Case Studies

section epub:type=”chapter” role=”doc-chapter”>


Chapter 20
Case Studies


We hope that the interest in including a chapter on case studies connected with high temperature corrosion in relevant industrial situations and related with theoretical, experimental, or other studies will contribute for the proper analysis and solution of problems with some complexity or difficulty. The case studies presented here have been adapted from recent literature on practical and fundamental problems that have been analyzed and subsequently solved using relatively innovative approaches. The references listed in Further Reading were the main sources for this examination.


20.1 Making Stainless Steels


Mild steel is an excellent structural material, but at low temperature it rusts, and at high it oxidizes rapidly. For many applications, there is a great demand of corrosion‐resistant steel. In response to this demand, a range of stainless irons and steels has been developed. When mild steel is exposed to hot air, it oxidizes quickly to form FeO (or higher oxides). But if one of the elements near the top of Table 20.1 with a large energy of oxidation is dissolved in the steel, then this element oxidizes preferentially (because it is much more stable than FeO), forming a layer of its oxide on the surface. And if this oxide is a protective one, such as Cr2O3, Al2O3, SiO2, or BeO, it stifles further growth and protects the steel.


Table 20.1 Energy of oxides formation at 273 K





image

A considerable quantity of this foreign element is needed to give adequate protection. The best is chromium, 18% of which gives a very protective oxide film: it cuts down the rate of attack at 900 °C, for instance, by more than 100 times.


Other elements, when dissolved in steel, also cut down the rate of oxidation. Al2O3 and SiO2 both form in preference to FeO (Table 20.1) and form protective films (Table 20.2). Note that 0.7 TM was chosen in Table 20.2 because this is a typical figure for the operating temperature of a turbine blade or similar component. Thus, 5% Al dissolved in steel decreases the oxidation rate by 30 times, and 5% Si by 20 times. The same principle can be used to impart corrosion resistance to other metals. We shall discuss nickel and cobalt in the next case study – they can be alloyed in this way – and so can copper; although it will not dissolve enough chromium to give a good Cr2O3 film, it will dissolve enough aluminum, giving a range of stainless steel called “aluminum bronzes.” Even silver can be prevented from tarnishing (reaction with sulfur) by alloying it with aluminum or silicon, giving protective Al2O3 or SiO2 surface films. And archeologists believe that the Delhi pillar – an ornamental pillar of cast iron that has stood, uncorroded, for some hundreds of years in a particularly humid spot – survived because the iron has some 6% silicon in its composition.


Table 20.2 Time, in hours, for material to be oxidized to a depth of 0.1 mm at 0.7 TM in air




















































































































Material Time Melting point (K)
Au >1020 1336
Ag Very long 1234
Al Very long  933
Si3N4 Very long 2173
SiC Very long 3110
Sn Very long  505
Si 2 × 106 1683
Be 106 1557
Pt 1.8 × 105 2042
Mg >105  923
Zn >104  692
Cr 1600 2148
Na >1000  371
K >1000  337
Ni 600 1726
Cu 25 1356
Fe 24 1809
Co 7 1765
Ti <6 1943
WC cermet <5 1700
Ba ≪0.5  983
Zr 0.2 2125
Ta Very short 3250
Nb Very short 2740
U Very short 1405
Mo Very short 2880
W Very short 3680

Data subject to considerable variability due to varying degrees of material purity, prior to surface treatment and presence of atmospheric impurities such as sulfur. TM is the absolute melting point temperature (K).


Ceramics themselves are sometimes protected in this way. Silicon carbide, SiC, and silicon nitride, Si3N4, both have large negative energies of oxidation (meaning that they oxidize easily). But when they do, the silicon in them turns to SiO2 that quickly forms a protective skin and prevents further attack.


This protection by alloying has one great advantage over protection by a surface coating (such as chromium plating or gold plating): it repairs itself when damaged. If the protective film is scored or abraded, fresh metal is exposed, and the chromium (or aluminum or silicon) it contains immediately oxidizes, healing the break in the film.


