18 pp. [89] C. Bombara, Review of surface factors in stress corrosion cracking of alloys, Metallurg. Sci. Technol. 2 (1984) 54. [90] J.P. Carter, S.D. Cramer, Materials of Construction for High-Salinity Geothermal Brines, United States Bureau of Mines, Report of Investigations no. 9402, 1992. [91] A.J. Sedricks, Stresscorrosion cracking of stainless steels and nickel alloys, J. Inst. Met. 101 (1973) 225. [92] H.R. Copson, G. Economy, Effect of some environmental conditions on stress corrosion behavior of NiCrFe alloys in pressurized water, Corrosion 24 (1968) 55. [93] J. Kolts, N. Sridhar, Temperature effects in localized corrosion, in: R.C. Scarberry (Ed.), Corrosion of NickelBase Alloys, American Society for Metals, Metals Park, OH, 1985, p. 191. [94] K.A. Lichi, H. Bijnen, P.G. McIlhone, A comparison of pitting corrosion and stress corrosion cracking characteristics of some engineering alloys, Met. Australasia 16 (5) (1984) 4. [95] L.F. Lin, G. Cragnolino, Z. Szklarska-Smialowska, D.D. Macdonald, Stress corrosion cracking of sensitized type 304 stainless steel in high temperature chloride solutions, Corrosion 37 (1981) 616. [96] L.G. Ljungberg, D. Cubicciotti, Effect of water impurities on stress corrosion cracking in a boiling water reactor, Corrosion 41 (1985) 290. [97] H.E. Hnninen, Inuence of metallurgical variables on environment-sensitive cracking of austenitic alloys, Int. Met. Rev. 3 (1979) 85. [98] D.D. Macdonald, On the modeling of stress corrosion cracking in iron and nickel base alloys in high temperature aqueous environments, Corrosion Sci. 38 (1996) 1003. [99] J.A., Van Kirk, Nelson Stress Corrosion Cracking Evaluation of Alloys in the Accelerated Wet Ox Environment, Corrosion 79, National Association of Corrosion Engineers, Houston, TX, Paper 43, 1979, 28 pp.

[65] B. Stellwag, Pitting resistance of alloy 800 as a function of temperature and prelming in high-temperature water, Corrosion 53 (1997) 120. [66] H. Yashiro, K. Tanno, S. Koshiyama, K. Akashi, Critical pitting potentials for type 304 stainless steel in hightemperature chloride solutions, Corrosion 52 (1996) 109. [67] A. Gebert, F. Schneider, K. Mummert, Inuence of oxide structure on the elevated temperature pitting behaviour of FeCrNi alloys, Nucl. Eng. Des. 174 (1997) 327. [68] Z. Szklarska-Smialowska, Pitting Corrosion of Metals, NACE National Association of Corrosion Engineers, Houston, TX, 1986, 430 pp. [69] J. Kruger, Breakdown of protective lms, in: E.L. Gaden (Ed.), Perspectives on Corrosion, AIChE Symp. Ser. No. 278, Vol. 86, American Institute of Chemical Engineers, New York, 1990, p. 14. [70] G.S. Frankel, Pitting corrosion of metalsa review of the critical factors, J. Electrochem. Soc. 145 (1998) 2186. [71] M.G.S. Ferreira, A.M.P. Simoes, Passivation and localized corrosion, in: M.G.S. Ferriera, C.A. Melendres (Eds.), Electrochemical and Optical Techniques for the Study and Monitoring of Metallic Corrosion, Kluwer Academic Publishers, Dordrecht, 1991, p. 485. [72] N. Sato, The stability of pitting dissolution of metals in aqueous solution, J. Electrochem. Soc. 129 (1982) 260. [73] N. Sato, Toward a more fundamental understanding of corrosion processes, Corrosion 45 (1989) 354. [74] N. Sato, The stability of localized corrosion, Corrosion Sci. 37 (1995) 1947. [75] H.H. Strehblow, Breakdown of passivity and localized corrosion: theoretical concepts and fundamental experimental results, Werkst. Korros. 35 (1984) 437. [76] H.H. Strehblow, Mechanisms of pitting corrosion, in: P. Marcus, J. Oudar (Eds.), Corrosion Mechanisms in Theory and Practice, Marcel Dekker, New York, 1995, p. 201. [77] M. Jayalakshimi, V.S. Muralidharan, Empirical and deterministic models of pitting corrosionan overview, Corrosion Rev. 14 (1996) 375. [78] J.R. Galvele Present state of understanding of the breakdown of passivity and repassivation, in: R.P. Frankenthal, J. Kruger (Eds.), Passivity of Metals, Proceedings of the 4th International Symposium on Passivity, Warrenton, VA, 1977, p. 285. [79] G. Cragnolino, A review of pitting corrosion in hightemperature aqueous solutions, in: H.: Isaacs (Ed.), Advances in: Localized Corrosion, International Corrosion Conference Series, NACE National Association of Corrosion Engineers, Houston, TX, 1990, p. 413. [80] S. Abe, M. Kaneko, Corrosion of grain boundaries in ironchromiumnickel alloys, in: S.M. Bruemmer, E.I. Meletis, R.H. Jones, W.W. Gerberich, F.P. Ford, R.W. Staehle (Eds.), Parkins Symposium on Fundamental Aspects on Stress Corrosion Cracking, The Minerals, Metals and Materials Society, Warrendale, PA, 1992, p. 389. [81] S.M. Bruemmer, Grain boundary chemistry and intergranular failure of austenitic stainless steels, Mater. Sci. Forum 46 (1989) 309.

