Abstract

In an effort to achieve worldwide decarbonization goals, a range of low-carbon fuels is being proposed for use in power-generation gas turbines. Two promising fuels are hydrogen and ammonia, which do not produce any CO2 upon combustion. While the use cases of these fuels differ, each has the potential to reduce the carbon intensity of power generation as compared to the current use of natural gas and fuel oil. Many studies have considered the impact that these fuels will have on combustion stability and emissions, as well as other practical considerations like balance of plant. However, potential challenges for the material systems in these engines, particularly high-temperature metal alloys and coatings, have not been sufficiently considered in preparation for the introduction of these fuels. The goal of this paper is to provide a review of the potential material issues associated with implementation of these hydrogen-containing fuels, with a focus on materials in the hot section of existing power-generation gas turbines. To date, relatively little research has considered these material issues at realistic gas turbine conditions, resulting in a need for new research. This paper provides a review of the literature, first-order analyses of the magnitude of potential issues, and avenues for research to facilitate the safe and reliable introduction of these fuels in power-generation gas turbines.

Introduction

Decarbonization of the power-generation sector will require an “all-of-the-above” approach to ensure reliability, affordability, and stability of the electrical grid. With many countries and regions adopting 2030, 2040, and 2050 range goals, the reality of these targets hinges on optionality. Neither renewables, energy storage, carbon capture, nor low-carbon fuels singularly can achieve net-zero carbon emissions, especially in light of rapid and widespread electrification. Currently in the United States, approximately 20% of all final energy is electricity and this number is projected to double (∼43%) or triple (∼59%) by 2050 [1]. In addition, models around capacity needs for 2050 vary widely; however, some things are certain: increased capacity is necessary and current installation trends are not sufficient. It is predicted that wind and solar will need 160% to 480% growth in capacity, respectively, to help meet the demand while decarbonizing the energy sector. Compared with 2023 installation trends, year over year to 2050, the country will be 20–87% short of these expansion needs—note this does not include retirements over the next 26 years [24]. Thus, gas turbines will continue to be a vital part of the electricity generation portfolio, whether with carbon capture or utilizing low-carbon fuels.

One of the key components of a stable, decarbonized grid is the availability of clean dispatchable power. Today, gas turbine engines play a vital role in providing a range of grid services, including base-load power, frequency stability, and peaking. Going forward, their role in grid balancing will become more critical. California is a key example of the rapidly changing needs in demand response with high renewable penetration, as shown in Fig. 1. These curves show the demand for nonrenewable generation as a function of hour throughout a representative day for three different years: 2013, 2018, and 2023. The “duck curve” of 2018 has been replaced by the “canyon curve” of 2023, which shows the significant cycling necessary for nonrenewable power generation resources.

Fig. 1
Change in demand response over a decade of increasing renewable penetration—transition from “duck curve” to “canyon curve” [1]
Fig. 1
Change in demand response over a decade of increasing renewable penetration—transition from “duck curve” to “canyon curve” [1]
Close modal

Moreover, as sunlight and wind availability are outside of human control, renewable energy cannot be the only means of generation. Likewise, the rate of additional energy storage capacity addition, including batteries and pumped hydro, is too slow to keep up with projected demand. Further, several regions require seasonal storage, which is not possible with modern battery technologies. Figure 2 is an example of a regional need for seasonal storage from the Electric Reliability Council of Texas (ERCOT), showing demand on peak summer and winter days. Note that ERCOT, as of the time of this publication, is the grid with the highest penetration of solar and wind capacity. From this chart, wind and solar provide a small part of the peak winter needs and only meet a portion of the summer demand. Hence, low-carbon, dispatchable generation is the key driver for meeting critical demand load levels.

Fig. 2
Electric Reliability Council of Texas winter and summer peak demand examples by generation type. Figure courtesy of Caravaggio, EPRI; data via [5].
Fig. 2
Electric Reliability Council of Texas winter and summer peak demand examples by generation type. Figure courtesy of Caravaggio, EPRI; data via [5].
Close modal

The use of hydrogen-containing fuels, including hydrogen and ammonia, provides a low-carbon pathway for gas turbines to play these key roles in power generation. Figure 3 describes the carbon intensity of H2 production based upon various production pathways. Some production pathways have very low carbon intensity, showing the potential role of these fuels for long-duration energy storage. For example, hydrogen could be produced using electrolysis when there is excess nuclear and/or renewable (wind or solar) energy available and then used for electricity production with gas turbines when demand is high.

Fig. 3
Carbon intensity of hydrogen for several production pathways [6]
Fig. 3
Carbon intensity of hydrogen for several production pathways [6]
Close modal

Significant research is being done to understand the impact of these fuels on engine operability and emissions. Several reviews [7,8] discuss the implications for hydrogen and ammonia on combustion. In general, the higher flame speeds of hydrogen and lower flame speeds of ammonia can dramatically impact flame stabilization, which in turn alters operability limits and can change engine control procedures. As a first step, engines will likely use a blend of gases rather than neat fuels. For example, demonstrations of over 40% hydrogen blended into natural gas [9] have shown the potential for the use of hydrogen in the current fleet. Further, several researchers have proposed the use of blends of hydrogen and ammonia to overcome the aforementioned flame speed issues [10]. In all these cases, a wide spectrum of research, from fundamental chemistry to plant-level demonstrations, is considering the operability effects of these renewable fuels.

However, equally important for gas turbine operators is the implication of these fuels for material durability, lifing, and repair. Traditional life management consists of developing an understanding of the impacts of operation, design, fabrication practices, and metallurgy of materials on overall performance. In many cases, failures occur due to a compounding effect of two or more of these aspects. A change in fuel or fuel blending may have an impact on any of these features, motivating our need to understand the effect of fuel on engine materials. Life management practices are also often built around knowledge developed from field experiences and, to date, only a limited number of field studies have been conducted with hydrogen blending in industrial land-based gas turbines [9] and there are no field demonstrations of ammonia in these machines.

Currently, the most comprehensive review of the material issues associated with the introduction of hydrogen fuels in gas turbine engines was published by Stefan et al. [11]. This paper provides a literature survey on the implications of hydrogen on gas turbine material systems, including high-temperature alloys and coatings. However, certain key elements are not considered, including changes to radiative heat transfer and the impact of changes in operability on mechanical stresses. Further, the only fuel considered is hydrogen, whereas in this review we expand our consideration to hydrogen, ammonia, and blends of these fuels. A more recent review on both hydrogen and ammonia by Alnaeli et al. [12] focuses mostly on combustion phenomenon, without detailed consideration of the impact of fuel composition on materials systems.

This paper provides an overview of the potential materials issues that may arise from the use of hydrogen-containing fuels in power-generation gas turbines. The focus is on materials systems in engines that are in operation today where fuel blending or retrofits are being considered. Thus, next-generation, high-temperature materials like ceramic matrix composites are not considered. We begin by discussing the interaction mechanisms that arise from the use of these fuels, including changes to heat transfer, mechanical stresses, and material compatibility (oxidation, corrosion, etc.). We use first-order analysis of the potential interactions to understand the expected magnitude of each effect as the fuel composition changes. The final impact of these changes on current designs is unfortunately unknown, as the margin for material degradation that is incorporated into current engine designs is highly protected intellectual property of each turbine manufacturer. As such, we highlight the potential changes that may occur and leave it to the reader to infer the potential effects on a given design. For each mechanism, we provide a literature review of previous work on the impacts of these fuels on high-temperature metal alloys and thermal barrier coatings (TBC) typically found in power-generation gas turbines. Finally, we outline research and development topics necessary for the successful introduction of these fuels into the fleet.

Impact of Fuels on Heat Transfer

In this section, we discuss the different ways in which fuels can impact heat transfer to hot section components. In this first-order analysis, we assume that the engines are running a typical Brayton cycle and we consider a baseline architecture to estimate the impact of fuel composition on heat transfer. The baseline is a GE 7 F frame engine with a combustor inlet condition of 408 °C (766.4 °F) and 1864.38 kPa (270.4 PSI) burning natural gas, which is modeled as methane. The 7 F was chosen given its prevalence in power generation in the United States [13].

Convective Heat Transfer.

Variations in fuel composition may affect convective heat transfer to hot-section components. Here, the heat capacity (cp) and thermal conductivity (k) of the gases are sensitive to changes in the products of combustion. To estimate the magnitude of these changes, we calculate the equilibrium composition of different blends of hydrogen and natural gas as well as hydrogen and ammonia at a constant adiabatic flame temperature, 1526.85 °C (1800 K, 2780.33 °F); this condition is consistent with previous calculations by Breer et al. [14] focusing on the impact of fuel blending on emissions. At this condition, the equivalence ratio of the 100% CH4 case is 0.508, of the 100% H2 case is 0.424, and of the 100% NH3 case is 0.542, representative of lean operation in a dry low-emissions combustion system [15]. Calculations are performed in Chemkin Pro [16] using GRIMech 3.0 [17] for mixtures of methane and hydrogen and the Mei 2019 mechanism [18] for mixtures of ammonia and hydrogen. For each fuel composition, the equivalence ratio in air is varied to maintain an adiabatic flame temperature within 2 °C of 1526.85 °C. From these compositions, the mixture-averaged cp and k are calculated using ideal gas, linear blending rules at a temperature of 1526.85 °C. The heat capacity for each species was determined using the NASA curve fits [19] and the thermal conductivity was determined from the NIST Chemistry Webbook [20]. Where data were not available from the Webbook, linear extrapolation to the desired temperature was used.

Table 1 shows the results of the equilibrium calculations for H2/CH4 and H2/NH3 mixtures. Note that in both cases, only the major product species are provided. In these calculations, the mole fraction of N2 is increased to account for all other minor species, ensuring that the sum of the mole fractions is unity. There are several interesting results from these calculations. First, the high heat capacity species, CO2 and H2O, vary significantly with fuel composition. In the case of H2/CH4 blends, CO2 mole fraction goes to zero at 100% hydrogen, whereas the mole fraction of water increases from ∼6.5% to ∼11% as the fuel composition is varied from 100% CH4 to 100% H2. While no CO2 is produced in any of the H2/NH3 flames, the trend in H2O is opposite with H2 content in these flames. The product mole fraction of H2O decreases from ∼13% to ∼11% as fuel composition is varied from 100% NH3 to 100% H2. In general, the heat capacity and thermal conductivity trend with the mole fraction of H2O in the products. The variation in these quantities is of the order of a few percent.

Table 1

Equilibrium product mole fractions, mixture-averaged heat capacity, and thermal conductivity, for mixtures of hydrogen in methane and ammonia at an adiabatic flame temperature of 1800 K

H2/CH4 mixtures
H200.10.20.30.40.50.60.70.80.91
H2O0.06460.06610.06770.06990.07200.07510.07840.08310.08900.09720.1097
CO0.00000.00000.00000.00000.00000.00000.00000.00000.00000.00000.0000
CO20.07900.07660.07360.07040.06600.06120.05480.04690.03630.02160.0000
O20.10960.11010.11110.11160.11310.11410.11610.11810.12100.12510.1306
N20.74670.74710.74760.74810.74890.74960.75070.75200.75370.75610.7597
cp (J/kg-K)1328.81329.913311332.71334.21336.613391342.61347.11353.51363.6
k (W/m-K)0.10960.10970.10980.10990.11000.11020.11030.11050.11080.11130.1119
H2/NH3 mixtures
H2O0.13030.12870.12710.12500.12340.12130.11920.11710.11450.11230.1097
O20.09620.09890.10170.10500.10790.11130.11480.11830.12250.12620.1305
N20.77350.77240.77120.77000.76880.76740.76610.76460.76310.76150.7598
cp (J/kg-K)1389.81387.71385.613831380.91378.21375.61372.91369.61366.91363.6
k (W/m-K)0.11330.11320.11310.11290.11280.11270.11250.11240.11220.11210.1119
H2/CH4 mixtures
H200.10.20.30.40.50.60.70.80.91
H2O0.06460.06610.06770.06990.07200.07510.07840.08310.08900.09720.1097
CO0.00000.00000.00000.00000.00000.00000.00000.00000.00000.00000.0000
CO20.07900.07660.07360.07040.06600.06120.05480.04690.03630.02160.0000
O20.10960.11010.11110.11160.11310.11410.11610.11810.12100.12510.1306
N20.74670.74710.74760.74810.74890.74960.75070.75200.75370.75610.7597
cp (J/kg-K)1328.81329.913311332.71334.21336.613391342.61347.11353.51363.6
k (W/m-K)0.10960.10970.10980.10990.11000.11020.11030.11050.11080.11130.1119
H2/NH3 mixtures
H2O0.13030.12870.12710.12500.12340.12130.11920.11710.11450.11230.1097
O20.09620.09890.10170.10500.10790.11130.11480.11830.12250.12620.1305
N20.77350.77240.77120.77000.76880.76740.76610.76460.76310.76150.7598
cp (J/kg-K)1389.81387.71385.613831380.91378.21375.61372.91369.61366.91363.6
k (W/m-K)0.11330.11320.11310.11290.11280.11270.11250.11240.11220.11210.1119

A canonical example is used to determine the change in heat transfer coefficient with fluid properties assuming that the flow velocity is held constant. The purpose of this analysis is not to provide realistic values for heat transfer coefficient in a combustor or turbine, but instead isolate the sensitivity of fundamental convective processes to gas composition. In these calculations, the heat transfer coefficient, h, is calculated for a laminar flat plate boundary layer and then normalized by hbaseline, which is at the 100% methane condition. Mixture-averaged densities and viscosities were used to calculate the mixture-averaged Prandtl number for all cases; density and viscosity values for each specie were obtained from the NIST Chemistry Webbook and linear extrapolation was used if the property values were not available at the desired temperature.

