Você está na página 1de 21

Corrosion Science 76 (2013) 626

Contents lists available at SciVerse ScienceDirect

Corrosion Science
journal homepage: www.elsevier.com/locate/corsci

Review

Corrosion in biomass combustion: A materials selection analysis and its interaction with corrosion mechanisms and mitigation strategies
Renato Altobelli Antunes a,, Mara Cristina Lopes de Oliveira b
a b

Engineering, Modeling and Applied Social Sciences Center (CECS), Federal University of ABC (UFABC), 09210-170 Santo Andr, SP, Brazil Electrocell Ind. Com. Equip. Elet. LTDA, Technology, Entrepreneurship and Innovation Center (CIETEC), 05508-000 So Paulo, SP, Brazil

a r t i c l e

i n f o

a b s t r a c t
Corrosion induced by chlorine during biomass combustion is a major drawback to the consolidation of this technology. The current literature brings valuable information on these issues. However, a complete assessment of the corrosion mechanisms, mitigation strategies and materials selection for biomass combustion is not encountered. The aim of this work is to develop a materials selection strategy for superheater tubes used in biomass combustion, assessing its interaction with both the corrosion mechanisms and mitigation methods. The Ashbys approach was used with this purpose. 2013 Elsevier Ltd. All rights reserved.

Article history: Received 4 January 2013 Accepted 7 July 2013 Available online 16 July 2013 Keywords: A. Stainless steel A. Nickel C. High temperature corrosion

1. Introduction Fossil fuels are the basis of energy production worldwide. Burning of coal, oil and natural gas represents near 80% of the total amount of energy used in the contemporary world [1]. The replacement of fossil fuels by less pollutant alternatives is as an inevitable step into the establishment of sustainable development policies based on a low-carbon economy. Biomass is a promising energy source to reduce greenhouse gas emissions [2]. Attractive issues regarding the use of biomass are its widespread availability and possibility of production and consumption in a near-neutral CO2 basis [3]. According to Bhutto et al. [4], biomass represents 9% of the global primary energy demand. If one considers the energy derived only from renewable sources, biomass represents 75% of the total amount [5,6]. Recently, Lim and Lee [7] pointed that the projections for the participation of biomass in the renewable energy demand is expected to increase up to 2030. Kirkels [8] described a scenario of the last 30 years where the application of biomass as an energy source in Western Europe sharply increased especially after 2000, as well as the budget for research and development in this area. In Brazil, the energy production from biomass increased 18.1% from 2009 to 2010 [9]. Energy outputs from biomass can be encountered in three different forms: solid fuels, liquid fuels such as biodiesel and bioethanol and gas fuels produced through the gasication of solid biomass [10]. Wood, agricultural crops, municipal solid wastes, animal wastes, algae and aquatic plants and waste from food
Corresponding author. Tel./fax: +55 11 4996 8241.
E-mail address: renato.antunes@ufabc.edu.br (R.A. Antunes). 0010-938X/$ - see front matter 2013 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.corsci.2013.07.013

processing are the most relevant sources for biomass energy from solid fuels [11]. Demirbas [12] reported that direct combustion and co-ring of these resources with coal are promising routes for generation of electricity. According to Dermibas [13], co-ring biomass with coal is an effective way of reducing greenhouse gas emissions in comparison with the combustion of pure coal. Moreover, fuel costs and waste generation are minimized whereas soil and water pollution can be diminished depending on the chemical composition of the biomass source. For both direct combustion and co-ring with coal, the preferred feedstocks for energy production are forest residues, bagasse and other agricultural crops [11]. Khalil et al. [14] highlighted that these types of biomass are continuously produced at specic locations which can be possibly in the vicinity of a power plant, being advantageous in regards to the cost reduction of fuel transportation. Biomass combustion can be described by the following sequence of events: heating-up, drying and devolatilization. During devolatilization, volatiles and char are produced. Next, both the volatiles and char are combusted [15]. Biomass composition plays a pivotal role in the combustion process. The relevance of this subject is thoroughly examined in the excellent reports by Vassilev et al. [16,17]. Carbon, oxygen and hydrogen are the main constituents of all biomass with little difference among different sources [18]. Inorganic elements such as N, P, K, Ca, Mg, Na, Si, S and Cl are always present. Heavy metals such as cadmium and lead can also be found as trace elements [19]. Biomass fuels such as straw are rich in alkali metals (K and Na) which are mainly present as simple salts and organic compounds [20] and chlorine. These species are promptly released to the gas phase during combustion, forming HCl and KCl. High amounts of KCl in the combustion gases are frequently associated with

R.A. Antunes, M.C.L. de Oliveira / Corrosion Science 76 (2013) 626

enhanced deposit formation. This, in turn, will lead to corrosion of superheater tubes in biomass-red boilers [21]. Sulfur is also involved in the corrosion-related phenomena [22]. Johansson et al. [23] considered superheaters corrosion as an important obstacle to the development of biomass as a competitive widespread source for green energy. As a response to this complex technological barrier, corrosion studies regarding both biomass ring and co-ring have proliferated in the past few years [2427]. In spite of the growing interest in these technologies, there is still much to be learned in order to accomplish optimized materials performance and combustion properties. In this context, corrosion phenomena are of crucial importance. This work provides a thorough analysis of materials selection for superheater tubes in biomass combustion, focusing its interaction with both the corrosion mechanisms and mitigation methods involved in this eld. The rst part of the article gives an overview of the corrosion mechanisms in biomass combustion and other related high temperature corrosion problems. In the second part, mitigation methods are explored. In the last part, the Ashbys approach is employed to propose a methodology for selecting materials for superheater tubes in biomass combustion. 2. Corrosion mechanisms 2.1. Active oxidation by chlorine The deleterious action of chlorine on the operation of superheater tubes for combustion of biomass is well documented [20,28]. Active oxidation of iron and steels occurs when the oxide scale on the metallic surface has no protective character and chloride-based contaminants strongly accelerate its oxidation rate. Grabke et al. [29] gave the basis for understanding this mechanism. The main reactions involved in the active oxidation mechanism can be summarized as follows [29]:

of HCl and O2. Next, solid metal chlorides are formed which evaporates at high temperatures. The subsequent step is the formation of the corresponding metal oxide through the reaction of the metal chloride with oxygen. By comparing the Gibbs free energies of CrCl2 (286.0 kJ/mol), FeCl2 (232.1 kJ/mol), and NiCl2 (174.2 kJ/mol) at the reference temperature of 600 C it is clearly seen that NiCl2 is the least reactive species. Regarding the conversion of the metal chlorides into metal oxides, the equilibrium oxygen partial pressures p(O2) for the different chloride/oxide reactions indicates the ease with which the reaction can proceed. NiO needs the highest oxygen partial pressure to be formed from the nickel chloride whereas chromium oxide is more easily converted from CrCl2 at lower oxygen partial pressures. The conversion of FeCl2 to Fe2O3 occurs at intermediate values of p(O2). The experimental data revealed that corrosion attack is more severe for iron and chromium than for nickel, conrming the validity of the active oxidation mechanism. Andersson and Norell [31,32] carried out eld tests in waste boilers with the ferritic T22 steel, the austenitic 310 stainless steel, a high alloy austenitic stainless steel (Sanicro 28) and a nickelbased alloy (Sanicro 65). The lowest corrosion loss was observed for the Sanicro 28 alloy which presented the most compact oxide scale. This was associated with the ability of the continuous oxide layer at preventing the penetration of chlorine inwards into the metal surface which supports the active oxidation mechanism. In this regard, the addition of chromium and nickel as alloying elements to iron seems to favor the formation of such a protective scale. A materials selection strategy for superheater tubes in biomass combustion should entail the application of this knowledge. The chemical composition of the metallic alloy is essential to the formation of a stable and corrosion resistant oxide scale. Section 4 deals with this problem by employing the Ashbys philosophy to select materials for superheater tubes in biomass combustion. 2.2. Alkali-containing species 2.2.1. Alkali chlorides The alkali content in the fuel has been has been recognized as an important factor in the corrosion process of superheater tubes used in biomass combustion and waste incineration plants. Krause [33] early described the role of alkali metal oxides in the corrosion of boiler tubes. These species, upon reaction with HCl, form alkali chlorides which can condensate and form deposits on the metal surface. Corrosion reactions involving solid phase Cl-species in deposits have been associated with alkali chlorides, especially KCl in the case of biomass ring. Nielsen et al. [20] described a mechanism based on the sulfation of alkali chlorides in deposits according to:

2HCl 1=2 O2 $ Cl2 H2 O 2NaCl Fe2 O3 1=2 O2 $ Na2 Fe2 O4 Cl2 4NaCl Cr2 O3 5=2 O2 $ 2NaCrO4 2Cl2 Fe Cl2 $ FeCl2 s FeCl2 s $ FeCl2 g 3FeCl2 2O2 $ Fe3 O4 3Cl2 2FeCl2 3=2 O2 $ Fe2 O3 2Cl2

1 2 3 4 5 6 7

Reactions (1)(3) refer to the formation of chlorine from the oxidation of HCl or upon reaction of condensed chlorides with the oxide scale. In a next step, if chlorine penetrates the oxide scale through pores or cracks, metal chlorides are able to form at the metal/scale interface as shown in reaction (4) for FeCl2(s). This species has a high vapor pressure at 500 C, which is a typical temperature of superheater materials employed for biomass combustion, and volatilizes, according to reaction (5). When it diffuses through the cracks and pores of the oxide scale, it can be oxidized, forming Fe3O4 and/or Fe2O3, according to reactions (6) and (7), releasing chlorine. The cycle can be repeated, thus sustaining the oxidation of the metallic surface beneath the non-protective oxide scale. The validity of the active oxidation mechanism has been experimentally investigated by Zahs et al. [30]. These authors performed a detailed study of the chloridation and oxidation of iron, chromium, nickel and their alloys in chloridizing and oxidizing atmospheres at temperatures ranging from 400 to 700 C. Under oxidizing atmospheres, chlorine is formed through the reaction

2KCls SO2 g 1=2 O2 g H2 Og ! K2 SO4 s 2HClg 8 2KCls SO2 g O2 g ! K2 SO4 s Cl2 g 9

The HCl(g) formed through Eq. (8) can be further oxidized, producing Cl2(g). Another possibility is that HCl diffuses through the deposit, reaching the metal surface and forming volatile metal chlorides, namely, FeCl2 or CrCl2. These species can diffuse out of the deposit toward areas with high partial pressures of oxygen, forming metal oxides and releasing Cl2(g) or HCl(g). After this step, the corrosion process proceeds as in the case of gaseous Cl-species. A similar mechanism has been postulated by Montgomery et al. [34] after conducting eld tests in straw-red boilers. The generation of Cl2(g) from solid phase Cl-species can occur by an alternative route. Alkali chlorides would react with the metal scale and not with SO2(g), as shown in Eqs. (10) and (11). Then,

R.A. Antunes, M.C.L. de Oliveira / Corrosion Science 76 (2013) 626

corrosion proceeds as described for the case of the sulfation of alkali chlorides by the action of Cl2(g).

2NaCls; l 1=2 Cr2 O3 5=4 O2 g ! NaCrO4 s; l Cl2 g 2NaCls; l Fe2 O3 s 1=2 O2 g ! NaFe2 O4 s; l Cl2 g

10 11

Oxidation reactions involving molten alkali chlorides are also of great relevance for the corrosion phenomena in biomass combustion. Nielsen et al. [20] stated that the corrosion rate increases in the presence of molten phases due to an acceleration of the chemical reactions in comparison with solidsolid reaction and to the formation of a conductive electrolyte for the onset of electrochemical attack. The feasibility of a corrosion mechanism based on the metal attack by molten Cl-phases was conrmed by Nielsen et al. [35]. They conducted laboratory investigations of high temperature corrosion in straw-red boilers, using ferritic alloys (X20 and AISI 347) as superheater tubes. The samples were covered with synthetic (KCl and/or K2SO4) and real deposits and heated to 550 C. The authors proposed that KCl forms a molten phase with K2SO4 and iron compounds such as FexOy and FeCl2 close to the oxide layer on the metal surface. Rapid sulfation of KCl in the melt produces HCl or Cl2 close to the metal surface. Chlorine easily diffuses through the oxide layer toward the metal surface, forming volatile iron chlorides. From this point, the mechanism follows the same steps of the gaseous Cl-species involved in the active oxidation mechanism. van lith et al. [36] gave experimental support to this mechanism. The presence of KCl in the deposit formed on boiler grade TP 347H FG stainless steel made the corrosion attack more severe, resulting in the formation of a more uniform and deeper internal attack as the KCl content increased. Cha [37] reported that the aggressiveness of deposited ashes on superheater materials tested at 535 C increased with the alkali content in the ash. Similar ndings have been reported by Sroda et al. [38] and Brossard et al. [39]. Jonsson et al. [40] showed that the corrosion process of a low alloyed steel was initiated below 400 C, leading to a fast redistribution of KCl particles and iron ions on the metal surface. The rapid transport of KCl and corrosion products was due to the formation of a thin liquid layer of a KCl/FeCl2 mixture whose eutectic temperature is at 355 C. The authors conrmed the very corrosive nature of KCl, observing that signicant areas of the steel surface were covered by thick corrosion products after only 1 h of exposure, while in the absence of this species only a thin oxide lm was formed. Pan et al. [41] observed that KCl increased the corrosion rate of FeNiAlCr alloys in air at 650 C due to the formation of potassium chromate and not of a protective Cr2O3 oxide. Pettersson et al. [42] identied a similar mechanism for the high temperature corrosion of 304 stainless steel samples at 600 C and exposed to KCl. However, at 400 C the chromium oxide layer is still protective during experiments conducted for 168 h [43]. This mechanism was also shown to be active for the high-nickel alloy Sanicro 28 [44,45]. The reaction of KCl with the protective chromium oxide led to the acceleration of the corrosion process under the formation of potassium chromate. Ishitsuka and Nose [46] observed that the protective Cr2O3 scale can be easily dissolved to CrO2 4 in alkali chloride mixtures, thus leading to accelerated corrosion of boiler tubes in waste incineration environments. Moreover, molybdenum and silicon improved the corrosion resistance of Fe CrNi during hot corrosion experiments. It is suggested that these elements and vanadium or tungsten can be effectively used as alloying elements to improve the high temperature corrosion performance of steels used in the manufacturing of boiler tubes. The remarkable effect of alkali chlorides in the corrosion process of superheater materials raises an issue on how different metals behave in contact with deposits consisting of these species. The propensity to form corrosion products in typical biomass combustion atmospheres can provide useful guidelines to properly select

the most resistant alloys. In this scenario, the reactions of iron, chromium and nickel surfaces with alkali chlorides have been investigated by Cha and Spiegel [47,48]. Alkali chlorides were shown to react with the iron surface at 300 C, enhancing local oxidation whereas this effect was not observed on the nickel surface. No visible local reactions between KCl and chromium surfaces were identied at 300 C. When the testing temperature was raised to 500 C, KCl particles were found to react locally with the chromium surface, accelerating its oxidation. This behavior was explained by an active oxidation process in which chloride diffused through the oxide layer to the chromium/oxide interface. These results imply that the presence of nickel as an alloying element would be benecial to avoid excessive oxidation of superheater tubes during biomass combustion as the biomass-based fuels are typically rich in alkali species [49,50]. In fact, the protective character of oxide scale formed on nickel-containing alloys has been highlighted by several authors [51,52]. Recent studies by Lehmusto et al. [53] attempted to elucidate the corrosion mechanisms of superheater materials used in biomass combustion by investigating the role of different metal chlorides. They studied the ability of BaCl2, CaCl2, KCl, LiCl, MgCl2, NaCl, PbCl2 and ZnCl2 to react with pure metallic chromium powder at 400 C, 500 C, 550 C and 600 C. The results revealed that BaCl2, CaCl2 and MgCl2 did not react with chromium upon heating whereas KCl, NaCl and PbCl2 reacted at 500 C and at higher temperatures. LiCl reacted only at 600 C. These indications suggested that the presence of chlorine in the chloride form does not sufce to explain the initiation of accelerated oxidation because not all the chlorides studied reacted with metallic chromium. The authors highlight that the cation plays a role in the corrosion mechanism. However, further studies are necessary to elucidate it. 2.2.2. Potassium carbonate and hydroxide KCl has been unequivocally correlated with the major corrosion mechanisms operating during biomass combustion. Recent research activity has been devoted to the study of other potassium-containing species such as carbonates, hydroxides and sulfates whose corrosion mechanisms are less explored in the literature. The corrosive action of potassium carbonate (K2CO3) on the 304L stainless steel at 500 C and 600 C has been found to be very similar to that of KCl. Both agents lead to chromium depletion of the protective oxide scale by the formation of potassium chromate according to Eqs. (12) and (13), thus making the steel surface less resistant to oxidation [54].
1=2 Cr2 O3 s 2KCls H2 Og 3=4 O2 g ! K2 CrO4 s 2HClg 12

