You are on page 1of 23

Notice: This manuscript has been authored by UT-Battelle, LLC under Contract No. DE-AC05-00OR22725 with the U.S.

Department of Energy. The United States Government retains and the publisher, by accepting the article for publication,
acknowledges that the United States Government retains a non-exclusive, paid-up, irrevocable, world-wide license to publish or
reproduce the published form of this manuscript, or allow others to do so, for United States Government purposes. The Department
of Energy will provide public access to these results of federally sponsored research in accordance with the DOE Public Access
Plan (http://energy.gov/downloads/doe-public-access-plan).

The Effect Of Shot Peening On Steam Oxidation Of 304H Stainless Steel


J. Matthew Kurley, Bruce A. Pint
Corrosion Science & Technology Group, Materials Science and Technology Division
Oak Ridge National Laboratory, P.O. Box 2008, Oak Ridge, TN 37831-6156 USA

Abstract
Shot peening is currently utilized in coal-fired power plant components to mitigate scale

exfoliation issues from steamside oxidation of austenitic stainless steel superheater and reheater

tubing. To focus on commercially available material, this study exposed quarter-ring specimens

cut from two different commercially shot peened type 304H tubes and exposed to 1 bar steam for

up to 15,000 hours at 550°, 600°, 625°, and 650°C. Specimens were removed at increments to

characterize the oxide thickness and microstructure. The shot peened inner surface generally

retained a thin, protective Cr-rich scale with occasional Fe-rich oxide nodules at 550°-625°C. The

increased oxidation resistance from shot peening began to degrade at 650°C after as little as 5,000

hours. Cut and polished faces of these specimens formed thick, Fe-rich oxides similar to polished

304H coupons. Surprisingly, the mechanically machined outer surface of the tube specimens

performed similarly to the shot peened inner diameter, suggesting it had sufficient cold work to

achieve a similar benefit. Electron backscatter diffraction and Vickers-hardness measurements

were used to characterize post-exposure changes in underlying microstructure and mechanical

properties, respectively, imparted by shot peening.

Key words: stainless steel, shot peening, cold work, steam oxidation, oxide spallation

1
Introduction
Long-term durability of hot section steels continues to limit the durability of coal-fired

boilers, despite decades of experience and research on fireside [1-3] and steamside [4-9] corrosion.

Fireside corrosion remains a complicated issue because of differences in coal chemistry and

variations in temperature and environment in the boiler. Steamside oxidation also is a complex

issue because of chemistry (alloy, oxide, water) and mechanical aspects, including stresses from

coefficient of thermal expansion (CTE) mismatch, oxide growth, thermal gradient [10], and

applied loads. There has been renewed interest in steamside oxidation in the past 20 years because

of boiler tube failures due to scale exfoliation: (1) large quantities of oxide flakes from austenitic

steels blocking steam flow and (2) localized overheating where the scale lifts off without spalling

and creates local hot spots in the underlying ferritic-martensitic (FM) steel [5,8]. Even without

flow blockage, oxide particulates cause erosion of the valves and steam turbine [11,12].

Such issues suggest that steamside oxidation is not fully understood. Changes in boiler

design or operating conditions could be the cause, but changes in alloy composition (within the

specification) may play a role, perhaps due to cleaner steels or leaner Cr and Ni contents. Recent

exfoliation modeling [13-15] has contributed to understanding the role of oxide growth, CTE

mismatch, and tube geometry, but more validation is needed.

Fundamentally, steamside oxidation issues arise from the effect of H2O in accelerating the

oxidation rate of FM and austenitic steels, effectively making them perform as if they have less Cr

[6,7,16-18]. Steels, such as FM Grade 91 (9Cr-1Mo) or type 304H or 347H stainless steel (18Cr,8-

10Ni), form a thin protective Cr-rich oxide when exposed to laboratory air for long periods of time.

However, oxidation in steam is different than oxidation in exhaust gas (i.e. air with water vapor).

In “wet air”, the formation of volatile Cr oxy-hydroxide can accelerate the attack [17,19]. The low

2
oxygen content in high purity steam limits the formation of CrO2(OH)2 [19]. Nevertheless, the

Fe-rich oxides formed on steels exposed to steam or air-water mixtures are significantly thicker

and more likely to exfoliate. Several strategies have been evaluated to inhibit the formation of

thick Fe-rich oxides, including increasing the amount of Cr and Ni in the steel within the

specification or by using a more highly alloyed steel such as type 310H (25Cr,20Ni). Coatings,

such as aluminides, offer a possible solution in the laboratory[20-24] but have not found

widespread commercial use with few field trials reported [25].

The most widely used commercial solution to prevent exfoliation in austenitic steels is to

shot peen (SP) the tube inner diameter (ID) prior to installation [26-32]. Shot peening causes plastic

deformation at the surface, which results in an increase in dislocation density. With more pathways

for Cr to diffuse to the surface, the Cr flux to supply the reaction front is increased, thereby

improving oxidation resistance similar to other cold working strategies studied previously [33-35].