Figure 4 shows the variation in the ratio of the heat transfer coefficient to the baseline for mixtures of H2/CH4 and H2/NH3 as the fuel mole fraction of H2 varies from 0 to 1. The mixture with the highest heat transfer coefficient is the 100% ammonia case, which is almost 4% higher than that of the baseline 100% methane case. Assuming a baseline wall temperature of 926.85 °C (1200 K, 1700.33 °F), this increase in the heat transfer coefficient could produce a 17 °C (62.6 °F) increase in wall temperature, resulting in the need for more cooling, material or design changes, or reduction of current part life if this increase falls outside the margin currently designed into the parts.

Fig. 4
Variation in normalized heat transfer coefficient with reactant H2 mole fraction
Fig. 4
Variation in normalized heat transfer coefficient with reactant H2 mole fraction
Close modal

Few studies have considered the impact of fuel composition on convective heat transfer in hot section components. An early study by Lefebvre [21] indicated that convective heat transfer to the liner was difficult to measure and model; he provided a relatively simplistic model for heat transfer in straight pipes to capture the effects of mass flow rate, gas properties, and gas temperature on the wall temperature. More recently, a program at the National Energy Technologies Laboratory (NETL) studied next-generation materials and cooling systems for syngas-based, integrated gasification combined-cycle (IGCC) systems [2226]. Syngas fuels can have up to 40% hydrogen in the fuel, with the remaining species including CO, CO2, CH4, and H2O [27], making the results relevant to the high-hydrogen fuel blends under consideration here. These studies considered the impact of gas composition on vane heat transfer using CFD, including the use of a P-1 radiation model. The mixtures considered both oxy-fuel combustion, where the product gases have significant mole fractions of H2O and CO2, as well as a regular air-based cycle with syngas fuels. The results show that the mixtures with high levels of H2O and CO2 have significantly higher heat transfer coefficients, as can be seen in Fig. 5, which shows heat transfer coefficient for a turbine blade. A first-principles analysis, similar to the one done in this paper, shows very good agreement with the high-fidelity simulation, indicating that the reason for the increase in heat transfer coefficient is the change in heat capacity and thermal conductivity in the product gases.

Fig. 5
Surface heat transfer coefficient at turbine vane half-span for three different gas compositions (bottom table) at 300 psi, 2600 °F, and 0.45 Mach number. Reprinted with permission from Chyu et al. [23].
Fig. 5
Surface heat transfer coefficient at turbine vane half-span for three different gas compositions (bottom table) at 300 psi, 2600 °F, and 0.45 Mach number. Reprinted with permission from Chyu et al. [23].
Close modal

Work by Chiesa et al. [28] used a correlation from Louis [29] to estimate the impact of the change in product composition on vane surface heat transfer, yielding similar conclusions to those in this work. A more fundamental study by Zou and Yuan [30] considered the change in gas composition on flat plate heat transfer. This experimental and computational study used a flat plate with a row of film cooling holes at the entrance of the channel, varying the water mole faction in air, static pressure, and static temperature; a baseline case showed very good agreement between the experiment and the simulation. The simulation used a gray gas model for the gas-phase radiation contribution. At all pressures and temperatures, increasing the mole fraction of H2O in the mainstream increased heat flux to the wall. Correlations for the dependence of the non-dimensional heat transfer on all three parameters were developed. Finally, some studies mention changes to the heat transfer with variations in fuel composition or product composition but did not provide any further investigation or discussion, including work by Gazzani et al. [31] and Goke et al. [32].

Radiative Heat Transfer.

Radiative heat transfer can account for a non-negligible amount of the total heat flux to the surface of gas turbine hot-section components. Work by Lefebvre showed the importance of radiation on combustor liner temperature in a study that considered the impact of fuel composition on gas turbine combustor liner temperature, pattern factor, and emissions [21]. This study used simplified scalings for radiation to the combustor liner to fit data from several engines, including a F101, a T56, and a J79. In general, liner temperature decreases with increasing hydrogen content of the hydrocarbon fuels, a result of the lower soot formation that occurs with less carbon in the fuel. As such, the authors point to flame radiation as one of the major sources of heat transfer to the wall.

In dry low NOx (DLN) gaseous-fueled combustors, however, there is no soot formation, so the source of radiation to the wall is from the gases. A study by Johnson and Zhao [33] considered the radiative heat transfer to a TBC-coated wall from a methane/air flame operating at 40 atm. The participating species in the radiation calculation were H2O, CO2, and CO; the results showed that a line-by-line calculation of the gas-phase radiation produced more realistic wall heating than using a gray-body assumption for the gases. In particular, the wall heat flux from gas-phase radiation was significant (of the order of 105 W/m2); this paper also discussed the particular impact on TBCs, which will be discussed in more detail later. This work, among others [3436], shows that gas-phase radiation should not be neglected in considering heat transfer in gas turbine hot sections.

We used a spectral line survey to estimate the variation in radiative forcing that could result from changes to the composition of the combustion products. This analysis is a first-order approximation that does not take into account any absorption or scattering of light, although it is expected that scattering will be minimal at these conditions. Further, it does not take into account any radiation from solid surfaces or the shape factor of the combustor. It simply compares the total radiative intensity of one mixture versus another over a given range of infrared wavelengths to understand how radiative heat transfer could potentially change with mixture composition. In the mixtures and temperature under consideration, the three species that contribute to radiative heat transfer are CO2, CO, and H2O, as was done by Johnson and Zhao [33].

The line survey was calculated using the online tool Spectraplot [37] between the wavelengths of 2 and 8 microns. Spectraplot's line survey tool provides information about spectroscopic transitions over a large range of wavelengths; other tools on the website can provide much more detailed absorption and emission profiles over smaller wavelength ranges. However, for the purpose of this exercise, the line survey tool was appropriate for estimating bulk radiative intensity. The wavelength range was chosen based on potential damage mechanisms present for typical thermal barrier coatings. Work by Eldridge et al. [38] showed that the topcoat of yttria-stabilized zirconia TBCs (8YSZ) is up to 50% transmissive between the wavelengths of 2 and 8 microns, which can lead to significant heating of the bond coat and substrate below. While the transmission level varies with composition (13.5YSZ can be up to 90% transmissive), the wavelength range over which IR radiation passes through the topcoat is relatively similar for most TBC compositions. As such, this analysis will apply to a wide range of coating chemistries.

Figure 6 shows the output from Spectraplot of the baseline spectrum for the products of the 100% CH4 flame at 1526.85 °C (2780.33 °F) and 1864.38 kPa (270.4 PSI), with H2O in green and CO2 in yellow; the levels of emission of CO are orders of magnitude smaller and so not plotted here. H2O has wide bands of radiation over most of the wavelength range, whereas CO2 has more concentrated regions of radiation with higher intensity. The strength of each of the bands is dependent on the mole fraction of each species as well as the pressure and temperature. In these calculations, the pressure and temperature are held constant and the mole fraction of the constituent species varies between cases.

Fig. 6
Radiative line survey between 2 and 8 microns for the products of a 100% CH4 flame at 1526.85 °C (2780.33 °F) and 1864.38 kPa (270.4 PSI) using Spectraplot [37]
Fig. 6
Radiative line survey between 2 and 8 microns for the products of a 100% CH4 flame at 1526.85 °C (2780.33 °F) and 1864.38 kPa (270.4 PSI) using Spectraplot [37]
Close modal

For each mixture, the line strength of each of the lines was summed and plotted as a function of hydrogen mole fraction in the reactants, as shown in Fig. 7. The units of the summed line strength are arbitrary. The results show that across the board, flames with CO2 in the products have much higher radiative power than those with just H2O. Although it is not obvious from the spectra in Fig. 6, the density of CO2 lines in the two bands around 2.8 microns and 4 microns is much higher than the density of H2O lines across the entire range of wavelengths. As a result, the radiative power of CO2 is higher than that of H2O, making the gas-phase radiation of any H2/NH3 flame at a given temperature less than that of a flame that has any amount of CH4. Thus, the radiative power of the H2/CH4 flames is largely controlled by CO2, decreasing by a factor of three as the fuel mixture varies from 100% CH4 to 100% H2. The variation in the radiative power for the H2/NH3 flames is much smaller, decreasing slightly as the H2O content in the mixture decreases.

Fig. 7
Summed line strength of IR spectra between 2 and 8 microns as a function of H2 mole fraction
Fig. 7
Summed line strength of IR spectra between 2 and 8 microns as a function of H2 mole fraction
Close modal

The literature on radiative heat transfer in gas turbine hot sections is limited, making the literature regarding fuel composition effects even more limited. A series of studies from Tsinghua University [34,3942] used both experiments and simulation to understand the role of gas composition on heat transfer, with explicit consideration of the role of radiation from CO2 and H2O. In these studies, a plate with a line of film cooling holes was tested and simulated at a range of gas compositions that mimic a variety of low-carbon fuel solutions, including natural gas, syngas, hydrogen, and oxy-fuel combustion of natural gas. The simulations include a discrete transfer model with a weighted sum of gray gases model to capture the spectral behavior of the gases. In all cases, the addition of the radiation to the heat transfer calculations significantly altered the temperature distribution on the wall and matched the experimental results more closely. Further, mixtures with higher levels of CO2 and H2O resulted in higher temperatures on the wall, a result of the increased heat transfer from radiation.

A number of studies in open flames have considered the level of flame radiation with different fuels, including hydrogen and ammonia [4348]. There is an emphasis on the radiation heat transfer in ammonia/air flames because this mode of heat transfer is known to significantly impact the flame speed much more than in methane or hydrogen flames. Simulations by Nakamura and Shindo [46] showed that the flame speed of an ammonia/air flame could reduce by 3% at stoichiometric conditions when radiation is taken into account and the reduction is greater at both higher and lower equivalence ratios. This heat loss suppresses the thermal decomposition of NH3 to H2 in the flame, slowing the flame speed as a result.

Another set of studies consider the impact of wall heat transfer on flame behaviors, particularly for ammonia [45], although few of these consider gas turbine configurations specifically. For example, Cai and Zhao [49] explored the impact of wall thermal conductivity on the performance of a micropower system, which consists of a thin channel for reactions of hydrogen/ammonia and oxygen. Greater heat loss through the wall significantly impacted the system temperature and flame location in the system. In another study, Lam et al. [50] experimentally tested blends of natural gas and hydrogen at a range of fuel-splits in the piloted injector of a Siemens SGT-400 combustion system. The results showed lower injector wall temperature with increasing H2 at a constant flame temperature; the reason given was the reduction in flame radiation with higher H2O contents in the product gases.

Some studies have pointed to the role of radiation without any detailed consideration. For example, a review by Takeishi [51] indicated that the use of hydrogen or ammonia in place of natural gas would increase the amount of H2O in the exhaust stream and hence the radiative heat transfer from H2O, but does not discuss the relative strength of radiation from CO2 versus H2O. The results shown here indicate that the total radiative output from high-H2O gases would be less than that with CO2 at a constant flame temperature.

Heat Transfer Discussion.

Put together, these heat transfer scalings show a mixed result as a function of fuel composition. Many of the effects are driven by the presence of H2O in the combustion products. There are significant differences in the equilibrium mole fraction of H2O for these fuel blends: 6.5% for stoichiometric CH4 flames, 11% for stoichiometric H2 flames, and 13% for stoichiometric NH3 flames. When H2 is blended into CH4, the convective heat transfer has the potential to increase as more H2O drives increases in heat capacity and thermal conductivity, but the radiative heat transfer can decrease significantly. When H2 is blended into ammonia, both the convective and radiative heat transfer decrease with increasing H2 as a result of decreasing H2O in the products, but the effects are less dramatic than in blends of H2 and CH4.