20.2 Corrosion Protection of Turbine Blades


The development of turbine blade materials to meet the challenge of increasing engine temperatures must satisfy the requirements listed in Table 20.3.


Table 20.3 Alloy requirements















(a)  Resistance to creep
(b)  Resistance to high temperature oxidation
(c)  Toughness
(d)  Thermal fatigue resistance
(e)  Thermal stability
(f)  Low density

These requirements severely limit our choice of very resistant materials. For example, ceramics, with their high softening temperatures and low densities, are ruled out for aeroengines because they are far too brittle (they are under evaluation for use in land‐based turbines, where the risks and consequences of sudden failure are less severe). Cermets offer no great advantage because their metallic matrices soften at a too much low temperature. The materials that best fill present needs are the nickel‐based superalloys.


The alloy used for turbine blades in the high pressure stage of an aircraft turbofan engine is a classic example of a material designed to be resistant to dislocation (power‐law) creep at high stresses and temperatures. At takeoff, the blade is subjected to stresses approaching 250 MN m−2, and the design specification requires that this stress be supported for 30 hours at 850 °C without more than 0.1% irreversible creep strain. In order to meet these stringent requirements, an alloy based on nickel has evolved with the rather mind‐boggling specification given in Table 20.4. The point of all these complicated additions of foreign atoms to the nickel is straightforward. It is (i) to have as many atoms in solid solution as possible (the cobalt, the tungsten, and the chromium); (ii) to form stable, hard precipitates of compounds like Ni3Al, Ni3Ti, MoC, and TaC to obstruct the dislocations; and (iii) to form a protective surface oxide film of Cr2O3 to protect the blade itself from attack by oxygen. These superalloys resist creep so well that they can be used at 850 °C, and they are so hard that cannot be machined easily by normal methods and must be precision‐cast (for example, by investment casting that produces a fine‐grained material that may undergo a fair amount of diffusion creep and that may fail rather soon by cavity formation) to their final shape. Due to this, the blades are expensive, the total cost of a rotor of 102 blades being €200 000. To stave off failure, the alloys are directionally solidified (DS) to give long grains with grain boundaries parallel to the applied stress. DS alloys are standard in high performance engines and are now also in use in civil aircrafts. The improved creep properties of the DS alloy will allow the engine to run at a flame temperature approximately 50 °C higher than before, for a doubling in production cost. Figure 20.1 shows how this evolutionary process has resulted in a continual improvement of creep properties of nickel alloys over time and shows how the amounts of the major foreign elements have been juggled to obtain these improvements – keeping a watchful eye on the remaining necessary properties. The figure also shows how improvements in alloy manufacture – in this case the use of directional solidification – have helped to increase the operating temperature. Nevertheless, it is clear from the graph that improvements in nickel alloys are now nearing the point of diminishing returns.


Table 20.4 Composition of typical creep‐resistance blade





























































Metals wt%
Ni 59
Co 10
W 10
Cr 9
Al 5.5   
Ta  2.5   
Ti  1.5   
Hf  1.5   
Fe  0.25  
Mo  0.25  
C  0.15  
Si  0.1   
Mn  0.1   
Cu  0.05  
Zr  0.05  
B  0.015 
S <0.008 
Pb <0.0005
c20f001

Figure 20.1 Superalloy developments by a sort of creepy Darwinism.


Internal cooling of the blade enabled the inlet temperature of the turbine to be increased immediately by 100 °C with no change in alloy composition. Later, film cooling constituted a continuous improvement in turbine blades (Figure 20.2). But by further ducting cold air through the blades will begin to reduce the thermal efficiency, and thus a unique characteristic of some alloys, called eutectics, was explored. These eutectics form spontaneously on aligned reinforced structure, which is usually a compound with high melting temperature, whose fibers improve the creep properties of the resulting composite alloy.

c20f002

Figure 20.2 Temperature evolution and future materials trends in turbine blades.