1=2 Cr2 O3 s K2 CO3 s 3=4 O2 g ! K2 CrO4 s CO2 g

13

Lehmusto et al. [55] conrmed the aggressive character of K2CO3 to Cr2O3-protected stainless steels. The proposed mechanism for the active oxidation of Cr2O3-protected stainless steels initiates with the destruction of this passive layer through the reaction with KCl or K2CO3, forming K2CrO4 or K2Cr2O7. In the next step, oxidation continues if there is chloride available to react with the exposed metal but it does not proceed in the absence of chloride. The investigation has been extended by the same authors by considering the combined effect of water vapor and potassium carbonate [56] on the high temperature corrosion behavior of 304L stainless steel which is a more realistic environment for biomass combustion processes than under the dry conditions employed in their previous work [55]. The main effect of humidity was related to the increase of the oxide layer thickness in comparison with the dry conditions whereas the oxide structure was not signicantly affected. In another publication [56] the same group showed that the corrosive character of KCl was higher than that of K2CO3 for a

R.A. Antunes, M.C.L. de Oliveira / Corrosion Science 76 (2013) 626

high nickel alloy (Alloy 625) and a ferritic stainless steel (10CrMo). The corrosion resistance of the nickel-based alloy was higher than that of the ferritic stainless steel, showing the benecial effect of nickel as an alloying element in order to produce a more homogeneous and protective oxide scale. The relevance of alkali carbonate to the high temperature corrosion during combustion processes have also been evidenced by other authors [57,58]. Blomberg [59] extended the analysis of the corrosion mechanisms involved in biomass combustion by considering the relevance of KOH in this process. The author suggests that KOH condensation and heterogeneous reactions with HCl, SO2, SO3 and CO2 would take place and should be included in experimental works and theoretical calculations concerning the high temperature corrosion in biomass red boilers. When KOH reacts in the gaseous phase forming K2CO3, KCl formation on the superheater tube surface would occur by condensation of KCl(g) or according to:

K2 CO3 s 2HClg ! 2KCls 2CO2 g H2 Og

14

The proposed route is the initial condensation of KOH on the tube surface, followed by its reaction with the ue gases, according to:

KOHl HClg ! KCls H2 Og 2KOHl SO2 g 1=2 O2 ! K2 SO4 s H2 Og 2KOHl SO3 g ! K2 SO4 s H2 Og 2KOHl CO2 g ! K2 CO3 s H2 Og

15 16 17 18

Petersson et al. [54] observed that the reaction of K2SO4 with the protective Cr2O3 scale formed on 304L stainless steel probes is not thermodynamically favored under simulated biomass combustion conditions. This would preserve the protective character of the oxide scale, preventing the depletion of chromium within it by the formation of potassium chromate according to reactions (12) and (13) which typically occur when KCl and K2CO3 are in contact with the stainless steel surface. This result points to a low aggressive character of K2SO4 when compared to KCl and K2CO3. In this regard, the sulfation reaction shown in Eq. (19) could be explored to reduce the corrosion rate of superheater tubes consisting of stainless steels. Yin and Wu [64] have found that SO2 had a positive effect on the control of the corrosion rate of a 316L stainless superheater tube in a biomass-red boiler operating at 500 C. This result was explained by the formation of the less aggressive K2SO4, according to the same route shown in Eq. (19). This reaction is accompanied by the formation of a compact scale consisting of FeSO4 which reduces the mass gain at high temperatures. Davidsson et al. [65] conrmed the validity of this mechanism in a commercial scale boiler. The addition of SO2(g) to the burning atmosphere led to the formation of K2SO4 according to Eq. (19) whereas the addition of HCl shifted this accelerated corrosion of the superheater tubes by the reaction shown in Eq. (12). The benecial effect of sulfur to the reduction of the corrosion rate of superheater materials in biomass combustion has practical implications. Mitigation methods based on this mechanism can be advantageously explored by the co-ring of biomass and coal as will be detailed in Section 3.1. 2.4. Temperature-related issues Corrosion of superheater materials in biomass combustion or waste incineration plants can be affected by both the metal and ue gas temperatures [66]. The thermodynamic stability of the corrosive species has a remarkable dependence on the combustion temperature and, therefore, strongly affects the corrosion process of superheater materials. Nielsen et al. [67] gave a useful analysis of the deposition of potassium salts on superheater tubes during straw ring. The thermodynamic stability of potassium species in straw combustion (Fig. 1) was shown to support and explain the experimental results. As seen in Fig. 1, potassium is present as KCl(s), K2SO4(s) and K2OSiO2(s) at low temperatures. KCl(g) and KOH(g) are the stable species at higher temperatures. K2SO4(s) would be the unique condensed species below 1080 C when sufcient amounts of sulfur are present in the system. K2SO4 condensed rst and, then, KCl condensed on these particles and directly on the deposition probe.

Yet, KOH(g) is stable at typical superheater surface temperatures (400 C to 600 C) if the K(g)/Cl(g) ratio in the gas phase exceeds unity. The condensation of KOH(g) to KOH(l) is likely to occur on the surface of the superheater tubes which is followed by a heterogeneous formation of K2CO3(s). This, in turn, can possibly induce the formation of basic alkali melts on the tube surface [60,61]. This implies that the protective Cr2O3 lm on the surface of stainless steels would be dissolved in contact with alkaline solutions typically formed in biomass combustion. In this regard, this mechanism would be of high relevance to the corrosion processes of superheater materials. 2.3. The role of sulfur In addition to chlorine, sulfur is also released during biomass combustion. Alkali metal sulfates or calcium sulfates are the main species in the condensed phase while SO2(g) and SO3(g) are found in the gas phase [62]. The effects of sulfur-containing species on the corrosion mechanisms involved in biomass combustion have been investigated by several authors. Hansen et al. [63] assessed the inuence of deposit formation on the corrosion of a straw-red boiler consisting of stainless steel tubes. Straw is a chlorine-rich fuel which has been found to cause severe corrosion problems in boilers used for biomass ring [49,50]. The formation of K2SO4 was due to the condensation of gaseous KCl which is sequentially sulfated to K2SO4. In this regard, KCl is initially formed and then reacts with a sulfur-containing compound (SO2 or SO3), yielding K2SO4. The sulfation reaction can be expressed as shown in:

2KCl SO2 1=2 O2 H2 O ! K2 SO4 2HCl

19

The thin and dense K2SO4 layer was believed to act as an effective barrier to progressive corrosion of the superheater. The compact nature of this layer would avoid the diffusion of gaseous chloride species (HCl or Cl2) to the metal surface. This reaction would be favored at high steam temperatures (520 C).

Fig. 1. Thermodynamically stable potassium compounds in straw combustion (reproduced from Nielsen et al. by permission of Elsevier Science, UK [67]).

10

R.A. Antunes, M.C.L. de Oliveira / Corrosion Science 76 (2013) 626

According to the experiments conducted at high temperatures (12001400 C), the addition of SO2 greatly increased the amount of K2SO4 in the condensed deposit with a proportional reduction of the KCl particles. As depicted in Sections 2.2 and 2.3, the corrosion rate of superheater materials depends on the formation of KCl or K2SO4, the lowest rates being associated with the formation of the latter. In this context, the corrosiveness of the combustion atmosphere is a function of its temperature. Theoretical thermodynamic calculations by Otsuka [68] showed that chloride salts condense from the vapor phase form the tube deposits on superheaters and melt at relatively low temperatures. When the tube wall temperature exceeds the melting point of the chloride salts corrosion of the superheater tubes becomes more intense. The most severe corrosion process occurs when the waste chemistry consists of high Cl, low S, moderate contents of Na and K. It is noticeable that the ue gas temperature inuence prevails over the tube wall temperature regarding the formation of potentially corrosive salts on the surface of superheater tubes. The molecular quantity of vapor-condensed KCl and NaCl increased with the ue gas temperature at 550750 C whereas it remained unchanged with the increase of the tube wall temperature. The study reveals that the severe reside corrosion of superheater tubes is a consequence of enhanced condensation of KCl and NaCl salts at high ue gas temperatures. Experimental evidence for the strong effect of ue gas temperature on the corrosion rate of boiler tubes made of carbon steel used in waste combustion plants has been given by Brossard et al. [39]. The corrosion rate increases with the ue gas temperature, especially above 650 C. Two concomitant corrosion mechanisms act in this situation. First, the formation of molten alkali phases above the protective oxide scale, which is dissolved in a uxing mechanism. Second, active oxidation by chlorine should not be disregarded. The melting temperature of the y ash particles produced during biomass combustion is also of prime importance [62] due to the increased corrosion rate in the presence of molten phases [20]. In this context, the composition of the y ash and its corresponding deposit on the surface of the superheater material determine if molten or solid phases will be present at a specic temperature. A survey of melting temperatures of relevant compounds and eutectic mixtures for biomass combustion can be found in [20,69]. The prediction of the melting properties of different y ashes is an important step toward the proper control of molten salt corrosion of superheater tubes [59]. 2.5. Formation of ne particles Biomass combustion releases ne particles (<1 lm) which contribute to the formation of deposits on the surface of superheater materials. The well-known high contents of alkali and chlorine in biomass fuels accelerate the corrosion rate of boiler tubes [70]. Hence, the aerosols produced during biomass combustion are closely related to the corrosion processes of superheater materials ensuing great interest in its formation mechanisms. Jimenez and Ballester have deeply explored this subject [7174]. The basic assumption is that particles form by nucleation/condensation of volatilized matter during the cooling process. Experimental evidence has been found that K2SO4 nucleates between 1300 C to 900 C and KCl is not found in this temperature range. KCl was detected at 560 C, forming a condensed layer on the already existing K2SO4 particles [71]. The composition of the deposits originated from ne particles can be tailored to yield less corrosive species. This can be accomplished by properly controlling the relative SO2(g) to O2(g) concentration in the ue gas as this ratio strongly affects the relative fraction of alkali chloride and sulfate in the ne particles. The formation of K2SO4 depends on the amount of SO3(g)

available in the ue gas and not SO2(g). Thus, the rate limiting factor would be the oxidation of SO2(g) toSO3(g) [72,75,76]. Additives that promote the formation of SO3(g) would therefore be benecial to slow down the corrosion processes of superheater materials used in biomass combustion. This mitigation method will be dealt with in Section 3.2. 2.6. Comparison to other types of hot corrosion Combustion processes associated with coal ring and gas turbine operation lead to hot corrosion of metallic materials. The corrosion mechanisms entailed in these cases have similarities with those typical of biomass ring. This section aims at highlighting such similarities in order to gain a deeper insight into the corrosion mechanisms involved in biomass combustion. 2.6.1. Coal combustion Corrosion problems of metallic components used in coal utilization and conversion systems arises from impurities typically encountered in this carbon-based fuel such as mineral ash, sulfur and chlorine [77]. The basic mechanisms of reside corrosion in coal-red boilers have been described by Nelson and Cain [78] and later reviewed by Harb and Smith [79]. Suldation or suldation-oxidation the main reactions involved in the corrosion attack [80]. The presence of sulfur at low oxygen partial pressures favors the corrosion process of superheater tubes because of the difcult formation of protective scales under this condition [79]. Low melting point deposits consisting of sodium iron trisulfate, potassium iron trisulfate or mixtures thereof form and become molten at the typical operating temperatures during coal combustion, forming slagging deposits on the surface of the superheater tubes [81]. The trisulfates would form by reaction of the alkali sulfate with the metal oxide and sulfur trioxide, according to:

3Na2 SO4 Fe2 O3 3SO3 ! 2Na3 FeSO4 3 3K2 SO4 Fe2 O3 3SO3 ! 2K3 FeSO4 3

20 21

The major fraction of SO3 in the combustion chamber arises from the oxidation of SO2. In this regard the concentration of SO2 would control the extent to which reactions (20) and (21) occur. The alkali sulfates (sodium or potassium) would be formed by the reaction of NaCl or KCl from the coal with H2O and SO2 and O2 in the gas phase, releasing HCl [82]. In this regard, the content of Na, K, Cl and S in the coal is of prime importance for the corrosion processes in coal-red boilers. The potassium and sodium content of coals have been correlated with increasing corrosion rates of superheater materials by the mechanisms depicted above [83]. This deleterious action of alkali metals in coal-red boilers is, therefore, similar to that of biomass combustion in regard to the high temperature corrosion of superheater materials. The alkali content, especially potassium, of biomass is higher and the species formed during biomass combustion are different than those formed during coal combustion. Nevertheless, low melting temperature phases consisting of alkali-based species are active corrosive agents in both cases. Sulfur and SO2 have been reported to accelerate the corrosion of superheater steels due to the formation of non-protective oxide scales [84]. Chlorine, even though recognized as playing a part in the corrosion process during coal ring, has not been clearly correlated with increasing corrosion rates [83]. The effects of S and Cl in biomass combustion are opposite to these. The stability of the oxide layer formed on the metal surface under combustion conditions is very important to withstand high temperature corrosion either in coal- or biomass-red boilers. The simultaneous addition of chromium and nickel as alloying elements has been shown to improve the oxidation resistance of

R.A. Antunes, M.C.L. de Oliveira / Corrosion Science 76 (2013) 626

11

ferrous alloys [79]. Even high corrosion resistant alloys can be attacked depending on the ue gas temperature and concentration of impurities in the coal [77]. Thus, a proper control of coal composition and a conscientious materials selection procedure must be implemented to avoid premature tube failure during coal combustion. The same recommendation is valid for biomass combustion, as will be discussed in Sections 3 and 4. 2.6.2. Hot corrosion in gas turbines Hot corrosion has been dened as the oxidative attack of metals and alloys upon contact of a thin lm of fused salt in an oxidizing atmosphere [85]. This process is of particular interest for gas turbines [86]. The corrosion mechanism is associated with the reaction of the metallic surface with a thin lm of molten Na2SO4 which is the dominant compound in the salt due to its high thermodynamic stability [87,88]. The sulde compounds resulting from this reaction are only produced by the interaction between the metal and the fused salt and not upon reaction with sulfur species in the gas phase. This mechanism diverges from those observed for coal and biomass combustion in which the sulfur species (SO2 and SO3) in the gas phase markedly participate in the corrosion process. The hot corrosion mechanism proposed by Rapp and Gotto [89] is based on the uxing of the protective oxide scale. The protective oxide scale dissolves at the oxide/salt interface and reprecitates within the salt lm in the form of non-protective particles. In this context, oxide solubilities in Na2SO4 have to be carefully assessed. Hot corrosion will proceed if a negative solubility gradient is established by the reduction of the solubility of the oxide in the salt/gas interface, thus leading to the precipitation of the non-protective particles within the salt lm [90]. Chromium has been found to play a major role in the protection ability of the oxide scale formed during hot corrosion. Cr2O3 does not satisfy the negative solubility gradient criterion in the salt lm and does not support the sustained hot corrosion process. Moreover, the formation of solid deposits consisting of Cr2O3 can block reducing sites such as grain boundaries and aws at the salt/metal interface, protecting the metallic material from hot corrosion [90]. In spite of the different mechanisms involved in the high temperature corrosion resulting from biomass combustion and the hot corrosion described by the RappGotto criterion, one common feature must be highlighted. The formation of Cr2O3 scales on the surface of the metallic alloys increases the ability to withstand the operating conditions in both cases. This feature can be explored to properly design and select the engineering alloys which will be used in these environments, irrespective if they are Fe-based or Nibased alloys. This issue will be developed in Section 4. 2.7. Overview and the need for protection methods The aim of the foregoing sections was to give the reader an overview of the main corrosion mechanisms involved in biomass combustion and other related high temperature processes. The complexity of this scenario can be promptly realized by considering the following points: (a) Although potassium plays a major role in the corrosion of metallic materials in biomass-red boilers, several other species can participate in this process such as Na, S, Cl, Si, Ca and Mg. The chemical composition of biomasses from different sources and locations can vary signicantly, leading to the formation of different compounds during the burning operation. (b) The temperatures at which the combustion process is carried out will determine if a species will be released into the gas phase or will be in the molten state at the surface

of superheater tubes, depending on its thermodynamic stability. The corrosion process will proceed with different kinetics in each case. (c) The formation of either aggressive species such as KCl and K2CO3 or more protective ones such as K2SO4 in solid deposits can be controlled by the ue gas composition. However, reliable thermodynamic models to predict the intermediate gas phase reactions that lead to the formation of each species are not yet readily available for real biomass combustion conditions. Thus, empirical data have to be obtained in order to conrm the formation of specic compounds. (d) The metallic alloy used as the superheater material gives rise to different interactions with the combustion environment depending on its chemical composition. Alloying elements such as chromium, nickel and molybdenum are known to have strong effects on the corrosion behavior of superheater materials in biomass-red boilers. However, the effect of other elements with potential to form stable oxide layers such niobium, tantalum, vanadium, titanium and aluminum is much less investigated. In spite of various unresolved questions ensuing from the points depicted above, the literature provides well-established data regarding the corrosion mechanisms in biomass combustion. Active oxidation by chlorine, the pernicious inuence of alkali chlorides and alkali carbonates, the benecial effect of adding SO2 to the ue gas, the protective character of the Cr2O3-based scales and the benecial action of nickel as an alloying element have been shown by several authors as described in the previous sections. The main question which is pursued in this work is how this information can be used to support materials selection for superheater tubes used in biomass combustion. Furthermore, how mitigation methods derive from the knowledge accumulated so far regarding the corrosion mechanisms involved in biomass corrosion? Next section deals with this question. It can be anticipated that the benecial effect of SO2 resulted in the development of co-combustion processes based on the concomitant burning of coal and biomass or on the addition of sulfur-based additives. Furthermore, modifying the surface of the superheater materials by depositing protective coatings based on the formation of corrosion resistant oxide scales is also a possible mitigation route. This route is based on the fact the Cr2O3-forming alloys and nickel-containing alloys have increased corrosion resistance under biomass ring conditions. Hence, the use of co-combustion, additives and coatings as mitigation methods for preventing corrosion during biomass combustion is explored in the next section. The analysis of the materials selection problem will be developed in Section 4.

