
To overcome the limitations of solid-liquid phase change materials, such as liquid leakage, unstable shape, and serious corrosion at elevated temperatures, the design of a solid-solid phase change material using Ni-Ti-Zr as a metal element was utilized for its high thermal conductivity and stability to provide new thermal storage solutions for downhole tools. The influence of Zr content on the thermophysical properties and microstructure of the material was also investigated. The results showed that the designed solid-solid phase change material could reach 4129.93 × 106 J2K−1s−1m−4 quality factor, which was much higher than most of the phase change materials. This was attributed to the increase in Zr content, where Zr atoms took the place of Ti atoms, creating vacancy defect and leading to the distortion of the alloy lattice where the Ti2Ni phase gradually became the (Ti, Zr)2Ni phase, the rise in density and enthalpy of the phase transition, and the increase in the temperature of the phase transition. When the Zr content was 12%, the phase transition temperature of the alloy matched the downhole working temperature, demonstrating excellent thermal cycling stability, and is expected to be a heat storage material for downhole tools.
Citation: Wentao Qu, Qian Zhang, Guibian Li, Boyang Pan, Haiying Liu. Microstructure and thermophysical properties of NiTiZr phase change alloys for downhole tool heat storage[J]. AIMS Materials Science, 2025, 12(1): 85-100. doi: 10.3934/matersci.2025007
[1] | Hugo F. Lopez . Microstructural features associated with the effect of temperature on the dimensional stability of an automotive Al-A319 alloy. AIMS Materials Science, 2016, 3(2): 634-644. doi: 10.3934/matersci.2016.2.634 |
[2] | Jian Qin, Zhan Zhang, X.-Grant Chen . Evolution of activation energy during hot deformation of Al–15% B4C composites containing Sc and Zr. AIMS Materials Science, 2019, 6(4): 484-497. doi: 10.3934/matersci.2019.4.484 |
[3] | Nuriya Mukhamedova, Yernat Kozhakhmetov, Mazhyn Skakov, Sherzod Kurbanbekov, Nurzhan Mukhamedov . Microstructural stability of a two-phase (O + B2) alloy of the Ti-25Al-25Nb system (at.%) during thermal cycling in a hydrogen atmosphere. AIMS Materials Science, 2022, 9(2): 270-282. doi: 10.3934/matersci.2022016 |
[4] | Giovanni Di Girolamo, Alida Brentari, Emanuele Serra . Some recent findings on the use of SEM-EDS in microstructural characterisation of as-sprayed and thermally aged porous coatings: a short review. AIMS Materials Science, 2016, 3(2): 404-424. doi: 10.3934/matersci.2016.2.404 |
[5] | Nur Farah Bazilah Wakhi Anuar, Mohd Shukor Salleh, Mohd Zaidi Omar, Saifudin Hafiz Yahaya . Mechanical properties and dry sliding wear behaviour of Al–Si–Mg alloy by equal channel angular pressing. AIMS Materials Science, 2022, 9(5): 733-749. doi: 10.3934/matersci.2022045 |
[6] | Mohamed Lokman Jalaluddin, Umar Al-Amani Azlan, Mohd Warikh Abd Rashid, Norfauzi Tamin . Effect of sintering temperatures on the physical, structural properties and microstructure of mullite-based ceramics. AIMS Materials Science, 2024, 11(2): 243-255. doi: 10.3934/matersci.2024014 |
[7] | Budhi Muliawan Suyitno, Dwi Rahmalina, Reza Abdu Rahman . Increasing the charge/discharge rate for phase-change materials by forming hybrid composite paraffin/ash for an effective thermal energy storage system. AIMS Materials Science, 2023, 10(1): 70-85. doi: 10.3934/matersci.2023005 |
[8] | Valeriy Ishchuk, Danil Kuzenko, Vladimir Sobolev . Piezoelectric and functional properties of materials with coexisting ferroelectric and antiferroelectric phases. AIMS Materials Science, 2018, 5(4): 711-741. doi: 10.3934/matersci.2018.4.711 |
[9] | D. Clayton John . Phase Field Theory and Analysis of Pressure-Shear Induced Amorphization and Failure in Boron Carbide Ceramic. AIMS Materials Science, 2014, 1(3): 143-158. doi: 10.3934/matersci.2014.3.143 |
[10] | G. A. El-Awadi . Review of effective techniques for surface engineering material modification for a variety of applications. AIMS Materials Science, 2023, 10(4): 652-692. doi: 10.3934/matersci.2023037 |
To overcome the limitations of solid-liquid phase change materials, such as liquid leakage, unstable shape, and serious corrosion at elevated temperatures, the design of a solid-solid phase change material using Ni-Ti-Zr as a metal element was utilized for its high thermal conductivity and stability to provide new thermal storage solutions for downhole tools. The influence of Zr content on the thermophysical properties and microstructure of the material was also investigated. The results showed that the designed solid-solid phase change material could reach 4129.93 × 106 J2K−1s−1m−4 quality factor, which was much higher than most of the phase change materials. This was attributed to the increase in Zr content, where Zr atoms took the place of Ti atoms, creating vacancy defect and leading to the distortion of the alloy lattice where the Ti2Ni phase gradually became the (Ti, Zr)2Ni phase, the rise in density and enthalpy of the phase transition, and the increase in the temperature of the phase transition. When the Zr content was 12%, the phase transition temperature of the alloy matched the downhole working temperature, demonstrating excellent thermal cycling stability, and is expected to be a heat storage material for downhole tools.