Figure 20.3 shows how eutectics are made, and Table 20.5 lists some typical high temperature composites under study for turbine blade applications.

c20f003

Figure 20.3 Directionally solidified eutectics for turbine blades.


Table 20.5 High temperature composites




























Matrix Reinforcing phase Reinforcing phase geometry
Ni TaC Fibers
Co TaC Fibers
Ni3Al Ni3Nb Plates
Co Cr7C3 Fibers
Nb Nb2C Fibers

As shown in Figure 20.2, if directionally solidified eutectics (DSEs) prove successful, they will allow the metal temperature to be increased by ≈100 °C above conventional DS nickel alloys and the inlet temperature by ≈200 °C (because of a temperature scaling effect caused by the blade cooling). Further improvements in alloy design are underway in which existing nickel alloys and DSEs are being blended to give a fiber‐reinforced structure with precipitates in the matrix.


The ceramics best suited for structural use at high temperatures (≥1000 °C) are listed in Table 20.6 and compared with nickel‐based superalloys. The comparison shows that all the ceramics have attractively low densities, high moduli, and high melting points (and thus excellent creep strength at 1000 °C); and many are completely resistant to oxidation – they are oxides already. But several have poor thermal conductivity (leading to high thermal stresses), and all have very low toughness.


Table 20.6 Ceramics for high temperature structures




















































Material Density (mg m−3) Melting or decomposition (D), temperature (K) Modulus (GN m−2) Expansion coefficient × 10+6 (K−1) Thermal conductivity at 1000 K (Wm−1 K−1) Fracture toughness Kc (MN m−3/2)
Alumina, Al2O3 4.0 2320 360 6.9 7 ≈5
Glass‐ceramics (pyroceramics) 2.7 >1700 ≈120 ≈3 ≈3 ≈3
Hot‐pressed silicon nitride, Si3N4 3.1 2173 (D) 310 3.1 16 ≈5
Hot‐pressed silicon carbide, SiC 3.2 3000 (D) ≈420 4.3 60 ≈3.5
Nickel alloys (Nimonics) 8.0 1600 200 12.5 12 ≈100

To overcome their major problems, ceramic matrix components are under development: they combine strong, exceptionally perfect, fibers (such as silicon carbide or alumina, grown by special techniques) in a matrix of ceramics (silicon carbide or alumina again), made by conventional means to give a material that you might think of as “high temperature wood.”


Any major materials development program, such as that on the DSEs, can only be undertaken if a successful outcome would be cost effective. The cost of the materials themselves must also be considered. For example, the use of greater quantities of exotic materials, such as hafnium, will drive the cost of blades up. But if new alloys offer improved life or inlet temperature, there is a strong incentive to pursue them.


As we saw above, the alloys used for turbine blades contain large amounts of chromium, dissolved in solid solution in the nickel matrix. Now, if we look at Table 20.1, we see that the formation of Cr2O3 releases much more energy (701 kJ mol−1 of O2) than NiO (439 kJ mol−1 of O2). This means that Cr2O3 will form in preference to NiO on the surface of the alloy. Obviously, the more Cr in the alloy, the greater the preference for Cr2O3. At the 20% level, enough Cr2O3 forms on the surface of the turbine blade to make the material act a bit as if it were chromium. Suppose for one moment that our material is chromium. Table 20.2 shows that Cr would lose 0.1 mm in 1600 hours at 0.7 TM. Of course, we have forgotten one thing, 0.7 TM for Cr is 1504 K (1231 °C), whereas, as already mentioned, for Ni it is 1208 K (935 °C). We should, therefore, consider how Cr2O3 would act as a barrier to oxidation at 1208 K rather than at 1504 K (Figure 20.4). The oxidation of Cr follows parabolic kinetics with an activation energy of 330 kJ mol−1. Then the ratio of the times required to remove 0.1 mm (from equation images, where Ap and Qp are constants) is


20.1equation
c20f004

Figure 20.4 Kp variation with temperature.