3. Mitigation strategies 3.1. Co-combustion Avoiding chlorine deposition on the surface of the superheater tubes is a traditional method of preventing corrosion during biomass combustion [91]. The sulfation of metal chlorides is an important reaction in this scenario. Sulfur dioxide reacts with alkali chlorides, entrapping the alkali and releasing HCl according to Eq. (8). In this regard, sulfur-rich fuels can be used to prevent chlorine deposition by properly mixing with chlorine-rich biomasses. Aluminumsilicates acts through a similar mechanism. Aho and Silvennoinen [92] studied the prevention of biomass ash-related corrosion by co-combusting biomass fuels. The fuels were chosen based on the elemental composition, aiming at achieving stoichiometric relations that would favor the sequestration of

12

R.A. Antunes, M.C.L. de Oliveira / Corrosion Science 76 (2013) 626

alkalis from the alkali chlorides, according to Eqs. (22) and (23), where M = K, Na:

2MCl SO2 1=2 O2 H2 O ! M2 SO4 2HCl Al2 O3 SiO2 2MCl H2 O ! M2 O Al2 O3 :2SiO2 2HCl

22 23

Pine bark, agricultural waste (AGW) and pulp sludge were selected to the study. Depending on the initial contents of alkali, sulfur, Al2O3, chlorine and SiO2 in the fuels the authors showed that co-combustion favored the inhibition of chlorine deposition especially by the mechanism shown in Eq. (23). In another investigation, Aho and Ferrer [93] studied the effect of coal ash composition in protecting the boiler against chlorine-induced corrosion during combustion of meat and bone meal (MBM). They observed that high aluminum and silicon concentrations in the coal were benecial to prevent chlorine deposition. The formation of alkali aluminosilicates (Eq. (23)) was considered to be the active path of alkali sequestration, avoiding alkali chloride deposition. The presence of sulfur did not prevent chlorine deposition evidencing the little effect of the sulfation reaction in comparison with the formation of alkali aluminosilicates. The protection mechanism based on Eq. (23) was also observed by other authors [9497]. Zheng et al. [98] have also conrmed the viability of this co-combustion strategy. They gave a further understanding on this mechanism by showing that Ca and Mg play important roles in the alkali sequestration ability of the coal/biomass blend. The Ca/Si mole ratio determines the availability of Ca to react with silicates according to the following reactions:

et al. [101] have also pointed out the effectiveness of this treatment at reducing chloride-induced corrosion of superheater materials in wood-red boilers. Recently, Kassman et al. [102,103] conrmed that the use of ammonium sulfate as an additive reduced the content of chlorine in deposits to negligible amounts during biomass combustion in a large-scale circulating uidized bed boiler. Viklund et al. [104] have found that ammonium sulfate greatly reduced the amounts of chlorine in the ue gas and deposits on cooled superheater probes during combustion experiments in which mixtures of municipal and industrial wastes were burned. Schoeld [105] showed that molybdenum salts can avoid the formation of pernicious alkali chlorides deposits in combustors by the formation of the more stable Na2Mo2O7 or K2Mo2O7 compounds. Kaolin has been reported as an additive for biomass combustion by several authors [106108]. The aluminosilicates in kaolin will bind potassium according to the possible reactions shown below [109]:

Al2 Si2 O5 OH4 2KCl ! 2KAlSiO4 H2 O 2HCl Al2 Si2 O5OH4 2KCl 2SiO2 ! 2KAlSi2 O6 H2 O 2HCl Al2 Si2 O5 OH4 K2 SO4 ! 2KAlSiO4 2H2 O SO3 Al2 Si2 O5 OH4 K2 SO4 2SiO2 ! 2KAlSi2 O6 2H2 O SO3

27 28 29 30

CaOs SiO2 s ! CaO SiO2 s CaOs SiO2 Al2 O3 s ! CaO SiO2 Al2 O3 s

24 25

A high content of calcium and magnesium has a negative effect on the binding of potassium and, consequently, on the formation of the potentially corrosive KCl compound. This occurs because of the possible reaction of calcium and magnesium with aluminosilicates, so that this last species becomes less available to react with KCl, releasing chlorine as HCl to the gas phase. In this regard, K/Si and K(S + Si) mole ratios are very important to reduce the chlorine-induced corrosion problems of the co-combustion process. The increase of the S/Cl mole ratio in the fuel favors the formation of the less sticky and higher melting point K2SO4 instead of KCl due to the sulfation reaction of KCl. These observations reveal that a proper control of the fuel composition is imperative to avoid chlorine-induced corrosion problems during the co-combustion of coal and biomass. 3.2. Additives Additives can reduce chlorine-induced corrosion during biomass combustion by two different routes: (i) prevention of gaseous KCl release or (ii) reaction with KCl forming less corrosive species [99]. Sulfur-containing additives are suitable to this purpose as they can react with alkali chlorides yielding less corrosive alkali sulfates. The traditional concept based on this approach is a patented process known as ChlorOut [100]. In this process an aqueous solution of ammonium sulfate is sprayed in the turbulent zone at the entrance of superheaters to induce the sulfation of gaseous alkali chlorides in the ue gas. The reactions proceed rstly by decomposition of ammonium sulfate into NH3 and SO3, according to the reaction shown below:

The compounds KAlSiO4 (kalsilite) and KAlSi2O6 (leucite) have high melting points (1600 C for pure kalsilite and 1500 C for pure leucite), leading to less corrosion problems in comparison with the low melting point KCl and K2SO4 compounds that are likely to create deposits on the surface of superheater [110]. The high chemical inertness and relatively small specic surface of kaolinite are limiting factors for binding the alkalis present in biomasses. Mroczek et al. [111] gave a valuable contribution toward the solution of this challenge. They replaced kaolinite by halloysite as an additive during combustion of different biomasses. The advantage of using halloysite, an aluminosilicate clay mineral (Al4(OH)8Si4O1010H2O), instead of kaolinite resides in its mineralogical structure. This structure is represented in Fig. 2 and consists of Si tetrahedra and Al octahedra, forming a single plate which is separated from the next layer by a space in which potassium cations can get intercalated. The high reactivity of halloysite can be explained by the presence of two hydroxyl groups: one located internally between tetrahedral and octahedral layers and another supercial group located on the octahedral layer. Such disposition gives rise to an interesting bonding structure where the bonds between ions are allowed to occur both on the surface layers and inside the crystal. Mroczek et al. showed that suitable blending of halloysite with biomass can reduce the high temperature corrosion during combustion, effectively diminishing KCl and NaCl concentrations in the ash deposits. 3.3. Coatings The troublesome situation of high temperature corrosion of metallic materials induced by chlorine during biomass combustion and co-combustion can be effectively managed by employing protective coatings. Schtze et al. [112] identied key factors for protective coatings used in high temperature applications. The phases that constitute the coating layer should be thermodynamically stable in the operating environment. Interdiffusion between coating and substrate plays a key role in the protective character of the coating layer. It should occur as slowly as possible to avoid premature coating failure. Moreover, the coefcients of thermal expansion of the substrate, coating and scale should be as close as possible in order to prevent thermal stresses during cooling/heating

NH4 2 SO4 aq ! 2NH3 g SO3 g H2 Og

26

The next step is the sulfation of alkali chlorides by the reactive SO3(g). The effectiveness of this method at reducing the corrosion rate and deposit was conrmed by Brostrm et al. [99]. Henderson

R.A. Antunes, M.C.L. de Oliveira / Corrosion Science 76 (2013) 626

13

Fig. 2. Scheme showing the structure of halloysite with the possible sites for absorbed ions (reproduced from Mroczek et al. by permission of Elsevier Science, UK [111]).

cycles. In addition to these aspects, the coating process is also of paramount importance to the performance of the protective layer. The deposited lm should have a dense and at splat structure to resist corrosive environments at high temperatures. As the corrosive species mostly diffuse through pores and along splat boundaries a dense and at splat structure increases the distance from the coating surface to the coating/substrate interface thereby slowing down the corrosion process of the coated substrate [113]. These general guidelines constitute a complex materials selection problem. The designer must deal with the interaction between substrate and coating, seeking for a compatible thermal expansion behavior without neglecting the thermodynamic stability of the phases that constitute the coating. Furthermore, the selected coating material must be produced as densely as possible to be an effective barrier against corrosion. Thus, a proper deposition method should be also carefully selected to achieve the required morphological architecture for the coating material. In light of this challenging scenario, several authors have pursued efcient coatings to prevent the high temperature corrosion of metallic materials in biomass fueled boilers and waste incinerators [114116]. Hearley et al. [117] showed that coatings produced by the high velocity oxygen fuel (HVOF) thermal spray process form effective barrier against high temperature corrosion. This deposition method is frequently reported for the production of coatings designed to resist high temperature oxidation in a variety of industrial applications [118120]. The attractive corrosion resistance of HVOF coatings comes from the intrinsic lm characteristics such as high density, low oxide levels and increased thickness [121] which make them superior to coatings obtained by other spraying processes such as plasma spraying [122]. The protection ability of HVOF coatings can be further improved by post-treatments that remelt the deposited layer, removing interconnected porosity and oxides from the splat boundaries. Laser melting is reported to be an effective post-treatment of HVOFdeposited Ni-based coatings under biomass combustion conditions [123,124]. In addition to the coating microstructure, its chemical composition is also of great relevance to the corrosion protection ability during biomass combustion. Studies by Mayoral et al. [125] and Torrell et al. [126] showed that the simultaneous presence of chromium and nickel in the coating layer is benecial to confer a reliable protective character. Coatings that allow the formation of homogeneous oxide scales consisting of Cr2O3 and Ni-based oxides decreases the corrosion rate of the metallic substrate. This is in accordance with oxidation behavior described by the corrosion

mechanisms involved in biomass combustion as depicted in Section 2 of the present work. Aluminide coatings produced by pack cementation have been explored as another protection route for superheater alloys [127]. This method is an in situ chemical vapor deposition (CVD) process in which the substrate surface is enriched with elements that are able to form a protective oxide layer in service such as Al, Cr, Si and combinations thereof. The coating is produced by diffusion of these elements at high temperatures. The pack cementation mixture consists of the metal elements to be deposited, halide activators and inert ller such as alumina [128]. The formation of Al2O3 on the surface of the treated substrate accounts for its enhanced oxidation resistance. Films obtained by physical vapor deposition (PVD) methods are hardly considered for the high temperature corrosion protection of metallic materials used in biomass ring or coal-biomass co-ring. Leyens et al. [129] showed that, depending on the microstructure obtained after deposition, the anticorrosion performance of the PVD layer can be sustained for long periods in contact with Na2SO4. However, intrinsic defects such as pinholes and macroparticles [130] which are typical of conventional PVD processes should be avoided. In this regard, multilayered coatings should provide enhanced resistance to high temperature corrosion. This coating architecture allows the formation of tailored lms with a dense and compact structure [131], improving the barrier effect against corrosive species. This approach can be advantageously explored to develop effective coatings to protect metallic materials against high temperature corrosion in biomass combustion applications. 3.4. Nanostructured materials Nanotechnology has emerged as a hot topic in corrosion research over the last two decades. The interest of developing nanostructured materials comes from the possibility of achieving improved mechanical and physical properties in comparison with the conventional polycrystalline counterparts. Fatigue and wear properties, as well as static mechanical strength have been reported to be enhanced by nanostructuring [132134]. The corrosion resistance of metallic materials can be also positively affected by nanostructuring. This can be achieved by two different approaches. The rst one is based on the top-down method in which an existing coarse-grained material is processed to produce a nanostructured alloy by means of a signicant grain renement. The second one is based on a bottom-up method in which a nanostructured material is assembled from nanoscale building blocks

14

R.A. Antunes, M.C.L. de Oliveira / Corrosion Science 76 (2013) 626

such as nanoparticles [135]. This is the case of nanostructured coatings. Both approaches are discussed inthis section, focusing the effect on the corrosion resistance of metallic alloys and their possible applicability for producing improved superheater materials for biomass combustion. 3.4.1. Top-down approach The top-down approach has been developed mainly by severe plastic deformation (SPD) processes such as equal channel angular pressing (ECAP), accumulative roll-bonding (ARB) and high pressure torsion (HPT) [136,137]. These processes are capable of producing bulk nanocrystalline metallic alloys for structural applications [138]. The conventional feedstock for the SPD methods is in the form of bars, rods or sheets. However, tubular components have also been produced using SPD processes [139,140]. Recently, Faraji et al. [141,142] proposed new SPD methods for obtaining tubular parts called tubular channel angular pressing (TCAP) and parallel tubular channel angular pressing (PTCAP). These processing routes would be of interest to the production of superheater tubes for application in biomass-red boilers. Notwithstanding, it is important to notice that most part of the investigations regarding the production of bulk nanocrystalline metallic materials by SPD methods are focused on magnesium or aluminum alloys [143,144]. As will be highlighted in Section 4, the superheater alloys employed for biomass combustion purposes are mostly stainless steels or nickel alloys. If one considers the processing of these materials by SPD methods, the available reports are much scarcer than those for either Al or Mg alloys. This is especially true for nickel alloys or even for pure nickel [145]. One exception can be found in the report by Kim and Kwun [146] who successfully processed a Ni-based alloy (IN718) by ECAP. SPD-processed stainless steels, in turn, are more frequently reported [147,148]. Zheng et al. [149] prepared nanocrystalline 304 stainless steel by ECAP and evaluated the resulting corrosion resistance in comparison to a conventional microcrystalline alloy. The authors underlined that corrosion studies of SPD-processed stainless steels are very scarce in the open literature. The ECAPed samples showed higher corrosion resistance than the as-received alloy. This result was attributed to the improvement of the compactness and stability of the passive lm after the ECAP processing. Enrichment of chromium in the passive lm was not observed. However, it is important to notice that the experiments were conducted at room temperature. Thus, if high temperature tests had been performed it would be possible that diffusion of chromium through the nanocrystalline grains would had been favored, thereby increasing the concentration of this element in the passive lm. Whilst room temperature corrosion is of little relevance to the metallic alloys employed as superheaters in biomass-red boilers, the improved corrosion resistance of SPD-processed metals can be regarded as an opportunity to develop new and optimized materials for this application. In this context, high temperature corrosion studies of SPD-processed metals are a fertile research eld yet to be explored. Since the compactness and stability of the passive lm is altered by the severe plastic deformation process, it is likely that the corrosion behavior of the metallic alloy can be improved even at high temperatures. Moreover, diffusion of passivating elements toward the passive lm could give a further enhancement for the corrosion resistance. Nevertheless, in addition to corrosion resistance, the mechanical strength at elevated temperatures is an important issue to be addressed when designing metallic materials for structural applications. Creep strength and ductility are the key properties for this situation. Creep deformation is a time-dependent phenomenon which is active at temperatures above 50% of the materials melting point (Tm). It occurs by diffusion and dislocation motion through mechanisms associated with grain boundary sliding or

bulk diffusion [150]. Atomic diffusion is facilitated for nanocrystalline materials [151]. Consequently, creep is likely to occur at lower temperatures and stress levels [152]. The lack of investigations on the creep behavior of nanocrystalline metals has been highlighted by several authors [153,154]. These reports generally conrm that the creep strength decreases with the grain size reduction whereas the ductility has an opposite trend. This behavior constitutes a further concern for the application of nanocrystalline metals as superheater tubes in biomass-red boilers. In addition to high temperature corrosion, experiments and reliable models of the creep behavior of nanocrystalline metallic alloys are still of limited availability in the current literature. In one hand, this scarceness of information may constitute a barrier for the prompt adoption of nanocrystallization through SPD processes as a viable method to avoid high temperature corrosion of superheater tubes. But, on the other hand, it also provides an irrefutable opportunity of conducting unprecedented research studies on the structural applications of bulk nanocrystalline metals. In parallel to the development of bulk nanocrystalline materials by SPD processes, the concept of surface nanocrystallization has been explored as well [155]. Different techniques are employed to produce nanocrystalline surfaces in metallic materials such as high energy shot peening, ultrasonic shot peening and surface mechanical attrition treatment (SMAT) [156158]. These processes, generally designated as surface severe plastic deformation (S2PD) methods, have been developed as a tentative to overcome manufacturing obstacles to the widespread industrial production of bulk nanocrystalline metals regarding both the size and cost of nished parts. SMAT has been particularly applied to attain surface nanocrystalline aluminum, magnesium, nickel, titanium and stainless steel components [159163]. In this technique surface nanocrystallization is achieved by shots or milling balls impacts, thus promoting grain renement induced by plastic deformation [164]. The effects of SMAT on the room temperature corrosion resistance of metallic alloys have received much attention [159,165,166]. Depending on the processing parameters corrosion resistance can be either improved or diminished [167,168]. Surface defects induced by plastic deformation can decrease corrosion resistance whereas the stability of the passive lm formed on the nanocrystalline surface can be enhanced due to the faster diffusion of passivating elements, thus increasing the corrosion resistance of the alloy. While room temperature investigations are commonly carried out to assess the performance of nanocrystalline metallic surfaces against corrosion, high temperature studies have been hardly reported. It is generally agreed that the properties of nanocrystalline surfaces depend on their thermal stabilities [169]. It is well-known that nanocrystalline materials have an excess of free energy which is associated with the high numbers of interfaces at grain boundaries. Thus, they are in a metastable state and, upon heating, thermally induced grain growth is likely to occur, affecting the physical properties of the treated surface [170]. This effect is dependent on both the metallic material and the method employed for nanocrystallization. To the best of our knowledge, these aspects are not investigated in the current literature related to the superheater materials used in biomass combustion. Attempts of clarifying the high temperature stability of S2PDtreated metals can be found in the recent report by Wen et al. [171]. These authors obtained nanocrystalline surface in commercial titanium plates using SMAT and studied their oxidation behavior between 500 and 700 C. The SMAT treatment formed numerous grain boundaries and dislocations on the surface of the titanium plates. The presence of these crystalline imperfections allowed the formation of a thicker and denser oxide layer in comparison to that formed on the coarse grained material. The formation of a thick and dense oxide layer can be benecial to avoid corrosion induced by aggressive ions such as chlorine, provided that it does