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129 [100] R.B. Rebak, Z. Szklarska-Smialowska, The mechanism of stress corrosion cracking of alloy 600 in high temperature water, Corrosion Sci. 38 (1996) 971. [101] S.M. Bruemmer, G.S. Was, Microstructural and microchemical mechanisms controlling intergranular stress corrosion cracking in light-water-reactor systems, J. Nucl. Mater. 216 (1994) 348. [102] M.B. Rockel, M. Renner, Pitting, crevice and stress corrosion resistance of high chromium and molybdenum alloy stainless steels, Wekst. Korros. 35 (1984) 537. [103] C.P. Sturrock, W.F. Bogaerts, Empirical learning investigations of the stress corrosion cracking of austenitic stainless steels in high-temperature aqueous environments, Corrosion 53 (1997) 333. [104] R. Garnsey, Corrosion of PWR steam generators, Nucl. Energy 18 (1979) 117. [105] T. Mukohara, S.I. Koshizuka, Y. Oka, Core design of a high-temperature fast reactor cooled by supercritical light water, Ann. Nucl. Energy 26 (1999) 1423. [106] K. Dobashi, A. Kimura, Y. Oka, S. Koshizuka, Conceptual design of a high temperature power reactor cooled and moderated by supercritical light water, Ann. Nucl. Energy 25 (1998) 487. [107] O. Wachter, G. Bruemmer, Experiences with austenitic steels in boiling water reactors, Nucl. Eng. Des. 168 (1997) 35. [108] M. Kehr, H. Breuer, H. Schettler, Die neuen 800-MWBraunkohlenbloecke der VEAG Vereinigte Energiewerke AG in Boxberg und Schwarze Pumpe, VGB Kraftwerkstechnik 73 (1993) 941. [109] G. Scheffknecht, Q. Chen, Material issues for supercritical boilers, in: A. Strang, W.M. Banks, R.D. Conroy, G.M. McColvin, J.C. Neal, S. Simpson (Eds.), Parsons 2000 ConferenceAdvanced Materials for 21st Century Turbines and Power Plant, Proceeding of the 5th International Charles Parsons Turbine Conference, Cambridge, UK, 2000, p. 249. [110] A.B Vajnman, O.I. Martynova, V.A. Ens, Failures of steam lines at thermal power stations due to corrosion surroundings, Teploenergetika 46 (5) (1999) 35. [111] T. Kondo, Y. Watanabe, Y.S. Yi, A. Hishinuma, An evaluation of potential material-coolant compatibility for applications in advanced fusion reactors, J. Nucl. Mater. 263 (1998) 2083. [112] N. Henriksen, O.H. Larsen, Corrosion in ultra supercritical boilers for straw combustion, Mater. High Temp. 14 (1997) 227. [113] Y. Katsumura, Y. Oka (Eds.), The 1st International Symposium on Supercritical-Cooled Water Reactors, Design and Technology, The University of Tokyo Press, Tokyo, Japan, 2000. [114] V.S. Mishra, V.V. Mahajani, J.B. Joshi, Wet air oxidation, Ind. Eng. Chem. Res. 34 (1995) 2. [115] J.B. Joshi, Y.T. Shah, S.J. Parulekar, Engineering aspects of the treatment of aqueous waste streams, Indian Chem. Eng. 27 (1985) 2/3. [116] M.J. Dietrich, T.L. Randall, P.J. Canney, Wet air oxidation of hazardous organics in wastewater, Environ. Prog. 4 (1985) 171.

