Yuyu Ma†
ab,
Meiyi Wang†ac,
Chunhua Tanga,
Hui Li*a,
Jie Fu*c and
Hengyong Xua
aDalian National Laboratory for Clean Energy, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, Dalian 116023, China. E-mail: hui.li@dicp.ac.cn
bUniversity of Chinese Academy of Sciences, Beijing 100049, China
cDalian Jiaotong University, Dalian 116028, China. E-mail: dicpfj@126.com
First published on 12th November 2021
It is known that hydrogen embrittlement could result in warping and destruction of pure Pd membranes, which limits the working temperatures to be above 293 °C. This study attempted to investigate the relationship between hydrogen embrittlement resistance and membrane geometry of ultrathin pure Pd membranes of 2.7–6.3 μm thickness. Thin tubular Pd membranes with an o.d. of 4 mm, 6 mm and 12 mm immediately suffered from structural destruction when exposed to H2 at room temperature. In contrast, thin hollow fiber membranes (outer diameter, 2 mm, thickness < 4 μm) exhibit strong resistance against hydrogen embrittlement at temperatures below 100 °C during repeated heating/cooling cycles at a rate up to 10 °C min−1 under H2 atmosphere. This is ascribed to reduced lattice strain gradients during α–β phase transition in cylindrical structures and lower residual stresses according to in situ XRD analysis, which shows a great prospect in low temperature applications.
The thermal and chemical stability of Pd membranes represent a great challenge for their wide applications, e.g. α to β phase transition at temperatures below 573 K and at pressures below 20 bar, leading to large difference in lattice parameters and significant structural deterioration to the membranes,6–9 and the presence of sulfur-containing species leading to the formation of palladium sulfide species as in the case of zeolite catalysts.10–13 Several strategies have been developed to improve the stability, e.g., Pd nanoparticles packed inside the porous support to suppress the α to β phase transition,14,15 and the formation of Pd–Ag, Pd–Cu or Pd/Y alloys.16,17
The application of pure Pd membranes often limits its operating temperature above 300 °C.18–21 There are two phases in the Pd–H system, the one with lower hydrogen content is called α-phase, and the hydrogen-rich phase is usually termed β-phase. When below the critical temperature about 293 °C and at hydrogen pressure < 20 bar, the β-phase nucleates and grows in α-phase and this system is possible for the two phases to coexist.22–25 The α- and β-phase have the same lattice symmetry but very different volumes, for example, at 25 °C the H/Pd ratio of α-phase is about 0.015, whereas β-phase has nH/nPd ≈ 0.7 at 1 bar.26,27 When phase transition occurs, the expansion of volume for the phase transition is over 10%, which is accompanied by considerable mechanical stress and leading to the fracture of membrane.22,28,29
It is reported that the shear stress generated during the hydrogen absorption process significantly decreases when reducing the radius of the tubular structure.29 On the other hand, residual stresses, which are defined as stresses remaining in material or body after processing, in the absence of external forces or thermal gradients, can exist after many manufacturing processes involving heat treatment, machining or processing operations. They influence the properties of the component and its lifetime.30 Moreover, residual stresses play an important role in the crack formation, and they can translate into stress intensity factors acting on cracks that nucleate and propagate around the imperfections.31 Because of these considerations, this study attempted to fabricate thin Pd membranes on capillary tubular support, which is expected to reduce the internal stress and correlated lattice strain gradients and thus suppress hydrogen embrittlement due to α–β phase transition.
(1) Thin tubular Pd membranes on porous support with a diameter of 2 mm (denoted as P-2), 4 mm (denoted as P-4), 6 mm (denoted as P-6) and 12 mm (denoted as P-12), respectively, were exposed to H2 at room temperature to evaluate the hydrogen resistance with pressure differential alternating between 1 bar and 4 bar.
(2) P-2 and P-4 were tested under repeated heating/cooling cycles between room temperature and 100 °C under a hydrogen atmosphere at a feed/permeate pressure differential up to 10 bar.
(3) To investigate the phase transition of P-6 at higher pressures, the feed pressure was increased from 2 bar to 10 bar while the permeate pressure remained as 1 bar with increasing the temperature from 200 °C to 400 °C.
The membrane surface and cross-sectional analysis of P-2 were carried out using scanning electron microscopy (SEM, JSM-7800F) equipped with energy-dispersive X-ray spectroscopy (EDX). The in situ X-ray diffraction patterns were analyzed using Empyrean-100. The stress was analyzed by X-ray Powder Diffractometer – SmartLab.