R.A. Antunes, M.C.L. de Oliveira / Corrosion Science 76 (2013) 626

15

not spall from the metallic surface. This effect is similar to that of a protective coating with a dense and compact structureas those described in Section 3.3 for the protection of superheater tubes against chlorine-induced high temperature corrosion during biomass combustion. However, these results are referred to titanium alloys. Typical superheater materials for biomass combustion are mainly stainless steels or nickel-based alloys. For these materials high temperature corrosion studies are not encountered for SMAT-treated samples. In this context, research activity in this area is needed, especially focusing typical biomass combustion environments. The potential of S2PD processes to improve the oxidation resistance of superheater materials used in biomass combustion is yet to be explored. 3.4.2. Bottom-up approach The bottom-up approach is focused on the development of nanocrystalline coatings from nanocrystalline powders such as in the HVOF process or spray-based processes or on nanocrystalline lms grown directly on the substrate such as physical vapor deposition or chemical vapor deposition methods (CVD). Nanostructured lms based on NiCr alloys have been reported to increase the oxidation resistance of metallic substrates. This is due to the development of a dense and compact lm structure which acts as an effective barrier against penetration of corrosive compounds such as Na2SO4 or Na2SO4K2SO4 mixtures. Moreover, the fast formation of a protective Cr2O3 oxide scale on the surface of the nanocrystalline coating is believed to be responsible for its increased hot corrosion resistance [172174]. According to Yu et al. [175] this is also valid for Al-containing Ni-based nanocrystalline coatings which form a Al2O3 protective scale more rapidly than their conventional microcrystalline counterparts. Furthermore, rareearth elements such as La and Ce and their oxides are reported to increase the hot corrosion resistance of Al2O3 and Cr2O3-forming coatings by the renement of the microstructure which facilitates the formation of protective scales [176,177]. Rahman et al. [178] studied the hot corrosion behavior of nanocrystalline Cr/CoAl coatings on a nickel super alloy substrate. The coatings were deposited using PVD-based method. The nanostructured coatings inhibited the movement of anions and cations, thus providing high hot corrosion resistance. This high performance was also supported by the dense structure of the coating. These results evidence the benecial effect of nanostructuring for improving the corrosion resistance of the Ni-based alloy. Another route for obtaining dense nanocrystalline coatings on metallic substrate has been developed by Mohammadnezad et al. [179]. They employed mechanical alloying to obtain nanostructured NiAl coatings on carbon steel substrates. The results showed that the proper control of processing parameters such as milling time can lead to the formation of dense nanostructured lms with enhanced hardness. The high temperature corrosion resistance of the coated-steel was not evaluated in this work. However, the dense structure of the nanostructured lm points toward a possible positive effect of the mechanical alloying process on this property. In the same way, DC/RF magnetron sputtering was successfully used by Rahman et al. [180] to produce nanocrystalline NiAl coatings with high corrosion resistance at elevated temperatures. The excellent performance of the NiAl was ascribed to a dense morphology of oxide layer developed during oxidation of the nanosized grains, forming a protective scale consisting of Cr2O3, Al2O3, NiO, Fe2O3 and NiCr2O4. Zhong et al. [181] showed that not only nanocrystalline NiAl but also nanocrystalline Ni AlN coatings can be produced by RF magnetron sputtering with dense structures tailored by the proper control of the substrate bias and showing increased oxidation resistance. It is interesting to note that substrate bias inuences the ion bombardment during deposition process and, consequently, controls the grain size of the

crystalline phases. Thus, the microstructure of the coating can be easily tailored during the deposition process. This can be advantageously used to protect superheater materials with nanocrystalline dense PVD lms used in biomass red-boilers.

3.5. Relationship between corrosion mechanisms and mitigation methods: the outputs for materials selection At this point, it is important to identify how the mitigation methods described in this section are related to the corrosion mechanisms reviewed in Section 2. Ultimately, this information will serve as the basis for the materials selection methodology developed in Section 4. In this respect, two major aspects regarding the corrosion mechanisms involved in biomass combustion have to be highlighted. The most critical issue is the remarkable effect of alkali chlorides, especially potassium chloride, in the enhancement of the corrosion rate of the metallic alloys used as superheater tubes. Reactions between the metallic surface and this species, or other alkali-containing species such as carbonates, have to be avoided in order to increase the corrosion resistance of the superheaters. The second aspect to be emphasized is that the formation of a dense and compact oxide lm decreases the corrosion rate of superheater materials under biomass combustion conditions and this is favored for Cr2O3-forming alloys. A further contribution to decreased corrosion rates can be obtained when nickel and chromium are concomitantly present as alloying elements. Mitigation methods based on the co-combustion of coal with biomasses are designed to prevent the reaction of the metallic surface with alkali chlorides by the formation of less aggressive alkali sulfates. The effectiveness of this method depends mainly on the sulfur content of the coal and also on the alkali and chlorine content of the biomass. The use of additives is also based on this philosophy. Either Sulfur-containing or aluminosilicates-containing additives can sequestrate the alkali species during biomass combustion, forming alkali sulfates or aluminosilicates, instead of alkali chlorides, thus avoiding corrosion. Coatings, in turn, rely on the formation of a dense layer on the surface of the metallic substrate to protect it from corrosion. The dense lm avoids penetration of aggressive species such as alkali chlorides that would otherwise react with the substrate. Furthermore, the stability of the coating layer depends on its chemical composition. In this regard, nickel-based coatings and Cr2O3-forming alloys are preferable to protect superheater materials against corrosion during biomass combustion. Nanostructured materials obtained from either the top-down or bottom-up approaches can also be used to mitigate corrosion problems in biomass combustion. In this case, the increased corrosion resistance of nanostructured surfaces arises from the more rapid formation of protective oxide scales. Obviously, the protective character of the scale will depend on its chemical composition and compactness. Once more, nickel and chromium would be mandatory for a satisfactory performance against corrosion. The effectiveness of protection methods can be improved if they are combined with the proper selection of superheater materials. This, in turn, can be affected by the combustion operating conditions, such as fuel type, fuel composition and temperature. In this regard, corrosion mechanisms, mitigation methods and materials selection are correlated. Indeed, materials selection itself could be envisaged as a mitigation method, as a proper choice of materials can greatly decrease the corrosion problems in specic environments. A direct consequence of the important points raised in this section, is that any materials selection strategy should encompass the chemical composition of the metallic alloy, in order to seek for those which will allow the formation of stable oxide scales under

16

R.A. Antunes, M.C.L. de Oliveira / Corrosion Science 76 (2013) 626

biomass combustion conditions. This topic will be unraveled in the next section. 4. Materials The proper selection of metallic materials is a must-attend issue focusing the successful long-term operation of structural engineering parts in a variety of industrial applications. One of the main concerns regarding biomass combustion resides in the corrosion behavior of superheater tube materials. In this context, we rstly present a list of the metallic alloys employedby different authors as superheater materials for biomass combustion, displaying the chemical composition of each alloy. This information, shown in Table 1, is important to guide the selection of the best candidates for the application. Next, a selection strategy based on technical criteria to the design of metallic alloys with improved corrosion resistance under the typical conditions of biomass combustion is discussed. 4.1. Materials selection strategy The development of a materials selection strategy depends on the denition of reliable technical criteria which will allow to clearly identifying the best candidates for a specic engineering application. Condensing the data reviewed above and establishing direct guidelines to properly select superheater materials for biomass combustion is a challenging task. The Ashby approach was chosen as the strategy to meet this goal. This methodology was developed by Prof. Ashby at the Cambridge University. It is based on the use of the CES (Cambridge Engineering System) software to construct materials property charts. The relevant aspects of this methodology are described below, supporting the arguments discussed thereafter. A detailed view of this strategy is found in [196]. 4.1.1. Materials selection using the Ashby approach The Ashby method is design-driven. It starts by identifying the desired materials properties prole. The next step is to compare this prole with real engineering materials, thus highlighting the best match. The identication of the attributes which are important to an specic application is based on a translation step consisting of four different actions: (i) to express the function(s) that the material will perform; (ii) to label the constraint(s) for the application; (iii) to dene the objective(s); (iv) to recognize which variables the designer is free to decide. After translation, the next step is to screen all the available materials using the constraints. By doing so, the number of candidates is signicantly reduced in comparison with the initial materials universe. Then, candidates that passed the screening step are ranked using the objective(s). Finally, the designer should search for supporting information regarding, for instance, the history of the top-ranked candidates owing to the application under consideration or some not-examined aspects such as recyclability or availability. This strategy is schematized in Fig. 3. The constraints are dened based on relevant materials properties to the focused application and also by dimension requirements of the component. The Ashby method proposes the graphical representation of these properties in Cartesian axes, using logarithmic scales. By doing so, two different properties can be simultaneously explored and the performance of a given material in respect to these properties can be promptly evaluated. This graphical representation is known as Ashby chart or material property chart. It is considered a powerful tool to the selection of engineering materials, granting for the analysis of different inter-related properties in a simple and rapid way. Then, the constraints are translated into attribute limits which are plotted as horizontal or vertical lines on

the material selection chart. The materials which do not meet the technical requirements established by the constraints are thereby easily screened. The objective(s) is used to rank the remaining candidates, always seeking to maximize a given performance. Ashby proposes that the objectives dene material indices, for which extreme values are sought. The denition of the material indices is done by the derivation of an objective function. This function has the general form presented in:

P f F ; G; M

31

P is the performance of the material which is a function (f) of three different elements: functional requirements (F) that are often some measurable quantity not directly expressing a material property. Examples are the heat ux through a bar, the load it carries, the temperature at which it is exposed, the electrical current owing through it and so on. The second element encompasses the geometric parameters (G) of the component such as length, diameter, thickness or width. The last element comprises the materials properties (M), forming the so-called material index. Deriving M involves the identication of proper equations relating the materials properties that are relevant to the application under study. Ashby gives a deep understanding of this procedure in [196]. Moreover, it is important to realize that the three elements in Eq. (31) are indeed separable and it can be written as follows:

P f1 F f2 G f3 M

32

According to Eq. (32) f1, f2 and f3 are separate functions that can be simply multiplied. Then, it is assumed that the materials properties are independent from the design-related F and G parameters. Consequently, the selection problem is greatly simplied as it is not necessary to solve for the complete picture, only for the materials attributes. When conicting objectives are involved, a tradeoff strategy can be undertaken. The chart method is classied as a screening method for materials selection. It advantageously applies as an initial screening of materials for a given engineering application. Moreover, it can be easily coupled to the eld of mechanical design, being considered a simple and quick method of evaluating the performance of specic candidates. However, it is frequently regarded as a useful tool for selecting materials when only two or three criteria govern the selection process. If, though, multiple criteria are necessary to exploit the case under study, then the method is of limited applicability. Multi-criteria decision making methods can be effectively used in these cases. An excellent introduction to such methods can be found in [197]. Although the Ashby approach has limitations at dealing with situations involving multiple objectives [197], it successfully applies to a variety of components [198200]. This methodology is developed in the present section to screen and rank candidates for superheater materials for biomass combustion plants. 4.1.2. Selection strategy for superheater materials for biomass-red boilers 4.1.2.1. Translation stage: function and constraints. Applying the procedure described above the rst step of the Ashby method is the translation stage, in which we dene the function, constraints, objective(s) and free variables of the materials selection process. The function is promptly dened as superheater tubes for biomass-red boilers. Next, the constraints must be specied. The rst constraint can be specied considering that the maximum service temperature (Tmax) of the superheater material should be at least 400 C. This criterion is envisaged as a must-pass stage. All the materials that do not pass this stage must be screened from the allowable alternatives. A second constraint is dened based on the fracture toughness of the material given as K1C values. This

Table 1 List of superheater materials tested by different authors. Reference Alloy steels Material Composition (wt.%) Cr [29,35,37,38,31,69,99,101,182,183] [25,26,29,31,32,37,38,40,99,101,104,182,184,185] [25] [101,186] [99,101,183] [99,101,182,187] Reference X20 (X20CrMoV121) 2.25Cr-1Mo (T22) T91 T92 15Mo3 13CrMo44 10.3 2.29 8.09.5 9.15 0.90 Ni 0.47 0.44 0.40 0.26 Mo 0.80 0.96 0.851.05 0.5 0.30 0.50 Nb 0.060.10 0.6 Ti Al 0.03 C 0.18 0.080.12 0.11 0.20 0.10 Si 0.30 0.23 0.50 0.22 0.20 0.30 Cu 0.20 V 0.30 0.01 0.20 W 1.70 Mn 0.80 0.59 0.300.60 0.46 0.60 0.70 Fe Bal. Bal. Bal. Bal. Bal. Bal. N 0.05 B 0.0003

Stainless steels Material Composition (wt.%) Cr Ni 21.1 10.1 8.16 10.7 12.5 0.15 12.0 9.65 9.0 14.5 20.0 15.517.5 12.014.0 Mo 0.10 1.90 0.78 0.94 3.1 1.62.0 Nb + Ta 0.56 <1.2 0.40 0.40 1.2 1.2 Ti Al C 0.02 0.01 0.02 0.05 0.07 0.08 0.03 0.03 0.10 0.041.0 0.04 Si 0.40 0.40 0.28 0.42 0.45 0.18 <0.75 0.58 0.20 0.40 0.75 0.300.60 0.300.60 Cu 0.10 3.0 V 0.28 0.22 W 1.72 Mn 1.70 1.20 1.03 1.84 1.91 0.44 2.0 6.25 0.80 0.40 62.0 1.5 1.5 Fe Bal. Bal. Bal. Bal. Bal. Bal. Bal. Bal. Bal. Bal. Bal. Bal. Bal. N 0.04 0.20 R.A. Antunes, M.C.L. de Oliveira / Corrosion Science 76 (2013) 626

[31,32,125,188] [31,4245,54,56] [189,125,188] [3638,69,101,190192] [34] [34,128,187] [35,186] [25,26,34,37,38,99,101,183,188,187,192] [187] [186] [193] [193] Reference

310 304L 304 TP 347H X3CrNiMoN17-13 HCM12 AISI 347 FG Esshete 1250 Super 304 317L HR3C X8CrNiMoNb 16 16 X8CrNiNb 16 13