25

[117] S.T. Kolaczkowski, P. Plucinski, F.J. Beltran, F.J. Rivas, D.B. McLurgh, Wet air oxidation: a review of process technologies and aspects in reactor design, Chem. Eng. J. 73 (1999) 143. [118] F. Luck, Wet air oxidation: past, present and future, Catal. Today 53 (1999) 81. [119] S. Huang, K. Daehling, T.E. Carleson, P. Taylor, C. Wai, A. Propp, Thermodynamic analysis of corrosion of iron alloys in supercritical water, in: Supercritical Science and Technology, ACS Symposium Series 406, American Chemical Society, Washington DC, 1989, p. 276. [120] L.B. Kriksunov, D.D. Macdonald, Corrosion in supercritical water oxidation systems: a phenomenological analysis, J. Electrochem. Soc. 142 (1995) 4069. [121] L.B. Kriksunov, D.D. Macdonald, Corrosion testing and prediction in SCWO environments, in: Proceedings of the ASME Heat Transfer Division, The American Society of Mechanical Engineers, New York, 1995, p. 281. [122] L.B. Kriksunov, Thermogalvanic effects in corrosion in supercritical water, Corrosion 98, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 418, 1998, 9 pp. [123] J.W. Tester, J.A. Cline, Hydrolysis and oxidation in subcritical and supercritical water: connecting process engineering science to molecular interactions, Corrosion 55 (1999) 1088. [124] M. Wagner, V. Kolarik, B. Michelfelder, M. del Mar Juez-Lorenzo, T. Hirth, N. Eisenreich, P. Eyerer, Werkstoffe fr SCWO-Prozesse zur Oxidation von chlorhaltigen Schadstoffen in berkritischem Wasser, Mater. Corrosion 50 (1999) 523. [125] C. Kaul, H. Vogel, H.E. Exner, Korrosionsverhalten anorganischer Materialien in nah- und berkritischen wssrigen Lsungen, Mat.-Wiss. Werkstofftech. 30 (1999) 326. [126] Y.S. Kim, D.B. Mitton, R.M. Latanision, Corrosion resistance of stainless steels in chloride containing supercritical water oxidating system, Korean J. Eng. 17 (2000) 58. [127] D.B. Mitton, J.H. Yoon, R.M. Latanision, Overview of corrosion phenomena in SCWO systems for hazardous waste destruction, Zairyo to Kankyo: Corrosion Eng. 49 (2000) 130. [128] E. Han, Materials degregation in supercritical water oxidation system, Corrosion Sci. Prot. Technol. China 11 (1999) 44. [129] D.B. Mitton, N. Eliaz, J.A. Cline, R.M. Latanision, An overview of the current understanding of corrosion in SCWO systems for the destruction of hazardous waste products, Mater. Technol. Adv. Perf. Mater. 16 (1) (2001) 30. [130] F. Matthews, E.F. Gloyna, Corrosion Behavior of Three High-Grade Alloys in Supercritical Water Oxidation Environments, Report CRWR 234, Center for Research in Water Resources, Bureau of Engineering Research, University of Texas in Austin, Austin, TX, 1992, 73 pp. [131] K.W. Downey, R.H. Snow, D.A. Hazlebeck, A.J. Roberts, Corrosion and chemical agent destruction, in: K.W. Hutchenson, N.R. Foster (Eds.), Research on Supercritical Water Oxidation Oxidation of Hazardous Military Wastes,