The variation in product composition with fuel blend, even at a constant firing temperature, has interesting implications for operation of gas turbines. Figure 8 shows the power output of a F-class engine with different fuel blends at a constant firing temperature. Burning with H2 has a higher power output because it has a higher energy per unit mass. Burning with ammonia has a higher power output because of the significantly higher fuel mass flow required to match the flame temperatures. As such, some of the heat transfer issues that arise with high-hydrogen fuels could be abated at a constant power output if the firing temperature was decreased. In this case, both convective and radiative heat transfer would decrease while maintaining the same load. For reasons like these, new fuels should be considered on a systematic level.

Fig. 8
Estimated performance from a generic F-class gas turbine based on various low-carbon fuel blends
Fig. 8
Estimated performance from a generic F-class gas turbine based on various low-carbon fuel blends
Close modal

Impact of Fuels on Mechanical Stresses

Because a change in the fuel composition will have an impact on both flame stabilization and heat transfer to the wall, these modifications will alter the temperature distribution in the hot section, resulting in changes to the mechanical stresses. The two main failure mechanisms of concern for these parts are creep and low-cycle fatigue [52]. Creep is a time-dependent plastic deformation process that occurs at elevated stresses and high temperatures, typically >50% of the melting temperature for gas turbine alloys. Low-cycle fatigue is driven by low-frequency load changes that cause oscillations in mechanical load. In gas turbines, the source of low-cycle fatigue is typically the cycling of the engine to vary load or the frequent starts and stops that might be experienced by a peaker machine.

Changes to the flame shape and heat transfer to the combustion liner could impact the mechanical stresses in this component. Several studies have shown the changes in flame shape with variations in fuel composition, which would drive the modification of liner temperature distribution. In swirl-stabilized flames, blends of hydrogen in natural gas show significant shortening of the flame [53]. For example, a study by Strollo et al. [54] blended hydrogen into natural gas at constant thermal power in an industry-relevant swirling fuel injector at atmospheric pressure. At each condition, the flame center of heat release was calculated from CH* chemiluminescence imaging, as shown in Fig. 9 (left). Variations in the H2 content in NG from 10% to 20% by volume could result in an 18% decrease in flame height and a 5% decrease in flame width, concentrating the high-temperature region on the combustion liner.

Fig. 9
Impact of fuel blending on flame shape and temperature with H2 (left, reprinted with permission from Strollo et al. [54]) and NH3 (right, reprinted with permission from Elbaz et al. [55])
Fig. 9
Impact of fuel blending on flame shape and temperature with H2 (left, reprinted with permission from Strollo et al. [54]) and NH3 (right, reprinted with permission from Elbaz et al. [55])
Close modal

Blending of ammonia into natural gas has the opposite effect as hydrogen, leading to slower flame speeds and typically longer flame structures [44]. For example, work by Elbaz et al. [55] in a swirl-stabilized flame with various blends of NH3 and CH4 in air at atmospheric pressure showed changes in the flame structure with fuel composition. Flame temperature distributions were measured for different conditions, including pure NH3 and blends of 50% NH3/50% CH4 at a constant equivalence ratio. Figure 9 (right) shows the change in temperature distribution, where the 100% NH3 flame shows slightly lower temperatures overall (as compared to 50% NH3), including a 50 K decrease in peak temperatures in the central region of the flame.

Changes to the structure and temperature of the flame could also impact the pattern factor, or temperature distribution at the exit of the combustor and entrance of the turbine first vane. Studies by Lefebvre [21] showed this trend for jet fuels, where relative changes in the fuel properties were significantly less severe than they will be for different blends of hydrogen, ammonia, and natural gas. For example, variations between liquid fuels (JP4, JP8, and DF2) could drive a 20% change in the pattern factor, defined as PF=(TmaxT4)/(T4T3), where Tmax is the maximum temperature in the profile, and T3 and T4 are the temperatures at the combustor inlet and exit, respectively.

Several studies have identified changes to pattern factor with variations in fuel composition for high-hydrogen fuels, particularly for microgas turbines [5661]. Most of these studies are computational, as few annular or can-annular burner rigs exist. Many of these studies also report the change in temperature distribution on the combustor liner, which shows similar trends to the experimental studies discussed previously. One of the most comprehensive studies was recently published by Park et al. [62], who considered the impact of hydrogen and ammonia blending in natural gas on the combustor liner temperature distribution, pattern factor, emissions, and combustion oscillations of a flame using a 501 F can combustor operated at atmospheric conditions. Up to 30% H2 and 20% NH3 were blended into NG at a constant thermal load. The injector, liner, and transition piece were instrumented with thermocouples to determine the impact of fuel composition on wall temperature profile. Figure 10 shows the results of the hydrogen blending (top) and ammonia blending (bottom) tests. Significant variations in temperature were measured with both tests, resulting in almost 60 °C local temperature increases with H2 and almost 50 °C local temperature decreases with NH3.

Fig. 10
Combustor liner temperature deviation along the length of the liner (see schematic at bottom of figure) for hydrogen (top) and ammonia (bottom) blending in NG. Reprinted with permission from Park et al. [62].
Fig. 10
Combustor liner temperature deviation along the length of the liner (see schematic at bottom of figure) for hydrogen (top) and ammonia (bottom) blending in NG. Reprinted with permission from Park et al. [62].
Close modal

Similarly, Fig. 11 shows the change in circumferential pattern factor, defined the same as in Lefebvre [21], with variations in fuel composition at 48% load. Companion CFD analysis and flame imaging in an optically-accessible rig showed that the flame shape changes, driven by changes in flame speed, were the root cause of these temperature variations. More compact flames with H2 and more diffuse flames with NH3 drove significant changes to liner temperatures and pattern factor.

Fig. 11
Circumferential pattern factor variations with fuel blending. Reprinted with permission from Park et al. [62].
Fig. 11
Circumferential pattern factor variations with fuel blending. Reprinted with permission from Park et al. [62].
Close modal

Combustor liners are typically made of high-temperature polycrystalline wrought alloys like Hasteloy or Inconel, coated with an yttria-stabilized zirconia thermal barrier coating. Turbine vanes and blades are made from similar materials but are typically cast, often as directionally solidified or single-crystal structures, which are then coated in TBC for early-stage blades and vanes. Changes in temperature distribution of the magnitudes shown in the literature could have significant implications for the stress states on the parts [63]. If the overall part temperature of early-stage blades is increased by ∼50 °C, then the typical creep life of these components would be reduced by at least one order of magnitude. The simplest way to represent this effect is through the Larson-Miller parameter [64], which is a method for estimating the lifetime of a component based on changes in either operational time or temperature. The Larson-Miller parameter is represented by: LMP=T×[log(tR)+C], where T is the temperature (in K), tR is the rupture time (in hours), and C is a constant (typically ∼20 for most alloys).

Further, these changes in temperature distribution could have a significant impact on the combustor liner and turbine vane cooling schemes. In the combustor, a combination of liner cooling strategies, including back-side cooling, impingement cooling, and effusion cooling, are designed to accommodate the high levels of heat transfer to the liner in the region of the flame [65]. In the turbine, both internal cooling passages and external cooling features, including showerhead and film cooling holes, are designed to accommodate a particular range of pattern factors [66]. Given the changes in flame shape and temperature distributions with variations in fuel composition, the efficacy of these cooling strategies, and hence part life, could significantly deteriorate.

Finally, the use of hydrogen-containing fuel also introduces the potential for higher mechanical load on the turbine components. These changes can be realized in two different scenarios. The first scenario involves the variation of fuel composition from 100% NG to 100% H2 with a constant turbine inlet temperature. Even though the inlet temperature is the same, the power generated by a blade row will increase proportionally with the increase in specific heat of the product gases with higher levels of H2O. For example, the cp of a 100% H2 flame products as compared to a 100% NG at Tad = 1800 K, as in Table 1, is 2.6% higher. Using specifications for a typical F-class engine (mass flow of 430 kg/s, combustor inlet temperature of 408 °C), the shift from CH4 to H2 fuel would result in a ∼3.8 MW increase in power for that stage, or a ∼30 kW increase in power per blade (assuming 90 blades). In addition to increased stresses, the actual stress profile and high stress locations may change because of the product composition. This variation could potentially lead to a need for design changes in early-stage blades to better redistribute stress or further considerations of changes in operating loads to account for the increased stress.

In the second scenario, hydrogen is used to increase the firing temperature of the machine, resulting in a turbine inlet temperature that could be as much as 200 °C (392 °F) higher than the baseline methane case. This increase in temperature also increases the specific heat, although the change in cp with temperature is less than the change due to the composition. The increasing power per stage would depend on the number of stages and the distribution of loading between them, making an estimate of the increased stresses difficult. Regardless of the details, the loading would increase, resulting in higher stresses and reduced part lifetime.

Fuels and Material Compatability

Several material issues arise throughout the gas turbine hot section with the use of any fuel, but several known problems could become worse and new ones could arise in the presence of high levels of hydrogen-containing fuels. In this discussion, we focus on two different materials systems: high-temperature metal alloys used in the fuel injector, combustor liner, and turbine, and thermal barrier coatings. The fuel injector is discussed separately from the high-temperature alloys used for the combustor liners and turbines because of the significantly different thermochemical conditions experienced by this component. The material compatibility issues in the fuel delivery systems are significant but not discussed here; however, a recent paper by Pigon et al. [67] reviews these issues.

Fuel Injectors.

Fuel injectors see a variety of gas compositions and temperatures across their different parts. Compressor discharge air is delivered at moderately high temperatures through the main flow passage, whereas high-pressure fuel (sometimes heated) is fed through small inner passages that exit into the main air stream. Finally, high-temperature combustion gases interact with the face of the injector at the flame stabilization point, particularly on features like centerbodies. The particular focus of this section will be the inner fuel passages, as many of the issues experienced by the combustion-facing parts will be covered in the next two sections, depending on whether the injector face is coated or not. Typical fuel system conditions consist of fuel pressures between ∼2070 kPa (300 psi) on older B- and E-class machines to ∼3450 kPa (500 psi) on newer H-class machines, and fuel temperatures that range from ambient to ∼175 °C (350 °F). At these conditions, two main material degradation mechanisms exist for hydrogen-containing fuels: hydrogen embrittlement and nitridation.

Hydrogen embrittlement is a process by which hydrogen diffuses into the alloy and, through several possible mechanisms, makes the material more brittle [68]. It typically occurs in high-pressure, low-temperature conditions, although the temperature ranges for embrittlement of Ni-based superalloys is typically higher than for stainless steels [69]. The high pressures drive hydrogen diffusion into the alloy, allowing for internal embrittlement to occur. Studies have shown that embrittlement occurs in almost all Ni-based superalloys at room temperature and pressures above 34 MPa (4931 psi) [69], although the rates of embrittlement decrease as the pressure decreases due to slower rates of H-atom diffusion.

Figure 12 shows the impact of embrittlement as a function of pressure for six Ni-based superalloys used in aerospace and power-generation applications. Here, the hydrogen environment embrittlement (HEE) index, or the ratio of the notched tensile strength (NTS) of the specimen in hydrogen to the NTS of that same specimen in air or helium, is plotted versus hydrogen pressure at 22 °C. For most alloys, a precipitous drop in HEE occurs as pressure first increases from atmospheric before leveling off at higher pressures. The initial drop is driven by faster diffusion of hydrogen into the alloy, whereas the plateau is a result of hydrogen saturation in the material. The impact of temperature on embrittlement is a strong function of the alloy. Figure 13 shows the impact of temperature on the HEE based on area reduction ratio, which is a measure of tensile ductility for hydrogen versus helium or air environments, for three Ni-based superalloys at 51.7 MPa for Inconel 718 and 45 MPa for the EP alloys. Inconel 718, a common material for fuel injectors, is highly impacted at low temperatures, but not as susceptible to embrittlement as temperature increases. However, the minimum HEE varies with alloy, where the EP 99 and EP 741 are two alloys with relatively high-temperature embrittlement behaviors. However, these alloys are typically not used in fuel injectors.

Fig. 12
Hydrogen environment embrittlement versus pressure in Ni-based superalloys, reprinted with permission from Lee [69]
Fig. 12
Hydrogen environment embrittlement versus pressure in Ni-based superalloys, reprinted with permission from Lee [69]
Close modal
Fig. 13
Hydrogen environment embrittlement versus temperature for Ni-based alloys, reprinted with permission from Lee [69]
Fig. 13
Hydrogen environment embrittlement versus temperature for Ni-based alloys, reprinted with permission from Lee [69]
Close modal

Based on these results and the typical line pressures and temperatures of fuel in gas turbine engines, we would expect some reduction in fracture properties for 100% hydrogen service, depending on material. With fuel pressures of up to 4 MPa at atmospheric temperature, the HEE (Fig. 12) would be near 1 for Astroloy, Hastelloy X, and Haynes 188, 0.8 for Inconel 718, and near 0.5 for Rene 41. The temperature range of the fuel (20–175 °C) is within the range where embrittlement may be relevant, and so these results show that the choice of alloy for the fuel injector is quite important. One way to potentially avoid embrittlement issues in high-hydrogen fuels would be to heat the fuel. As shown in Fig. 13, HEE for Inconel 718 increases with temperature above 100 °C, where even moderate increases in fuel temperature could meaningfully reduce embrittlement. Typically, fuel heating also results in higher cycle efficiencies and so this strategy could have multiple benefits.