Shot peening was used to protect the 347H tubing in one of the first ultra-supercritical steam plants

in the U.S. [36]. However, relatively little long-term data on steam oxidation of commercially shot

peened austenitic steel is available. This study explored the evolution of two different

commercially shot peened type 304H tubes exposed to 1 bar high purity steam, ranging from 550°

to 650°C and 500 to 15,000 hr.

Experimental Procedure
Compositions of the two tubes studied in the project are shown in Table 1. The detailed

process conditions of the tubes in the current study, including the shot peening conditions, is not

publicly available. To study the behavior of commercial shot peened 304H tubes, a typical

oxidation coupon (Figure 1a) could not be used. Instead, quarter-ring specimens were cut from

the two different (Tube 1 and Tube 2) tubes (Figure 1b) with a 1cm width. The outside diameter
3
(OD) was electro-discharge machined (EDM) to reduce the wall thickness and the other faces were

hand polished to a 600-grit finish to remove the EDM surface residue. Mass change of these

specimens was compared to conventional coupons (11x19x1.5mm) cut from Tube 1 and polished

to a 600-grit finish.

Following polishing, all specimens were cleaned ultrasonically in acetone followed by

methanol. They were dried, dimensioned using a Keyence model VR-3000 microscope, and

weighed with a Mettler Toledo model XP205 balance (±0.04 mg accuracy or ±0.01 mg/cm2) before

exposure. The specimens were placed in an alumina boat and exposed in an alumina reaction tube

with metal flanges on the ends. The specimens were heated in a 3-zone furnace with an Ar cover

gas for ~4 h until equilibrated at the exposure temperature before the introduction of steam. Water

was filtered, deaerated, and deionized to ~0.065µS/cm conductivity and ~10 ppb oxygen before

atomization. The specimens were exposed to 1 bar steam at a temperature range between 550° and

650°C and for total times of 500 to 15,000 h using 500-h cycles. The specimens were cooled to

room temperature in Ar and weighed after each cycle. Following exposure, the specimens were

Cu-plated to protect the surface oxide, metallographically mounted in epoxy, polished, and imaged

with light optical microscopy (LOM). Select specimens were evaluated for Vickers-hardness using

a LECO model M-400-H2 hardness tester with 50 and 500 g loads. Additional specimens were

carbon-coated and imaged with scanning electron microscopy (SEM) and energy dispersive

spectroscopy (EDS) using a TESCAN Mira3 GMH with Oxford Instruments Ultim Max 170 EDS

detector and electron backscatter diffractometry (EBSD) using a Next Generation Oxford

Instruments Symmetry EBSD.

Results

4
Figure 2a summarizes the mass change data in 1 atm steam for the two shot peened 304H

tube specimens and compares the results to one polished 304H coupon exposed at each

temperature. For clarity, the tube specimen mass changes shown are averages for the multiple

specimens exposed with a very small standard deviation. The values for 10 and 15 kh were from

single specimens only. Multiple specimens were exposed at each condition, and one specimen of

each material was removed after 500, 1000, 2500, 5000, 7500, 10,000, and 15,000 h for

metallographic characterization. At 550°C, the 304H coupon exhibited small mass gains followed

by higher mass gains associated with Fe-rich oxide nodule formation. Recall that the shot peened

specimens are only shot peened on one surface, so the mass change data are a composite of the

shot peened ID, the machined OD and the polished sides. At 550°C, the shot peened specimens

initially showed a higher mass gain than the 304H coupon. The small mass loss after ~2,000 h

suggests spallation, perhaps on the polished faces. Tube 2 was exposed for up to 10 kh at 550°C.

At temperatures 600°C or above, the 304H coupon consistently showed higher mass losses than

the tube specimens due to scale spallation. The two sets of shot peened tube specimens showed

smaller mass losses and very similar behavior at 600° and 625°C. At 650°C, the mass losses were

higher for Tube 1 specimens compared to Tube 2 specimens.

For the Tube 1 specimens, Figure 3 shows cross-sections of the shot peened ID for three

different exposure times. At 600°C, a thin, continuous, protective oxide formed under all

conditions (Figures 3a-c). At 625°C, oxide nodules formed after the 7.5 kh exposure with the

distinctive duplex structure associated with an inward- (L1) and outward-growing (L2) oxide layer

(Figure 3f).[37] Larger nodules were observed at 650°C with some observed after only 1 kh

(Figure 3i). Figures 3d, 3h and 3l represent the oxide formed on the machined OD after 15 kh

exposure, which showed similarly protective behavior, particularly at 600°C. On the polished

5
faces of each specimen (Figure 1b) where no cold work was applied by shot peening or machining,

a thicker duplex oxide formed. Figure 4 shows that this oxide was very similar to that formed on

the polished 304H coupons at each temperature. In both cases, a similar duplex oxide formed,

typical of steam-grown oxides on stainless steels at these temperatures. At 550°C, the outer L2

layer remained largely adherent, while at higher temperatures, it was mostly spalled, consistent

with the mass change data in Figure 2a.