26

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129 ACS Symposium Series 608, ACS, Washington, DC, 1995, p. 313. D.M. Harradine, S.J. Buelow, P.C. Dell Orco, R.B. Dyer, B.R. Foy, J.M. Robinson, Oxidation chemistry of energetic materials in supercritical water, Hazardous Waste Hazardous Mater. 10 (1993) 233. M.J. Drews, M. Williams, M. Barr, The corrosion of solgel coated type 316 SS in chlorinated SC water, Ind. Eng. Chem. Res. 39 (2000) 4772. C. Liu, D.D. Macdonald, E. Medina, J.J. Villa, J.M. Bueno, Probing corrosion activity in high subcritical and supercritical water through electrochemical noise analysis, Corrosion 50 (1994) 687. D.B. Mitton, J.C. Orzalli, R.M. Latanision, Corrosion phenomena associated with SCWO systems, in: G. Brunner, M. Perrut (Eds.), Proceedings of the 3rd International Symposium on Supercritical Fluids, Strasbourg, France, 1994, p. 43. S.F. Rice, R.R. Steeper, C.A. LaJeunesse, Destruction of Representative Navy Wastes Using Supercritical Water Oxidation, Final Report, Sandia National Laboratories, Livermore, CA, 1994, 35 pp. K.M. Garcia, R.E. Mizia, Corrosion investigation in supercritical water oxidation process environments, in: Proceedings of the ASME Heat Transfer Division, The American Society of Mechanical Engineers, New York, 1995, p. 299. R.M. Latanision, Corrosion science, corrosion engineering and advanced technologies, Corrosion 51 (1995) 270. D.B. Mitton, P.A. Marrone, R.M. Latanision, Interpretation of the rationale for feed modication in SCWO systems, J. Electrochem. Soc. 143 (1996) L59. V.A. Zilberstein, J.A. Bettinger, D.W. Ordway, G.T. Hong Evaluation of materials performance in a supercritical wet oxidation system, Corrosion 95, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 558, 1995, 19 pp. N. Boukis, G. Franz, C. Friedrich, W. Habicht, K. Ebert, Corrosion screening tests with ni-base alloys in supercritical water containing hydrochloric acid and oxygen, Proceedings of the ASME Heat Transfer Division, The American Society of Mechanical Engineers, New York, Vol. 335, vol. 4, 1996, p. 159. A.G. Robertson, Corrosion evaluation of selected materials under acidic hydrothermal oxidation conditions, Master thesis, University of Texas in Austin, Austin, TX, 1996, 148 pp. C. Fill, H. Tiltscher, Untersuchungen zur Werkstoff- und Korrosionsproblematik bei der Oxidation von Schadstoffen in berkritischem Wasser, Mater. Corrosion 48 (1997) 146. S. Fodi, J. Konys, J. Hausselt, H. Schmidt, V. Casal, corrosion of high temperature alloys in a supercritical water oxidation process, EUROCORR 97, European Federation of Corrosion, Event no. 208, Trondheim, Norway, vol. 1, 1997, p. 629. R.M. Latanision, D.B. Mitton, S.H. Zhang, J.A. Cline, N. Caputy, T.A. Arias, A. Rigos, Corrosion and corrosion mechanisms in supercritical water oxidation systems for hazardous waste diposal, Proceedings of the 4th International Symposium on Supercritical Fluids, Sendai, Japan, 1997, p. 865. D.B. Mitton, S.H. Zhang, K.E. Hautanen, J.A. Cline, E.H. Han, R.M. Latanision, Evaluating stress corrosion and corrosion aspects in supercritical water oxidation systems for the destruction of hazardous waste, Corrosion 97, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 203, 1997, 6 pp. S. Fodi, J. Konys, J. Hausselt, H. Schmidt, V. Casal, Corrosion of high temperature alloys in supercritical water oxidation systems, Corrosion 98, NACE National Association of Corrosion Engineers, Houston, TX, Paper No. 416, 1998, 10 pp. K.S. Lin, H.P. Wang, M.C. Li, Oxidation of 2,4-dichlorophenol in supercritical water, Chemosphere 36 (1998) 2075. D.B. Mitton, S.H. Zhang, M.S. Quintana, J.A. Cline, N. Caputy, P.A. Marrone, R.M. Latanision, Corrosion mitigation in SCWO systems for hazardous waste disposal, Corrosion 98, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 414, 1998, 15 pp. A. Ruck, J. Novotny, J. Konys, J. Hausselt, Effect of iron in nickel based alloys for SCWO plants, Eurocorr 98, European Federation of Corrosion, Event no. 221, Utrecht, The Netherlands, 1998, electronic proceedings. A. Ruck, J. Novotny, J. Konys, J. Hausselt, Comparison of nickel, iron and cobalt based alloys for supercritical water oxidation plants, Eurocorr 98, European Federation of Corrosion, Event no. 221, Utrecht, The Netherlands, 1998, electronic proceedings. J. Konys, S. Fodi, J. Hausselt, H. Schmidt, V. Casal, Corrosion of high-temperature alloys in chloride-containing supercritical water oxidation systems, Corrosion 55 (1999) 45. N. Boukis, N. Claussen, K. Ebert, R. Janssen, M. Schacht, Corrosion screening tests of high-performance ceramics in supercritical water containing oxygen and hydrochloric acid, J. Eur. Ceram. Soc. 17 (1997) 71. M. Schacht, N. Boukis, E. Dinjus, K. Ebert, R. Janssen, F. Meschke, N. Claussen, Corrosion of zirconia ceramics in acidic solutions at high pressures and temperatures, Key Eng. Mater. 132136 (1997) 1677. M. Schacht, Das Korrosionsverhalten von Werkstoffen auf Aluminiumoxid- und Zirkondioxid-Basis in wrigen Lsungen unter hydrothermalen Bedingungen, Report FZKA 6112, Forschungszentrum Karlsruhe GmbH, Karlsruhe, Germany, 1998, 125 pp. C. Friedrich, P. Kritzer, N. Boukis, G. Franz, E. Dinjus, The corrosion of tantalum in oxidizing sub- and supercritical aqueous solutions of HCl, H2 SO4 and H3 PO4 , J. Mater. Sci. 34 (1999) 3137. G.L. DArcy, Corrosion behavior of titanium alloys exposed to supercritical water oxidation conditions, Master thesis, University of Texas in Austin, Austin, TX, 1996, 174 pp. N. Boukis, C. Friedrich, W. Habicht, M. Schacht, E. Dinjus, Corrosion screening tests in supercritical water containing hydrochloric acid and oxygen, EUROCORR 97, European