With the development and exploitation of oil and gas fields gradually moving deeper, and the depletion of shallow oil and gas fields in the ground, oil companies are gradually shifting their focus to the exploration and exploitation of ultra-deep wells. However, ultra-deep well surveys are constrained by various factors such as high temperature and pressure [1,2], complex formations, ultra-long wellbores, corrosion [3,4], and higher demands on downhole tool performance. The logging instrument is an indispensable tool for the oil and gas resources industry to obtain downhole information by measuring downhole temperature, pressure, porosity, and other parameters to determine the location of the oil field and reserves; however, it often faces problems such as heat leakage from the external high-temperature environment and defects in the internal self-heating of the electronic devices, resulting in the temperature of the electronics in the logging instrument to exceed the design temperature rapidly and failing to work [5].
Downhole tools usually use solid-liquid phase change materials (SL-PCM) for temporary thermal storage [6,7], which effectively extends the uptime of downhole electronics [8], but SL-PCM has the disadvantages of low thermal storage density, phase separation, and low heat absorption rate [9], which largely limits the temperature control effect of the thermal management system of logging instrument; this greatly limits the temperature control effect of the thermal management system of the logging instrument; Song et al. used graphite to reduce the supercooling of MgCl2·6H2O from 40.6 to 29.4 ℃, while the thermal conductivity was significantly improved [10]; Li et al. successfully solved the supercooling and phase separation problems of CaCl2·6H2O-MgCl2·6H2O eutectic salt by adjusting the ratios of the nucleating agent and the thickening agent [11]; and Wu et al. prepared Na2SO4·10H2O-Na2HPO4·12H2O stoichiometric composite phase change materials using porous SiO2 and expanded graphite [12,13], which effectively solved the phenomena of overcooling and phase segregation of phase change materials. However, SL-PCM has serious defects such as liquid leakage and easy reacts with containers when in the liquid state; thus, SL-PCM must be encapsulated when in use, which inevitably leads to a reduction in the volumetric storage capacity of the system and an increase in the production cost [14].
The solid-solid phase change materials (SS-PCM) have the advantages of high heat storage density, simple heat storage structure, high reliability, no power drive, and are not affected by other factors, which are particularly suitable for temperature control in high-temperature fields such as aerospace and petroleum industries [15,16]. Fukahori et al. found that compared with the conventional molten-salt phase change materials [17], Al-Si alloys have a higher heat storage capacity and a higher thermal conductivity, which is more suitable for high-temperature superconducting heat storage materials; Shharar D et al. found that NiTi can be used as a high power material substitute for 1-octadecanoic without sacrificing the bulk heat capacity [18]; and Kato1 et al. investigated the latent heat values of Ni-Ti alloys with different compositions and found that the latent heat values of Ni-Ti alloys are greater than those of other SMAs and can reach 35 J/g [19]. However, the isotonic ratio of the Ni-Ti alloys has phase transition temperatures below 100 ℃ and cannot be applied in high-temperature environments [20,21]. Li et al. developed ultra-high performance high-temperature Ni-Mn-Ti PCMs using a martensitic phase transition strategy [22]. The latent heat of Ni-Mn-Ti SS-PCMs was in the range of 43–65 J/g. Hite et al. utilized a shape memory alloy Ni-Ti-Hf as a solid-solid phase transition material since Ni50.3Ti29.7Hf20 exhibits a high enthalpy of transformation of 32.5 J/g and a thermal conductivity of 11.19 Wm−1K−1 [23]. Nevertheless, the phase transition temperatures of these ternary alloys are too high to be suitable as heat storage materials for downhole high-temperature environments. It is well known that adding an appropriate amount of Zr element to the Ni-Ti alloy can significantly improve the mechanical properties and martensitic phase transition temperature. The Ni-Ti-Zr ternary alloy not only has a lower cost but also has the advantages of high thermal conductivity and no risk of leakage as a phase change material, which makes it an ideal thermal energy storage material. Moreover, solid-solid phase change materials are relatively lacking in downhole applications; thus, it is of great significance to utilize the Ni-Ti-Zr ternary alloy as a solid-solid phase change material for the thermal control of downhole tools.
We focus on the high-temperature environment of the downhole, Zr as a substitute of Ti element, designing an Ni49.5Ti50.5−xZrx (x = 8, 10, 12, 15 at.%) alloy, and exploring new thermal storage materials for downhole tools. We also investigate the effect of different Zr content on the thermophysical properties and microstructure of the material. Finally, we use the thermophysical properties of alloys after multiple alternations of hot and cold environments to verify the thermal cycling stability of materials.
Ni49.5Ti50.5−xZrx (x = 8, 10, 12, 15 at.%) alloy button ingots were fabricated by arc melting under argon protection using high purity Ni (99.99%), Ti (99.98%) and Zr (99.99%) elements as raw materials. To ensure homogeneity of the material composition, the ingots were repeatedly melted at least six times. The button ingots were then sealed into quartz tubes filled with argon gas and homogenized at 1223 K for 2 h, followed by water cooling. Samples of different sizes were cut from the ingots by electro-discharge machining for various experimental measurements.