To quantify the observed kinetics, Figure 5 shows box and whisker plots of the oxide

thicknesses measured on the ID, OD, and polished faces of each Tube 1 specimen. The whiskers

show the maximum and minimum values measured and the boxes are defined by the 25 and 75%

values with the median value inside the box. As noted in Figure 3, the ID and OD thicknesses were

similar at 600° and 625°C while the polished faces formed much thicker oxide scales. However,

the measured values were more similar at 650°C. At each temperature the thickest oxides were

observed on the polished faces. However, Figure 4 shows that the outer layer on the polished faces

spalled in most cases. The total oxide thickness should be twice the thickness measured because

L1 and L2 should grow at roughly the same rate. In contrast, Figure 3 shows that the nodules that

formed at longer times and higher temperatures did not tend to spall. This difference needs to be

considered when interpreting the results in Figure 5. The SP benefit is actually much larger than

indicated. Especially at 650°C, the median ID and polished oxide thickness shown are closer in

magnitude because of this spallation.

Figure 6 compares the reaction products formed on the two SP tubes after 7.5 or 10 kh

exposures at each temperature. Small Fe-rich oxide nodules were observed in most cases,

particularly at 650°C where the SP benefit was breaking down. Expanding the y-axis in Figure 5,

6
Figure 7 quantifies the ID oxide thickness observations for specimens from both tubes. There was

not a systematic difference in behavior between the two tubes.

Based on the thick oxide formed on the polished specimens, the effect of cold working

(i.e., shot-peening the ID and machining the OD) on the mass change could be calculated by

subtracting the relative contribution (Equation 1) of the polished cut faces as shown in Figure 2b.

𝛥𝑚 ― 𝛥𝑀𝑝𝑜𝑙·𝐴𝑝𝑜𝑙
𝛥𝑀𝑐𝑜𝑙𝑑 = (1)
𝐴𝑐𝑜𝑙𝑑
where ΔMcold is the normalized mass change for the cold worked surfaces (mg/cm2), Δm is the

mass change for all surfaces (mg), ΔMpol is the normalized mass change for the coupons (mg/cm2),

Apol is the surface area of the polished surface (cm2), and Acold is the surface area of the cold-

worked surfaces (cm2). The results were a little more difficult to isolate at 550°C, perhaps because

of the limited dataset where spallation had just begun. However, at 600°C and higher, both Tubes

1 and 2 showed a mass gain with Tube 2 typically showing a higher mass gain, suggesting more

nodule formation, consistent with the comparisons in Figure 6. Increasing temperature resulted in

larger mass gain for the cold-worked surfaces, likely because of increased nodule formation

(Figures 3 & 6) and thicker oxides (Figure 5).

Figures 8-10 show SEM-EDS results from representative areas of the reaction products to

identify the chemical compositions, which were consistent with many observations in the literature

[7,18,19,38-40]. Figure 8 shows a typical duplex oxide nodule. The outer L2 layer predominantly

contained Fe and O with a small amount of Mn (Figures 8c, 8d, 8f), either Fe2O3 and/or Fe3O4.

The inner L1 oxide layer was Cr-rich (Figure 8b) M3O4 [39] and also included Mn (Figure 8f) and

Ni (Figure 8g), likely in metallic form [38,41]. In the substrate adjacent to the scale, Cr and Mn

were depleted. Figure 9 shows an example of the thin oxide formed, which was rich in Cr and Mn

(Figure 9b and 9f). Some Si enrichment was evident at the metal-scale interface in Figure 9e, but

7
the signal was weak. In both cases, Cr-rich precipitates (likely carbides) deeper in the alloy could

be observed (Figures 8b and 9b).

To quantify the composition, example line profiles through a nodule and a thin oxide region

are shown in Figures 10a and 10b, respectively. Consistent with the maps in Figures 8 and 9, the

line profiles show the L2 layer being primarily Fe and O and the inner layer being a mixed oxide

rich in Cr and Mn. Likewise, the thin oxide in Figure 10b also was rich in Cr and Mn. Rather

than showing examples from each temperature, Figure 11 attempts to summarize the composition

observations after 15 kh exposures at 600°-650°C. The Cr/Fe ratios in Figure 11a suggest that the

Cr depletion beneath a nodule was increasing with temperature. Additionally, the Cr enrichment

in the L1 layer was increasing with temperature. The Mn/Cr ratios in Figure 11b indicated that

the Mn content was higher at 650°C both at the metal-oxide interface and at the oxide-gas interface

but the Mn content was highest at the gas interface at all temperatures.

To study the evolution of the microstructure, EBSD analysis was performed on polished

cross-sections of Tube 1 specimens near the ID before and after steam exposure. The as-peened

specimen (Figure 12a) exhibited a highly deformed, non-uniform surface region, with a depth of

20-30 µm from the surface. After exposures in steam for 15 kh, a recrystallized fine grain structure

appeared near the surface with the extent and depth increasing with temperature (Figure 12 b-d).

Some coarsening of the recrystallized grains was evident as the exposure temperature increased.