Nitridation is a process by which atomic nitrogen dissolves into the alloy and forms nitrides [7072]. Several components of Ni-based superalloys have a propensity to form nitrides, including aluminum, titanium, and tantalum. That said, nickel-based alloys are typically more nitridation resistant than iron-based alloys. The rate at which nitridation occurs is proportional to the concentration of molecular nitrogen. In air atmospheres, molecular nitrogen is present in trace levels except at very high temperatures. In gaseous mixtures with ammonia, however, significantly more atomic nitrogen can be present at a wider range of temperatures, which increases the potential for nitridation. Further, the diffusivity of nitrogen into alloys increases significantly with temperature, which can both increase the rate as well as the penetration depth of nitridation [70]. Despite the moderately high pressures and potentially high levels of ammonia in fuel injectors, there is likely a low rate of nitridation at the temperatures present in the fuel-side surfaces of gas turbine fuel injectors. Most “low-temperature” nitridation studies in superalloys consider temperatures of ∼400 °C (752 °F), which is far above the temperature in even a heated fuel system.

High-Temperature Alloys.

Nickel-base superalloys are commonly used in gas turbine hot sections due to their exceptional high-temperature mechanical properties. Typical alloys include trade names such as Rene, Inconel, Hasteloy, PWA, and CMSX [52]. In the combustor and first turbine stages, these materials are typically coated with thermal barrier coatings, which protects them from direct exposure to the combustion gases. However, further downstream in the turbine, several uncoated stages still interact with combustion products at relatively high temperatures. This discussion focuses on these conditions.

The greatest potential hazard for these components is enhanced oxidation in the presence of higher water concentration in the combustion products of high-hydrogen fuels. Surface oxidation of nickel-base alloys is expected in post-combustion atmospheres in gas turbines. Such alloys, especially those with higher aluminum or chromium content, form a protective oxide scale on the surface of the part that can slow diffusion of gases into the metal and reduce the rate of internal corrosion [73]. The effect of higher H2O concentrations on alloy oxidation behavior is dependent on the oxide chemistry of an alloy [74]. Most high-temperature nickel alloys typically form alumina or chromia scales and data show differing impacts of H2O on oxidation rate, scale thickness, and scale adhesion depending on the oxide chemistry. Several studies have considered the impact of water-containing environments on oxidation of Ni-base superalloys [7584], although the literature is not as extensive or complete as is needed to understand the impact of hydrogen fuels at gas turbine conditions. In particular, no fundamental studies of material oxidation have been identified that were conducted in atmospheres relevant to the products of high-hydrogen combustion; most studies consider different levels of water vapor in air.

A brief review by Pint et al. [85] highlighted several initial studies of a range of Ni-base superalloys and stainless steels in various H2O-containing atmospheres. In general, alumina-forming alloys are less affected by the presence of H2O as compared to chromia-forming alloys. For example, work by Pillai and Pint [86] examined the difference in mass change for alloys 740H, 282, X4, and 1483 during cyclic testing (100 h cycles for 500 h at 800 °C) in atmospheres of air with 10% and 60% H2O. Alloy 282 experienced some increase in mass loss at the higher H2O level, while X4 had a significantly lower mass loss at the higher H2O level. For each material, two specimens were tested and there were significant differences between the behaviors of the two samples over the course of the testing. High-resolution imaging suggested that the structure and chemistry of the scale formation can have a significant impact on the internal oxidation rates.

Further, the impact that trace elements in the alloy have on scale adhesion can change the oxidation behavior as the presence of a stable scale can significantly reduce the incidence of internal oxidation [87]. Oxidation rates can increase with cyclic behaviors, which can also drive the cracking and spallation of the protective scales, leading to part damage. In addition to material loss, the potential for cyclic spallation behavior poses the risk of foreign object damage (FOD) in the hot section as scales are shed during cooldown.

Additionally, there can be a strong coupling between oxidation and the presence of material stresses [88]. Several studies [8993] have shown more rapid oxidation rates in the presence of both tensile and compressive strain in superalloy materials. For example, work by Barnard et al. [89] considered the impact of stress on the isothermal and cyclic oxidation rate of two Haynes superalloys: alloy 75 and alloy 230. Figure 14 shows a summary of the results of the study after 1000 h of testing for both isothermal oxidation (iso) and cyclic (cyc) testing at two temperatures, where the cycles were performed at 100-hour intervals. Overall, the dog-bone samples under stress had a lower mass change than the unloaded samples. However, the rate of oxide layer growth was faster under stress, a result of the faster diffusion of Cr atoms to the surface to form oxide scales.

Fig. 14
Oxidation mass gain for two Haynes alloys with and without applied stresses for both isothermal and cyclic oxidation testing, reprinted with permission from Barnard et al. [89]
Fig. 14
Oxidation mass gain for two Haynes alloys with and without applied stresses for both isothermal and cyclic oxidation testing, reprinted with permission from Barnard et al. [89]
Close modal

Inversely, the presence of oxidation can change the fracture behavior of a superalloy [9497]. For example, fatigue experiments from Bache et al. [94] showed that fatigue failure occurred earlier for samples in air than in vacuum in a sample that was machined from a RR1000 turbine rotor. Inspection of the notched samples indicated that oxide formation in the notches initiated crack growth, resulting in poorer fatigue performance. Given the enhanced oxidation due to higher H2O content as well as potentially higher mechanical stresses due to increased specific heats, changes to the fuel composition could have measurable impacts on both corrosion and mechanical failure modes.

The incidence of nitridation and hydrogen embrittlement in the combustor and turbine will likely not be any different with high-hydrogen fuels than they are for typical fuels. The rates of these mechanisms are all highly dependent on both the partial pressure of the attacking species (atomic nitrogen, atomic hydrogen, etc.) as well as the exposure time of the metal to these species. Assuming reasonable combustion efficiencies (>95%) and rarity of events like blowout where significant amounts of unburned fuel could propagate downstream past any coated parts, then the rate of these damage mechanisms should be negligibly small as compared to oxidation.

Despite the unfavorable conditions for nitridation, there are some initial studies on nitridation in ammonia flame environments. Work by Wang et al. [98] considered the nitriding effects of ammonia flames on iron-based metal walls, but the materials systems are more relevant for industrial systems rather than gas turbines. Recent work by Laws et al. [99] described an effort to design an experimental facility for understanding nitridation processes in combustion-relevant environments with high levels of ammonia.

Thermal Barrier Coatings.

Thermal barrier coatings are ultralow conductivity, multilayered ceramic coatings that reduce the heat transferred from the hot gas to the metallic part. The typical structure of a TBC includes a bond coat, typically made of mullite or a nickel- or cobalt-modified aluminide, that is applied to allow the topcoat, typically made of yittria-stabilized zirconia (YSZ), to better adhere to the substrate. The bond coat also provides corrosion and oxidation protection for the substrate. Between these two layers is a thermally grown oxide (TGO) layer that develops as the coating is subjected to high service temperatures. These coatings can be applied through various methods, including electron beam physical vapor deposition or air plasma spray processes. Regardless of the application method, both the material and structure are designed to minimize the thermal conductivity of the coating. Typically, the structure of the topcoat is optimized for both heat transfer and the difference in thermal expansion coefficient between the substrate and the TBC, allowing it to flex as the parts heat and cool.

The main damage mechanism to TBCs is failure of the thermally grown oxide layer between the topcoat and the bond coat [100]. The growth of this layer is driven by the diffusion of oxygen through the topcoat and the oxidation rate of the aluminum in the bond coat. As the layer forms, significant residual stresses can arise as a result of the difference in the thermal expansion coefficient between the substrate, the bond coat, and the topcoat; these stresses tend to increase with the thickness of the TGO [101]. These stresses can cause a number of potential failure mechanisms, all of which can result in cracking of the TGO and then cracking and spallation of the topcoat [102104]. Stresses are exacerbated in parts with high levels of curvature, like turbine vane leading edges, leading to earlier failure in these critical locations [105]. Once the topcoat is cracked or spalled, significant oxidation of the bond coat and damage to the metal substrate are possible. The addition of Pt-modified coatings has been shown to extend the lifetime of the TBC [103], but are typically only used in aircraft engines, not ground-based turbines, and so are not discussed in detail here.

Thermally grown oxide growth and failure is a known durability issue in hot-section coatings; the coating is typically applied with the knowledge that it will be stripped and reapplied during regular service intervals [106]. However, the use of high-hydrogen fuels can accelerate the TGO growth rate due to the increase in H2O concentration [104,107110]. Fundamental studies of oxide formation rates identified two mechanisms by which H2O accelerates oxide growth and damage. First, dissociation of H2O to form oxygen radicals has a lower energy barrier than dissociation of O2, resulting in more rapid oxidation [74]. Second, the presence of hydrogen and hydroxyl radicals can accelerate the oxidation kinetics through several pathways [74,111,112], particularly for aluminum oxides. Further, oxides formed in the presence of water have a higher volatility than those formed in oxygen atmospheres [74], resulting in lower damage thresholds in harsh hot section environments.

Figure 15 shows work by Zhou et al. [109], which plots weight gain as a function of time for a plasma-sprayed TBC with a NiCrAlY bond coat and a 7.5YSZ topcoat. The samples were exposed to two different atmospheres—O2 and O2 with 5% H2O—at isothermal conditions (1050 °C) for up to 400 h. While the initial TGO growth rate is similar for the two atmospheres, the oxidation continues in the presence of H2O, whereas it levels off after 100 h in the O2 atmosphere.

Fig. 15
Thermally grown oxide (TGO) weight gain of a NiCrAlY bond coat at 1050 °C with two different atmospheres, reprinted with permission from Maris-Sida et al. [109]
Fig. 15
Thermally grown oxide (TGO) weight gain of a NiCrAlY bond coat at 1050 °C with two different atmospheres, reprinted with permission from Maris-Sida et al. [109]
Close modal

Further, the presence of water can result in non-ideal TGO structure and the formation of spinels, which are mechanically weak layers of Cr, Co, and Ni oxides that grow between the topcoat and the alumina TGO [113]. Spinels typically form during the initial transient TGO growth [114,115] and can continue to grow at longer times [116118]. Sullivan and Mumm [107,108] used isothermal oxidation experiments on several bond coats without their topcoats to better understand spinel growth. Figure 16 shows cross-sectional micrographs of the bond coat, alumina TGO, and spinel formation at four different levels of H2O in the ambient after 300 h of exposure at 1125 °C. The thickness of the spinel layer increases significantly with H2O concentration in the air. Further, it is evident that the spinel layer is much less uniform than the TGO layer, which can promote the wrinkled-interface failure mechanism of TBCs [100].

Fig. 16
Cross-sectional micrograph of a NiCoCrAlRe bond coat in three different atmospheres to identify spinel growth. Reprinted with permission from Sullivan and Mumm [108].
Fig. 16
Cross-sectional micrograph of a NiCoCrAlRe bond coat in three different atmospheres to identify spinel growth. Reprinted with permission from Sullivan and Mumm [108].
Close modal

The composition of the substrate alloy may also play a role in the impact of H2O on TGO growth rate. A study by Haynes et al. [119] compared the failure of a high-velocity oxy-fuel (HVOF) sprayed NiCoCrAlYHfSi bond coat and an air plasma sprayed 8YSZ topcoat, typical of power-generation gas turbine coatings, with two nickel superalloy substrates: superalloy 1483 and X4. The 1483 had a significantly shorter lifetime than the X4 in an environment of 10% H2O in O2, failing after 267 one-hour cycles at 1100 °C as compared to 380 cycles for the X4. Scanning electron microscopy (SEM) showed no significant difference in the thickness or structure of the TGO in both cases but X-ray elemental mapping using energy dispersive spectroscopy (EDS) identified significant differences in the composition of the oxide layers. The TGO on alloy 1483 showed lower levels of Al and higher levels of Ti and Cr, both of which have been shown to reduce the interfacial strength of the TGO [107,111,118]. These elemental differences could be connected back to the alloy itself, such that X4 is comprised of 7.5% Cr, 13.3% Al, and 1.2% Ti, and 1483 is comprised of 13.6% Cr, 7.3% Al, and 4.9% Ti. The lack of Al, which forms stronger TGOs, and abundance of metals that form weaker TGOs led to earlier TBC failure.