25.4 18.3 17.8 17.6 17.01 11.67 18.0 15.0 18.0 18.5 25.0 15.517.5 15.017.0

Ni-based alloys Material Composition (wt.%) Cr Ni 30.4 61.3 39.5 61.0 38.042.0 Bal. 30.3 Bal. Bal. Bal. 20.0 Bal. Mo 3.3 8.30 3.2 9.0 0.51.5 8.5 9.0 15.016.5 3.0 15.7 Nb + Ta 3.60 3.5 0.60 Ti 0.7 0.2 Al 0.10 0.2 0.100.40 C 0.01 0.01 0.01 0.05 60.03 60.03 0.07 0.10 0.01 Si 0.40 0.40 0.40 0.20 60.60 60.50 <0.10 0.40 Cu 0.90 2.0 1.50 V W 2.5 Mn 1.60 0.40 0.20 0.20 62.0 60.50 <0.50 1.0 0.20 Fe Bal. Bal. Bal. 2.0 Bal. 3.0 Bal. 8.3 <1.0 <1.50 Bal. 1.20 N 0.100.20 Co

[25,31,32,38,44,45,183,187] [31,32] [30,31] [26,37,184,187,194196] [25] [25] [30,193] [30] [128] [195] [195] [187]

Sanicro 28 Sanicro 65 Alloy 825 Alloy 625 HR11N Sanicro 63 Alloy 800 Alloy 600 IN617 Alloy 59 Alloy 556 Hastelloy C-2000

27.6 20.7 22.4 21.5 27.030.0 21.0 20.3 16.0 22.0 22.024.0 22.0 22.5

17

18

R.A. Antunes, M.C.L. de Oliveira / Corrosion Science 76 (2013) 626

All available materials

Translation (definition of function, constraints, objectives and free variables)

temperature. The materials that meet both constraints simultaneously are concentrated in the upper right part of the chart. These candidates are mainly ferrous metals and some non-ferrous alloys. In order to facilitate the visual inspection of the remaining candidates, the chart K1CTmax was replotted in Fig. 4b, showing only the materials that survived the screening stage. Some representative materials are identied in the chart. Refractory metals such as tantalum alloys, nickel alloys, stainless steels, cast irons, low alloy and high alloy steels are the surviving candidates. 4.1.2.2. Translation stage: objectives. The next step of the translation process is to dene the objective(s) of the selection process. In order to do this, the following question should be made: what should be minimized or maximized? The answer to this question is not absolute and different designers can choose different criteria. In the present selection process, we have based our reasoning on the information collected from the relevant literature concerning the biomass combustion research eld. In this context, a must-attend issue to validate our selection process is to derive our objectives from materials attributes clearly related to the corrosion mechanisms reviewed in Section 2 of the present text. As highlighted in Section 2.7 the chemical composition of the metallic alloy plays a major role in the corrosion process during biomass combustion. The formation of a stable oxide scale decreases the corrosion rate in biomass combustion environments. This is mainly accomplished with Cr2O3-forming alloys and nickelbased alloys. The simultaneous addition of Cr and Ni as alloying elements is reported to greatly enhance the corrosion resistance of superheater materials [79]. Thus, our rst objective will encompass this well-established experimental evidence. Hence, from the information reviewed so far, our rst objective is to maximize the high temperature corrosion resistance in typical biomass combustion environments. A reliable technical criterion is needed to guide the selection and allow different candidates to be compared. At a rst approximation, the pitting resistance equivalent number (PREN) is a common parameter used to rank different stainless steels and nickel-based alloys according to their resistance to pitting corrosion by chloride attack [201]. This number is calculated from the chemical composition of the passivating alloy and depends on the contents of the three major elements associated with the formation of a passive oxide lm with high resistance to localized corrosion, that is, Cr, Mo and N for stainless steels and Cr, Mo and W for nickel alloys. The expressions used to calculate PREN for stainless steels and nickel alloys are shown in Eqs. (33) and (34), respectively [202,203]. In both equations the percentages refer to the mass percentages of the alloying elements.

Screening (materials which do not meet the constraints are screened)

Ranking (classification of the surviving materials according to the objectives)

Seek documentation (further information about the top-ranked candidates)


Fig. 3. Materials selection based on the Ashbys approach.

property is a measure of the resistance to brittle fracture. According to Ashby [196], 20 MPa m1/2 is considered a minimum value of K1C for conventional engineering design. Materials with K1C values below this limit must be screened from the selection process. This stage is performed using the information shown in Fig. 4. This is the Ashby chart for fracture toughness (K1C) versus maximum service temperature. The chart was constructed using the CES Edupack 2009 software developed by Granta Design. The constraints are marked as solid lines in Fig. 4a which shows the materials universe available in the database of the CES software. The materials classes are represented as envelopes with specic colors. In this regard, technical ceramics are represented as yellow,1 vitreous ceramics as magenta, non-technical ceramics as dark yellow, elastomers as cyan, thermosetting resins as dark-blue (navy), thermoplastics as blue, polymer foams as green, wood and natural materials as olive, ferrous metals as dark cyan, and non-ferrous metals as red bubbles. The bigger envelopes represent each family of materials in a more general way and can be used to guide the reader even in the absence of colors. The chart shows a range of values for each property and material. Such spanning results from different porosity levels for ceramics, different heat treatments and cold working conditions for metals, different degree of crystallinity for polymers and so on. As clearly seen, most materials fail at meeting both constraints simultaneously. All the plastics and polymer foams are below the limit for maximum service temperature. Ceramics (both technical and non-technical) are below the limit for fracture toughness and various non-ferrous metals are below the limit for maximum service

PREN %Cr 3:3%Mo 16%N PREN %Cr 3:3%Mo 0:5%W

33 34

1 For interpretation of color in Fig. 4, the reader is referred to the web version of this article.

It is important to realize, though, that PREN is an empiric expression, resulting from data obtained under specic experimental conditions such as immersion in ferric chloride solutions with controlled concentrations of Cl- [204]. Moreover, the expressions used to calculate PREN do not take the nickel content into account. As reviewed in Section 2, the addition of nickel is necessary to develop a more stable oxide scale under biomass combustion conditions [31,32]. Furthermore, Phongphiphat et al. [195] showed that alloying elements which avoid the loss of chromium due to intergranular corrosion, such as Nb, Ta and Ti, are also desirable to improve the corrosion resistance of superheater materials during biomass combustion. The corrosion rate was found to be reduced up to six times in the presence of these elements. Yet, according to Ishitsuka and Nose [46] Mo and W could also improve the stability of the oxide scale. In this regard, we will dene a modied criterion, based on the same philosophy used for PREN, i.e., to give

R.A. Antunes, M.C.L. de Oliveira / Corrosion Science 76 (2013) 626

19

Fig. 4. Fracture toughness maximum service temperature chart.

a relative measure of the oxide layer stability, but focused on the specic case of biomass combustion and use it as one objective of our materials selection process. Thus, we dene this corrosion resistance index (CI) as follows where the percentages are referred to the mass percentages of each element in the alloy:

CI %Ni %Cr 5:0%Nb %Ta %Ti %W %Mo

35

Then, CI must be maximized to maximize the high temperature corrosion resistance in typical biomass combustion environments.

20

R.A. Antunes, M.C.L. de Oliveira / Corrosion Science 76 (2013) 626

Table 2 Desgin requirements for superheater tubes for biomass-red boilers. Design requirements Function Constraints Objectives Free variables Superheater tubes for biomass-red boilers Maximum service temperature above 400 C; fracture toughness (K1C) above 20 MPa m1/2 To maximize the corrosion resistance index (CI); to maximize the yield strength (0.2% offset limit) Tube wall thickness; choice of material

Now, our second objective must be dened. According to Jones [205], creep fractures of boiler tubes cause 10% of all power-plant breakdowns and 30% of all tube failures in boilers. Creep is dened as a time-dependent plastic deformation under static stresses usually signicant at temperatures above one 0.30.5Tm where Tm is the melting temperature of metals [206]. The relevance of creep for the mechanical stability of boiler tubes has been highly recognized by many authors [207209]. Hence, creep strength can be a valuable criterion to rank the best candidates for superheater tubes in biomass-red boilers. Nevertheless, comparing different materials based on this property is a challenging task. Creep behavior depends on both the temperature and the static stress level employed during the tests [210]. Thus, in spite of the availability of engineering data regarding the creep strength of typical boiler materials, several data are obtained according to different experimental conditions regarding either the temperature or the static stress or both parameters. Consequently, creep data of different superheater materials obtained under the same experimental conditions are not readily available from different sources. In this scenario, the tensile yield strength at 0.2% offset obtained according to the procedure described in the ASTM E8 standard [211] can be regarded as a reliable parameter for comparing the strength of different metallic materials to the onset of plastic deformation. It must be clear, though, that the yield strength and the creep strength are different properties, quantifying the response of the metallic material to distinct phenomena. Creep relates to time-dependent plastic deformation at high temperatures whereas the yield strength relates to plastic deformation at room temperature and higher strain rates. Thus, different deformation mechanisms apply to each case. However, the room temperature yield strength is determined directly according to the standardized procedure indicated in ASTM E8. It does not depend on experimental parameters such as the stress level and temperature employed in creep tests. Thus, it can be considered as a more easily comparable parameter to indicate the

resistance of metallic materials to plastic deformation than the creep strength. This means that the yield strength of different candidates which are available in different sources will be comparable, since they are obtained according to a unique procedure indicated in ASTM E8. This led us to choose the yield strength as our second objective. It will be used as a merit index to rank the candidates of the selection process. However, creep strength will not be neglected. As described in Section 4.1.1 the Ashbys approach recommends that, after ranking, supporting information should be seek in order to cope with some not-examined characteristics of the top-ranked materials. In our case, the creep strength of the topranked materials will be examined after the ranking stage. This should give consistent supporting to the conclusions obtained throughout the whole selection process. The translation stage can be now completed by dening the free variables of the selection process. In our case, these will be the wall thickness of the superheater tube and the material used to manufacture it. The design requirements dened in the translation stage are summarized in Table 2. 4.1.2.3. Screening and ranking stages. The screening stage has already been done, as shown in Fig. 4. The materials which passed this stage were easily identied, using the constraints presented in Table 2. The possible candidates have been identied as refractory metals such as tantalum alloys, nickel alloys, stainless steels, cast irons, low alloy and high alloy steels. To proceed with the selection process, these candidates must be ranked according to the objectives dened in the translation stage. Before this step, we will further restrict the surviving materials, thus obtaining a minor set of candidates to the ranking stage. In order to carry out this further restriction we will consider only the materials shown in Table 1 to our ranking stage. These materials have been collected from the relevant literature reviewed in this article and can be envisaged as traditional options for superheater tubes. Hence, cast irons and refractory metals will not be considered to our purposes since these materials are not reported as viable candidates to superheater tubes of biomass-red boilers. This can be a consequence of a relatively low K1C in the case of cast irons or high cost in the case of refractory metals. Thus, only the nickel alloys, stainless steels, low alloy steels and high alloy steels shown in Table 1 will be ranked according to the objectives dened in the translation stage. Fig. 5 is an Ashby chart of the CI values of the alloys shown in Table 1 as the ordinates whereas the yield strength (0.2% offset

120
X8CrNiNb 16 13 232 Hastelloy C-2000 221 Alloy 59 222

Sanicro 63 224

Alloy 625 223

100

X8CrNiMoNb 16-16 232 600 223 Alloy 825 223

Sanicro 65 224 Alloy IN617 223

80

HR11N 230

CI

60

310 228

Sanicro 28 225 Alloy 800 223 317L 228

Alloy 556 223

40

347FG 228 TP347H 228 HR3C 220 X3CrNiMoN17-13 232 T92 220 HCM12 229 X20 233 T22 234

20

304 228 304L 228

Super 304 231 Esshete 1250 220 15Mo3 227 13CrMo44 226

T91 220

0 180

280

380

480

580

Yield strength (MPa)


Fig. 5. CI yield strength chart (See above-mentioned references for further information.).

R.A. Antunes, M.C.L. de Oliveira / Corrosion Science 76 (2013) 626

21

limit) is plotted in the abscissas. These values were obtained from the references indicated as superscripts, next to the alloy designation shown in the chart. The CI values were calculated using Eq. (35). It is worth mentioning that the chart in Fig. 5 is based on the same principles of the CES software but was not plotted using it. The materials displayed in the chart are those investigated by top research groups throughout the world. The CES software version used to plot the chart in Fig. 4 is an educational version and does not contain such information. Then, we chose to use the representation shown in Fig. 5. It follows the same philosophy of the CES software and is perfectly suited to the materials selection process developed in this text. The disposition of the different candidates shown in Fig. 5 reveals that the alloys can be ranked according to three distinct approaches. If one considers that the most important objective is to maximize the yield strength, the best candidates could be ranked according to the following order, considering only the top ten materials: Alloy 625 > X20 > HCM12 > T92 > Sanicro 63  Sanicro 65 > T91  T22 > Alloy 556 > IN617. If one considers that the most important objective is to maximize the CI value, the best candidates are ranked according to the following order, considering only the top 10 materials: Alloy 625 > Sanicro 63 > Hastelloy C2000  Alloy 59 > Alloy 600 > Sanicro 65 > IN617 > HR11N > Alloy 825 > Sanicro 28. The third approach is based on the candidates that give the best compromise between both objectives, but is not necessarily optimal by either of the objectives. In this case, the best candidate is Alloy 625 which is the ultimate choice from both objectives, followed by Sanicro 63 and Sanicro 65 which lie more internally in the chart. It is concluded, thus, that the best candidates based on the CI values are not those that maximize the yield strength. This congures an approach based on the conicting objectives philosophy developed by Ashby [196]. In our selection process we will consider that both objectives have the same relative importance. Hence, the top-ranked candidates are those which give the best compromise between the two objectives. Consequently, Alloy 625 is the best solution to the selection process, followed by Sanicro 63, Sanicro 65. IN617 is also an interesting solution, regarding its excellence in corrosion resistance based on the CI value. This result reveals the remarkable inuence of nickel and chromium in the formation of a stable oxide layer capable of withstanding the corrosion processes taking place during biomass combustion. This is in agreement with results published bydifferent authors. Moreover, our materials selection strategy is further supported by the corrosion mechanisms by Phongphiphat et al. [195] owing to the benecial role of elements that avoid the loss of chromium due to intergranular corrosion such as niobium and titanium, thus forming a more stable and protective Cr2O3 scale. The importance of considering the presence of such alloying elements is promptly recognized when the relative positions of other high-CI alloys such as Hastelloy C-2000 and Alloy 59 (which do not contain either Nb or Ti but are highly alloyed with Ni, Mo and Cr) are compared with the positions of Alloy 625 and Sanicro 63 which contain Nb and are also highly alloyed with Ni, Cr and Mo. Alloy 625 and Sanicro 63 have a higher corrosion resistance index (CI) and should be preferable to the manufacturing of superheater tubes used in biomass combustion. 4.1.2.4. Supporting information. The last stage of the Ashby method is to seek supporting information of the top-ranked materials. As mentioned above, creep strength is an important property for superheater materials used in biomass-red boilers. Thus, creep data of Alloy 625, Sanicro 63, Sanicro 65 and IN617have been consulted. Sanicro 63 and Sanicro 65 are trademarks of Sandvik Materials Technology. They consist of composite tubes developed for

high temperature applications with an outer layer (based on Alloy 625) co-extruded to inner ferritic steel [212]. Sanicro 65 is similar to Sanicro 63 except that it does not contain niobium in its chemical composition [213]. We did not nd information regarding the creep behavior of these materials in the current literature. However, they are recommended for high temperature applications [214] and can be regarded as having at least creep strength comparable to Alloy 625. Finally, comparable creep data could be found for IN617 and Alloy 625 [215,216]. The static tensile stress to produce rupture of Alloy 625 within 100 h at 649 C is reported to be 490 MPa whereas for IN617 it is only 379 MPa. Furthermore, the high creep strength of these materials has been evidenced in the literature [217,218]. Thus, the supporting information focused on creep data conrmed that the best candidate to superheater tubes for biomass-red boilers is Alloy 625. In light of the results encountered in the current literature, it is evident that the corrosion performance of the superheater materials used in biomass combustion or co-combustion is strongly related to the alloy composition. High levels of nickel and chromium increase the stability of the oxide lms and the corrosion resistance of the alloys. Molybdenum and carbide former elements (Nb, Ta and Ti) are also highly desirable. However, it should be noted that, depending on the metal temperature and fuel composition, even highly alloyed materials can be severely corroded. In this regard, any selection strategy for superheater materials should encompass the previous analysis of existing data regarding the compatibility of the intended superheater material with the environment conditions envisaged for its nal use. Frequently, a corrosion protection method will have to be employed to guarantee that the superheater will have a longer service life. Coatings are attractive options in this case. Coated alloys can be regarded as new materials, giving rise to a broader set of candidates for a given engineering application. The materials selector should be aware of this possibility. Yet, in addition to the objectives considered in our selection process, a cost analysis is also mandatory. A highly alloyed material can withstand corrosion during an extended lifetime but its cost is frequently high. A low alloyed material with a proper protective coating could resist in a similar or even better manner with a more attractive nal cost. The literature shows that bare metallic alloys are affected by high temperature corrosion. The extent of corrosion damage can be low for the highly alloyed materials but it is always present. Only coated alloys have been reported to be unaffected by high temperature corrosion [192]. 5. Conclusions A materials selection strategy for superheater materials used in biomass combustion has been developed in this work. Corrosion mechanisms and mitigation methods regarding corrosion in biomass combustion were reviewed to support the selection procedure. A direct interaction between these subjects could be identied. The main ndings of this work can be summarized as follows: (a) Alkali chlorides, especially potassium chloride, increase the corrosion rate of the materials used as superheater tubes. Reactions between the metallic surface and this species, or other alkali-containing species such as carbonates, have to be avoided in order to increase the corrosion resistance of the superheaters. (b) The formation of a dense and compact oxide lm decreases the corrosion rate of superheater materials under biomass combustion conditions. This is favored for Cr2O3-forming alloys. A further contribution to increased corrosion resistance can be obtained when nickel and chromium are

22

R.A. Antunes, M.C.L. de Oliveira / Corrosion Science 76 (2013) 626

concomitantly present as alloying elements. Moreover, niobium, tantalum and titanium prevent the loss of chromium due to intergranular corrosion further improving the corrosion resistance of superheater alloys. (c) Mitigation methods are designed to avoid the reaction of the metallic surface with the aggressive alkali-containing containing species. This can be accomplished by co-combustion strategies based on the simultaneous burning of coal and biomass or by the use of sulfur-containing additives or species capable of sequestrating the alkalis in the biomass, forming less aggressive compounds such as alkali sulfates of aluminosilicates, instead of alkali chlorides. Coatings based on nickel and chromium as the major elements can form stable oxide layers to withstand the corrosion processes during biomass combustion. (d) Nanostructured materials, although little explored in the current literature, are promising alternatives to mitigate corrosion problems in biomass combustion due to the more rapid formation of a stable oxide layer. (e) The Ashbys approach successfully applied to the selection of materials for superheater tubes used in biomass combustion. A corrosion resistance index (IC) was proposed as a technical criterion to selection the best candidates for this application. This index was derived based on the information reviewed from the literature concerning the corrosion mechanisms involved in biomass combustion. The best candidates identied in this work were those which gave the best compromise between corrosion resistance (determined using the IC value of the metallic alloy) and the mechanical strength (considering the tensile yield strength at 0.2% offset obtained according to the procedure described in the ASTM E8 standard). The relevance of the chemical composition of the superheater material was evidenced. In this context, nickel, chromium and niobium have the most striking inuence on the nal choice. The development of metallic alloys to withstand corrosion in biomass combustion can be successfully designed based on the materials selection strategy described in this work.