Finally, micro Vickers-hardness was measured on polished sections as a function of depth

beneath the ID before and after steam exposure. As-peened Tube 1 (Figure 12a, black squares)

showed higher hardness values than as-peened Tube 2 (Figure 12b, black squares). After 10-15

kh, Tube 1 showed a decrease in hardness to ~300 to 350 Hv but some increase in hardness was

retained in all cases. Previously, short anneals at ≥850°C were shown to eliminate the increased

8
hardness in Tube 1 [26]. In contrast, Tube 2 which started with a lower hardness, showed less

decrease in hardness after 5-10 kh exposures compared to Tube 1.

Discussion

Given the importance of SP in commercial coal-fired boilers, SP has not been examined

extensively in the laboratory. Previous reports studied in-house shot peened specimens

continuously exposed to steam [32] or commercial tubes for short periods of time[28]. To avoid

studying SP processing parameters and their relevance to the ID of long superheater tubes, this

study focused on commercially SP tubes. Unfortunately, it was not possible to obtain commercial

tubes of a different alloy and these tubes showed similar performance (Figures 6 and 7). However,

the adjusted mass change data in Figure 2b did suggest that Tube 2 formed more oxide than Tube

1, consistent with the observation that the starting surface hardness was higher in Tube 1 (Figure

13). Interestingly, both tubes relaxed to similar hardness values and those were similar to literature

values [33,42], including after annealing for 4h at 700°C in Ar [26]. The fine recrystallized grains

may increase the hardness due to Hall-Petch strengthening. There was no definite trend associated

with time and temperature, most likely because of variability arising from the shot peening itself.

The hardness measurement can also be affected by the compositional change associated with the

oxidation process, further clouding the trends. Higher temperature annealing could completely

eliminate any increase in hardness [26,42].

The combination of mass change and sequential characterization of the reaction products

provides an indication of the long-term behavior and benefit of SP for these tubes. Above 550°C,

the mass change data was dominated by mass loss associated with the spallation of the outer L2

9
layer from the polished edges of the specimens. The spallation of this layer is central to the

exfoliation problem that SP addresses.

Comparing Figures 3, 5, and 6 illustrates the SP benefit and that it was retained for long

times, particularly at 550-625°C. Unintentionally, the quarter-tube specimens nicely illustrated

the SP benefiting by comparing the cut and polished faces to the SP ID (Figure 5). Surprisingly,

the cold work in the machined OD provided a similar benefit as SP. The machined OD appeared

to exhibit a more uniform reaction product (Figure 3), perhaps because machining imparted a more

uniform surface than the SP process.

The images in Figures 3 and 6 indicate that the shot peening did not completely prevent

the formation of Fe-rich oxide nodules and that the benefit was lost after longer times at 650°C.

This may indicate the commercial SP tubes are not as heavily peened as laboratory specimens,

possibly from difficulties in SP long superheater tubes. It also may reflect the economics of this

process. Incorporating more cold work would require more time or energy. It is possible that this

level of peening was chosen as being sufficient to impart the protection needed to avoid

catastrophic scale exfoliation blockage. However, the small amount of nodule formation is likely

to spall at some point and the spalled oxide could lead to downstream erosion issues.

The reaction products formed were very typical of other studies [7,18,19,38-40] with

similar structures observed (Figures 8-10). The attempt to quantify the Cr/Fe and Mn/Cr ratios is

part of a larger effort to understand the effect of temperature and alloy composition on steam

oxidation products. The results also do not necessarily help to understand the role of H2O in

accelerating the oxidation rate of 304H. It can be argued that shot peening accelerates Cr transport

in the adjacent substrate and prevents a critical Cr depletion from occurring that leads to breakaway

oxidation by chemical or mechanical failure [41]. It is difficult to compare the SP results to the

10
polished 304H coupon results. Any depleted metal in the latter would be consumed by the reaction

front. As 500-h cycles were used in these exposures, there were very few thermal cycles that may

have caused scale cracking and mechanical failure.

The measurements in Figure 7 are somewhat scattered and do not necessarily show a

progression in median oxide thickness with time. This may be due to sampling bias, i.e. the

tendency to take images where more nodules are apparent rather than selecting random areas to

image. An effort is in progress to compare these median values to those obtained from 10

sequential images stitched together.

The EBSD results were used to study the SP microstructure and its evolution after 15 kh at

600°-650°C (Figure 12). In the as-peened specimen, the deformation surface appeared to be non-

uniform with 20-30 µm deep dislocations. The non-uniformity may reflect the difficulty in shot

peening the ID of a long tube and may explain the non-uniform benefit of shot-peening (e.g.

Figures 3). After 15 kh exposures, the deformed region recrystallized to form a fine-grain structure

that coarsened at higher exposure temperatures. The microstructure appeared to be affected to a

greater depth at 650°C (Figure 12d) but this could be just in the area that was examined. In contrast

to the 20-30 µm depth observed here, Nam et al. showed heavy shot peening reaching ~100 µm

into the material and that recrystallization was incomplete after 1 h at 600°C [42].