Finally, changes to the radiative emission from the gases could change the damage rates in TBCs. Our previous analysis showed that reduction of CO2 levels in the product gases could reduce the radiative intensity of the gases, potentially resulting in less radiative heat transfer to the wall. Given the transmissivity of TBC topcoats in the midwave infrared spectrum [38] and the reduction in intensity in these same ranges, the radiative heat transfer to the TBC bond coats would likely decrease with the use of high-hydrogen fuels as compared to natural gas. The combination of these effects—increased oxidation damage and reduced radiative heat transfer—due to the presence of H2O need to be considered together. The relative contribution of these effects is an important topic of future research.

However, one drawback of many of previous studies is that the effect of water addition is tested in air or oxygen, rather than in realistic combustion product compositions where the level of O2 in the atmosphere would be substantially less. Lance et al. [120] tested the impact of replacing air with CO2 on TBC lifetimes in mixtures with 10% H2O and showed that CO2 has little influence on the TBC lifetime with 100-hour cycles. However, further testing in realistic environments is necessary to better identify the effects of H2O concentration in product gases on TBC behaviors and lifetimes.

Future Research Directions

This initial analysis shows that there could be significant implications for the high-temperature material systems found in gas turbine hot sections with the introduction of hydrogen and ammonia fuels. The impact of these fuels on existing hardware must be inter-rogated by each manufacturer given the unique operating and lifing margin that is inherent to each design. Incorporating these lessons learned and future research in this area into industrial design and analysis tools will be an important step toward integration of these new fuel blends into operation in existing machines. Further, use of these fuels will have implications for other parts of the power-generation system as well, including fuel delivery hardware and the heat recovery steam generators downstream of the engine [121].

Several lessons learned should be considered going forward. First, heat transfer from the gases to the walls can change significantly because of the change in product composition. The higher proportion of H2O in the products of both ammonia and hydrogen combustion as compared to natural gas has two opposing effects. Convective heat transfer can be augmented as a result of the increase in heat capacity and thermal conductivity of the product gases. However, reduction of CO2 in the product gases reduces the radiative emission at a given temperature. As such, more detailed research on the relative impact of these two modes of heat transfer on combustor liner temperatures, durability, and cooling strategies in realistic, reacting environments is critical to supporting the transition to high-hydrogen fuels.

These same changes to the product gas composition and temperature distributions will have significant implications for the mechanical stresses on hot-section components, particularly rotating components. Further research is necessary to understand the impact on part life and performance. For example, cycle-level analysis can help identify the ways in which these gas property changes can be leveraged to improve efficiency and power output. The results of these analyses can feed into better modeling of mechanical stresses in the rotating components. Further, better understanding of how fuel composition will alter gas temperatures and temperature distributions will provide input to research on improved cooling schemes and creep life prediction. Better component- and system-level research and integration is necessary to take full advantage of the benefits that can arise from the implementation of new fuels.

Material compatibility issues may be significant and a better understanding of material behaviors and processes for choosing appropriate materials for these new fuels and operating conditions will be critical. The initial analyses in this paper suggest that issues related to hydrogen embrittlement on the fuel-side of fuel injectors and enhanced oxidation of gas-side components could significantly reduce part life. However, many of the experiments done to understand corrosion and degradation in both metallic and coatings systems are not done at appropriate operating conditions or with relevant mixtures. Better experiments to understand the long-duration effects of high-hydrogen fuel combustion on these materials are required, and companion theory and simulation will be critical to understanding the engineering solutions necessary to enhance part life with these new fuels.

Finally, new advances in materials and manufacturing will support the transition to high-hydrogen fuels. Additive manufacturing (AM) has been a key enabling technology in development of novel combustion hardware (fuel mixers, burners, etc.) allowing for design freedom combined with simplified manufacturing of complex components. Manufacturers of high-hydrogen and fuel-flexible combustors are already taking advantage of technologies such as powder bed fusion AM. However, turbine components, traditionally made with superalloys, have seen less application of AM. The optimization of structural performance using directionally solidified (DS) and single crystal (SX) technology coupled with advanced TBCs and cooling is now reaching limits for the most advanced machines [122]. Only recently have industrial gas turbines seen the benefit of AM in the hot section with static guide vanes resulting in improvement in performance [123,124].

As discussed previously, the combustion of hydrogen-containing fuels will either increase temperatures or stresses in the hot-section structural components. To maintain current service intervals or improve durability under this increased demand for the structural vanes and blades, existing polycrystalline and directionally solidified superalloys will likely need to be upgraded to higher performing superalloys or alternatively take advantage of novel cooling and design features in AM coupled with new AM-specific superalloys that have the requisite high-temperature performance [125]. However, as demands for even higher efficiency are placed on new components where single-crystal superalloys may not provide adequate life, alternative material systems such as oxide dispersion strengthened materials or refractory alloys may provide new avenues if the challenges with processing can be overcome. Programs such as arpa-e'sultimate program are studying such approaches that address the significant processing challenges and new coatings that may be needed for these radically different material classes [126]. It is important to note that all these materials will still require improved coating systems for both thermal and environmental protection. As highlighted in this review, new coating strategies to mitigate spinel formation and delay TGO formation could provide improved durability with these high-hydrogen fuels.

Finally, ceramic matrix composites (CMCs) have already been demonstrated for robust operation in aerospace turbines and ultrahigh temperature CMCs could provide another alternative solution, but research is needed to further drive costs down [127]. Finally, it should be noted that not only the hottest section components will be impacted by increased temperature and/or loading; later stages of the turbine made today with more traditional cast superalloys will also need to be upgraded offering a plethora of opportunities for materials and manufacturing innovation.

Acknowledgment

The authors would like to thank Jim Harper, Tom Sambor, and Michael Caravaggio at EPRI, Chris Perullo at Turbine Logic, and Wenting Sun and David Wu at Georgia Tech for their discussions about the paper.

Data Availability Statement

The authors attest that all data for this study are included in the paper.