X-ray diffraction patterns (XRD) were obtained utilizing a Rigaku diffractometer (D8 Advance), equipped with Cu-Kα radiation, to analyze the phase compositions of the investigated alloys. The chemical composition from selected sample areas was analyzed with an EDS detector. Scanning electron microscopy (SEM, Zeiss Gemini 360) was used to analyze the surface features. The SEM samples were incrementally ground using SiC papers (400–3000 grit) to obtain a scratch-free surface before ultrasonic cleaning with acetone and ethanol for 10 min and drying with cold air, which was then chemically etched with Kroll's reagent (5 mL of HF, 25 mL of HNO3, and 50 mL of H2O).
The densities of samples with different compositions were measured using the Archimedes' method. The samples of size 10 × 10 × 1.5 mm3 were first weighed in air (denoted as wa) and then immersed in water and weighed again (denoted as ws). The density (ρ) of the measured samples was determined by ρ = ρwwa/(wa − ws), where ρw is the density of water.
The phase transition temperature and the enthalpy change of the phase transition were determined by differential scanning calorimetry measurements (DSC25) at a heating/cooling rate of 20 ℃min−1 over a temperature range of −40 to 450 ℃. The specific heat capacity was measured by the modulated DSC technique, utilizing sapphire with a well-known specific heat capacity as a standard sample. The thermal cycling stability of Ni49.5Ti38.5Zr12 PCM was examined by DSC measurements during 100 cycles of heating and cooling between −40 and 130 ℃ at 10 ℃ min−1.
Thermal diffusivity was measured using the laser flash method (NETZSCH, LFA467). Square samples with dimensions of 10 × 10 × 1.5 mm3 were used. Measurements were carried out at several temperatures during the heating for the phase change to occur. The thermal conductivity k was obtained by k = Cp·ρ·α, where Cp, ρ, and α are the specific heat capacity, density, and thermal diffusivity. The specimen was measured by an HV-50A hardness tester with a load of 100 N and a retention time of 15 s.
Figure 1a–d shows the scanning electron microscopy images of Ti-Ni-Zr alloys with different Zr contents. The smooth (Ti, Zr)Ni matrix and austenite phase can be seen in all the images, in addition to some Ti2Ni-type phases. With the increase of Zr content, the Ti2Ni-type phases gradually decrease, and the Ti2Ni-type phase around the second phase gradually turns into a gray color [24]. The gray zone should originate from the amplitude-modulated decomposition of the single-phase solid solution at high temperatures during cooling and the corresponding consists of the elements Ti, Ni, and Zr. From the EDS results in Table 1, the gray zone may be (Ti, Zr)2Ni. On the other hand, with the increase of Zr content, the austenite phase is gradually connected into flakes from a single vesicle, and the (Ti, Zr)2Ni phase is mostly distributed in the grain boundaries or within the grains. This is mainly due to the displacement of small-sized Ti atoms by large-sized Zr atoms, creating vacancy defects and causing redistribution of Zr and Ti atoms in the B2 phase, which in turn leads to distortion of the lattice, and the small vesicular matrix phase is squeezed and connected into sheets.
Zr content (at.%) | P1 zone (matrix phase) (at.%) | P2 zone (gray zone) (at.%) | ||||||
Ni | Ti | Zr | Ti + Zr/Ni | Ni | Ti | Zr | Ti + Zr/Ni | |
8 | 49.81 | 42.52 | 7.67 | 1.0076 | 30.52 | 65.72 | 3.76 | 2.2765 |
10 | 49.62 | 40.66 | 9.72 | 1.0153 | 34.67 | 56.61 | 8.72 | 1.8843 |
12 | 49.71 | 38.23 | 12.06 | 1.0117 | 33.59 | 52.34 | 14.07 | 1.9771 |
15 | 49.69 | 35.53 | 14.78 | 1.0125 | 30.88 | 49.79 | 19.33 | 2.2383 |
Figure 2 shows the XRD patterns of Ni49.5Ti50.5−xZrx alloys (x = 8, 10, 12, 15 at.%) at room temperature. It can be seen that all the alloys have a (Ti, Zr)2Ni phase and austenite B2 and martensite B19' phases, with the observed diffraction peaks of the Ni-Ti binary alloys exhibiting a high degree of similarity to the aforementioned position. The crystal surface indices corresponding to the main diffraction peak of XRD are (210), (220), (020), and (132) crystal planes, respectively. When the Zr content is 8% and 10%, the Ni atoms preferentially bind with the Ti atoms to form the Ti2Ni phase. With the addition of the Zr element, the Ni-Ti lattice is distorted [25], and when the Zr content exceeds 12%, the residual Ni atoms diffuse into the TiNi matrix to form a Ni-rich region, which combines with Zr atoms to form the NiZr phase [26]. This promotes the formation of the B2 phase and the Ti2Ni phase, and the Ti2Ni phase gradually becomes the (Ti, Zr)2Ni phase, and the diffraction peak is shifted to the left. Furthermore, the (Ti, Zr)2Ni and Ti2Ni phases are body-centered cubic (BCC) lattice structures, and the NiZr phase is a face-centered cubic (FCC) lattice structure.