Conclusions

Specimens from two commercially shot peened type 304H stainless steel tubes were

exposed to 1 bar steam at 550°, 600°, 625°, and 650°C for up to 15 kh to observe the effect on

steam oxidation. A time series of specimens was exposed to examine the microstructure and

reaction product evolution in addition to mass change. The results showed the shot peening was

11
very effective at improving the steam oxidation resistance of 304H up to 625°C. The shot peened

surface formed a thin, protective Cr- and Mn-rich oxide with only limited Fe-rich oxide nodules

forming after ≥7500 h exposures at 625°C. At 650°C, the benefit began to break down after ~5,000

h. In addition to a shot peened ID on the specimens, the machined OD showed improved oxidation

resistance. However, the mass change was dominated by the growth of oxide on the specimen

edges that were not cold-worked. Oxide thickness measurements as a function of time and

temperature quantified these observations. The two tubes showed similar behavior with a slightly

higher hardness measured in one tube corresponding to slightly less mass gain at the highest

temperatures. Characterization of the reaction products showed similar results observed in other

studies and minor changes in Cr depletion in the metal with temperature. EBSD results from the

higher hardness tube showed that the surface recrystallized and the fine grains coarsened at higher

temperatures. This was consistent with a measured drop in hardness profiles at the peened surface.

Acknowledgments

The authors would like to thank G. W. Garner, M. Howell, T. M. Lowe and T. Jordan for

assistance with the experimental work. American Electric Power donated the shot peened 304H

tubes for this study. R. Pillai and Y. Yamamoto at ORNL provided helpful comments on the

manuscript. This research was sponsored by the U. S. Department of Energy, Office of Fossil

Energy, Cross-cutting Technology Program.

References

1. A. J. B. Cutler and E. Raask, External corrosion in coal-fired boilers: Assessment from


laboratory data. Corrosion Science 21, 1981 (789-800).
2. D. B. Meadowcroft, High temperature corrosion of alloys and coatings in oil- and coal-
fired boilers. Materials Science and Engineering 88, 1987 (313-320).

12
3. J. Henry, G. Zhou and T. Ward, Lessons from the past: materials-related issues in an
ultra-supercritical boiler at Eddystone plant. Materials at High Temperatures 24, 2007
(249-258).
4. I. M. Rehn, W. R. Apblett Jr and J. Stringer, Controlling steamside oxide exfoliation in
utility boiler superheaters and reheaters. Materials Performance 20, 1981 (27-31).
5. N. Nishimura, N. Komai, Y. Hirayama and F. Masuyama, Japanese experience with
steam oxidation of advanced heat-resistant steel tubes in power boilers. Materials at High
Temperatures 22, 2005 (3-10).
6. S. R. J. Saunders and N. L. McCartney, Current understanding of steam oxidation -
Power plant and laboratory experience. Materials Science Forum 522-523, 2006 (119-
128).
7. I. G. Wright and R. B. Dooley, A review of the oxidation behaviour of structural alloys in
steam. International Materials Reviews 55, 2010 (129-167).
8. 2nd EPRI-NPL workshop on scale exfoliation from steam-touched surfaces. EPRI, Palo
Alto, CA: 2012. 1026663.
9. J. P. Shingledecker, B. Pint, A. Sabau, A. Fry and I. G. Wright, Managing steam-side
oxidation and exfoliation in USC Boiler Tubes. Advanced Materials & Processes 171,
2013 (23-25).
10. A. T. Fry, L. J. Brown and J. P. Banks, The effect of heat flux on the steam oxidation
kinetics and scale morphology of low alloy materials. Proceedings from the 6th
International Conference on Advances in Materials Technology for Fossil Power Plants
2010 pp. 171-184 (2011).
11. M. M. Stack, J. Chacon-Nava and F. H. Stott, Relationship between the effects of
velocity and alloy corrosion resistance in erosion-corrosion environments at elevated
temperatures. Wear 180, 1995 (91-99).
12. S.-S. Wang, L. Cai, J.-R. Mao, J.-J. Zhang and Y.-T. Xu, Mechanisms of steam turbine
blade particle erosion and crucial parameters for minimizing blade erosion. Proceedings
of the Institution of Mechanical Engineers, Part A: Journal of Power and Energy 227,
2013 (546-556).
13. I. G. Wright, M. Schutze, P. F. Tortorelli and R. B. Dooley, Towards improved prediction
of scale exfoliation from steam tubes. Materials at High Temperatures 24, 2007 (265-
274).
14. I. G. Wright, A. S. Sabau and R. B. Dooley, Development of strain in oxides grown in
steam tubes. Materials Science Forum 595-598, 2008 (387-395).
15. A. S. Sabau and I. G. Wright, Influence of oxide growth and metal creep on strain
development in the steam-side oxide in boiler tubes. Oxidation of Metals 73, 2010 (467-
492).
16. C. T. Fujii and R. A. Meussner, The mechanism of the high-temperature oxidation of
iron-chromium alloys in water vapor. Journal of The Electrochemical Society 111, 1964
(1215-1221).
17. H. Asteman, J. E. Svensson, M. Norell and L. G. Johansson, Influence of water vapor and
flow rate on the high-temperature oxidation of 304L; Effect of chromium oxide
hydroxide evaporation. Oxidation of Metals 54, 2000 (11-26).
18. W. J. Quadakkers, J. Żurek and M. Hänsel, Effect of water vapor on high-temperature
oxidation of FeCr alloys. JOM 61, 2009 (44-50).