References
[1] A.A. Khan, W. de Jong, P.J. Jansens, H. Spliethoff, Biomass combustion in uidized bed boilers: potential problems and remedies, Fuel Process. Technol. 90 (2009) 2150. [2] G. Taylor, Biofuels and the biorenery concept, Energy Policy 36 (2008) 4406 4409. [3] K. Openshaw, Biomass energy: employment generation and its contribution to poverty alleviation, Biomass Bioenergy 34 (2010) 365378. [4] A.W. Bhutto, A.A. Bazmi, G. Zahedi, Greener energy: issues and challenges for Pakistan biomass energy prospective, Renew. Sustain. Energy Rev. 15 (2011) 32073219. [5] H. Haberl, T. Beringer, S.C. Bhattacharya, K.-H. Erb, M. Hoogwijk, The global technical potential of bio-energy in 2050 considering sustainability constraints, Curr. Opin. Environ. Sustain. 2 (2010) 394403. [6] K. Srirangan, L. Akawi, M.-Y. Murray, C.P. Chou, Towards sustainable production of clean energy carriers from biomass resources, Appl. Energy 100 (2012) 172186. [7] S. Lim, K.T. Lee, Leading global energy and environmental transformation: unied ASEAN biomass-based bio-energy system incorporating the clean development mechanism, Biomass Bioenergy 35 (2011) 24792490. [8] A.F. Kirkels, Discursive shifts in energy from biomass: a 30 year European overview, Renew. Sustain. Energy Rev. 16 (2012) 41054115. [9] Brasil. Empresa de Pesquisa Energtica. Balano Energtico Nacional 2011 Ano base 2010: Resultados Preliminares Rio de Janeiro: EPE, 2011. [10] P. McKendry, Energy production from biomass (part 2): conversion technologies, Bioresour. Technol. 83 (2002) 4754. [11] M.F. Demirbas, M. Balat, H. Balat, Potential contribution of biomass to the sustainable energy development, Energy Convers. Manage. 50 (2009) 1746 1760. [12] A. Demirbas, Potential applications of renewable energy sources, biomass combustion problems in boiler power systems and combustion related environmental issues, Prog. Energy Combust. Sci. 31 (2005) 171192.

[13] A. Demirbas, Combustion characteristics of different biomass fuels, Prog. Energy Combust. Sci. 30 (2004) 219230. [14] R.A. Khalil, E. Houshfar, W. Musinguzi, M. Becidan, O. Skreiberg, F. Goile, T. Lovas, L. Sorum, Experimental investigation on corrosion abatement in straw combustion by fuel mixing, Energy Fuels 25 (2011) 26872695. [15] A. Williams, J.M. Jones, L. Ma, M. Pourkashanian, Pollutants from the combustion of solid biomass fuels, Prog. Energy Combust. Sci. 38 (2012) 113137. [16] S.V. Vassilev, D. Baxter, L.K. Andersen, C.G. Vassileva, An overview of the chemical composition of biomass, Fuel 89 (2010) 913933. [17] S.V. Vassilev, D. Baxter, L.K. Andersen, C.G. Vassileva, T.J. Morgan, An overview of the organic and inorganic phase composition of biomass, Fuel 94 (2012) 1 33. [18] A. Faaij, J. van Doorn, T. Curvers, L. Waldheim, E. Olsson, A. van Wijk, C. DaeyOuwens, Characteristics and availability of biomass waste and residues in the Netherlands for gasication, Biomass Bioenergy 12 (1997) 225240. [19] S. Vassilev, C. Braekman-Danheux, P. Laurent, Characterization of refusederived char from municipal solid waste: 1. Phase mineral and chemical composition, Fuel Process. Technol. 59 (1999) 95134. [20] H.P. Nielsen, F.J. Frandsen, K. Damp-Johansen, L.L. Baxter, The implications of chlorine-associated corrosion on the operation of biomass-red boilers, Prog. Energy Combust. Sci. 26 (2000) 283298. [21] M. hman, A. Nordin, B.J. Skrifvars, R. Backman, M. Hupa, Bed agglomeration characteristics during uidized bed combustion of biomass fuels, Energy Fuels 14 (2000) 169178. [22] L.L. Baxter, Inuence of ash deposit chemistry and structure on physical and transport properties, Fuel Process. Technol. 56 (1998) 8188. [23] L.-G. Johansson, J.-E. Svensson, E. Skog, J. Pettersson, C. Pettersson, N. Folkeson, H. Asteman, T. Jonsson, M. Halvarsson, Critical corrosion phenomena on superheaters in biomass and waste-red boilers, in: Proceedings of the Sino-Swedish Structural Materials Symposium, 2007, pp. 3539. [24] N. Folkeson, J. Pettersson, C. Pettersson, L.-G. Johansson, E. Skog, J.-E. Svensson, Fireside corrosion of stainless and low alloyed steels in a wastered CFB boiler: the effect of adding sulphur to thefuel, Mater. Sci. Forum 595598 (2008) 289297. [25] B.-J. Skrifvars, M. Westen-Karlsson, M. Hupa, K. Salmenoja, E. Vakkilainen, Corrosion of superheater steel materials under alkali salt deposits. Part 1: the effect of salt deposit composition and temperature, Corros. Sci. 50 (2008) 12741282. [26] B.-J. Skrifvars, M. Westen-Karlsson, M. Hupa, K. Salmenoja, Corrosion of super-heater steel materials under alkali salt deposits, Corros. Sci. 52 (2010) 10111019. [27] K.O. Davidsson, L.-E. Amand, B.-M. Steenari, A.L. Elled, D. Eskilsson, B. Leckner, Contermeasures against alkali-related problems during combustion of biomass in a circulating uidized bed boiler, Chem. Eng. Sci. 63 (2008) 53145329. [28] D.A. Tillman, D. Duong, B. Miller, Chlorine in solid fuels red in pulverized fuel boilers sources, forms, reactions and consequences: a literature review, Energy Fuels 23 (2009) 33793391. [29] H.J. Grabke, E. Reese, M. Spiegel, The effects of chlorides, hydrogen chloride and sulfur dioxide in the oxidation of steels below deposits, Corros. Sci. 37 (1995) 10231043. [30] A. Zahs, M. Spiegel, H.J. Grabke, Chloridation and oxidation of iron, chromium, nickel and their alloys in chloridizing and oxidizing atmospheres at 400 700 C, Corros. Sci. 42 (2000) 10931122. [31] P. Andersson, M. Norell, Field test of superheater corrosion in a CFB waste boiler: part 1 metal loss characteristics, Mater. Corros. 56 (2005) 449 458. [32] P. Andersson, M. Norell, Field test of superheater corrosion in a CFB waste boiler. Part II scale formation characteristics, Mater. Corros. 56 (2005) 550 560. [33] H.H. Krause, High temperature corrosion problems in waste incineration systems, J. Mater. Energy Syst. 7 (1986) 322332. [34] M. Montgomery, A. Karlsson, O.H. Larsen, Field test corrosion experiments in Denmark with biomass fuels. Part 1: straw-ring, Mater. Corros. 53 (2002) 121131. [35] H.P. Nielsen, F.J. Frandsen, K. Dam-Johansen, Lab-scale investigations of hightemperature corrosion phenomena in straw-red boilers, EnergyFuels 13 (1999) 11141121. [36] S.C. van Lith, F.J. Frandsen, M. Montgomery, T. Vilhelmsen, S.A. Jensen, Labscale investigation of deposit-induced chlorine corrosion of superheater materials under simulated biomass-ring conditions. Part 1: exposure at 560 C, Energy Fuels 23 (2009) 34573468. [37] C.C. Cha, High temperature corrosion of superheater materials below deposited biomass ashes in biomass combusting atmospheres, Corros. Eng. Sci. Technol. 42 (2007) 5060. [38] S. Sroda, M. Mkipa, S.-C. Cha, M. Spiegel, The effect of ash deposition on corrosion behaviour of boiler steels in simulated combustion atmospheres containing carbon dioxide (CORBI PROJECT), Mater. Corros. 57 (2006) 176 181. [39] J.-M. Brossard, I. Diop, X. Chaucherie, F. Nicol, C. Rapin, M. Vilasi, Superheater reside corrosion mechanisms in MSWI plants: lab-scale study and on-site results, Mater. Corros. 62 (2011) 543548. [40] T. Jonsson, N. Folkeson, J.-E. Svensson, L.-G. Johansson, M. Halvarsson, An ESEM in situ investigation of initial stages of the KCl induced high

R.A. Antunes, M.C.L. de Oliveira / Corrosion Science 76 (2013) 626 temperature corrosion of a Fe2.25Cr1Mo steel at 400 C, Corros. Sci. 53 (2011) 22332246. T.J. Pan, Y.S. Li, Q. Yang, R.F. Feng, A. Hirose, Internal oxidation and phase transformations of multi-phase FeNiAl and FeNiAlCr alloys induced by KCl corrosion, Corros. Sci. 53 (2011) 21152121. J. Pettersson, H. Asteman, J.-E. Svensson, L.-G. Johansson, KCl induced corrosion of a 304-type austenitic stainless steel at 600 C; the role of potassium, Oxid. Met. 64 (2005) 2341. J. Pettersson, J.-E. Svensson, L.-G. Johansson, KCl-induced corrosion of a 304type austenitic stainless steel in O2 + H2O environment: the inuence of tremperature, Oxid. Met. 72 (2009) 159177. C. Pettersson, J. Pettersson, H. Asteman, J.-E. Svensson, L.-G. Johansson, KClinduced high temperature corrosion of the austenitic FeCrNi alloys 304L and Sanicro 28 at 600 C, Corros. Sci. 48 (2006) 13681378. C. Pettersson, L.-G. Johansson, J.-E. Svensson, The inuence of small amounts of KCl (s) on the initial stages of the corrosion of alloy Sanicro 28 at 600 C, Oxid.Met. 70 (2008) 241256. T. Ishitsuka, K. Nose, Stability of protective oxide lms in waste incineration environment solubility measurement of oxides in molten chlorides, Corros. Sci. 44 (2002) 247263. S.C. Cha, M. Spiegel, Local reactions between NaCl and KCl particles and metal surfaces, Corros. Eng. Sci. Technol. 40 (2005) 249254. S.C. Cha, M. Spiegel, Local reactions of KCl particles with iron, nickel and chromium surfaces, Mater. Corros. 57 (2006) 159164. H.P. Michelsen, O.H. Larsen, F.J. Frandsen, K. Dam-Johansen, Deposition and high temperature corrosion in a 10 MW straw red boiler, Fuel Process. Technol. 54 (1998) 95108. C. Yin, L.A. Rosendahl, S.K. Kaer, Grate-ring of biomass for heat and power production, Prog. Energy Combust. Sci. 34 (2008) 725754. Y. Kawahara, Y. Kaihara, Recent Trends in Corrosion-resistant Tube Materials and Improvements of Corrosion Environments in WTE Plants, Corrosion 2001, Paper No. 01173, Houston, TX, 2001. B.A. Baker, G.D. Smith, L.D. Shoemaker, Performance of Commercial Alloys in Simulated Waste Incineration Environments, Corrosion 2001, Paper 01183, Houston, TX, 2001. J. Lehmusto, B.-J. Skrifvars, P. Yrjas, M. Hupa, High temperature oxidation of metallic chromium exposed to eight different metal chlorides, Corros. Sci. 53 (2011) 33153323. J. Pettersson, N. Folkeson, L.-G. Johansson, J.-E. Svensson, The effects of KCl, K2SO4 and K2CO3 on the high temperature corrosion of a 304-type austenitic stainless steel, Oxid. Met. 76 (2011) 93109. J. Lehmusto, D. Lindberg, P. Yrjas, B.-J. Skrifvars, M. Hupa, Thermogravimetric studies of high temperature reactions between potassium salts and chromium, Corros. Sci. 59 (2012) 5562. J. Lehmusto, B.-J. Skrifvars, P. Yrjas, M. Hupa, Comparison of potassium chloride and potassium carbonate with respect to their tendency to cause high temperature corrosion of stainless 304L steel, Fuel Process. Technol. 105 (2013) 98105. M. Steinberg, K. Schoeld, The controlling chemistry of surface deposition from sodium and potassium seeded ames free of sulfur or chlorine impurities, Combust. Flame 129 (2002) 453470. K. Schoeld, The chemical nature of combustion deposition and corrosion: the case of alkalichlorides, Combust. Flame 159 (2012) 19871996. T. Blomberg, Which are the right test conditions for the simulation of high temperature alkali corrosion in biomass combustion?, Mater Corros. 57 (2006) 170175. T. Blomberg, A thermodynamic study of the gaseous potassium chemistry in the convection sections of biomass red boilers, Mater. Corros. 62 (2011) 635641. T. Blomberg, Correlation of the corrosion rates of steels in a straw red boiler with the thermodynamically predicted trend of KOH(g) in the ue gases, Biomass Bioenergy 39 (2012) 489493. T. Lind, E.I. Kauppinen, J. Hokkinen, J.K. Jokiniemi, M. Orjala, M. Aurela, R. Hillamo, Effect of chlorine and sulfur on ne particle formation in pilot-scale CFBF of biomass, Energy Fuels 20 (2006) 6168. L.A. Hansen, H.P. Nielsen, F.J. Frandsen, K. Dam-Johansen, S. Horlyck, A. Karlsson, Inuence of deposit formation on corrosion at a straw-red boiler, Fuel Process. Technol. 64 (2000) 189209. J. Yin, Z. Wu, Corrosion behavior of TP316L of superheater in biomass boiler with simulated atmosphere and deposit, Chin. J. Chem. Eng. 17 (2009) 849 853. K.O. Davidsson, L.-E. Amand, B. Leckner, Potassium, chlorine, and sulfur in ash, particles, deposits, and corrosion during wood combustion in a circulating uidized-bed boiler, Energy Fuels 21 (2007) 7181. G. Sorell, The role of chlorine in high temperature corrosion in waste-toenergy plants, Mater. High Temp. 14 (1997) 137150. H.P. Nielsen, L.L. Baxter, G. Sclippab, C. Morey, F.J. Frandsen, K. Dam-Johansen, Deposition of potassium salts on heat transfer surfaces in straw-red boilers: a pilot-scale study, Fuel 79 (2000) 131139. N. Otsuka, A thermodynamic approach on vapor-condensation of corrosive salts from ue gas on boiler tubes in waste incinerators, Corros. Sci. 50 (2008) 16271636. M. Montgomery, T. Vilhelmsen, S.A. Jensen, Potential high temperature corrosion problems due to co-ring of biomass and fossil fuels, Mater. Corros. 59 (2008) 783793.