The method of measuring residual stresses with XRD is based on the measurement of lattice strains by studying the variations of lattice spacing induced by compressive or tensile stresses and to calculate the stresses from the strains.33 The sin2ψ method can determine the stress in any direction along the plane XY (Fig. 1), and when the angle ψ between the normal of the sample and the normal of the diffracting plane changes, the diffraction angle 2θ of the plane will also change.21 Therefore, the measurement of planes at an angle ψ can be made by changing the tilt of the sample within the diffractometer, and the strains along that direction can be calculated from the variation of lattice spacing d: d0, determined by the position of the Bragg peak of stressed (θ) and stress-free (θ0) material:
![]() | (1) |
The direction Φ is the angle between a direction fixed in the plane and the projection of the normal to the plane of diffraction in that plane. Using the strains to evaluate the stress σϕ, which can be given by:
![]() | (2) |
Diameter (mm) | Porous support | Thickness (μm) | N2 fluxa (mol s−1 m−2) | N2 fluxb (mol s−1 m−2) | Multiple |
---|---|---|---|---|---|
a Before feeding hydrogen.b After feeding hydrogen. | |||||
2.00 | Alumina | 4.00 | 2.12 × 10−10 | 3.72 × 10−11 | 0.18 |
2.00 | Alumina | 2.70 | 2.70 × 10−10 | 3.12 × 10−10 | 1.16 |
4.00 | Alumina | 3.50 | 7.69 × 10−9 | 1.66 × 10−8 | 2.16 |
6.00 | Stainless-steel | 3.90 | 4.43 × 10−10 | 1.10 × 10−8 | 24.86 |
12.00 | Alumina | 6.30 | 1.25 × 10−9 | 2.02 × 10−8 | 16.15 |
![]() | ||
Fig. 2 (a) SEM images of P-2 surface and (b) energy-dispersive X-ray (EDX) spectroscopy line analysis of the cross section. |
To further investigate the hydrogen embrittlement resistance of P-4 and P-2, it was tested under repeated heating/cooling cycles between room temperature and 100 °C under a hydrogen atmosphere at a feed/permeate pressure differential up to 10 bar. The N2 leak rate was measured at a 4.5 bar pressure differential between the temperature cycles as shown in Fig. 3 to evaluate the integrity of the membrane. It can be seen that both H2 and N2 flux of P-2 remained stable after 8 repeated cycles, after exposure to hydrogen atmosphere below 100 °C for about 157 h in total, while N2 flux of P-4 increased steadily and tripled after 6 cycles. Note that the H2 flux obtained was at an appreciable level of 10−7 mol s−1 m−2 Pa−1 under the operating temperature of 100 °C and 10 bar pressure differential, which exhibits great prospects for low-temperature applications.
![]() | ||
Fig. 3 H2 and N2 flux of (a) P-2, (b) P-4 during repeated heating/cooling cycles between room temperature and 100 °C at a pressure differential up to 10 bar. |
Meanwhile, as shown in Fig. 4, the surface morphology of P-2 before and after heating/cooling cycles is basically unchanged, and there is no major change in the surface grain size (Fig. 4a and b).
![]() | ||
Fig. 4 SEM images of P-2 surface (a) and (c) before heating/cooling cycles, (b) and (d) after heating/cooling cycles. |
Subsequently, in situ XRD analysis was carried out to clarify the different behavior of these tubular membranes in the coexistence of α and β phase at room temperatures. The scanning procedure is as below: the scan was first conducted under N2 atmosphere, and then switched to H2 atmosphere, with each scan lasting for 1.5 min. Fig. 5 shows the in situ XRD patterns of Pd membranes with a diameter of 2 mm, 6 mm, and 12 mm, respectively. Peaks appearing on the right correspond to (111) of Pd and α-phase, and the peaks on the left can be indexed to β-phase. Interestingly, it is found that the peaks of α-phase of P-2 and P-6 completely disappeared after 4.5 min of feeding hydrogen whereas that of P-12 disappeared after 6 min of feeding hydrogen. Fig. 5d shows the comparison of the peak intensities of P-2, P-6 and P-12 at 3rd min, which can be seen that the α-phase of P-2 is only 14.4% of the β-phase, the α-phase of P-6 is 44.7% of the β-phase, and the β-phase of P-12 is only 27.3% of the α-phase, implying that it takes a longer time to complete phase transition as to P-6 and P-12 than P-2.