The differential scanning calorimetry (DSC) curves of Ni49.5Ti50.5−xZrx (x = 8, 10, 12, 15 at.%) are shown in Figure 3a. It can be seen that all materials have heat absorption peaks during the temperature rise and exothermic peaks during the temperature reduction, corresponding to the reversible phase transformation processes of the low-temperature martensitic phase and the high-temperature austenitic phase [22]. With the increase of the Zr content, the heat absorption peaks of these solid-solid phase transition alloys shift to the right, which implies an increase in the onset temperature of the phase transformation. When the Zr content is in the range of 8%–10%, the phase transformation onset temperatures are increased slightly, and when the Zr content is 10%–15%, the phase transition onset temperature increases significantly. Based on the results of DSC, the onset and end temperatures of the forward and reverse phase transitions of these PCMs (Ms, Mf, As, and Af) are determined using the tangent method; the results are shown in Figure 3b. The As for x = 8 is 11 ℃ and that for x = 15 is 144 ℃. This shows that the phase transition temperature is increased by about 20 ℃ for each 1% increase in Zr content, indicating that the replacement of Ti by Zr can increase the phase transition temperature of the alloy [27,28].
Figure 3c shows the variation curves of average valence electron number e/a and average valence electron concentration Cv with Zr content. The average valence electron number (ev/a) and the average valence electron concentration (Cv) can be expressed using Eqs 1 and 2, respectively: fNi, fTi, and fZr represent the atomic fraction of the Ni, Ti, and Zr atoms in the Ni-Ti-Zr alloy; evNi, evTi and evZr represent the valence electron number of Zr atoms of Ti atoms of Ni, respectively; ZNi, ZTi and ZZr represent the atomic number of the Ni, Ti, and Zr atoms, respectively. With the increase of Zr content, the average valence electron number e/a of each alloy does not change, but the average valence electron concentration Cv of the alloys gradually decreases. This is mainly due to Ti and Zr belonging to the same family, with the same number of outermost electrons and different atomic numbers, which lead to an increase in the phase transition temperature.
(1) |
(2) |
The enthalpy changes (which equals the latent heat L) associated with the forward and reverse phase transitions, denoted as ΔHM and ΔHA, respectively, are obtained by integrating the areas of the exothermic and absorptive peaks on the DSC curves (Figure 3a), and the results are shown in Figure 3d. It is seen that the enthalpy change value increases gently with the increase of Zr content x when the Zr content is 8% to 10%. When the Zr content exceeds 12%, the enthalpy change value grows more rapidly, up to about 30 J g−1, which shows a more favorable performance.
Figure 4 shows the density as a function of Zr content x. It can be seen that the density of the alloys increases linearly with increasing Zr content x. The density of all compositions ranges from 6.55–6.78g·cm−3 due to the replacement of the low-density Ti (4.54 g·cm−3) by the high-density Zr (6.49g·cm−3). The densities of these metallic SS-PCMs have much higher densities than the organic, inorganic, and hydrated salt PCMs (typically 0.10–2.50 g·cm−3). In addition, the red dashed line shows a linear regression model of the trend of the density fitted with the Zr content, which is given by the following equation: ρ = 0.03356 × Zr + 6.26159, where ρ is density in g/cm3 and Zr is hafnium content in at.%. Due to the small variation of the Zr content, there is some error between the measured density values and the calculated prediction from the fitted curve.
Figure 5a–d shows the curves of specific heat capacity Cp and thermal conductivity k versus temperature T for different materials. From the left axis, it can be seen that for different material compositions, the Cp versus temperature T curves exhibit similar shapes. When the phase transformation occurs, the Cp rises with the temperature up to the peak temperature, and when it exceeds the peak temperature, the Cp starts to show a downward trend, which is the same as that exhibited by the DSC images of each material. The Cp value rapid increase is due to the increase in the enthalpy of the phase transition during the reverse martensitic phase transformation, so the specific heat capacity Cp in the range of the phase transformation does not represent the true Cp value [29]. After the completion of the phase transformation, the Cp value of each material is almost the same at high versus low temperatures, which is 0.49–0.52 Jg−1K−1. This overall trend is consistent with the specific heat capacity trend in NiTi SMA [30], which suggests that the Cp values may be independent of the composition of the high and low temperature phases.
As seen from the right axis of Figure 5a–d, the thermal conductivity k exhibits the same trend as the specific heat capacity Cp at different temperatures. Before the phase transition occurs, the thermal conductivity k is about 8–10 W m−1K−1 at low temperatures, whereas during heating to undergo the phase transition, k exhibits a linear increase with increasing temperature. This could be due to the increase in electronic thermal conductivity due to the change in carrier mobility during the phase transition [31]. After the completion of the phase transition, it decreases slightly as the temperature increases, and the value of k at high temperatures is about 12 W m−1K−1, which is much higher than that of the thermal conductivity of the conventional organic materials (commonly 0.1–2.0 W m−1K−1). From Table 2, it can be seen that the thermal conductivity of the Ni-Ti-Zr alloy is higher compared to traditional organic paraffin. The reason for this is that the molecular interactions in paraffin are relatively weak, and the molecular chains are long, leading to a lower efficiency in heat transfer. The high thermal conductivity of the Ni-Ti-Zr alloy can be attributed to the following reasons: The Ti2Ni phase in the alloy has a BCC lattice structure, which is relatively dense. Moreover, heat transfer primarily relies on lattice vibrations. The (Ti, Zr)2Ni phase has a FCC lattice structure, which has a higher thermal conductivity. This lattice structure is less sparse than the Ti2Ni phase, which facilitates better heat transfer. According to the XRD pattern, the increase in Zr content promotes the formation of the (Ti, Zr)2Ni phase, thereby enhancing heat transfer. As a result, the thermal conductivity K of the alloy increases progressively. As the temperature increases, the lattice vibrations in the (Ti, Zr)2Ni phase are more strongly excited by heat, increasing the lattice vibrations and further improving the thermal conductivity of the alloy.