Nomenclature

C =

constant in the Larson Miller equation

LMP =

Larson Miller parameter

PF =

pattern factor

T =

temperature

cp =

heat capacity

h =

heat transfer coefficient

k =

thermal conductivity

tR =

rupture time

References

1.
EPRI,
2023
,
Leading Economy-Wide Carbon Reduction: The Practical Potential of Energy Supply Resources
,
EPRI
,
Palo Alto, CA
.
2.
EIA,
2024
, “
Preliminary Monthly Electric Generator Inventory
,” Energy Information Administration, Washington, DC, accessed Dec. 1, 2023, https://www.eia.gov/electricity/data/eia860m/
3.
EPRI,
2022
,
LCRI Net-Zero 2050: U.S. Economy-Wide Deep Decarbonization Scenario Analysis
,
Palo Alto, CA
.
4.
EIA,
2024
, “
Electricity Generation, Capacity, and Sales in the United States
,” Energy Information Administration, Washington, DC, accessed Dec. 1, 2023, https://www.eia.gov/energyexplained/electricity/electricity-in-the-us-generation-capacity-and-sales.php
5.
EIA,
2024
, “
Hourly Electric Grid Monitor
,” Energy Information Administration, Washington, DC, accessed Dec. 1, 2023, https://www.eia.gov/electricity/gridmonitor/dashboard/electric_overview/US48/US48
6.
Wang
,
M.
,
Elgowainy
,
A.
,
Lee
,
U.
,
Bafana
,
A.
,
Banerjee
,
S.
,
Benavides
,
P. T.
,
Bobba
,
P.
,
Burnham
,
A.
,
Cai
,
H.
, and
Gracida-Alvarez
,
U. R.
,
2021
,
Summary of Expansions and Updates in GREET® 2021
,
Argonne National Lab. (ANL)
,
Argonne, IL
.
7.
Noble
,
D.
,
Wu
,
D.
,
Emerson
,
B.
,
Sheppard
,
S.
,
Lieuwen
,
T.
, and
Angello
,
L.
,
2021
, “
Assessment of Current Capabilities and Near-Term Availability of Hydrogen-Fired Gas Turbines Considering a Low-Carbon Future
,”
ASME J. Eng. Gas Turbine Power
,
143
(
4
), p.
041002
.10.1115/1.4049346
8.
Valera-Medina
,
A.
,
Xiao
,
H.
,
Owen-Jones
,
M.
,
David
,
W. I. F.
, and
Bowen
,
P. J.
,
2018
, “
Ammonia for Power
,”
Prog. Energy Combust. Sci.
,
69
, pp.
63
102
.10.1016/j.pecs.2018.07.001
9.
Harper
,
J.
,
Cloyd
,
S.
,
Pigon
,
T.
,
Thomas
,
B.
,
Wilson
,
J.
,
Johnson
,
E.
, and
Noble
,
D. R.
,
2023
, “
Hydrogen Co-Firing Demonstration at Georgia Power's Plant McDonough: M501G Gas Turbine
,”
ASME
Paper No. GT2023-102660.10.1115/GT2023-102660
10.
Valera-Medina
,
A.
,
Morris
,
S.
,
Runyon
,
J.
,
Pugh
,
D. G.
,
Marsh
,
R.
,
Beasley
,
P.
, and
Hughes
,
T.
,
2015
, “
Ammonia, Methane and Hydrogen for Gas Turbines
,”
Energy Procedia
,
75
, pp.
118
123
.10.1016/j.egypro.2015.07.205
11.
Stefan
,
E.
,
Talic
,
B.
,
Larring
,
Y.
,
Gruber
,
A.
, and
Peters
,
T. A.
,
2022
, “
Materials Challenges in Hydrogen-Fuelled Gas Turbines
,”
Int. Mater. Rev.
,
67
(
5
), pp.
461
486
.10.1080/09506608.2021.1981706
12.
Alnaeli
,
M.
,
Alnajideen
,
M.
,
Navaratne
,
R.
,
Shi
,
H.
,
Czyzewski
,
P.
,
Wang
,
P.
,
Eckart
,
S.
,
Alsaegh
,
A.
,
Alnasif
,
A.
, et al.,
2023
, “
High-Temperature Materials for Complex Components in Ammonia/Hydrogen Gas Turbines: A Critical Review
,”
Energies (Basel)
,
16
(
19
), p.
6973
.10.3390/en16196973
13.
Hetherington
,
D.
,
2022
, “
McCoy Power Reports
,” accessed Dec. 1, 2023, https://www.mccoypower.net/
14.
Breer
,
B.
,
Rajagopalan
,
H.
,
Godbold
,
C.
,
Johnson
,
H.
, II
,
Emerson
,
B.
,
Acharya
,
V.
,
Sun
,
W.
,
Noble
,
D.
, and
Lieuwen
,
T.
,
2023
, “
Numerical Investigation of NOx Production From Premixed Hydrogen/Methane Fuel Blends
,”
Combust. Flame
,
255
, p.
112920
.10.1016/j.combustflame.2023.112920
15.
Angello
,
L.
,
Noble
,
B.
,
Steele
,
R.
,
Cohen
,
M.
, and
Emerson
,
B.
,
2022
, “
Dry Low NOx (DLN) Combustion System Operability Considerations
,”
ASME
Paper No. GT2022-82667.10.1115/GT2022-82667
16.
Kee
,
R. J.
,
Rupley
,
F. M.
,
Miller
,
J. A.
,
Coltrin
,
M. E.
,
Grcar
,
J. F.
,
Meeks
,
E.
,
Moffat
,
H. K.
,
Lutz
, A. E.
,
DixonLewis
, G.
, et al.,
2022
, “
Ansys Chemkin-Pro 2022 R1
,” Ansys, Canonsburg, PA.
17.
Smith
,
G. P.
,
Golden
,
D. M.
,
Frenklach
,
M.
,
Moriarty
,
N. W.
,
Eiteneer
,
B.
,
Goldenberg
,
M.
,
Bowman
,
C. T.
, et al., “
GRIMech 3.0
,” accessed Dec. 1, 2023, http://combustion.berkeley.edu/gri-mech/
18.
Mei
,
B.
,
Zhang
,
X.
,
Ma
,
S.
,
Cui
,
M.
,
Guo
,
H.
,
Cao
,
Z.
, and
Li
,
Y.
,
2019
, “
Experimental and Kinetic Modeling Investigation on the Laminar Flame Propagation of Ammonia Under Oxygen Enrichment and Elevated Pressure Conditions
,”
Combust Flame
,
210
, pp.
236
246
.10.1016/j.combustflame.2019.08.033
19.
Srinivasan
,
S.
,
Tannehill
,
J. C.
, and
Weilmuenster
,
K. J.
,
1987
, “
Simplified Curve Fits for the Thermodynamic Properties of Equilibrium Air
,” NASA, Washington, DC, Report No.
CFD-15
.https://ntrs.nasa.gov/api/citations/19870016876/downloads/19870016876.pdf
20.
Linstrom
,
P. J.
, and
Mallard
,
W. G.
,
2001
, “
The NIST Chemistry WebBook: A Chemical Data Resource on the Internet
,”
J. Chem. Eng. Data
,
46
(
5
), pp.
1059
1063
.10.1021/je000236i
21.
Lefebvre
,
A. H.
,
1984
, “
Fuel Effects on Gas Turbine Combustion-Liner Temperature, Pattern Factor, and Pollutant Emissions
,”
J. Aircr.
,
21
(
11
), pp.
887
898
.10.2514/3.45059
22.
Alvin
,
M. A.
,
2010
, “
Materials and Component Development for Advanced Turbine Systems: Project Summary
,”
ASME
Paper No. GT2010-23173.10.1115/GT2010-23173
23.
Chyu
,
M. K.
,
Mazzotta
,
D. W.
,
Siw
,
S. C.
,
Karaivanov
,
V. G.
,
Slaughter
,
W. S.
, and
Alvin
,
M. A.
,
2009
, “
Aerothermal Challenges in Syngas, Hydrogen-Fired, and Oxyfuel Turbines—Part I: Gas-Side Heat Transfer
,”
ASME J. Therm. Sci. Eng. Appl.
,
1
(
1
), p.
011002
.10.1115/1.3159479
24.
Mazzotta
,
D. W.
,
Chyu
,
M. K.
, and
Alvin
,
M. A.
,
2007
, “
Airfoil Heat Transfer Characteristics in Syngas and Hydrogen Turbines
,”
ASME
Paper No. GT2007-28296.10.1115/GT2007-28296
25.
Mazzotta
,
D. W.
,
Karaivanov
,
V. G.
,
Chyu
,
M. K.
,
Slaughter
,
W. S.
, and
Alvin
,
M. A.
,
2008
, “
Gas-Side Heat Transfer in Syngas, Hydrogen-Fired and Oxy-Fuel Turbines
,”
ASME
Paper No. GT2008-51474.10.1115/GT2008-51474
26.
Chyu
,
M. K.
,
Siw
,
S. C.
,
Karaivanov
,
V. G.
,
Slaughter
,
W. S.
, and
Alvin
,
M. A.
,
2009
, “
Aerothermal Challenges in Syngas, Hydrogen-Fired, and Oxyfuel Turbines—Part II: Effects of Internal Heat Transfer
,”
ASME J. Therm. Sci. Eng. Appl.
,
1
(
1
), p.
011003
.10.1115/1.3159480
27.
Moliere
,
M.
,
2002
, “
Benefiting From the Wide Fuel Capability of Gas Turbines: A Review of Application Opportunities
,”
ASME
Paper No. GT2002-30017.10.1115/GT2002-30017
28.
Chiesa
,
P.
,
Lozza
,
G.
, and
Mazzocchi
,
L.
,
2005
, “
Using Hydrogen as Gas Turbine Fuel
,”
ASME J. Eng. Gas Turbines Power
,
127
(
1
), pp.
73
80
.10.1115/1.1787513
29.
Louis
,
J. F.
,
1978
, “
Systematic Studies of Heat Transfer and Film Cooling Effectiveness
,”
AGARD High Temperature Problems in Gas Turbine Engines
, SEE N 78-21118 12-07.
30.
Zou
,
X.
, and
Yuan
,
X.
,
2011
, “
Investigation of Gas Composition Effects on the Aerothermal Characteristics of a Film-Cooled Flat Plate at High-Temperature Conditions
,”
ASME
Paper No. GT2011-45429.10.1115/GT2011-45429
31.
Gazzani
,
M.
,
Chiesa
,
P.
,
Martelli
,
E.
,
Sigali
,
S.
, and
Brunetti
,
I.
,
2014
, “
Using Hydrogen as Gas Turbine Fuel: Premixed Versus Diffusive Flame Combustors
,”
ASME J. Eng. Gas Turbine Power
,
136
(
5
), p.
051504
.10.1115/1.4026085
32.
Gö Ke
,
S.
,
Terhaar
,
S.
,
Schimek
,
S.
,
Gö Ckeler
,
K.
, and
Paschereit
,
C. O.
,
2011
, “
Combustion of Natural Gas, Hydrogen and Bio-Fuels at Ultra-Wet Conditions
,”
ASME
Paper No. GT2011-45696.10.1115/GT2011-45696
33.
Johnson
,
A. L.
, and
Zhao
,
X.
,
2021
, “
Analysis of the Heat Transfer Within Combustor Liners Using a Combined Monte Carlo and Two-Flux Method
,”
ASME J. Turbomach.
,
143
(
3
), p.
031006
.10.1115/1.4049913
34.
Yin
,
H.
,
Li
,
M.
,
Chi
,
Z.
,
Ren
,
J.
, and
Jiang
,
H.
,
2014
, “
Effect of Radiative Heat Transfer Factor on the Temperature Distribution of a First Stage Vane
,”
ASME
Paper No. GT2014-25711.10.1115/GT2014-25711
35.
Sun
,
P.
,
Li
,
X.
,
Ren
,
J.
, and
Jiang
,
H.
,
2020
, “
Analysis of Radiation-Convection Coupled Effects on the Turbine Vane With Different Gas Compositions
,”
ASME
Paper No. GT2020-14376.10.1115/GT2020-14376
36.
Wang
,
M.
,
You
,
R.
,
Li
,
H.
, and
Tao
,
Z.
,
2023
, “
Correction Formula for Film-Cooling Effectiveness Considering the Influence of Thermal Radiation in High-Temperature Environments
,”
Appl. Therm. Eng.
,
222
, p.
119903
.10.1016/j.applthermaleng.2022.119903
37.
Goldenstein
,
C. S.
,
Miller
,
V. A.
,
Spearrin
,
R. M.
, and
Strand
,
C. L.
,
2017
, “
SpectraPlot. Com: Integrated Spectroscopic Modeling of Atomic and Molecular Gases
,”
J. Quant. Spectrosc. Radiat. Transf.
,
200
, pp.
249
257
.10.1016/j.jqsrt.2017.06.007
38.
Eldridge
,
J. I.
,
Spuckler
,
C. M.
,
Street
,
K. W.
,
Markham
,
J. R.
, and
Hartford
,
E.
,
2008
, “
Infrared Radiative Properties of Yttria–Stabilized Zirconia Thermal Barrier Coatings
,”
26th Annual Conference on Composites, Advanced Ceramics, Materials, and Structures: B: Ceramic Engineering and Science Proceedings
, John Wiley & Sons, Inc., Hoboken, NJ, pp.
417
430.
10.1002/9780470294758.ch47
39.
Li
,
M.
,
Ren
,
J.
, and
Jiang
,
H.
,
2016
, “
Conjugate Heat Transfer Characteristics of a Highly Thermal-Loaded Film Cooling System in Hot and Multi-Composition Gas Condition
,”
ASME J. Heat Transfer
, 138(2), p.
024503
.10.1115/1.4031172
40.
Wang
,
W.
,
Sun
,
P.
,
Ren
,
J.
, and
Jiang
,
H.
,
2011
, “
Radiative Effectiveness on the Aero-and Thermodynamics in a Highly Thermally Loaded Film Cooling System
,”
ASME
Paper No. GT2011-45592.10.1115/GT2011-45592
41.
Sun
,
P.
,
Wang
,
W.
,
Ren
,
J.
, and
Jiang
,
H.
,
2011
, “
Effect of the Hot Gas Composition on a Highly Thermally Loaded Film Cooling System
,”
ASME
Paper No. GT2011-45593.10.1115/GT2011-45593
42.
Ren
,
J.
,
Li
,
X.
, and
Jiang
,
H.
,
2019
, “
Conjugate Heat Transfer Characteristics in a Highly Thermally Loaded Film Cooling Configuration With TBC in Syngas
,”
Aerospace
,
6
(
2
), p.
16
.10.3390/aerospace6020016
43.
Murai
,
R.
,
Omori
,
R.
,
Kano
,
R.
,
Tada
,
Y.
,
Higashino
,
H.
,
Nakatsuka
,
N.
,
Hayashi
,
J.
,
Akamatsu
,
F.
,
Iino
,
K.
, and
Yamamoto
,
Y.
,
2017
, “
The Radiative Characteristics of NH3/N2/O2 Non-Premixed Flame on a 10 KW Test Furnace
,”
Energy Procedia
,
120
, pp.
325
332
.10.1016/j.egypro.2017.07.232
44.
Kobayashi
,
H.
,
Hayakawa
,
A.
,
Somarathne
,