It is known that the coexistence of two unequal phases with different specific volume causes internal stress.2 Since P-2 completes the α–β phase transition at the highest rate, the risk of lattice distortion due to internal stresses is diminished. This provides an explanation for the strong resistance of thin tubular P-2 in the coexistence of α and β phase, i.e., the internal stress is significantly suppressed due to a faster phase transition process with the decrease of the radius. The hydrogen embrittlement resistance of thin hollow fiber Pd membranes, as compared to conventional tubular membranes, is attributed to tubular structure effects originating from the suppressed internal stress, which coincides with modeling studies.29
Fig. 6a–c shows that the phase transition of pure Pd membranes occur in the temperature range of 150 °C to 200 °C at a feed pressure of 2 bar which has no relation to varying diameters, i.e. 2 mm, 6 mm and 12 mm. Fig. 6d indicates that the β–α phase transition is completed at temperatures above 170 °C for these membranes. Further detailed analysis (Fig. 7) shows that the peak change is completed at 150 °C for P-2 while there is certain α phase remaining as to P-6 at 150 °C, corroborating the faster phase transition process and shorter-term coexistence of two phases in P-2 with a lower diameter.
![]() | ||
Fig. 6 Phase transition temperatures of membranes with different diameters (a–d) at a feed pressure of 2 bar between 25 °C and 400 °C. |
![]() | ||
Fig. 8 H2 and N2 flux of P-6 during repeated heating/cooling cycles between 200 °C and 400 °C at a pressure differential of 1 bar. |
To further investigate the phase transition of P-6 at higher pressures, the feed pressure was increased from 2 bar to 10 bar while the permeate pressure remained as 1 bar. Fig. 9 shows the activation energy of H2 permeation between 200 °C and 400 °C at a feed/permeate pressure of 10/1 bar, which exhibits two stages in the temperature range of 200–250 °C and 250–400 °C. Note that the N2 permeation of P-6 tripled during the following 10 repeated cycles between 200 °C and 400 °C, while an obvious increase in N2 permeation was observed after only 4 repeated cycles between RT and 400 °C (Fig. 10). This implies the hydrogen embrittlement resistance at temperatures above 200 °C for P-6 even at a high feed pressure of 10 bar.
![]() | ||
Fig. 9 Activation energy curve of P-6 at a pressure differential of 9 bar between 200 °C and 400 °C. |
![]() | ||
Fig. 10 H2 and N2 flux of P-6 during repeated heating/cooling cycles between 200 °C/RT and 400 °C at a pressure differential of 9 bar. |
According to the equation for stress measurement by XRD method (eqn (2)), the stress σϕ can be obtained from the slope of the line by measuring the diffraction line displacement at multiple ψ angles and using 2θ as the vertical coordinate and sin as the horizontal coordinate to calculate and plot the line closest to each experimental value using the least squares. ψ angles are taken as 0°, 17°, 25°, 30°, 35°, 40°, 45° respectively, Young's modulus E takes 117 GPa and Poisson's ratio υ takes 0.39 (from MatWeb). The results of P-2 and P-4 are plotted in Fig. 12, and the residual stress of P-2 is 248.91 MPa with an uncertainty of ±23.86, and that of P-4 is 701.66 MPa with an uncertainty of ±57.47. It is observed from Fig. 12 that P-2 and P-4 show tensile stress and the value of P-4 is greater than P-2, which means that P-2 is more resistant to fatigue than P-4.34 Given that the other fabrication conditions of P-2 and P-4 are the same, the greater tensile stress in P-4 than in P-2 is attributed to the difference in diameter. This may offer a new route for the development of metal materials against hydrogen embrittlement from the structural design point of view.
![]() | ||
Fig. 12 Graphical plot of 2θ as a function of sin2![]() |
In situ XRD analysis indicates a faster phase transition process and shorter-term coexistence of two phases in Pd membranes with a lower diameter. On the other hand, the stress measurement by XRD method presents a higher tensile stress for Pd membranes with a higher diameter. This is ascribed to reduced internal stress and lattice strain gradients with the decrease of the radius in cylindrical structures and lower residual stress, which may provide a route for the suppression of hydrogen embrittlement of other metal materials.
Footnote |
† These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2021 |