Title | Density/g·cm−3 | Specific heat capacity/J g−1 K−1 | Thermal conductivity/W m−1K−1 | Latent heat/J g−1 |
organic paraffin | 0.82 | 2.011 | 0.15 | 210 |
Ni49.5Ti42.5Zr8 | 6.55 | 0.492 | 13.80 | 12.6 |
Ni49.5Ti40.5Zr10 | 6.58 | 0.443 | 10.59 | 17.8 |
Ni49.5Ti38.5Zr12 | 6.65 | 0.431 | 15.65 | 28.3 |
Ni49.5Ti35.5Zr15 | 6.78 | 0.423 | 20.17 | 30.2 |
We compare this with the commonly used thermal storage material (organic paraffin) to investigate the feasibility of Ni-Ti-Zr alloy as a heat storage material for downhole tools. The thermophysical parameters of paraffin and the alloy are shown in Table 2. It can be seen that the latent heat value of the paraffin is higher than that of the Ni-Ti-Zr alloy, but the Ni-Ti-Zr alloy has a high density, and more heat is stored under the same mass, which is favorable for use in a limited space and is expected to be used for downhole heat storage tools. Moreover, the high thermal conductivity of the Ni-Ti-Zr alloy indicates that the heat is transferred faster inside the alloy. The quality factor (FOM) as a quantifiable measure of PCM performance is usually given by the Eq 3:
(3) |
where k, ρ, and L are the thermal conductivity, density, and latent heat of phase transition, respectively [32,33]. Calculating the FOM values of Ni-Ti-Zr alloys with different compositions according to Eq 3 gives a quality factor of up to 4129.93 × 106 J2K−1s−1m−4, which is much higher than existing organic paraffins. This provides a theoretical basis for thermal energy storage in metal-solid-solid phase-change materials.
The thermal cycling stability of the thermophysical properties of PCMs is critical for their practical thermal storage applications when working in downhole high-temperature environments [34]. The thermal cycling stability of Ni49.5Ti38.5Zr12 SS-PCM is investigated by measuring the DSC curves during 100 heating and cooling processes of the material. Figure 6a shows the results of Ni49.5Ti38.5Zr12 SS-PCM after 100 rapid heating and cooling cycles at 10 ℃min−1. It can be seen that the peak of DSC gradually shifts to the left with the increase in the number of thermal cycles and stabilizes after several thermal cycles in which the reverse phase transition temperatures (As and Af) slightly decrease during the thermal cycles; the forward phase transition temperatures (Ms and Mf) remain almost unchanged during the cycles. Its thermophysical parameters after multiple thermal cycles are shown in Figure 6b, where the enthalpy change values ΔHM (for forward phase change) and ΔHA (for reverse phase change) remain unchanged during multiple thermal cycles. Ni-Ti-Zr alloy heat is released during the transition to martensite, but when cooled by the reverse transition from martensite down to autensite, the heat is absorbed. In this process, the heat is caused by the internal phase transition of the alloy, which involves atomic rearrangement, structural change, lattice dislocation, and other processes that are symmetrical, resulting in roughly the same heat absorbed or released by each phase transition without the loss or increase of heat. Thus, the latent heat of the forward/reverse martensitic phase transition, in this study, is almost equivalent. This indicates the superior thermal cycling stability of Ni-Ti-Zr SS-PCM, making it a strong candidate for heat storage materials for downhole tools.
From Figure 7, the hardness of Ni-Ti-Zr alloy gradually increases in the process of 1 to 20 thermal cycles, which promotes the precipitation of the enhanced phase or the refinement of the grains. When the number of thermal cycles reaches 30 times, the precipitate may dissolve or redistribute, and the alloy shows a higher hardness. However, when the number of thermal cycles exceeds 30 times, the hardness of the alloy does not increase because the grain has reached a certain uniformity, and the precipitation reinforcement tends to balance. When the number of thermal cycles continues to increase, the hardness of the Ni-Ti-Zr alloy begins to decrease, possibly because the excessive thermal cycles lead to excessive dissolution of the precipitate or the formation of an unstable phase, which reduces the hardness of the alloy [35,36].
The thermophysical parameters of Ni49.5Ti50.5−xZrx (x = 8, 10, 12, 15 at%) SS-PCMs were investigated due to the advantages of metallic phase change materials such as high thermal conductivity and density. The effect of different contents of Zr on their thermophysical parameters was analyzed, and our major conclusions are as follows:
1. The microscopic morphology reveals the presence of a (Ti, Zr)Ni matrix and a small amount of Ti2Ni phase, and with the increase of Zr, the Ti2Ni phase gradually becomes the (Ti, Zr)2Ni phase. Moreover, the Zr content is in the range of 8%~15%, the enthalpy of the phase transition of the alloy with the increase in the Zr content of x increases, and when the Zr content is more than 12%, the enthalpy of phase transition can reach around 30Jg−1. This is attributed to the fact that the addition of the Zr element induces distortion in the Ni-Ti lattice and hinders the martensitic transformation, thus promoting the increase of the enthalpy of phase transition.
2. Ni49.5Ti50.5−xZrx SS-PCMs quality factor FOM have a maximum value of 4129.93 × 106J2K−1s−1m−4, and the phase transition temperature of these SS-PCMs can be adjusted by controlling the elemental composition ratio. When the Zr content is 12%, the temperature range required for the phase transformation meets the requirements of downhole working conditions. After 100 rapid heating and cooling cycles, the phase transformation temperature of the material decreases slightly, but the enthalpy change value is almost unchanged, which indicates that the material has good thermal cycling stability.