K. D. K. A.
, and
Okafor
,
E. C.
,
2019
, “
Science and Technology of Ammonia Combustion
,”
Proc. Combust. Inst.
,
37
(
1
), pp.
109
133
.10.1016/j.proci.2018.09.029
45.
Zhang
,
M.
,
Xu
,
W.
,
Wang
,
R.
,
Wei
,
X.
,
Wang
,
J.
,
Huang
,
Z.
, and
Tan
,
H.
,
2023
, “
Wall Heat Loss Effect on the Emission Characteristics of Ammonia Swirling Flames in a Model Gas Turbine Combustor
,”
Combust. Flame
,
256
, p.
112955
.10.1016/j.combustflame.2023.112955
46.
Nakamura
,
H.
, and
Shindo
,
M.
,
2019
, “
Effects of Radiation Heat Loss on Laminar Premixed Ammonia/Air Flames
,”
Proc. Combust. Inst.
,
37
(
2
), pp.
1741
1748
.10.1016/j.proci.2018.06.138
47.
Okafor
,
E. C.
,
Tsukamoto
,
M.
,
Hayakawa
,
A.
,
Somarathne
,
K. D. K. A.
,
Kudo
,
T.
,
Tsujimura
,
T.
, and
Kobayashi
,
H.
,
2021
, “
Influence of Wall Heat Loss on the Emission Characteristics of Premixed Ammonia-Air Swirling Flames Interacting With the Combustor Wall
,”
Proc. Combust. Inst.
,
38
(
4
), pp.
5139
5146
.10.1016/j.proci.2020.06.142
48.
Xiao
,
S.
,
Zhang
,
Y.
,
Lu
,
Q.
,
Wang
,
Y.
,
Quaye
,
E. K.
, and
Pan
,
J.
,
2022
, “
Numerical Simulation and Emission Analysis of Ammonia/Oxygen Premixed Combustion Process in Micro-Combustor With Baffle
,”
Chem. Eng. Process.-Process Intensif.
,
174
, p.
108871
.10.1016/j.cep.2022.108871
49.
Cai
,
T.
, and
Zhao
,
D.
,
2020
, “
Effects of Fuel Composition and Wall Thermal Conductivity on Thermal and NOx Emission Performances of an Ammonia/Hydrogen-Oxygen Micro-Power System
,”
Fuel Process. Technol.
,
209
, p.
106527
.10.1016/j.fuproc.2020.106527
50.
Lam
,
K.-K.
,
Geipel
,
P.
, and
Larfeldt
,
J.
,
2015
, “
Hydrogen Enriched Combustion Testing of Siemens Industrial SGT-400 at Atmospheric Conditions
,”
ASME J. Eng. Gas Turbine Power
,
137
(
2
), p.
021502
.10.1115/1.4028209
51.
Takeishi
,
K.
,
2022
, “
Evolution of Turbine Cooled Vanes and Blades Applied for Large Industrial Gas Turbines and Its Trend Toward Carbon Neutrality
,”
Energies (Basel
),
15
(
23
), p.
8935
.10.3390/en15238935
52.
Pollock
,
T. M.
, and
Tin
,
S.
,
2006
, “
Nickel-Based Superalloys for Advanced Turbine Engines: Chemistry, Microstructure and Properties
,”
J. Propuls. Power
,
22
(
2
), pp.
361
374
.10.2514/1.18239
53.
Taamallah
,
S.
,
Vogiatzaki
,
K.
,
Alzahrani
,
F. M.
,
Mokheimer
,
E. M. A.
,
Habib
,
M. A.
, and
Ghoniem
,
A. F.
,
2015
, “
Fuel Flexibility, Stability and Emissions in Premixed Hydrogen-Rich Gas Turbine Combustion: Technology, Fundamentals, and Numerical Simulations
,”
Appl. Energy
,
154
, pp.
1020
1047
.10.1016/j.apenergy.2015.04.044
54.
Strollo
,
J.
,
Peluso
,
S.
, and
O'Connor
,
J.
,
2021
, “
Effect of Hydrogen on Steady-State and Transient Combustion Instability Characteristics
,”
ASME J. Eng. Gas Turbine Power
,
143
(
7
), p.
071023
.10.1115/1.4049481
55.
Elbaz
,
A. M.
,
Albalawi
,
A. M.
,
Wang
,
S.
, and
Roberts
,
W. L.
,
2023
, “
Stability and Characteristics of NH3/CH4/Air Flames in a Combustor Fired by a Double Swirl Stabilized Burner
,”
Proc. Combust. Inst.
,
39
(
4
), pp.
4205
4213
.10.1016/j.proci.2022.06.004
56.
Neilson
,
C. E.
,
Shafer
,
D. G.
, and
Carpentieri
,
E.
,
1999
, “
LM2500 Gas Turbine Fuel Nozzle Design and Combustion Test Evaluation and Emission Results With Simulated Gasified Wood Product Fuels
,”
ASME J. Eng. Gas Turbine Power
,
121
(
4
), pp.
600
606
.10.1115/1.2818514
57.
Shih
,
H.-Y.
, and
Liu
,
C.-R.
,
2014
, “
A Computational Study on the Combustion of Hydrogen/Methane Blended Fuels for a Micro Gas Turbines
,”
Int. J. Hydrogen Energy
,
39
(
27
), pp.
15103
15115
.10.1016/j.ijhydene.2014.07.046
58.
Masi
,
M.
,
Gobbato
,
P.
,
Toffolo
,
A.
,
Lazzaretto
,
A.
, and
Cocchi
,
S.
,
2011
, “
Numerical and Experimental Analysis of the Temperature Distribution in a Hydrogen Fuelled Combustor for a 10 MW Gas Turbine
,”
ASME J. Eng. Gas Turbines Power
, 133(2), p.
021506
.10.1115/1.4002017
59.
Bastani
,
M.
,
Tabejamaat
,
S.
, and
Ashini
,
H.
,
2024
, “
Numerical and Experimental Study of Combustion and Emission Characteristics of Ammonia/Methane Fuel Mixture in Micro Gas Turbine Combustor
,”
Int J Hydrogen Energy
,
49
, pp.
1399
1415
.10.1016/j.ijhydene.2023.09.319
60.
Tamang
,
S.
, and
Park
,
H.
,
2023
, “
Numerical Investigation of Combustion Characteristics for Hydrogen Mixed Fuel in a Can-Type Model of the Gas Turbine Combustor
,”
Int. J. Hydrogen Energy
,
48
(
30
), pp.
11493
11512
.10.1016/j.ijhydene.2022.05.273
61.
Shih
,
H.-Y.
,
Ho
,
S.-P.
, and
Cheng
,
C.
,
2023
, “
Combustion Analysis of a Can Combustor With CH4/NH3 Blended Fuels for a Micro Gas Turbine
,”
ASME
Paper No. GT2023-102359.10.1115/GT2023-102359
62.
Park
,
S.
,
Shin
,
J.
,
Park
,
J.
,
Lee
,
S.
, and
Choi
,
N.
,
2023
, “
Comparison of Hydrogen and Ammonia Co-Firing With Natural Gas Using a Practical Gas Turbine Combustor (501F) Under Atmospheric Conditions: Changes in Metal Temperature, Pattern Factor, and NOx Emission
,”
ASME
Paper No. GT2023-103364.10.1115/GT2023-103364
63.
Karaivanov
,
V. G.
,
Mazzotta
,
D. W.
,
Chyu
,
M. K.
,
Slaughter
,
W. S.
, and
Alvin
,
M. A.
,
2008
, “
Three-Dimensional Modeling of Creep Damage in Airfoils for Advanced Turbine Systems
,”
ASME
Paper No. GT2008-51278.10.1115/GT2008-51278
64.
Larson
,
F. R.
, and
Miller
,
J.
,
1952
, “
A Time-Temperature Relationship for Rupture and Creep Stresses
,”
ASME Trans. Am. Soc. Mech. Eng.
,
74
(
5
), pp.
765
771
.10.1115/1.4015909
65.
Han
,
J. C.
,
Dutta
,
S.
, and
Ekkad
,
S.
,
2013
,
Gas Turbine Heat Transfer and Cooling Technology
,
CRC Press
,
Boca Raton, FL
.
66.
Shih
,
T. I. P.
, and
Yang
,
V.
,
2014
,
Turbine Aerodynamics, Heat Transfer, Materials, and Mechanics
,
American Institute of Aeronautics and Astronautics, Inc
., Reston, VA.
67.
Pigon
,
T.
,
Cloyd
,
S.
,
Springer
,
C.
,
Boggs
,
J.
,
Shiraiwa
,
T.
, and
Yamazaki
,
S.
,
2023
, “
Best Practices From Hydrogen Fuel System Retrofit
,”
ASME
Paper No. GT2023-101368.10.1115/GT2023-101368
68.
Popov
,
B. N.
,
Lee
,
J.-W.
, and
Djukic
,
M. B.
,
2018
, “
Hydrogen Permeation and Hydrogen-Induced Cracking
,”
Handbook of Environmental Degradation of Materials
,
William Andrew Applied Science Publishers, Oxford, UK
, pp.
133
162
.
69.
Lee
,
J. A.
,
2012
, “
Hydrogen Embrittlement of Nickel, Cobalt and Iron-Based Superalloys
,”
Gaseous Hydrogen Embrittlement Mater. Energy Technol.
,
2
, pp.
624
667
.10.1533/9780857093899.3.624
70.
Lai
,
G.
,
2007
, “
Chapter 4: Nitridation
,”
High-Temperature Corrosion and Materials Applications
,
ASM International
, Materials Park, OH.
71.
Krupp
,
U.
, and
Christ
,
H.
,
1999
, “
Internal Nitridation of Ni‐Based Alloys During High‐Temperature Corrosion by Nitrogen
,”
Adv. Eng. Mater.
,
1
(
3–4
), pp.
194
198
.10.1002/(SICI)1527-2648(199912)1:3/4%3C194::AID-ADEM194%3E3.0.CO;2-N
72.
Hao
,
L.-Y.
,
Wen
,
X.-Z.
,
Lei
,
X.-W.
,
Yao
,
W.-J.
, and
Wang
,
N.
,
2022
, “
Distinct Nitriding Behaviors of γ and Γ′ Phases in Ni-Based Superalloys: A First-Principles Study
,”
J. Alloys Compd.
,
920
, p.
165996
.10.1016/j.jallcom.2022.165996
73.
Lai
,
G. Y.
,
2007
, “
Chapter 3: Oxidation
,”
High-Temperature Corrosion and Materials Applications
,
ASM International, Materials Park, OH
.
74.
Saunders
,
S. R. J.
,
Monteiro
,
M.
, and
Rizzo
,
F.
,
2008
, “
The Oxidation Behaviour of Metals and Alloys at High Temperatures in Atmospheres Containing Water Vapour: A Review
,”
Prog. Mater. Sci.
,
53
(
5
), pp.
775
837
.10.1016/j.pmatsci.2007.11.001
75.
Wright
,
I. G.
, and
Dooley
,
R. B.
,
2010
, “
A Review of the Oxidation Behaviour of Structural Alloys in Steam
,”
Int. Mater. Rev.
,
55
(
3
), pp.
129
167
.10.1179/095066010X12646898728165
76.
Abbasi
,
M.
,
Kim
,
D.-I.
,
Shim
,
J.-H.
, and
Jung
,
W.-S.
,
2016
, “
Effects of Alloyed Aluminum and Titanium on the Oxidation Behavior of INCONEL 740 Superalloy
,”
J. Alloys Comp.
,
658
, pp.
210
221
.10.1016/j.jallcom.2015.10.198
77.
Zheng
,
L.
,
Zhang
,
M.
,
Chellali
,
R.
, and
Dong
,
J.
,
2011
, “
Investigations on the Growing, Cracking and Spalling of Oxides Scales of Powder Metallurgy Rene95 Nickel-Based Superalloy
,”
Appl. Surf. Sci.
,
257
(
23
), pp.
9762
9767
.10.1016/j.apsusc.2011.06.005
78.
Smialek
,
J. L.
,
2010
, “
Moisture-Induced Alumina Scale Spallation: The Hydrogen Factor
,” NASA, Washington, DC, Report No.
TM-2010-216260
.https://ntrs.nasa.gov/api/citations/20100021167/downloads/20100021167.pdf
79.
Yan
,
K.
,
Guo
,
H.
, and
Gong
,
S.
,
2013
, “
High-Temperature Oxidation Behavior of Minor Hf Doped NiAl Alloy in Dry and Humid Atmospheres
,”
Corros. Sci.
,
75
, pp.
337
344
.10.1016/j.corsci.2013.06.017
80.
Smialek
,
J. L.
, and
Morscher
,
G. N.
,
2002
, “
Delayed Alumina Scale Spallation on Rene'N5+ Y: Moisture Effects and Acoustic Emission
,”
Mater. Sci. Eng.: A
,
332
(
1–2
), pp.
11
24
.10.1016/S0921-5093(01)01712-9
81.
Onal
,
K.
,
Maris-Sida
,
M. C.
,
Meier
,
G. H.
, and
Pettit
,
F. S.
,
2003
, “
Water Vapor Effects on the Cyclic Oxidation Resistance of Alumina Forming Alloys
,”
Mater. High Temp.
,
20
(
3
), pp.
327
337
.10.1179/mht.2003.039
82.
Janakiraman
,
R.
,
Meier
,
G. H.
, and
Pettit
,
F. S.
,
1999
, “
The Effect of Water Vapor on the Oxidation of Alloys That Develop Alumina Scales for Protection
,”
Metall. Mater. Trans. A
,
30
(
11
), pp.
2905
2913
.10.1007/s11661-999-0128-3
83.
Unocic
,
K. A.
,
Unocic
,
R. R.
,
Pint
,
B. A.
, and
Hayes
,
R. W.
,
2010
, “
Effect of Microstucture and Environment on the High-Temperature Oxidation Behavior of Alloy 718Plus
,”
Proceedings of the 7th International Symposium on Superalloy
, 718, The Minerals, Metals, & Materials Society, Pittsburgh, PA, Oct. 10–13.https://www.tms.org/Superalloys/10.7449/2010/Superalloys_2010_977_991.pdf
84.
Ghule
,
B.
,
Sundaresan
,
C.
,
Ningshen
,
S.
, and
Raja
,
V. S.
,
2023
, “
Oxidation Behaviour of Ni-Base Superalloy 617 in Simulated Advanced Ultra Supercritical (AUSC) Steam
,”
Corros. Sci.
,
224
, p.
111503
.10.1016/j.corsci.2023.111503
85.
Pint
,
B. A.
,
Pillai
,
R.
,
Romedenne
,
M.
, and
Dryepondt
,
S.
,
2023
, “
The Effect of Higher Water Vapor Content in H2-Fired Turbines on High Temperature Durability
,”
ASME
Paper No. GT2023-103866.10.1115/GT2023-103866
86.
Pillai
,
R.
, and
Pint
,
B. A.
,
2021
, “
The Role of Oxidation Resistance in High Temperature Alloy Selection for a Future With Green Hydrogen