13
19. D. J. Young and B. A. Pint, Chromium volatilization rates from Cr2O3 scales into flowing
gases containing water vapor. Oxidation of Metals 66, 2006 (137-153).
20. R. Sivakumar and E. J. Rao, An investigation of pack-aluminide coating on steel.
Oxidation of Metals 17, 1982 (391-405).
21. N. V. Bangaru and R. C. Krutenat, Diffusion coatings of steels: Formation mechanism
and microstructure of aluminized heat‐resistant stainless steels. Journal of Vacuum
Science & Technology B: Microelectronics Processing and Phenomena 2, 1984 (806-
815).
22. D. M. Miller, S. C. Kung, S. D. Scarberry and R. A. Rapp, Simultaneous chromizing-
aluminizing coating of austenitic stainless steels. Oxidation of Metals 29, 1988 (239-254).
23. D. Wang, Corrosion behavior of chromized and/or aluminized 214Cr-1Mo steel in
medium-BTU coal gasifier environments. Surface and Coatings Technology 36, 1988
(49-60).
24. A. Agüero, R. Muelas, M. Gutiérrez, R. Van Vulpen, S. Osgerby and J. P. Banks, Cyclic
oxidation and mechanical behaviour of slurry aluminide coatings for steam turbine
components. Surface and Coatings Technology 201, 2007 (6253-6260).
25. J. Allenou, J. Piron Abellan, E. Hugon, C. Landier and T. Le Guevel, Influence of
aluminium diffusion layer on T/P92 steam oxidation resistance - A laboratory and field
study. Proceedings from the 8th International Conference on Advances in Materials
Technology for Fossil Power Plants 2016 (2016).
26. B. M. Tossey, H. Khan and T. Andress, Steam oxidation resistance of shot peened
austenitic stainless steel superheater tubes. NACE Corrosion International Conference
2011 pp. 186 (2011).
27. B. A. Pint, J. P. Shingledecker and I. G. Wright, Characterization of steam oxidation
products from field-exposed tubes. International Conference on Advances in Condition
and Remaining Life Assessment for Fossil Power Plants – Coal, Gas and Heat Recovery
Steam Generator (2012).
28. J. C. Rosser, M. I. Bass, C. Cooperet al. , Steam oxidation of Super 304H and shot-
peened Super 304H. Materials at High Temperatures 29, 2012 (95-106).
29. R. Naraparaju, H. J. Christ, F. U. Renner and A. Kostka, Effect of shot-peening on the
oxidation behaviour of boiler steels. Oxidation of Metals 76, 2011 (233).
30. R. Naraparaju, H. J. Christ, F. U. Renner and A. Kostka, Dislocation engineering and its
effect on the oxidation behaviour. Materials at High Temperatures 29, 2012 (116-122).
31. V. Bystrianský, A. Krausová, J. Macák, J. Děd, E. Eltai and A. M. Hamouda, Beneficial
effect of shot peening on steamside oxidation of 300-series austenitic steels: An
electrochemical study. Applied Surface Science 427, 2018 (680-685).
32. H. Ma, Y. He, S.-Y. Bae and K. Shin, Effect of shot peening on oxidation behavior of
SS304H upon long term steam exposure. Journal of Nanoscience and Nanotechnology
18, 2018 (6167-6172).
33. D. Caplan, Effect of cold work on the oxidation of FeCr alloys in water vapour at 600°C.
Corrosion Science 6, 1966 (509-515).
34. D. Caplan and M. Cohen, Effect of cold work on the oxidation of iron from 400–650 °C.
Corrosion Science 6, 1966 (321-335).
35. A. S. Khanna and J. B. Gnanamoorthy, Effect of cold work on the oxidation resistance of
2.25Cr-1 Mo steel. Oxidation of Metals 23, 1985 (17-33).

14
36. B. A. Pint, High-temperature corrosion in fossil fuel power generation: Present and
future. JOM 65, 2013 (1024-1032).
37. S. R. J. Saunders, M. Monteiro and F. Rizzo, The oxidation behaviour of metals and
alloys at high temperatures in atmospheres containing water vapour: A review. Progress
in Materials Science 53, 2008 (775-837).
38. N. Otsuka, Y. Shida and H. Fujikawa, Internal-external transition for the oxidation of Fe-
Cr-Ni austenitic stainless steels in steam. Oxidation of Metals 32, 1989 (13-45).
39. B. A. Pint, The effect of shot peening on steam oxidation of stainless steel at 600°-650°C.
NACE Corrosion International Conference 2017 11 (2017).
40. J. Ehlers, D. J. Young, E. J. Smaardijket al. , Enhanced oxidation of the 9%Cr steel P91
in water vapour containing environments. Corrosion Science 48, 2006 (3428-3454).
41. R. Peraldi and B. A. Pint, Effect of Cr and Ni contents on the oxidation behavior of
ferritic and austenitic model alloys in air with water vapor. Oxidation of Metals 61, 2004
(463-483).
42. K. Nam, Y. He and K. Shin, Microstructural evolution of Super304H upon ultrasonic
shot peening and subsequent annealing. Journal of Nanoscience and Nanotechnology 18,
2018 (6274-6277).