Thus, the time at 1208 K is t2 = 0.65 × 1.6 × 106 hours = 1.04 × 106 hours. Note that in Eq. 20.1, t1 and t2 are the oxidation times with t2 > t1.


Now, as we have said, there is only at most 20% Cr in the alloy, and the alloy behaves only partly as if it were protected by Cr2O3. In fact, experimentally, we find that 20% Cr increases the time for a given metal loss by only about 10 times, i.e. the time taken to lose 0.1 mm at blade working temperature becomes 600 × 10 hours = 6000 hours rather than 106 hours.


Why this large difference? Well, whenever you consider an alloy rather than a pure material, the oxide layer – whatever its nature (NiO, Cr2O3, etc.) – has foreign elements contained in it, too. Some of these will greatly increase either the diffusion coefficients in, or electrical conductivity of, the layer and make the rate of oxidation through the layer much more than it would be in the absence of foreign elements contamination.


This 0.1 mm loss in 6000 hours from a 20% Cr alloy at 950 °C, though better than pure nickel, is still not good enough. The obvious way out of this problem is to coat the blades with a protective layer (Figure 20.5). This is usually carried out by spraying molten droplets of aluminum onto the blade surface to form a layer, some microns thick. The blade is then heated in a furnace to allow the Al to diffuse into the surface of the Ni. During this process, some of the Al forms compounds such as AlNi with the nickel – which are themselves good barriers to oxidation of the Ni – while the rest of the Al becomes oxidized up to give Al2O3.

c20f005

Figure 20.5 Protection of turbine blades by sprayed‐on aluminum.


Because oxides are usually quite brittle at the temperatures encountered on a turbine blade surface, they can crack, especially when the temperature of the blade changes and differential thermal contraction and expansion stresses are set up between alloy and oxide. These can act as ideal nucleation centers for thermal fatigue cracks, and because oxide layers in nickel alloys are well stuck to the underlying alloy (they would be useless if they were not), the crack can spread into the alloy itself.


The properties of the oxide film are thus very important in affecting the fatigue properties of the whole component. The pure refractory metals Nb, Ta, Mo, and W have high melting temperatures (2740, 3250, 2880, and 3680 K, respectively) and should therefore have very good creep properties. But they oxidize very rapidly indeed (see Table 20.2) and are utterly useless without coatings. The problem with coated refractory metals is that if a break occurs in the coating (e.g. by thermal fatigue, erosion by dust particles, etc.), catastrophic oxidation of the underlying metal will take place, leading to rapid failure. The “unsafeness” of this situation is a major problem that has to be solved before we can use these on‐other‐counts potentially excellent materials.


The ceramics SiC and Si3N4 do not share this problem. They oxidize readily (Table 20.1); but, in doing so, a surface film of SiO2 forms that gives adequate protection up to 1300 °C. And because the film forms by oxidation of the material itself, it is self‐healing.


20.3 Oxidation of Silicides for VLSI Applications


Electronic materials are those used in electrical industries, electronics and microelectronics, and the substances for the building up of integrated circuits (ICs), circuit boards, packaging materials, communication cables, optical fibers, displays, and various controlling and monitoring devices. Discovery, development, and application of new materials are the robust power for the development of human society.


Processing of electronic materials mainly includes crystal growth, epitaxy, lithography, deposition, annealing, doping (diffusion or ion implantation), etching (dry or wet), and metallization. In ICs, the fundamental electronic switching device is the (complementary) metal oxide semiconductor field effect transistor (MOSFET), which has at least three terminals – the gate, source, and drain. The gate electrode is separated electronically from the source and drain by a thin dielectric film, which is commonly SiO2. Formation of this film could be accomplished by using sputtering, evaporation, chemical vapor deposition, or atomic layer deposition. Thermal oxidation is also used to grow the gate dielectric films due to its ease in processing control, but it results in many problems in association with material degradation at high temperature; these problems lead to the formation of oxides, carbides, sulfides, nitrides, etc., but when these oxidation processes are well controlled, they have been finding many applications in the modern electronics industry for surface cleaning, film growth, preparation of new materials, and building up of novel structures for devices. To date, Si is still the backbone of the modern semiconductor industry due to its (i) availability in a wide variety of sizes and shapes, (ii) mature material preparation and property control, (iii) native oxide films on its surface, and (iv) compatibility to planar IC technology.