3. Metal phase change materials for heat storage have no risk of leakage and exhibit better heat storage performance than conventional SL-PCMs, and Ni-Ti-Zr alloys offer a potential option as SS-PCMs for heat storage in downhole tools. The phase change material studied in this paper can be applied to the special thermos, mainly to absorb the heat generated by the electronic devices when the logging instrument works.
4. Thermal cycle times can increase the hardness of the alloy within a certain range. Thus, the Ni-Ti-Zr alloy has 30 thermal cycles with the highest hardness and the strongest deformation resistance. Therefore, this material is applied to the log instrument to withstand a large load and high underground pressure.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
This work was supported by the National Natural Science Foundation of China (No. 52071261).
The authors are grateful for the financial support from the National Natural Science Foundation of China (No. 52071261).
Conceptualization: W.Q.; methodology: Q.Z.; validation: H.L. and G.L.; formal analysis: Q.Z.; investigation: Q.Z. and B.P.; resources: H.L.; data curation: Q.Z.; writing—original draft preparation: Q.Z. and H.L; writing—review and editing: Q.Z. and B.P.; visualization: H.L.; supervision: W.Q. and G.L. All authors have read and agreed to the published version of the manuscript.
The authors declare no conflict of interest.
[1] |
Zhang Y, Xu M, Song J, et al. (2022) Study on the corrosion change law and prediction model of cement stone in oil wells with CO2 corrosion in ultra-high-temperature acid gas wells. Constr Build Mater 323: 125879. https://doi.org/10.1016/j.conbuildmat.2021.125879 doi: 10.1016/j.conbuildmat.2021.125879
![]() |
[2] |
Li H, Zhang X, Qi D, et al. (2015) Failure analysis of the adhesive metal joint bonded on anticorrosion plastic alloy composite pipe. Eng Fail Anal 47: 49–55. https://doi.org/10.1016/j.engfailanal.2014.09.009 doi: 10.1016/j.engfailanal.2014.09.009
![]() |
[3] |
Peng Z, Lv F, Feng Q, et al. (2022) Enhancing the CO2-H2S corrosion resistance of oil well cement with a modified epoxy resin. Constr Build Mater 326: 126854. https://doi.org/10.1016/j.conbuildmat.2022.126854 doi: 10.1016/j.conbuildmat.2022.126854
![]() |
[4] |
Qu W, Pan B, Gong H, et al. (2023) Electrochemical corrosion and impedance studies of Ti-30Zr-xNb (x = 7, 10, 13 at.%) alloy in simulated downhole environment. J Solid State Electr 27: 1155–1164. https://doi.org/10.1007/s10008-023-05430-z doi: 10.1007/s10008-023-05430-z
![]() |
[5] |
Xue Z, Zhang C, Li Z, et al. (2022) An on-site temperature prediction method for passive thermal management of high-temperature logging apparatus. Meas Control 55: 1180–1189. https://doi.org/10.1177/00202940221076809 doi: 10.1177/00202940221076809
![]() |
[6] |
Liu Y, Zheng R, Li J (2022) High latent heat phase change materials (PCMs) with low melting temperature for thermal management and storage of electronic devices and power batteries: Critical review. RSER 168: 112783. https://doi.org/10.1016/j.rser.2022.112783 doi: 10.1016/j.rser.2022.112783
![]() |
[7] |
Sharma RK, Ganesan P, Tyagi V (2016) Long-term thermal and chemical reliability study of different organic phase change materials for thermal energy storage applications. J Therm Anal Calorim 124: 1357–1366. https://doi.org/10.1007/s10973-016-5281-5 doi: 10.1007/s10973-016-5281-5
![]() |
[8] |
Peng J, Deng C, Wei F, et al. (2023) A hybrid thermal management system combining liquid cooling and phase change material for downhole electronics. J Energy Storage 72: 108610. https://doi.org/10.1016/j.est.2023.108610 doi: 10.1016/j.est.2023.108610
![]() |
[9] |
Yang X, Li C, Ma Y, et al. (2023) High thermal conductivity of porous graphite/paraffin composite phase change material with 3D porous graphite foam. Chem Eng J 473: 145364. https://doi.org/10.1016/j.cej.2023.145364 doi: 10.1016/j.cej.2023.145364
![]() |
[10] |
Song Z, Deng Y, Li J, et al. (2018) Expanded graphite for thermal conductivity and reliability enhancement and supercooling decrease of MgCl2·6H2O phase change material. Materials Res Bull 102: 203–208. https://doi.org/10.1016/j.materresbull.2018.02.024 doi: 10.1016/j.materresbull.2018.02.024
![]() |
[11] |