[132]

[146]

[133]

[134]

[147]

[135]

[148] [149]

[136]

[150]

[137]

[151]

[138] [139]

[152]

[140]

[153]

[141]

[154]

[155]

[142]

[143]

[156]

[144]

[157]

[145]

[158]

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129 Federation of Corrosion, Event no. 208, Trondheim, Norway, vol. 1, 1997, p. 617. N. Boukis, C. Friedrich, E. Dinjus, Titanium as reactor material for SCWO applications. First Experimental Results, CORROSION 98, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 417, 1998, 7 pp. H. Glasbrenner, R. Kraft, S. Leistikow, V. Casal, M. Gegenheimer, H. Schmidt, Experimentelle Untersuchungen zur Korrosion potentieller Anlagen-Werkstoffe fr die oxidative Schadstoffzersetzung in berkritischem Wasser bei 500 C, 270 bar, Report KfK 5401, Kernforschungszentrum Karlsruhe GmbH, Karlsruhe, Germany, 1994, 60 pp. N. Boukis, P. Kritzer, Corrosion phenomena on alloy 625 in aqueous solutions containing hydrochloric acid and oxygen under subcritical and supercritical conditions, Corrosion 97, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 10, 1997, 11 pp. P. Kritzer, N. Boukis, E. Dinjus, Change of corrosion phenomena in sub- and supercritical water, EUROCORR 97, European Federation of Corrosion, Event no. 208, Trondheim, Norway, vol. 2, 1997, p. 229. P. Kritzer, N. Boukis, E. Dinjus, Transpassive dissolution of nickel-base alloys and stainless steels in oxygen and chloride containing high-temperature water, Mater. Corrosion 48 (1997) 799. P. Kritzer, N. Boukis, E. Dinjus, Corrosion phenomena of alloy 625 in aqueous solutions containing sulfuric acid and oxygen under subcritical and supercritical conditions, Corrosion 98, National Association of Corrosion Engineers, Houston, TX, Paper no. 415, 1998, 8 pp. P. Kritzer, N. Boukis, E. Dinjus, Corrosion of alloy 625 in high temperature sulfate solutions, Corrosion 54 (1998) 689. P. Kritzer, N. Boukis, E. Dinjus, Corrosion of alloy 625 in aqueous chloride and oxygen containing solutions, Corrosion 54 (1998) 824. P. Kritzer, N. Boukis, E. Dinjus, The corrosion behaviour of nickel-base alloy 625 (NiCr22Mo9Nb, 2.4856) and ceria stabilized tetragonal zirconia polycrystal (CeTZP) against oxidizing aqueous solutions of hydrouoric acid (HF), hydrobromic acid (HBr), and hydriodic acid (HI) at sub- and supercritical temperatures, Mater. Corrosion 50 (1999) 505. P. Kritzer, N. Boukis, E. Dinjus, The corrosion of nickel-base alloy 625 in sub- and supercritical aqueous solutions containing oxygen. A long time study, J. Mater. Sci. Lett. 17 (1999) 1845. P. Kritzer, N. Boukis, E. Dinjus, Investigations of the corrosion of reactor materials during the process of supercritical water oxidation (SCWO), in: Proceedings of the 6th Meeting on Supercritical Fluids, Nottingham, UK, Institute National Polytechnique De Lorraine (INPL), Vandoeuvre Cedex, France, 1999, p. 433. P. Kritzer, N. Boukis, E. Dinjus, Factors controlling corrosion in high-temperature aqueous solutions. A contribution to the dissociation and solubility data inuencing corrosion processes, J. Supercritical Fluids 15 (1999) 205. P. Kritzer, N. Boukis, E. Dinjus, Transpassive dissolution of alloy 625, chromium, nickel, and molybdenum in