23

[41]

[42]

[43]

[44]

[45]

[46]

[47] [48] [49]

[50] [51]

[52]

[53]

[54]

[55]

[56]

[57]

[58] [59]

[60]

[61]

[62]

[63]

[64]

[65]

[66] [67]

[68]

[69]

[70] B.M. Jenkins, L.L. Baxter, T.R. Miles Jr, T.R. Miles, Combustion properties of biomass, Fuel Proc. Technol. 54 (2002) 1746. [71] S. Jimnez, J. Ballester, Formation and emission of submicron particles in pulverized olive residue (Orujillo) combustion, Aerosol Sci. Technol. 38 (2004) 707723. [72] S. Jimnez, J. Ballester, Inuence of operating conditions and the role of sulfur in the formation of aerosols from biomass combustion, Combust. Flame 140 (2005) 346358. [73] S. Jimnez, J. Ballester, Particulate matter formation and emission in the combustion of different pulverized biomass fuels, Combust. Sci. Technol. 178 (2006) 655683. [74] S. Jimnez, J. Ballester, Formation of alkali sulphate aerosols in biomass combustion, Fuel 86 (2007) 486493. [75] K.A. Christensen, M. Stenholm, H. Livbjerg, The formation of submicron aerosol particles, HCl and SO2 in straw-red boilers, J. Aerosol Sci. 29 (1998) 421444. [76] K.A. Christensen, H. Livbjerg, A plug ow model for chemical reactions and aerosol nucleation and growth in an alkali-containing ue gas, Aersol Sci. Technol. 33 (2000) 470489. [77] H.S. Hsu, Review of the development and breakdown of protective oxide scales on alloys exposed to coal-derived atmospheres, Oxid. Met. 28 (1987) 213235. [78] W. Nelson, C. Cain, Corrosion of superheaters and reheaters of pulverizedcoal-red boilers, Trans. ASME, Ser. A., J. Eng. Power 82 (1960) 194204. [79] J.N. Harb, E.E. Smith, Fireside corrosion in PC-red boilers, Prog. Energy Combust. Sci. 16 (1990) 169190. [80] J. Stringer, A.J. Minchener, High temperature corrosion in uidized bed combustion systems, J. Mater. Energy Syst. 7 (1986) 333344. [81] E.M. Labuda, D.A. Cline Jr., K.J. Shields, Fireside corrosion in coal and oil-red units: failure mechanisms and methods of prevention, in: NACE Conference, Paper 00234, Corrosion 2000, 2000. [82] A.J.B. Cutler, E. Raask, External corrosion in coal-red boilers: assessment from laboratory data, Corros. Sci. 21 (1981) 789800. [83] L. Li, Y. Duan, Y. Cao, P. Chu, R. Carty, W.-P. Pan, Field corrosion tests for a low chromium steel carried out at superheater area of a utility boiler with three coals containing different chlorine contents, Fuel Process. Technol. 88 (2007) 387392. [84] X. Peng, K. Liu, W.-P. Pan, J.T. Riley, High-temperature corrosion of A210-C carbon steel in simulated coal-combustion atmospheres, Oxid. Met. 60 (2003) 117135. [85] R.A. Rapp, Chemistry and electrochemistry of hot corrosion of metals, Mater. Sci. Eng. 87 (1987) 319327. [86] J. Stringer, Hot corrosion of high-temperature alloys, Ann. Rev. Mater. Sci. 7 (1977) 477509. [87] R.A. Rapp, Hot corrosion of materials, High Temp. Sci. 27 (1990) 355367. [88] R.A. Rapp, Y.S. Zhang, Hot corrosion of materials: fundamental studies, JOM 46 (1994) 4755. [89] R.A. Rapp, K.S. Goto, in: J. Braunstein, J.R. Selman (Eds.), Hot Corrosion of Metals by Molten Salts, Molten Salts 1, Electrochem. Soc., Pennington, NJ, 1981, pp. 159177. [90] R.A. Rapp, N. Otsuka, The role of chromium in the hot corrosion of metals, ECS Trans. 16 (2009) 271282. [91] M. Aho, P. Vainikka, R. Taipale, P. Yrjas, Effective new chemicals to prevent corrosion due to chlorine in power plant superheaters, Fuel 87 (2008) 647 654. [92] M. Aho, J. Silvennoinen, Preventing chlorine deposition on heat transfer surfaces with aluminium-silicon rich biomass residue and additive, Fuel 83 (2004) 12991305. [93] M. Aho, E. Ferrer, Importance of coal ash composition in protecting the boiler against chlorine deposition during combustion of chlorine-rich biomass, Fuel 84 (2005) 201212. [94] M. Aho, A. Gil, R. Taipale, P. Vainikka, H. Vesala, A pilot-scale reside deposit study of co-ring Cynara with two coals in a uidised bed, Fuel 87 (2008) 58 69. [95] C. Bartolom, A. Gil, I. Ramos, Ash deposition behavior of cynaracoal blends in a PF pilot furnace, Fuel Process. Technol. 91 (2010) 15761584. [96] M. Aho, P. Yrjas, R. Taipale, M. Hupa, J. Silvennoinen, Reduction of superheater corrosion by co-ring risky biomass with sewage sludge, Fuel 89 (2010) 23762386. [97] M. Aho, T. Envall, J. Kauppinen, Corrosivity of ue gases during co-ring Chinese biomass with coal at uidized bed conditions, Fuel Process. Technol. 105 (2013) 8288. [98] Y. Zheng, P.A. Jensen, A.D. Jensen, B. Sander, H. Junker, Ash transformation during co-ring coal and straw, Fuel 86 (2007) 10081020. [99] M. Brostrm, H. Kassman, A. Helgesson, M. Berg, C. Andersson, R. Backman, A. Nordin, Sulfation of corrosive alkali chlorides by ammonium sulfate in a biomass red CFB boiler, Fuel Process. Technol. 88 (2007) 11711177. [100] ChlorOut, European Patent EP 1354167, International Patent Application: PCT/SE 02/00129, 2002. [101] P. Henderson, P. Szaklos, R. Pettersson, C. Andersson, J. Hgberg, Reducing superheater corrosion in wood-red boilers, Mater. Corros. 57 (2006) 128 134. [102] H. Kassman, M. Brostrm, M. Berg, L.-E. mand, Measures to reduce chlorine in deposits: application in a large-scale circulating uidised bed boiler ring biomass, Fuel 90 (2011) 13251334.

24

R.A. Antunes, M.C.L. de Oliveira / Corrosion Science 76 (2013) 626 [132] T. Roland, D. Retraint, K. Lu, J. Lu, Fatigue life improvement through surface nanostructuring of stainless steel by means of surface mechanical attrition treatment, Scr. Mater. 54 (2006) 19491954. [133] M. Jafari, M.H. Enayati, M.H. Abbasi, F. Karimzadeh, Compressive and wear behaviors of bulk nanostructured Al2024 alloy, Mater. Des. 31 (2010) 663 669. [134] K.A. Padmanabhan, Mechanical properties of nanostructured materials, Mater. Sci. Eng. A 304306 (2001) 200205. [135] Y.T. Zhu, T.C. Lowe, T.G. Langdon, Performance and applications of nanostructured materials produced by severe plastic deformation, Scr. Mater. 51 (2004) 825830. [136] R.Z. Valiev, R.K. Islamgaliev, I.V. Alexandrov, Bulk nanostructured materials from severe plastic deformation, Prog. Mater. Sci. 45 (2000) 103189. [137] V. Maier, T. Hausl, C.W. Schmidt, W. Bhm, H. Nguyen, M. Merklein, H.W. Hppel, M. Gken, Tailored heat treated accumulative roll bonded aluminum blanks: microstructure and mechanical behavior, Metall. Mater. Trans.A 43 (2012) 30973107. [138] Z. Horita, T. Fujinami, T.G. Langdon, The potential for scaling ECAP: effect of sample size on grain renement and mechanical properties, Mater. Sci. Eng. A 318 (2001) 3441. [139] M.S. Mohebbi, A. Zkbarzadeh, Accumulative spin-bonding (ASB) as a novel SPD process for fabrication of nanostructured tubes, Mater. Sci. Eng. A 528 (2010) 180188. [140] L.S. Toth, M. Arzaghi, J.J. Fundenberger, B. Beausir, O. Bouaziz, R. ArruffatMassion, Severe plastic deformation of metals by high-pressure tube twisting, Scr. Mater. 60 (2009) 175177. [141] G. Faraji, M.M. Mashhadi, H.S. Kim, Tubular channel angular pressing (TCAP) as a novel severe plastic deformation method for cylindrical tubes, Mater. Lett. 65 (2011) 30093012. [142] G. Faraji, A. Babaei, M.M. Mashhadi, K. Abrinia, Parallel tubular channel angular pressing (PTCAP) as a new severe plastic deformation method for cylindrical tubes, Mater. Lett. 77 (2012) 8285. [143] G.B. Hamu, D. Eliezer, L. Wagner, The relation between severe plastic deformation microstructure and corrosion behavior of AZ31 magnesium alloy, J. Alloys Compd. 468 (2009) 222229. [144] E.A. El-Danaf, M.S. Soliman, A.A. Almajid, M.M. El-Rayes, Enhancement of mechanical properties and grain size renement of commercial purity aluminum 1050 processed by ECAP, Mater. Sci. Eng. A 458 (2007) 226234. [145] K. Neishi, Z. Horita, T.G. Langdon, Grain renement of pure nickel using equal-channel angular pressing, Mater. Sci. Eng. A 325 (2002) 5458. [146] I.-H. Kim, S.I. Kwun, Precipitation behavior of c00 in severely plastic deformed Ni-base alloys, Mater. Sci. Forum 534536 (2007) 213216. [147] H. Ueno, K. Kakihata, Y. Kaneko, S. Hashimoto, A. Vinogradov, Enhanced fatigue properties of nanostructured austenitic SUS 316L stainless steel, Acta Mater. 59 (2011) 70607069. [148] C.X. Huang, G. Yang, Y.L. Gao, S.D. Wu, Z.F. Zhang, Inuence of processing temperature on the microstructures and tensile properties of 304L stainless steel byECAP, Mater. Sci. Eng. A 485 (2008) 643650. [149] Z.J. Zheng, Y. Gao, Y. Gui, M. Zhu, Corrosion behaviour of nanocrystalline 304 stainless steel prepared by equal channel angular pressing, Corros. Sci. 54 (2012) 6067. [150] P.C. Millett, T. Desai, V. Yamakov, D. Wolf, Atomistic simulations of diffusional creep in a nanocrystalline body-centered cubic material, Acta Mater. 56 (2008) 36883698. [151] I.A. Ovidko, A.G. Sheinerman, Enhanced ductility of nanomaterials through optimization of grain boundary sliding and diffusion processes, Acta Mater. 57 (2009) 22172228. [152] B. Arab, S. Sanjabi, A. Shokuhfar, Size dependency of self-diffusion and creep behavior of nanostructured metals, Mater. Lett. 65 (2011) 712715. [153] S. Gollapudi, K.V. Rajulapati, I. Charit, C.C. Koch, R.O. Scattergood, K.L. Murty, Creep in nanocrystalline materials: role of stress assisted grain growth, Mater. Sci. Eng. A 527 (2010) 57735781. [154] V. Slenicka, J. Dvorak, P. Kral, M. Svoboda, M. Kvapilova, T.G. Langdon, Factors inuencing creep ow and ductility in ultrane-grained metals, Mater. Sci. Eng. A 558 (2012) 403411. [155] K.Y. Zhu, A. Vassel, F. Brisset, K. Lu, J. Lu, Nanostructure formation mechanism of a-titanium using SMAT, Acta Mater. 52 (2004) 41014110. [156] G. Liu, S.C. Wang, X.F. Lou, J. Lu, K. Lu, Low carbon steel with nanostructured surface layer induced by high-energy shot peening, Scr. Mater. 44 (2001) 17911795. [157] G. Liu, J. Lu, K. Lu, Surface nanocrystallization of 316L stainless steel induced by ultrasonic shot peening, Mater. Sci. Eng. A 286 (2000) 9195. [158] B. Arifvianto, M. Mahardika, P. Dewo, P.T. Iswanto, U.A. Salim, Effect of surface mechanical attrition treatment (SMAT) on microhardness, surface roughness and wettability of AISI 316L, Mater. Chem. Phys. 125 (2011) 418 426. [159] L. Wen, Y. Wang, Y. Zhou, L.X. Guo, J.H. Ouyang, Iron-rich layer introduced by SMAT and its effect on corrosion resistance and wear behavior of 2024 Al alloy, Mater. Chem. Phys. 126 (2011) 301309. [160] M. Laleh, F. Kargar, Effect of surface nanocrystallization on the microstructural and corrosion characteristics of AZ91D magnesium alloy, J. Alloys Compd. 509 (2011) 91509156. [161] L.L. Shaw, J.-W. Tian, A.L. Ortiz, K. Dai, J.C. Villegas, P.K. Liaw, R. Ren, D.L. Klarstrom, A direct comparison in the fatigue resistance enhanced by surface severe plastic deformation and shot peening in a C-2000 superalloy, Mater. Sci. Eng. A 527 (2010) 986994.

[103] H. Kassman, J. Pettersson, B.-M. Steenari, L.-E. mand, Two strategies to reduce gaseous KCl and chlorine in deposits during biomass combustion injection of ammonium sulphate and co-combustion with peat, Fuel Process. Technol. 105 (2013) 170180. [104] P. Viklund, R. Pettersson, A. Hjrnhede, P. Henderson, P. Sjvall, Effect of sulphur containing additive on initial corrosion of superheater tubes in waste red boiler, Corres. Eng. Sci. Technol. 44 (2009) 234240. [105] K. Schoeld, New method to minimize high temperature corrosin resulting from alkali sulfate and chloride deposition in combustion systems. II. Molybdenum salts, Energy Fuels 19 (2005) 18981905. [106] X. Wei, U. Schnell, K.R.G. Hein, Behaviour of gaseous chlorine and alkali metals during biomass thermal utilization, Fuel 84 (2005) 841848. [107] M. hman, D. Bostrm, A. Nordin, H. Hedman, Effect of kaolin and limestone addition on slag formation during combustion of wood fuels, Energy Fuels 18 (2004) 13701376. [108] M. Aho, J. Silvennoinen, Preventing chlorine deposition on heat transfer surfaces with aluminiumsilicon rich biomass residue and additive, Fuel 83 (2004) 12991305. [109] B.M. Steenari, A. Lundberg, H. Pettersson, M. Wilewska-Bien, D. Andersson, Investigation of ash sintering during combustion of agricultural residues and the effect of additives, Energy Fuels 23 (2009) 5655 5662. [110] M. Theis, B.J. Srifvars, M. Zevenhoven, M. Hupa, H. Tran, Fouling tendency of ash resulting from burning mixtures of biofuels. Part 2: deposit chemistry, Fuel 85 (2006) 19922001. [111] K. Mroczek, S. Kalisz, M. Pronobis, J. Soltys, The effect of halloysite additive on operation of boilers ring agricultural biomass, Fuel Process. Technol. 92 (2011) 845855. [112] M. Schtze, M. Malessa, V. Rohr, T. Weber, Development of coatings for protection in specic high temperature environments, Surf. Coat. Technol. 201 (2006) 38723879. [113] T.S. Sidhu, S. Prakash, R.D. Agrawal, Hot corrosion studies of HVOF NiCrBSi and Stellite-6 coatings on a Ni-based superalloy in an actual industrial environment of a coal red boiler, Surf. Coat. Technol. 201 (2006) 1602 1612. [114] D. Bendix, G. Tegeder, P. Crimmann, J. Metschke, M. Faulstich, Development of thermal sprayed layers for high temperature areas in waste incineration plants, Mater. Corros. 59 (2008) 389392. [115] R.A. Mahesh, R. Jayaganthan, S. Prakash, Evaluation of hot corrosion behaviour of HVOF sprayed Ni5Al and NiCrAl coatings in coal red boiler environment, Surf. Eng. 26 (2010) 413421. [116] H. Naganuma, N. Ikeda, T. Ito, H. Satake, M. Matsuura, Y. Ueki, R. Yoshiie, I. Naruse, Control of ash deposition in solid fuel red boiler, Fuel Process. Technol. 105 (2013) 7781. [117] J.A. Hearley, C. Liu, J.A. Little, A.J. Sturgeon, Corrosion of NiAl high velocity oxyfuel (HVOF) thermal spray coating by y ash and synthetic biomass ash deposits, Br. Corros. J. 36 (2001) 111120. [118] H.Y. Al-Fadhli, J. Stokes, M.S.J. Hashmi, B.S. Yilbas, The erosion-corrosion behavior of high velocity oxy-fuel (HVOF) thermally sprayed Inconel-625 coatings on different metallic surfaces, Surf. Coat. Technol. 200 (2006) 5782 5788. [119] M. Mohammadi, S. Javadpour, S.A.J. Jahromi, K. Shirvani, A. Kobayashi, Characterization and hot corrosion performance of LVPS and HVOFCoNiCrAlYSi coatings, Vacuum 86 (2012) 14581464. [120] M.M. Verdian, K. Raeissi, M. Salehi, Characterization and electrochemical properties of Ni(Si)/Ni5Si2 multiphase coatings prepared by HVOF spraying, Appl. Surf. Sci. 261 (2012) 493498. [121] T.S. Sidhu, R.D. Agrawal, S. Prakash, Hot corrosion of some superalloys and role of high-velocity oxy-fuel spray coatingsa review, Surf. Coat. Technol. 198 (2005) 441446. [122] H. Herman, S. Sampath, R. Mccune, Thermal spray: current status and future trends, MRS Bull. 25 (2000) 1725. [123] M.A. Uusitalo, P.M.J. Vuoristo, T.A. Mntyl, High temperature corrosion of coatings and boiler steels in reducing chlorine-containing atmosphere, Surf. Coat. Technol. 161 (2002) 275285. [124] M.A. Uusitalo, P.M.J. Vuoristo, T.A. Mntyl, High temperature corrosion of coatings and boiler steels in oxidizing chlorine-containing atmosphere, Mater. Sci. Eng. A 346 (2003) 168177. [125] M.C. Mayoral, J.M. Andrs, J. Belzunce, V. Higuera, Study of sulphidation and chlorination on oxidized SS310 and plasma-sprayed NiCr coatings as simulation of hot corrosion in fouling and slagging in combustion, Corros. Sci. 48 (2006) 13191336. [126] M. Torrell, S. Dosta, J.R. Miguel, J.M. Guilemany, Optimisation of HVOF thermal spray coatings fortheir implementation as MSWI superheater protector, Corros. Eng. Sci. Technol. 45 (2010) 8493. [127] J. Kalivodova, D. Baxter, M. Schtze, V. Rohr, High temperature corrosion of coatings and boiler steels in oxidizing chlorine-containing atmosphere, Mater. Corros. 56 (2005) 882889. [128] V. Rohr, M. Schtze, Diffusion coatings for heat exchanger materials, Surf. Eng. 20 (2004) 266274. [129] C. Leyens, I.G. Wright, B.A. Pint, Hot corrosion of an EB-PVD thermal-barrier coating system at 950 C, Oxid. Met. 54 (2000) 401424. [130] P. Panjan, M. Cekada, M. Panjan, D. Kek-Merl, Growth defects in PVD hard coatings, Vacuum 84 (2010) 209214. [131] M. Bromark, M. Larsson, P. Hendenqvist, S. Hogmark, Wear of PVD Ti/TiN multilayer coatings, Surf. Coat. Technol. 90 (1997) 217223.