Li G, Zhang B, Li X, et al. (2014) The preparation, characterization and modification of a new phase change material: CaCl2·6H2O-MgCl2·6H2O eutectic hydrate salt. Sol Energ Mat Sol C 126: 51–55. https://doi.org/10.1016/j.solmat.2014.03.031 doi: 10.1016/j.solmat.2014.03.031
![]() |
[12] |
Wu Y, Wang T (2015) Hydrated salts/expanded graphite composite with high thermal conductivity as a shape-stabilized phase change material for thermal energy storage. Energy Convers Manage 101: 164–171. https://doi.org/10.1016/j.enconman.2015.05.006 doi: 10.1016/j.enconman.2015.05.006
![]() |
[13] |
Wu Y, Wang T (2014) Preparation and characterization of hydrated salts/silica composite as shape-stabilized phase change material via sol-gel process. Thermochim Acta 591: 10–15. https://doi.org/10.1016/j.tca.2014.07.012 doi: 10.1016/j.tca.2014.07.012
![]() |
[14] |
Zhang N, Yuan Y, Cao X, et al. (2018) Latent heat thermal energy storage systems with solid-liquid phase change materials: A review. Adv Eng Mater 20: 753. https://doi.org/10.1002/adem.201700753 doi: 10.1002/adem.201700753
![]() |
[15] |
Shamberger PJ, Bruno NM (2020) Review of metallic phase change materials for high heat flux transient thermal management applications. Appl Energy 258: 113955. https://doi.org/10.1016/j.apenergy.2019.113955 doi: 10.1016/j.apenergy.2019.113955
![]() |
[16] |
Fernandes D, Pitié F, Cáceres G, et al. (2012) Thermal energy storage: "How previous findings determine current research priorities". Energy 39: 246–257. https://doi.org/10.1016/j.energy.2012.01.024 doi: 10.1016/j.energy.2012.01.024
![]() |
[17] |
Fukahori R, Nomura T, Zhu C, et al. (2016) Thermal analysis of Al-Si alloys as high-temperature phase-change material and their corrosion properties with ceramic materials. Appl Energy 163: 1–8. https://doi.org/10.1016/j.apenergy.2015.10.164 doi: 10.1016/j.apenergy.2015.10.164
![]() |
[18] |
Sharar DJ, Leff AC, Wilson A, et al. (2021) High-capacity high-power thermal energy storage using solid-solid martensitic transformations. Appl Therm Eng 187: 116490. https://doi.org/10.1016/j.applthermaleng.2020.116490 doi: 10.1016/j.applthermaleng.2020.116490
![]() |
[19] |
Kato H (2021) Latent heat storage capacity of NiTi shape memory alloy. J Mater Sci 56: 8243–8250. https://doi.org/10.1007/s10853-021-05777-6 doi: 10.1007/s10853-021-05777-6
![]() |
[20] |
Li G, Bao J, Yu T, et al. (2024) An atomistic study of effects of temperature and Ni element on the phase transition and wear behavior of NiTi shape memory alloy. Tribol Int 192: 109309. https://doi.org/10.1016/j.triboint.2024.109309 doi: 10.1016/j.triboint.2024.109309
![]() |
[21] |
Frenzel J, George EP, Dlouhy A, et al. (2010) Influence of Ni on martensitic phase transformations in NiTi shape memory alloys. Acta Mater 58: 3444–3458. https://doi.org/10.1016/j.actamat.2010.02.019 doi: 10.1016/j.actamat.2010.02.019
![]() |
[22] |
Li S, He L, Lu H, et al. (2023) Ultrahigh-performance solid-solid phase change material for efficient, high-temperature thermal energy storage. Acta Mater 249: 118852. https://doi.org/10.1016/j.actamat.2023.118852 doi: 10.1016/j.actamat.2023.118852
![]() |
[23] |
Hite N, Sharar DJ, Trehern W, et al. (2021) NiTiHf shape memory alloys as phase change thermal storage materials. Acta Mater 218: 117175. https://doi.org/10.1016/j.actamat.2021.117175 doi: 10.1016/j.actamat.2021.117175
![]() |
[24] |
Li J, Yi X, Sun K, et al. (2018) The effect of Zr on the transformation behaviors, microstructure and the mechanical properties of Ti-Ni-Cu shape memory alloys. J Alloys Compd 747: 348–353. https://doi.org/10.1016/j.jallcom.2018.03.053 doi: 10.1016/j.jallcom.2018.03.053
![]() |
[25] |
Evirgen A, Karaman I, Santamarta R, et al. (2016) Relationship between crystallographic compatibility and thermal hysteresis in Ni-rich NiTiHf and NiTiZr high temperature shape memory alloys. Acta Mater 121: 374–383. https://doi.org/10.1016/j.actamat.2016.08.065 doi: 10.1016/j.actamat.2016.08.065
![]() |
[26] |
Hsieh SF, Hsue AWJ, Chen SL, et al. (2013) EDM surface characteristics and shape recovery ability of Ti35.5Ni48.5Zr16 and Ni60Al24.5Fe15.5 ternary shape memory alloys. J Alloys Compd 571: 63–68. https://doi.org/10.1016/j.jallcom.2013.03.111 doi: 10.1016/j.jallcom.2013.03.111