List of Figures
Figure 1. Images of 304H coupon and tube specimens: 1.5mm thick rectangular coupon (a) and
quarter-ring (b) specimen. The coupons were polished to a 600-grit finish on all faces.
The quarter-ring specimens were polished to a 600-grit finish on all the cut-faces with a
shot peened ID and a machined OD.
Figure 2. (a) Mass change data of 304H specimens exposed to 1 bar steam at 550-650oC for up to
15kh. Specimens were polished coupons (black squares) and quarter-ring sections of two
different shot peened tubes (Tube 1, red circles and Tube 2, blue triangles). (b) Mass
change data showing the cold-worked surfaces (shot-peened and machined) only.
Figure 3. LOM images of the shot peened ID and machined OD of 304H quarter-ring specimens
from Tube 1 exposed to 1 bar steam for different times at 600oC (a-d), 625oC (e-h), and
650oC (i-l). Samples were exposed for 1kh (a,e,i), 7.5kh (b,f,j), and 15kh (c,d,g,h,k,l)
Figure 4. LOM images of cross-sectioned 304H specimens exposed to 1 bar steam for 5kh at
550o (a,b), 600o (c,d), 625o (e,f), and 650oC (g,h). Images focused on the polished surface
of coupons (a,c,e,g) and quarter-ring specimens (b,d,f,h)
Figure 5. Box and whisker plot of measured oxide thickness of the cut face (black), shot peened
ID (red), and machined OD (blue) for Tube 1 exposed to 1 bar steam at 600o (top), 625o
(middle), and 650oC (bottom)

15
Figure 6 LOM images of shot peened 304H quarter-ring specimens from Tube 1 (a,c,e,g) and
Tube 2 (b,d,f,h) exposed to 1 bar steam at 550oC (a,b) for 7.5kh and 600oC (c,d), 625oC
(e,f), and 650oC (g,h) for 10kh
Figure 7. Box and whisker plot of measured oxide thickness of the shot peened ID for Tube 1
(black, wider box) and Tube 2 (red, thinner box) exposed to 1 bar steam at 600o (top),
625o (middle), and 650oC (bottom)
Figure 8. SEM image and EDS maps of a nodule formed on the shot peened ID from the quarter-
ring specimen exposed to 1 bar steam at 650oC for 2.5kh. (a) Secondary electron image
and Cr (b), O (c), Fe (d), Si (e), Mn (f), Ni (g), and Cu (h) EDS maps
Figure 9. SEM image and EDS maps of the protective oxide between nodules on the shot peened
ID from the quarter-ring specimen exposed to 1 bar steam at 650oC for 2.5kh. Secondary
electron image (a) and Cr (b), O (c), Fe (d), Si (e), Mn (f), Ni (g), and Cu (h) EDS maps
Figure 10. EDS line scans of a nodule (a) and the protective oxide (b) between nodules formed
on the shot peened ID from the quarter-ring specimen exposed to 1 bar steam at 650oC
for 2.5kh. Mapped elements: Cr (black), Fe (red), O (blue), Mn (purple), and Ni (green)
Figure 11. EDS quantification after 15,000 h exposures of (a) Cr/Fe ratios in or below the
nodules and (b) Mn/Cr ratio of the thin oxide. (a) Quantification was performed at the
level of maximum Cr depletion and at the level of maximum Cr-enrichment in L1. (b)
Quantification was performed at the metal/oxide interface and at the oxide/steam
interface
Figure 12. EBSD Euler map of shot-peened Tube 1 304H quarter-ring specimens as-peened (a)
and exposed to 1 bar steam at 600o (b), 625o (c), and 650oC (d) for 15kh
Figure 13. Hardness as a function of distance from the shot peened surface for Tube 1 (a) and
Tube 2 (b). Specimens were measured as-peened (black squares) and after exposure to 1
bar steam for 5kh (red diamonds), 10kh (green circles), and 15kh (purple triangles) at
600oC to 650oC

16
Table 1. Elemental compositions of the 3 different 304H samples in weight percent measured by ICP and combustion analyses.

Sample Fe Cr Ni Mn Mo Si C S(ppm) Other


Tube 1 70.0 18.3 8.6 1.8 0.3 0.3 0.07 51 0.4Cu,0.06N
Tube 2 70.7 18.2 8.4 1.2 0.3 0.4 0.08 21 0.4Cu,0.1Co,0.06N

Fig. 1 Images of 304H coupon and tube specimens: 1.5mm thick rectangular coupon (a) and quarter-ring (b) specimen.
The coupons were polished to a 600-grit finish on all faces. The quarter-ring specimens were polished to a 600-grit finish
on all the cut-faces with a shot peened ID and a machined OD.