R.A. Antunes, M.C.L. de Oliveira / Corrosion Science 76 (2013) 626 [162] M. Wen, G. Liu, J.-F. Gu, W.-M. Guan, J. Lu, The tensile properties of titanium processed by surface mechanical attrition treatment, Surf. Coat. Technol. 202 (2008) 47284733. [163] T. Balusamy, S. Kumar, T.S.N.S. Narayanan, Effect of surface nanocrystallization on the corrosion behaviour of AISI 409 stainless steel, Corros. Sci. 52 (2010) 38263834. [164] K. Lu, J. Lu, Nanostructured surface layer on metallic materials induced by surface mechanical attrition treatment, Mater. Sci. Eng. A 375377 (2004) 3845. [165] Y.-W. Hao, B. Deng, C. Zhong, Y.-M. Jiang, J. Li, Effect of surface mechanical attrition treatment on corrosion behavior of 316 stainless steel, J. Iron Steel Res. Int. 16 (2009) 6872. [166] H. Han, L. Zhang, J. Lu, W. Zhang, Thermal stability and corrosion resistance of nanocrystallized zirconium formed by surface mechanical attrition treatment, J. Mater. Res. 24 (2009) 31363145. [167] A.Q. Lu, Y. Li, G. Liu, Q.H. Zhang, C.M. Liu, Effect of nanocrystalline and twin boundaries on corrosion behavior of 316L stainless steel using SMAT, Acta Metall. Sin. 19 (2006) 183189. [168] T. Wang, J. Yu, B. Dong, Surface nanocrystallization induced by shot peening and its effect on corrosion resistance of 1Cr18Ni9Ti stainless steel, Surf. Coat. Technol. 200 (2006) 47774781. [169] H.-W. Chang, P.M. Kelly, Y.-N. Shi, M.-X. Zhang, Thermal stability of nanocrystallized surface produced by surface mechanical attrition treatment in aluminum alloys, Surf. Coat. Technol. 206 (2012) 39703980. [170] Y.J. Mai, X.H. Jie, L.L. Liu, N. Yu, X.X. Zheng, Thermal stability of nanocrystalline layers fabricated by surface nanocrystallization, Appl. Surf. Sci. 256 (2010) 19721975. [171] M. Wen, C. Wen, P. Hodgson, Y. Li, Thermal oxidation behaviour of bulk titanium with nanocrystalline surface layer, Corros. Sci. 59 (2012) 352359. [172] J. He, J.M. Schoenung, Nanostructured coatings, Mater. Sci. Eng. A 336 (2002) 274319. [173] K. Tao, X.-L. Zhou, H. Cui, J.-S. Zhang, Oxidation and hot corrosion behaviors of HVAF-sprayed conventional and nanostructured NiCrC coatings, Trans. Nonferrous Met. Soc. China 19 (2009) 11511160. [174] A. Keyvani, M. Saremi, M.H. Sohi, An investigation on oxidation, hot corrosion and mechanical properties of plasma-sprayed conventional and nanostructured YSZ coatings, Surf. Coat. Technol. 206 (2011) 208216. [175] P. Yu, W. Wang, F. Wang, S. Zhu, High-temperature corrosion behavior of sputtered K38 nanocrystalline coatings with and without yttrium addition in molten sulfate at 900 C, Surf. Coat. Technol. 206 (2011) 6874. [176] T. Sundararajan, S. Kuroda, J. Kawakita, S. Seal, High temperature corrosion of nanoceria coated 9Cr1Mo ferritic steel in air and steam, Surf. Coat. Technol. 201 (2006) 21242130. [177] H. Wang, D. Zuo, G. Chen, G. Chen, G. Sun, X. Li, X. Cheng, Hot corrosion behaviour of low Al NiCoCrAlY cladded coatings reinforced by nano-particles on a Ni-base super alloy, Corros. Sci. 52 (2010) 35613567. [178] A. Rahman, V. Chawla, R. Jayaganthan, R. Chandra, R. Ambardar, Study of cyclic hot corrosion of nanostructured Cr/CoAl coatings on superalloy, Mater. Chem. Phys. 126 (2011) 253261. [179] M. Mohammdnezhad, M. Shamanian, M.H. Enayati, Formation of nanostructured NiAl coating on carbon steel by usingmechanical alloying, Appl. Surf. Sci. 263 (2012) 730736. [180] A. Rahman, R. Jayaganthan, S. Prakash, V. Chawla, R. Chandra, High temperature oxidation behavior of nanostructured NiAl coatings on superalloy, J. Alloys Compd. 472 (2009) 478483. [181] D. Zhong, J.J. Moore, E. Sutter, B. Mishra, Microstructure, composition and oxidation resistance of nanostructured NiAl and NiAlN coatings produced by magnetron sputtering, Surf. Coat. Technol. 200 (2005) 12361241. [182] I. Pisa, G. Lazaroiu, Inuence of co-combustion of coal/biomass on the corrosion, Fuel Process. Technol. 104 (2012) 356364. [183] K. Persson, M. Brostrm, J. Carlsson, A. Nordin, R. Backman, High temperature corrosion in a 65 MW waste to energy plant, Fuel Process. Technol. 88 (2007) 11781182. [184] J. Lehmusto, P. Yrjas, B.-J. Skrifvars, M. Hupa, High temperature corrosion of superheater steels by KCl and K2CO3 under dry and wet conditions, Fuel Process. Technol. 104 (2012) 253264. [185] K. Chandra, V. Kain, G.K. Dey, Failure of 2.25Cr1Mo steel superheater tubes in a uidized bed combustor due to reside corrosion, Mater. Charact. 62 (2011) 6269. [186] A.U. Syed, N.J. Simms, J.E. Oakey, Fireside corrosion of superheaters: effects of air and oxy-ring of coal and biomass, Fuel 101 (2012) 6273. [187] P. Viklund, A. Hjrnhede, P. Henderson, A. Stlenheim, R. Pettersson, Corrosion of superheater materials in a waste-to-energy plant, Fuel Process. Technol. 105 (2013) 106112. [188] P. Viklund, R. Pettersson, HCl-induced high temperature corrosion of stainless steels in thermal cycling conditions and the effect of peroxidation, Oxid. Met. 76 (2011) 111126. [189] H. Asteman, M. Spiegel, Investigation of the HCl(g) attack on pre-oxidized pure Fe, Cr, Ni and commercial 304 steel at 400 C, Corros. Sci. 49 (2007) 36263637. [190] C. Liu, J.A. Little, P.J. Henderson, P. Ljung, Corrosion of TP347H FG stainless steel in a biomass red PF utilityboiler, J. Mater. Sci. 36 (2001) 10151026. [191] M. Montgomery, S.A. Jensen, U. Borg, O. Biede, T. Vilhelmsen, Experiences with high temperature corrosion at straw-ring power plants in Denmark, Mater. Corros. 62 (2011) 593605.

25

[192] A. Nafari, A. Nylund, Field study on superheater tubes in the loop seal of a wood red CFB plant, Mater. Corros. 55 (2004) 909920. [193] I. Klevtsov, H. Tallermo, T. Bojarinova, R. Crane, High temperature corrosion of boiler steels under chlorine-containing surface deposits, J. Press. Ves. Technol. 127 (2005) 106111. [194] C. Cuevas-Arteaga, D. Verhelst, A. Alfantazi, A. Alfantazi, Performance of alloy 625 under combustion gas environments: a review, ECS Trans. 28 (2010) 61 76. [195] A. Phongphiphat, R. Ryu, Y.B. Yang, K.N. Finney, A. Leyland, V.N. Shari, J. Swithenbak, Investigation into high-temperature corrosion in a large-scale municipal waste-to-energy plant, Corros. Sci. 52 (2010) 38613874. [196] M.F. Ashby, Materials Selection in Mechanical Design, fourth ed., Elsevier, Oxford, UK, 2010. [197] A. Jahan, M.Y. Ismail, S.M. Sapuan, F. Mustapha, Material screening and choosing methods a review, Mater. Des. 31 (2010) 696705. [198] A.K. Sharma, N. Gupta, Material selection of RF-Mems switch used for recongurable antenna using Ashbys methodology, Prog. Electromagn. Res. Lett. 31 (2012) 147157. [199] A. Chauhan, R. Vaish, A comparative study on material selection for microelectromechanical systems, Mater. Des. 41 (2012) 177181. [200] B.N. Aditya, N. Gupta, Material selection methodology for gate dielectric material in metaloxidesemiconductor devices, Mater. Des. 35 (2012) 696 700. [201] Z. Szklarska-Smialowska, Pitting and Crevice Corrosion of Metals, NACE Intl., Houston, TX, 2005. [202] H.-B. Li, Z.-H. Jiang, Y. Yang, Y. Cao, Z.-R. Zhang, Pittingcorrosion and crevice corrosion behaviors of high nitrogen austenitic stainless steels, Int. J. Miner. Metall. Mater. 16 (2009) 517524. [203] N.S. Zadorozne, C.M. Giordano, M.A. Rodrguez, R.M. Carranza, R.B. Rebak, Crevice corrosion kinetics of nickel alloys bearing chromium and molybdenum, Electrochim. Acta 76 (2012) 94101. [204] ASTM G48, Standard Test Methods for Pitting and Crevice Corrosion Resistance of Stainless Steels and Related Alloys by Use of Ferric Chloride Solution 03, 2009. [205] D.R.H. Jones, Creep failures of overheated boiler, superheater and reformer tubes Eng, Fail. Anal. 11 (2004) 873893. [206] H.-T. Yao, F.-Z. Xuan, Z. Wang, S.-T. Tu, A review of creep analysis and design under multi-axial stress states, Nucl. Eng. Des. 237 (2007) 19691986. [207] K. Maruyama, H.G. Armaki, R.P. Chen, K. Yoshimi, M. Yoshizawa, M. Igarashi, Cr concentration dependence of overestimation of long term creep life in strength enhanced high Cr ferritic steels, Int. J. Press. Ves. Piping 87 (2010) 276281. [208] M.W. Spindler, S.L. Spindler, Creep deformation, rupture and ductility of Esshete 1250, Int. J. Press. Ves. Piping 85 (2008) 8998. [209] P.P. Psyllaki, G. Pantazopoulos, H. Lefakis, Metallurgical evaluation of creepfailed superheater tubes, Eng. Fail. Anal. 16 (2009) 14201431. [210] W.F. Hosford, Mechanical Behavior of Materials, second ed., Cambridge University Press, New York, 2010. [211] ASTM E8/E8M-11 Standard Test Methods for Tension Testing of Metallic Materials, ASTM International, 2011. [212] G.Y. Lai, High Temperature Corrosion and Materials Applications, ASM International, 2007. [213] J.R. Kish, D.L. Singbeil, J.R. Keiser, North american experience with composite tubes in kraft recovery boilers, Pulp Paper Canada 106 (2005) T75T81. [214] M. Hamaguchi, E. Vakkilainen, Corrosion of superheater tubes in recovery boilers: a challenge, O Papel 71 (2010) 5771. [215] <www.haynes.ch/doc/HAYNES_617.pdf> (accessed 22.05.13). [216] <www.hightempmetals.com/techdata/hitempInconel625data.php> (accessed 22.05.13). [217] R.W. Swindeman, M.J. Swindeman, A comparison of creep models for nickel base alloys for advanced energysystems, Int. J. Press. Ves. Piping 85 (2008) 7279. [218] M.D. Mathew, K.B.S. Rao, S.L. Mannan, Creep properties of service-exposed Alloy 625 after re-solution annealing treatment, Mater. Sci. Eng. A 372 (2004) 327333. [219] ASTM A213/213M Standard Specication for Seamless Ferritic and Austenitic Alloy-steel Boiler, Superheater, and Heat-exchanger Tubes, ASTM International, 2011. [220] ASTM B564-11 Standard Specication for Nickel Alloy Forgings, ASTM International, 2011. [221] ASTM B626-10 Standard Specication for Welded Nickel and NickelCobalt Alloy Tube, ASTM International, 2010. [222] ASTM International Volume 2: Properties and Selection: Nonferrous Alloys and Special Purpose Materials, 2000. [223] <www.smt.sandvik.com/en/materials-center/material-datasheets/weldingproducts/welding-wire/sanicro-60> (accessed 22.05.13). [224] ASTM B668-05 Standard specication for UNS N08028 Seamless Pipe and Tube, ASTM International, 2010. [225] ASTM A387/A387M-11 Standard Specication for Pressure Vessel Plates, Alloy Steel, ChromiumMolybdenum, ASTM International, 2011. [226] ASTM A53/A53M-12 Standard Specication for Pipe, Steel, Black and Hotdipped, Zinc-coated, Welded and Seamless, ASTM International, 2012. [227] ASTM A240/A240M-12a Standard Specication for Chromium and ChromiumNickel Stainless Steel Plate, Sheet, and Strip for Pressure Vessels and for General Applications, ASTM International, 2012.

26

R.A. Antunes, M.C.L. de Oliveira / Corrosion Science 76 (2013) 626 [231] EN 10216-5 European Standard. Seamless Steel Tubes for Pressure Purposes Technical Delivery Conditions Part 5: Stainless Steel Tubes, European Committee for Standardization, 2004. [232] Z.-F. Hu, Heat-resistant steels, microstructure evolution and life assessment in power plants, in: M. Rasul (Ed.), Thermal Power Plants, In Tech Publications, Croatia, 2012, pp. 195226. [233] ASTM A250/A250M-05 Standard Specication for Electric-resistance-welded Ferritic-alloy Steel Boiler and Superheater Tubes, ASTM International, 2009.

[228] A. Hernas, G. Moskal, K. Rodak, J. Pasternak, Properties and microstructure of 12% CrW steels after long-term service, J. Achiev. Mater. Manufact. Eng. 17 (2006) 6972. [229] ASTM B407-08a Standard Specication for NickelIronChromium Alloy Seamless Pipe and Tube, ASTM International, 2008. [230] K. Zhan, C.H. Jiang, V. Ji, Surface mechanical properties of S30432 austenitic steel after shot peening, Appl. Surf. Sci. 258 (2012) 95599563.

Você também pode gostar