![]() |
[27] |
Kornegay SM, Kapoor M, Hornbuckle BC, et al. (2021) Influence of H-phase precipitation on the microstructure and functional and mechanical properties in a Ni-rich NiTiZr shape memory alloy. Mater Sci Eng A 801: 140401. https://doi.org/10.1016/j.msea.2020.140401 doi: 10.1016/j.msea.2020.140401
![]() |
[28] |
Karakoc O, Atli KC, Evirgen A, et al. (2020) Effects of training on the thermomechanical behavior of NiTiHf and NiTiZr high temperature shape memory alloys. Mater Sci Eng A 794: 139857. https://doi.org/10.1016/j.msea.2020.139857 doi: 10.1016/j.msea.2020.139857
![]() |
[29] |
Zhu J, Wu H, Wu Y, et al. (2021) Influence of Ni4Ti3 precipitation on martensitic transformations in NiTi shape memory alloy: R phase transformation. Acta Mater 207: 116665. https://doi.org/10.1016/j.actamat.2021.116665 doi: 10.1016/j.actamat.2021.116665
![]() |
[30] |
Cheng F, Hu L, Reddy JN, et al. (2014) Temperature-dependent thermal properties of a shape memory alloy/MAX phase composite: Experiments and modeling. Acta Mater 68: 267–278. https://doi.org/10.1016/j.actamat.2013.12.014 doi: 10.1016/j.actamat.2013.12.014
![]() |
[31] |
Zheng Q, Zhu G, Diao Z, et al. (2019) High contrast thermal conductivity change in Ni-Mn-In heusler alloys near room temperature. Adv Eng Mater 21: 1342. https://doi.org/10.1002/adem.201801342 doi: 10.1002/adem.201801342
![]() |
[32] |
Trehern W, Hite N, Ortiz-Ayala R, et al. (2023) NiTiCu shape memory alloys with ultra-low phase transformation range as solid-state phase change materials. Acta Mater 260: 119310. https://doi.org/10.1016/j.actamat.2023.119310 doi: 10.1016/j.actamat.2023.119310
![]() |
[33] |
Liu G, Li S, Song C, et al. (2024) High-entropy Ti-Zr-Hf-Ni-Cu alloys as solid-solid phase change materials for high-temperature thermal energy storage. Intermetallics 166: 108177. https://doi.org/10.1016/j.intermet.2023.108177 doi: 10.1016/j.intermet.2023.108177
![]() |
[34] |
Long Z, Tao P, Kong L, et al. (2023) Effect of cryogenic thermal cycling on the microstructure and mechanical properties of Zr-based bulk metallic glasses. Mater Sci Eng A 863: 144513. https://doi.org/10.1016/j.msea.2022.144513 doi: 10.1016/j.msea.2022.144513
![]() |
[35] |
Wu SK, Chang YC (2019) Thermal cycling effect on transformation temperatures of different transformation sequences in TiNi-based shape memory alloys. Materials 12: 12–16. https://doi.org/10.3390/ma12162512 doi: 10.3390/ma12162512
![]() |
[36] |
Swaminathan G, Sampath V, Santosh S (2024) Effect of cobalt addition on thermal cycling behaviour of Ti50Ni(50−x)cox shape memory alloys. Phys Scr 99: 23–28. https://doi.org/10.1088/1402-4896/ad23b5 doi: 10.1088/1402-4896/ad23b5
![]() |
Zr content (at.%) | P1 zone (matrix phase) (at.%) | P2 zone (gray zone) (at.%) | ||||||
Ni | Ti | Zr | Ti + Zr/Ni | Ni | Ti | Zr | Ti + Zr/Ni | |
8 | 49.81 | 42.52 | 7.67 | 1.0076 | 30.52 | 65.72 | 3.76 | 2.2765 |
10 | 49.62 | 40.66 | 9.72 | 1.0153 | 34.67 | 56.61 | 8.72 | 1.8843 |
12 | 49.71 | 38.23 | 12.06 | 1.0117 | 33.59 | 52.34 | 14.07 | 1.9771 |
15 | 49.69 | 35.53 | 14.78 | 1.0125 | 30.88 | 49.79 | 19.33 | 2.2383 |
Title | Density/g·cm−3 | Specific heat capacity/J g−1 K−1 | Thermal conductivity/W m−1K−1 | Latent heat/J g−1 |
organic paraffin | 0.82 | 2.011 | 0.15 | 210 |
Ni49.5Ti42.5Zr8 | 6.55 | 0.492 | 13.80 | 12.6 |
Ni49.5Ti40.5Zr10 | 6.58 | 0.443 | 10.59 | 17.8 |
Ni49.5Ti38.5Zr12 | 6.65 | 0.431 | 15.65 | 28.3 |
Ni49.5Ti35.5Zr15 | 6.78 | 0.423 | 20.17 | 30.2 |
Zr content (at.%) | P1 zone (matrix phase) (at.%) | P2 zone (gray zone) (at.%) | ||||||
Ni | Ti | Zr | Ti + Zr/Ni | Ni | Ti | Zr | Ti + Zr/Ni | |
8 | 49.81 | 42.52 | 7.67 | 1.0076 | 30.52 | 65.72 | 3.76 | 2.2765 |
10 | 49.62 | 40.66 | 9.72 | 1.0153 | 34.67 | 56.61 | 8.72 | 1.8843 |
12 | 49.71 | 38.23 | 12.06 | 1.0117 | 33.59 | 52.34 | 14.07 | 1.9771 |
15 | 49.69 | 35.53 | 14.78 | 1.0125 | 30.88 | 49.79 | 19.33 | 2.2383 |
Title | Density/g·cm−3 | Specific heat capacity/J g−1 K−1 | Thermal conductivity/W m−1K−1 | Latent heat/J g−1 |
organic paraffin | 0.82 | 2.011 | 0.15 | 210 |
Ni49.5Ti42.5Zr8 | 6.55 | 0.492 | 13.80 | 12.6 |
Ni49.5Ti40.5Zr10 | 6.58 | 0.443 | 10.59 | 17.8 |
Ni49.5Ti38.5Zr12 | 6.65 | 0.431 | 15.65 | 28.3 |
Ni49.5Ti35.5Zr15 | 6.78 | 0.423 | 20.17 | 30.2 |