Fig. 2 (a) Mass change data of 304H specimens exposed to 1 bar steam at 550-650oC for up to 15kh. Specimens were
polished coupons (black squares) and quarter-ring sections of two different shot peened tubes (Tube 1, red circles and
Tube 2, blue triangles). (b) Mass change data showing the cold-worked surfaces (shot peened and machined) only.

17
Fig. 3 LOM images of the shot peened ID and machined OD of 304H quarter-ring specimens from Tube 1 exposed to 1
bar steam for different times at 600oC (a-d), 625oC (e-h), and 650oC (i-l). Samples were exposed for 1kh (a,e,i), 7.5kh
(b,f,j), and 15kh (c,d,g,h,k,l)

Fig. 4 LOM images of cross-sectioned 304H specimens exposed to 1 bar steam for 5kh at 550o (a,b), 600o (c,d), 625o
(e,f), and 650oC (g,h). Images focused on the polished surface of coupons (a,c,e,g) and quarter-ring specimens (b,d,f,h)

18
0.0 2.5 5.0 7.5 10.0 12.5 15.0
60

Oxide Thickness (m) 40 Polished


Shot-peened
Shot-Peened 550oC
20 Shot-peened
Machined

0
100
75
50 600oC
25
0
100
75
50 625oC
25
0
100
75
50 650oC
25
0
0.0 2.5 5.0 7.5 10.0 12.5 15.0
Time (kh)
Fig. 5 Box and whisker plot of measured oxide thickness of the cut face (black), shot peened ID (red), and machined OD
(blue) for Tube 1 exposed to 1 bar steam at 550° (top), 600o (upper middle), 625o (lower middle), and 650oC (bottom)

Fig. 6 LOM images of shot peened 304H quarter-ring specimens from Tube 1 (a,c,e,g) and Tube 2 (b,d,f,h) exposed to 1
bar steam at 550oC (a,b) for 7.5kh and 600oC (c,d), 625oC (e,f), and 650oC (g,h) for 10kh

19
0.0 2.5 5.0 7.5 10.0 12.5 15.0
6
5 5.0 Tube 1
4 Tube 2
Oxide Thickness (m)
3.3
3 3.0
2.6
2.5 2.3
2
1 0.8 1.5 550oC
0
12
10
8
6 6.4 600oC 5.4
4 1.3
3.9 3.4 4.0
2.5 2.8
2 1.8 2.0
0
20
15 15.5
625oC
14.0

10 9.4
8.3
5 2.7 6.7 4.9 6.1
1.2 2.0 1.2
0
30
25 25.2
20
650oC
20.3
15 13.8
11.5
10 9.7 9.7 10.2
2.0 7.2
5 3.8
6.5
0
0.0 2.5 5.0 7.5 10.0 12.5 15.0
Time (kh)
Fig. 7 Box and whisker plot of measured oxide thickness of the shot peened ID for Tube 1 (black, left box) and Tube 2
(red, right box) exposed to 1 bar steam at 550° (top), 600o (upper middle), 625o (lower middle), and 650oC (bottom)

Fig. 8 SEM image and EDS maps of a nodule formed on the shot peened ID from the quarter-ring specimen exposed to
1 bar steam at 650oC for 2.5kh. (a) Secondary electron image and Cr (b), O (c), Fe (d), Si (e), Mn (f), Ni (g), and Cu (h)
EDS maps

20
Fig. 9 SEM image and EDS maps of the protective oxide between nodules on the shot peened ID from the quarter-ring
specimen exposed to 1 bar steam at 650oC for 2.5kh. Secondary electron image (a) and Cr (b), O (c), Fe (d), Si (e), Mn
(f), Ni (g), and Cu (h) EDS maps

Fig. 10 EDS line scans of a nodule (a) and the protective oxide (b) between nodules formed on the shot peened ID from
the quarter-ring specimen exposed to 1 bar steam at 650oC for 2.5kh. Mapped elements: Cr (black), Fe (red), O (blue),
Mn (purple), and Ni (green)

21
Fig. 11 EDS quantification after 15 kh exposures of (a) Cr/Fe ratios in or below the nodules and (b) Mn/Cr ratio of the
thin oxide. (a) Quantification was performed at the level of maximum Cr depletion and at the level of maximum Cr-
enrichment in L1. (b) Quantification was performed at the metal/oxide interface and at the oxide/steam interface

Fig. 12 EBSD Euler map of shot peened Tube 1 304H quarter-ring specimens as-peened (a) and exposed to 1 bar steam
at 600o (b), 625o (c), and 650oC (d) for 15kh

22
Fig. 13 Hardness as a function of distance from the shot peened surface for Tube 1 (a) and Tube 2 (b). Specimens were
measured as-peened (black squares) and after exposure to 1 bar steam for 5kh (red diamonds), 10kh (green circles), and
15kh (purple triangles) at 600oC to 650oC

23

You might also like