27

[159]

[172]

[160]

[173]

[174]

[161]

[175]

[162]

[176]

[163]

[177]

[178]

[164]

[179]

[165] [166]

[180]

[167]

[181]

[168]

[182]

[169]

[183]

[170]

[184]

[171]

high-temperature solutions containing hydrochloric acid and oxygen, Corrosion 56 (2000) 265. P. Kritzer, N. Boukis, E. Dinjus, Review of the corrosion of nickel base alloys and stainless steels in strongly oxidizing pressurized high-temperature solutions at sub- and supercritical temperatures, Corrosion 56 (2000) 1093. P. Kritzer, Die Korrosion der Nickel-Basis-Legierung 625 unter hydrothermalen Bedingungen, Report FZKA 6168, Forschungszentrum Karlsruhe GmbH, Karlsruhe, Germany, 1998, 180 pp. P. Kritzer, N. Boukis, E. Dinjus, The corrosion of nickel-base alloy 625 in oxidizing sub- and supercritical aqueous solutions of HNO3 , J. Mater. Sci. Lett. 17 (1999) 771. P. Kritzer, N. Boukis, E. Dinjus, The corrosion of alloy 625 (NiCr22Mo9Nb, 2.4856) in high-temperature, high-pressure aqueous solutions of phosphoric acid and oxygen. Corrosion at sub- and supercritical temperatures, Mater. Corrosion 49 (1998) 831. P. Kritzer, N. Boukis, G. Franz, E. Dinjus, The corrosion of niobium in oxidizing sub- and supercritical aqueous solutions of HCl and H2 SO4 , J. Mater. Sci. Lett. 17 (1999) 25. P. Kritzer, N. Boukis, Verfahren zur Durchfhrung von chemischen Reaktionen in berkritischen wssrigen Systemen, German Patent DE 197 47 696, 1999. Fournier, D. Delafosse, C. Bosch, T. Magnin, Stress corrosion cracking of nickel base superalloys in aerated supercritical water, Corrosion 2001, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 1361, 2001. J.A. Cline, P.A. Marrone, D.B. Mitton, R.M. Latanision, J.W. Tester, Corrosion mechanisms of alloy N10276 in hydrothermal hCl solutions: failure analysis and expoure studies, Corrosion 2001, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 1362, 2001. S. Iimura, H. Yakuwa, M. Miyasaka, Y. Kuriki, J. Sakai, Corrosion behavior of several commercial alloys in supercritical water including high concentration of chloride ions, Corrosion 2001, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 1363, 2001. Y. Ikushima, M. Son, H. Kim, Y. Kurata, K. Hatakeda, Corrosion behavior of metals in supercritical water solutions in the presence of salts, Corrosion 2001, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 1364, 2001. N. Hara, S. Tanaka, K. Sugimoto, Corrosion behavior of constituent metals of stainless alloys in SCWO environments, Corrosion 2001, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 1365, 2001. J. Konys, A. Ruck, J. Novotny, J. Hausselt, Corrosion of nickel based alloy G-30 between 100 and 420 C, pH values from 2 to 6 and pressures ranging from 250 to 480 bar under conditions of supercritical water oxidation, Corrosion 2001, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 1366, 2001. X.Y. Zhou, S.N. Lvov, X.J. Wie, L.G. Benning, D.D. Macdonald, Measuring corrosion rate of type 304 stainless steel in subcritical and supercritical aqueous solutions via