,”
JOM
,
73
(
12
), pp.
3988
3997
.10.1007/s11837-021-04972-9
87.
Pint
,
B. A.
,
Haynes
,
J. A.
,
Zhang
,
Y.
,
More
,
K. L.
, and
Wright
,
I. G.
,
2006
, “
The Effect of Water Vapor on the Oxidation Behavior of Ni–Pt–Al Coatings and Alloys
,”
Surf. Coat. Technol.
,
201
(
7
), pp.
3852
3856
.10.1016/j.surfcoat.2006.07.244
88.
Evans
,
H. E.
,
1995
, “
Stress Effects in High Temperature Oxidation of Metals
,”
Int. Mater. Rev.
,
40
(
1
), pp.
1
40
.10.1179/imr.1995.40.1.1
89.
Barnard
,
B. R.
,
Liaw
,
P. K.
,
Buchanan
,
R. A.
, and
Klarstrom
,
D. L.
,
2010
, “
Affects of Applied Stresses on the Isothermal and Cyclic High-Temperature Oxidation Behavior of Superalloys
,”
Mater. Sci. Eng.: A
,
527
(
16–17
), pp.
3813
3821
.10.1016/j.msea.2010.03.050
90.
Lee
,
C. M.
,
Han
,
Y.-S.
,
Jeong
,
J. S.
,
Jung
,
W. S.
,
Song
,
S. H.
, and
Kim
,
J.-S.
,
2022
, “
Effect of Pre-Strain on Oxidation Behaviour of Nickel-Based Superalloys in Air at 700 °C
,”
Corros. Sci.
,
200
, p.
110229
.10.1016/j.corsci.2022.110229
91.
Ramsay
,
J. D.
,
Evans
,
H. E.
,
Child
,
D. J.
,
Taylor
,
M. P.
, and
Hardy
,
M. C.
,
2019
, “
The Influence of Stress on the Oxidation of a Ni-Based Superalloy
,”
Corros. Sci.
,
154
, pp.
277
285
.10.1016/j.corsci.2019.04.023
92.
Foss
,
B. J.
,
Hardy
,
M. C.
,
Child
,
D. J.
,
McPhail
,
D. S.
, and
Shollock
,
B. A.
,
2014
, “
Oxidation of a Commercial Nickel-Based Superalloy Under Static Loading
,”
JOM
,
66
(
12
), pp.
2516
2524
.10.1007/s11837-014-1196-4
93.
Karabela
,
A.
,
Zhao
,
L. G.
,
Tong
,
J.
,
Simms
,
N. J.
,
Nicholls
,
J. R.
, and
Hardy
,
M. C.
,
2011
, “
Effects of Cyclic Stress and Temperature on Oxidation Damage of a Nickel-Based Superalloy
,”
Mater. Sci. Eng.: A
,
528
(
19–20
), pp.
6194
6202
.10.1016/j.msea.2011.04.029
94.
Bache
,
M. R.
,
O'Hanlon
,
J.
,
Child
,
D. J.
, and
Hardy
,
M. C.
,
2016
, “
High Temperature Fatigue Behaviour in an Advanced Nickel Based Superalloy: The Effects of Oxidation and Stress Relaxation at Notches
,”
Theor. Appl. Fract. Mech.
,
84
, pp.
64
71
.10.1016/j.tafmec.2016.03.007
95.
Zhao
,
L. G.
,
Tong
,
J.
, and
Hardy
,
M. C.
,
2010
, “
Prediction of Crack Growth in a Nickel-Based Superalloy Under Fatigue-Oxidation Conditions
,”
Eng. Fract. Mech.
,
77
(
6
), pp.
925
938
.10.1016/j.engfracmech.2010.02.005
96.
Cervellon
,
A.
,
Yi
,
J. Z.
,
Corpace
,
F.
,
Hervier
,
Z.
,
Rigney
,
J.
,
Wright
,
P. K.
,
Torbet
,
C. J.
,
Cormier
,
J.
,
Jones
,
J. W.
, and
Pollock
,
T. M.
,
2020
, “
Creep, Fatigue, and Oxidation Interactions During High and Very High Cycle Fatigue at Elevated Temperature of Nickel-Based Single Crystal Superalloys
,”
Proceedings of the 14th International Symposium on Superalloys
, Superalloys 2020,
Springer
, Cham, Switzerland, pp.
185
195
.
97.
Cruchley
,
S.
,
Li
,
H. Y.
,
Evans
,
H. E.
,
Bowen
,
P.
,
Child
,
D. J.
, and
Hardy
,
M. C.
,
2015
, “
The Role of Oxidation Damage in Fatigue Crack Initiation of an Advanced Ni-Based Superalloy
,”
Int. J. Fatigue
,
81
, pp.
265
274
.10.1016/j.ijfatigue.2015.08.016
98.
Wang
,
D.
,
Lee
,
M.
, and
Suzuki
,
Y.
,
2023
, “
Nitriding Effects of Ammonia Flames on Iron-Based Metal Walls
,”
J. Ammonia Energy
,
1
(
1
), pp. 1–10.10.18573/jae.8
99.
Laws
,
N.
,
Campuzano
,
F.
,
Dawood
,
Z.
, and
Roberts
,
W. L.
,
2023
, “
Design and Performance Assessment of an Experimental Rig to Conduct Material Nitridation Studies at Extreme Ammonia Conditions
,”
ASME
Paper No. GT2023-10079.10.1115/GT2023-10079
100.
Evans
,
A. G.
,
Mumm
,
D. R.
,
Hutchinson
,
J. W.
,
Meier
,
G. H.
, and
Pettit
,
F. S.
,
2001
, “
Mechanisms Controlling the Durability of Thermal Barrier Coatings
,”
Prog. Mater. Sci.
,
46
(
5
), pp.
505
553
.10.1016/S0079-6425(00)00020-7
101.
Trice
,
R.
, and
Trumble
,
K.
,
2014
, “
Thermal Barrier Coatings
,”
Turbine Aerodynamics, Heat Transfer, Materials, and Mechanics
,
American Institute of Aeronautics and Astronautics
,
Reston, VA
, pp.
467
494
.
102.
Rudolphi
,
M.
,
Renusch
,
D.
,
Zschau
,
H.-E.
, and
Schütze
,
M.
,
2009
, “
The Effect of Moisture on the Delayed Spallation of Thermal Barrier Coatings: VPS NiCoCrAlY Bond Coat+ APS YSZ Top Coat
,”
Mater. High Temp.
,
26
(
3
), pp.
325
329
.10.3184/096034009X12506025935911
103.
Pint
,
B. A.
,
Unocic
,
K. A.
, and
Allen Haynes
,
J.
,
2016
, “
The Effect of Environment on Thermal Barrier Coating Lifetime
,”
ASME J. Eng. Gas Turbine Power
,
138
(
8
), p.
082102
.10.1115/1.4032438
104.
Duan
,
W.
,
Song
,
P.
,
Li
,
C.
,
Huang
,
T.
,
Ge
,
Z.
,
Feng
,
J.
, and
Lu
,
J.
,
2019
, “
Effect of Water Vapor on the Failure Behavior of Thermal Barrier Coating With Hf-Doped NiCoCrAlY Bond Coating
,”
J. Mater. Res.
,
34
(
15
), pp.
2653
2663
.10.1557/jmr.2019.51
105.
Pint
,
B. A.
,
Lance
,
M. J.
, and
Allen Haynes
,
J.
,
2019
, “
The Effect of Coating Composition and Geometry on Thermal Barrier Coatings Lifetime
,”
ASME J. Eng. Gas Turbine Power
,
141
(
3
), p.
031004
.10.1115/1.4041309
106.
Clarke
,
D. R.
,
Oechsner
,
M.
, and
Padture
,
N. P.
,
2012
, “
Thermal-Barrier Coatings for More Efficient Gas-Turbine Engines
,”
MRS Bull.
,
37
(
10
), pp.
891
898
.10.1557/mrs.2012.232
107.
Sullivan
,
M. H.
, and
Mumm
,
D. R.
,
2014
, “
Transient Stage Oxidation of MCrAlY Bond Coat Alloys in High Temperature, High Water Vapor Content Environments
,”
Surf. Coat Technol.
,
258
, pp.
963
972
.10.1016/j.surfcoat.2014.07.048
108.
Sullivan
,
M. H.
, and
Mumm
,
D. R.
,
2014
, “
Vapor-Phase-Mediated Phenomena Associated With High Temperature, High Water Content Oxidation of MCrAlX Bond Coats
,”
Oxid. Met.
,
82
(
1–2
), pp.
1
20
.10.1007/s11085-014-9473-8
109.
Zhou
,
C.
,
Yu
,
J.
,
Gong
,
S.
, and
Xu
,
H.
,
2003
, “
Influence of Water Vapor on the High Temperature Oxidation Behavior of Thermal Barrier Coatings
,”
Mater. Sci. Eng.: A
,
348
(
1–2
), pp.
327
332
.10.1016/S0921-5093(02)00742-6
110.
Chen
,
K.
,
Seo
,
D.
, and
Canteenwalla
,
P.
,
2021
, “
The Effect of High-Temperature Water Vapour on Degradation and Failure of Hot Section Components of Gas Turbine Engines
,”
Coatings
,
11
(
9
), p.
1061
.10.3390/coatings11091061
111.
Maris-Sida
,
M. C.
,
Meier
,
G. H.
, and
Pettit
,
F. S.
,
2003
, “
Some Water Vapor Effects During the Oxidation of Alloys That Are α-Al2O3 Formers
,”
Metall. Mater. Trans. A
,
34
(
11
), pp.
2609
2619
.10.1007/s11661-003-0020-5
112.
Angle
,
J. P.
,
Morgan
,
P. E. D.
, and
Mecartney
,
M. L.
,
2013
, “
Water Vapor‐Enhanced Diffusion in Alumina
,”
J. Am. Ceram. Soc.
,
96
(
11
), pp.
3372
3374
.10.1111/jace.12610
113.
Üstündag
,
E.
,
Zhang
,
Z.
,
Stocker
,
M. L.
,
Rangaswamy
,
P.
,
Bourke
,
M. A. M.
,
Subramanian
,
S.
,
Sickafus
,
K. E.
,
Roberts
,
J. A.
, and
Sass
,
S. L.
,
1997
, “
Influence of Residual Stresses on the Evolution of Micro-Structure During the Partial Reduction of NiAl2O4
,”
Mater. Sci. Eng.: A
,
238
(
1
), pp.
50
65
.10.1016/S0921-5093(97)00433-4
114.
Giggins
,
C. S.
, and
Pettit
,
F. S.
,
1971
, “
Oxidation of Ni ‐ Cr ‐ Al Alloys Between 1000° and 1200 °C
,”
J. Electrochem. Soc.
,
118
(
11
), p.
1782
.10.1149/1.2407837
115.
Wallwork
,
G. R.
, and
Hed
,
A. Z.
,
1971
, “
Some Limiting Factors in the Use of Alloys at High Temperatures
,”
Oxid. Met.
,
3
(
2
), pp.
171
184
.10.1007/BF00603485
116.
Nijdam
,
T. J.
,
Kwakernaak
,
C.
, and
Sloof
,
W. G.
,
2006
, “
The Effects of Alloy Microstructure Refinement on the Short-Term Thermal Oxidation of NiCoCrAlY Alloys
,”
Metall. Mater. Trans. A
,
37
(
3
), pp.
683
693
.10.1007/s11661-006-0040-z
117.
Sloof
,
W. G.
, and
Nijdam
,
T. J.
,
2009
, “
On the High-Temp Oxidation of MCrAlY Coatings
,”
Int. J. Mater. Res.
,
100
(
10
), pp.
1318
1330
.10.3139/146.110201
118.
Shillington
,
E. A. G.
, and
Clarke
,
D. R.
,
1999
, “
Spalling Failure of a Thermal Barrier Coating Associated With Aluminum Depletion in the Bond-Coat
,”
Acta Mater.
,
47
(
4
), pp.
1297
1305
.10.1016/S1359-6454(98)00407-8
119.
Haynes
,
J. A.
,
Unocic
,
K. A.
,
Lance
,
M. J.
, and
Pint
,
B. A.
,
2013
, “
Impact of Superalloy Composition, Bond Coat Roughness and Water Vapor on TBC Lifetime With HVOF NiCoCrAlYHfSi Bond Coatings
,”
Surf. Coat Technol.
,
237
, pp.
65
70
.10.1016/j.surfcoat.2013.09.062
120.
Lance
,
M. J.
,
Unocic
,
K. A.
,
Haynes
,
J. A.
, and
Pint
,
B. A.
,
2014
, “
The Effect of Cycle Frequency, H2O and CO2 on TBC Lifetime With NiCoCrAlYHfSi Bond Coatings
,”
Surf. Coat Technol.
,
260
, pp.
107
112
.10.1016/j.surfcoat.2014.08.082
121.
Dillon
,
J. J.
,
Desch
,
P. B.
, and
Lai
,
T. S.
,
2011
,
The Nalco Water Guide to Boiler Failure Analysis
,
McGraw-Hill
,
New York
.
122.
National Academies of Sciences, Division on Engineering, Physical Sciences, Aeronautics, Space Engineering Board, & Committee on Advanced Technologies for Gas Turbines,
2020
,
Advanced Technologies for Gas Turbines
, National Academies of Sciences Engineering, and Medicine,
The National Academies Press
,
Washington, DC
.
123.
EPRI,
2022
, “
Vistra Sponsors Groundbreaking Research in Additive Manufactured Turbine Vanes
,” Electric Power Research Institute, Inc., Palo Alto, CA, accessed Dec. 1, 2023, https://www.epri.com/research/products/000000003002023344
124.
Torkaman
,
A.
,
Vogel
,
G.
, and
Houck
,
L.
,
2021
, "
Design, Development and Validation of Additively Manufactured First Stage Turbine Vane for F Class Industrial Gas Turbine
,"
ASME
Paper No. GT2021-60201.10.1115/GT2021-60201
125.
Bridges
,
A.
,
Shingledecker
,
J.
,
Clark
,
J.
,
Crudden
,
D.
,
Kirka
,
M.
, and
Fernandex-Zelaia
,
P.
,
2024
, “
Options for Improving Performance of Additively Manufactured Nickel-Base Superalloys for Gas Turbine Applications
,”
ASME J. Eng. Gas Turbine Power
,
146
(
3
), p.
031017
.10.1115/1.4063631
126.
ARPA-E,
2024
, “
Ultimate: Ultrahigh Temperature Impervious Materials Advancing Turbine Efficiency
,” ARPA-E, Washington, DC, accessed Dec. 1, 2023, https://arpa-e.energy.gov/technologies/programs/ultimate
127.
Binner
,
J.
,
Porter
,
M.
,
Baker
,
B.
,
Zou
,
J.
,
Venkatachalam
,
V.
,
Diaz
,
V. R.
,
D'Angio
,
A.
,
Ramanujam
,
P.
,
Zhang
,
T.
, and
Murthy
,
T. S. R. C.
,
2020
, “
Selection, Processing, Properties and Applications of Ultra-High Temperature Ceramic Matrix Composites, UHTCMCs
,”
Int. Mater. Rev.
,
65
(
7
), pp.
389
444
.10.1080/09506608.2019.1652006