Silicides of transition metals are used in two distinct ways in very large scale integrated (VLSI) circuits: (i) in the contact pads of individual devices on a chip and (ii) in the interconnection lines. For the latter applications, two particular properties of the silicides make them desirable: their metallic electrical conductivity and the possibility to produce a protective coating of SiO2 under heat treatment in an oxidizing ambient. Of those two attributes, the second one is particularly attractive because it offers inherent advantages in multilevel metallization schemes. To use this feature of silicides successfully, an understanding of the oxidation process is needed. It is convenient to subdivide the subject into two parts, depending on whether the silicide rests on a substrate of excess silicon or whether the substrate is inert, that is, the substrate is in a fully oxidized state, such as SiO2. Other reasons for the use of silicides in VLSI applications are their very low resistance; good process compatibility with Si, e.g. ability to withstand high temperatures; oxidizing ambients; various chemical cleans used during processing; little or no electronization; easy to dry etch; and good contacts to other materials. But these benefits rise many problems in integrating silicides in an IC, and under this point of view, the actual preferred silicides are WSi2, TiSi2, NiSi, and CoSi2. Table 20.7 indicates preferred silicides for VLSI application. The main techniques used for the formation of these silicides are (i) metal deposition on Si and formation by thermal heating, laser irradiation, or ion beam mixing (these are sensitive to interface cleanliness and heavy doping, selective silicidation on Si possible, and widely used for silicides of Pt, Pd, Co, and Ti), (ii) co‐evaporation (E‐gun) of metal and Si (poor process control, poor step coverage, and good tool for research but not used in manufacturing), and (iii) sputtering from a composite target (possibility of high level of contaminants like C, O, Na, and Ar; poor step coverage; used for MoSi2 and WSi2).


Table 20.7 Overall properties of preferred silicides for VLSI applications







































































































Silicide Thin‐film resistivity (μΩ cm) Sintering temperature (°C) Stable on Si up to (°C) Reaction with Al at (°C) nm of Si consumed per nm of metal nm of resulting silicide per nm of metal Barrier height to n‐Si (eV)
PtSi 28–35 250–400 ∼750 250 1.12 1.97 0.84
TiSi2 (C54) 13–16 700–900 ∼900 450 2.27 2.51 0.58
TiSi2 (C49) 60–70 500–700

2.27 2.51
Co2Si ∼70 300–500

0.91 1.47
CoSi 100–150 400–600

1.82 2.02
CoSi2 14–20 600–800 ∼950 400 3.64 3.52 0.65
NiSi 14–20 400–600 ∼650
1.83 2.34
NiSi2 40–50 600–800

3.65 3.63 0.66
WSi2 30–70 1000 ∼1000 500 2.53 2.58 0.67
MoSi2 40–100 800–1000 ∼1000 500 2.56 2.59 0.64
TaSi2 35–55 800–1000 ∼1000 500 2.21 2.41 0.59

20.3.1 Oxidation of Silicides on Silicon


Typically, silicides are not used alone as interconnections, but with an underlayer of polycrystalline Si. Upon thermal treatment in an oxidizing ambient, most silicides form a SiO2 overcoat while the silicide layer below is morphologically preserved. The rate of growth of the SiO2 layer is different for the different silicides, and these rates also differ from that of pure Si. They can be represented by the linear–parabolic growth law


20.2equation

Only gold members can continue reading. Log In or Register to continue

Aug 11, 2021 | Posted by in Fluid Flow and Transfer Proccesses | Comments Off on Case Studies
Premium Wordpress Themes by UFO Themes