28

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129 electrochemical noise analysis, Corrosion 2001, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 1368, 2001. Y. Watanabe, T. Kobayashi, T. Adschiri, Signicance of water density on corrosion behavior of alloys in supercritical water, Corrosion 2001, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 1369, 2001. N. Boukis, G. Franz, W. Habicht, E. Dinjus, Corrosion resistant materials for SCWO-applications. Experimental results from long-time experiments, Corrosion 2001, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 1353, 2001. T. Adschiri, K. Sue, K. Arai, Y. Watanabe, Estimation of metal oxide solubility and understanding corrosion in supercritical water, Corrosion 2001, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 1354, 2001. D.B. Mitton, N. Eliaz, J.A. Cline, R.M. Latanision, Assessing degradation mechanisms in supercritical water oxidation systems, Corrosion 2001, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 1352, 2001. E.H. Han, L. Zhang, H. Guan, W. Ke, Corrosion behavior of materials in supercritical water oxidation systems, Corrosion 2001, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 1356, 2001. Y. Tsuchiya, N. Saito, K. Hiruta, Y. Arai, Corrosion behavior of metals for SCWO reactors for organic waste processing plants, Corrosion 2001, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 1357, 2001. D.D. Macdonald, An overview of the chemical and electrochemical conditions that exist in SCWO systems, Corrosion 2001, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 1358, 2001. D.B. Mitton, H. Kim, J. Zhang, N. Eliaz, C.R. Sydnor, R.M. Latanision, An examination of the corrosion phenomena of potential constructional materials for SCWO system fabrication, Corrosion 2002, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 2353, 2002. K. Sue, N. Tsujinaka, T. Adschiri, K. Arai, Y. Watanabe, Relationship between corrosion rate and metal oxide solubility in supercritical water, Corrosion 2002, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 2354, 2002. Y. Watanabe, K. Shoji, T. Adschiri, K. Sue, Effects of oxygen concentration on corrosion behavior of alloys in a acidic supercritical water, Corrosion 2002, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 2355, 2002. M. Son, Y. Kurata, Y. Ikushima, Corrosion behavior of metals in SCW environments containing salts and oxygen, Corrosion 2002, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 2357, 2002. N. Hara, S. Tanaka, S. Soma, K. Sugimoto, Corrosion resistance of FeCr and NiCr Alloys in oxidizing supercritical Cl solutions, Corrosion 2002, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 2358, 2002. [197] C. Schroer, J. Konys, J. Novotny, J. Hausselt, Corrosion investigations on binary and ternary NiCr alloys in SCWO systems, Corrosion 2002, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 2359, 2002. [198] H. Yakuwa, M. Miyasaka, S. Iimura, Effect of Cr and Mo contents on the corrosion properties of Ni-base alloys in the supercritical water including chloride, Corrosion 2002, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 2361, 2002. [199] J.A. Cline, K. Takahashi, T.W. Marin, C.D. Jonah, D.M. Bartels, Kinetics of electron transfer reactions in hydrothermal and supercritical water, Corrosion 2002, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 2362, 2002. [200] C. Frayret, Th. Jaszay, B. Lestienne, M.H. Delville, Development of a model for the anodic behavior of T60 titanium in chlorinated and oxygenated aqueous media. Application to the specic conditions of hydrothermal oxidation (1 MPa < pressure < 30 MPa, 20 C < temperature < 400 C), Electrochim. Acta, 48 (2003) in press. [201] M. Weber, B. Wellig, K. Lieball, P. Rudolf von Rohr, Operating transpiring-wall SCWO reactors: characteristics and quantitative aspects, Corrosion 2001, NACE National Association of Corrosion Engineers, Houston, TX, Paper no. 1370, 2001. [202] X. Chen, R.M. Izatt, J.L. Oscarson, Thermodynamic data for ligand interaction with protons and metal ions in aqueous solutions at high temperatures, Chem. Rev. 94 (1994) 467. [203] J.D. Frantz, W.L. Marshall, Electrical conductances and ionization constants of salts, acids, and bases in supercritical aqueous uids: I. Hydrochloric acid from 100 to 700 C and at pressures to 4000 bars, Am. J. Sci. 284 (1984) 651. [204] G. Barbery, Y. Brub, An automated method for analysis of dissolved gases in water at high temperatures and pressures, Ind. Eng. Chem. Fundam. 10 (1971) 632. [205] M.B. Reynolds, W.L. Clarke Jr, Measurement and control of dissolved O2 content of water at elevated temperatures and pressures, Corrosion Sci. 10 (1970) 883. [206] S.D. Cramer, The solubility of oxygen in brines from 0 to 300 C, Ind. Eng. Chem. Process Des. Dev. 19 (1980) 300. [207] H.A. Pray, C.E. Schweickert, B.H. Minnich, Solubility of hydrogen, oxygen, nitrogen, and helium in water at elevated temperatures, Ind. Eng. Chem. 44 (1952) 1146. [208] L.M. Zoss, S.N. Suciu, W.L. Sibbitt, The solubility of oxygen in water, Trans. ASME (1954) 69. [209] C.M. Chen, K. Arai, G.J. Theus, Computer-Calculated Potential pH Diagrams to 300 C, EPRI-Report NP-3137, Electric Power Research Institute, Palo Alto, CA, 1983, 130 pp. [210] R.J. Biernat, R.G. Robins, High temperature potential/pH diagrams for the sulphurwater system, Electrochim. Acta 14 (1969) 809. [211] R.C. Murray, D. Cubicciotti Jr, Thermodynamics of aqueous sulfur species to 300 C and potentialpH diagrams, J. Electrochem. Soc. 130 (1983) 866.

[185]

[186]

[187]

[188]

[189]

[190]

[191]

[192]

[193]

[194]

[195]

[196]

P. Kritzer / J. of Supercritical Fluids 29 (2004) 129 [212] S. Oana, H. Ishikawa, Sulfur isotopic fractionation between sulfur and sulfuric acid in the hydrothermal solution of sulfur dioxide, Geochem. J. 1 (1966) 45. [213] A.J. Ellis, W. Giggenbach, Hydrogen sulphide ionization and sulphur hydrolysis in high temperature solution, Geochim. Cosmochim. Acta 35 (1971) 247. [214] R.P. Rafalsky, L.S. Medvedeva, V.A. Alexeev, Interaction of sulfur and water under high temperatures, Geokhimia 5 (1983) 665. [215] T.P. Dadze, V.I. Sorokin, Experimental determination of the concentrations of H2 S, HSO4 , SO2aq , H2 S2 O3 , S0 , aq and Stot in the aqueous phase in the SH2 O system at elevated temperatures, Geochem. Int. 30 (8) (1993) 36. [216] G.V. Bondarenko, Y.E. Gorbaty, In situ raman spectroscopic study of sulfur-saturated water at 1000 bar between 200 and 500 C, Geochim. Cosmochim. Acta 61 (1997) 1413. [217] P. Marcus, E. Protopopoff, Potential pH diagrams for sulfur and oxygen adsorbed on nickel in water at 25 and 300 C, J. Electrochem. Soc. 140 (1993) 1571. [218] P. Marcus, E. Protopopoff, PotentialpH diagrams for sulfur and oxygen adsorbed on chromium in water, J. Electrochem. Soc. 144 (1997) 1586. [219] C.H. Shen, P.G. Shewmon, A mechanism for hydrogeninduced intergranular stress corrosion cracking in alloy 600, Met. Trans. A 21A (1990) 1261.

29

[220] N. Totsuka, E. Lunarska, G. Cragnolino, Z. SzklarskaSmialowska, Effect of hydrogen on the intergranular stress corrosion cracking of alloy 600 in high-temperature aqueous environments, Corrosion 43 (1987) 505. [221] G. Sui, J.M. Titchmarsh, G.B. Heys, J. Congleton, Stress corrosion cracking of alloy 600 and alloy 690 in hydrogen/ steam at 380 C, Corrosion Sci. 39 (1997) 565. [222] J. Robertson, The mechanism of high temperature aqueous corrosion of steel, Corrosion Sci. 29 (1989) 1275. [223] P. Elliot, C.J. Tyreman, R. Prescott, High temperature alloy corrosion by halogens, J. Met. (1985) 20. [224] M. Schtze, Mechanical properties of oxide scales, Oxidation Met. 44 (1995) 29. [225] T. Link, A. Rahmel, M. Schtze, The inuence of gaseous impurities in air on the high temperature corrosion of coated and uncoated nickel-based superalloys, Mater. High Temp. 13 (1995) 55. [226] R.A. Rapp, Fundamental aspects of high-temperature corrosion, J. Phys. IV 3 (1993) 1. [227] H.J. Grabke, E.M. Mller-Lorenz, J. Klwer, D.C. Agarwal, Metal dusting of nickel-based alloys, Mater. Perform. 37 (7) (1998) 58. [228] J. Matsuda, N. Shimono, K. Tamura, Supercritical fossil red power plants: design and developments, in: Y. Katsumura, Y. Oka (Eds.), The 1st International Symposium on Supercritical-Cooled Water Reactors, Design and Technology, The University of Tokyo Press, Tokyo, Japan, 2000, p. 79.

You might also like