Introduction

Cr2AlC belongs to the Mn+1AXn phases, where M is a transition metal, A an A group element and X represents either C or N, a class of nanolaminates which have attracted considerable attention within the last decade due to their unusual combination of properties1,2. These characteristics are caused by the alternation of layers with metallic bonds in-between M and A elements and covalent/ionic bonds in the case of M-X layers2,3,4. Thus, MAX phases exhibit ceramic properties as for instance high stiffness5,6 as well as typically metallic features such as good machinability3 and high thermal and electrical conductivity5,7,8,9,10. MAX phases are compensated conductors exhibiting a metal-like behavior of the electrical resistivity ρ upon heating2,7. Density functional theory calculations show that the density of states at the Fermi level and thus the conductivity is dominated by the d states of the transition metal M11. The MAX phase Cr2AlC additionally exhibits excellent oxidation resistance due to the formation of a dense alumina scale12,13,14,15. The electrical resistivity of bulk Cr2AlC MAX phases has been studied at temperatures below room temperature2,5,8 as well as up to 900 °C9,10. Reported values of ρ at room temperature for Cr2AlC MAX phase range from ρ = 0.60 to 0.74 μΩ m2,5,8,9,10. Furthermore, it was shown that Cr2AlC is a self-healing material as cracks can be filled by selective oxidation of aluminum16,17,18,19,20. Thus, Cr2AlC may be of interest e.g. for nuclear applications21,22, heat exchangers23 or aero-engines20.

X-ray amorphous Cr2AlC powder samples synthesized by physical vapor deposition (PVD) have been analyzed by differential scanning calorimetry (DSC) by Walter et al.24 and Abdulkadhim et al.25. Based on correlative DSC and ex-situ X-ray powder diffraction (XRD) measurements Abdulkadhim et al.25 identified the formation of Cr2AlC MAX phase at 610 °C. Additionally, the presence of the disordered solid solution (Cr,Al)2Cx was observed at 560 °C. Thin film synthesis of this phase by sputter deposition was previously reported at substrate temperatures of 300 °C by Shtansky et al.26. Ex-situ annealing of the sample at 800 °C resulted in the formation of Cr2AlC MAX phase26. (Cr,Al)2Cx is structurally similar to the Cr2AlC MAX phase, which was suggested to consist of three perfectly ordered (Cr,Al)2Cx unit cells with a stacking sequence of Cr-Cr-Al-Cr-Cr-Al exhibiting a non-metal sub lattice order of C-vac-vac-C-vac-vac25.

It is the ambition of this work to detect phase changes by in-situ resistivity measurements during heat treatment. To this end, amorphous Cr-Al-C thin films are annealed to various temperatures up to 800 °C in a vacuum furnace, while in-situ resistivity measurements are performed. Ex-situ structural analysis reveals the formation of the disordered solid solution between 540 °C and 560 °C and a subsequent phase change to Cr2AlC MAX phase in the temperature range from 580 °C to 600 °C. Here it is shown that these structural transitions result in characteristic resistivity changes which enable remote tracking of phase changes by in-situ resistivity measurements.

Methods

Synthesis

Cr2AlC thin films were synthesized by direct current magnetron sputtering in an industrial chamber (CC800/8, CemeCon AG, Wuerselen, Germany). A Cr:Al:C compound target exhibiting a stoichiometry of 2:1:1 was employed (provided by PLANSEE Composite Materials GmbH, Lechbruck am See, Germany). The base pressure was below 0.1 mPa and the argon pressure during the deposition was set to 190 mPa. The target power density was 2.3 W cm−2. 10 × 10 × 0.5 mm single crystalline MgO substrates (Crystal GmbH, Berlin, Germany) were located at a distance of 5 cm to the plasma source and were kept at floating potential. No substrate heating was applied during deposition.

Thin films for in-situ heating transmission electron microscopy (TEM) measurements were deposited with the same deposition parameters on polycrystalline NaCl substrates. The deposition time was adjusted to grow an approx. 85 nm thick thin film. The as deposited film was separated from the NaCl substrate by dissolving in distilled water. The obtained film flakes were cleaned in distilled water, isopropanol and acetone prior to TEM analysis.

Characterization

Electrical resistivity measurements during annealing experiments were performed in a vertical tube high vacuum furnace. 14 samples were heated to various temperatures in between 400 and 800 °C. No holding time was applied and a heating and cooling rate of 5 K min−1 was used. The base pressure was below 3∙10−6 mbar. The resistivity was measured in-situ employing a Van-Der-Pauw setup including a Keithley 2611B System SourceMeter using a current of 5 mA.

For determination of the chemical composition an as deposited sample was characterized by time-of-flight elastic recoil detection analysis (TOF-ERDA) and elastic backscattering spectroscopy (EBS) at the tandem accelerator laboratory of Uppsala University. For TOF-ERDA, 127I8+ projectiles with a primary energy of 36 MeV were employed. Further details can be found in reference27 and its supplements. EBS was carried out using 4.5 MeV 4He+ ions, a detection angle of 170° and thus employing the strong 12C(4He,4He)12C elastic resonance at ~4.260 MeV28.

Structural analysis by XRD was carried out in a Siemens D5000 system equipped with a Cu radiation source. The 2θ range from 10 to 90° was scanned in unlocked coupled setup with a tilt angle of 2° with respect to the substrate surface to avoid substrate peaks. The step size was set to 0.02° at a scan time of 10 s per step. Further structural characterization of the thin films was carried out using a Tecnai F20 TEM operated at a voltage of 200 kV. Cross-sectional TEM foils were prepared using a standard lift-out method in a DualBeam focused ion beam (FIB) system (Helios NanoLab 660). A thin layer of platinum was deposited on the film surface for protection against the beam damage.

Flakes obtained from NaCl substrates were positioned on a Nano-Chip (DENSsolutions). The in-situ heating experiment was conducted with a double tilt heating holder from DENSsolutions in a JEOL JSM 2200F field emission gun TEM. The selected area electron diffraction (SAED) patterns were obtained using an acceleration voltage of 200 kV. The sample was heated to a nominal temperature of 700 °C at a heating rate of 5 K min−1. Temperatures were obtained from the software, which is used to read out the chip. For a similar chip Niekiel et al. measured deviations between experimentally determined temperatures and the read out value of the software depending on the position of the sample on the chip29. The position of the window employed in this work corresponds to a temperature which is approx. 13 °C lower than the read out value. All temperatures obtained from the in-situ TEM heating experiments are corrected for this deviation.

Results and Discussion

The diffractogram of the as deposited sample, see Fig. 1a, exhibits a wide hump around 42° indicating an X-ray amorphous thin film. The chemical composition of the as deposited thin film was measured by TOF-ERDA and EBS to be Cr: 49.9 ± 2.5 at.%, Al: 24.9 ± 1.2 at.%, C: 24.7 ± 1.2 at.% and O: 0.5 ± 0.3 at.%. Thus, the film is stoichiometric. The oxygen contamination is expected to stem from residual gas incorporation during the deposition process30. The venting temperature was lower than 50 °C to minimize modification of the surface composition by air exposure31.

Figure 1
figure 1

In-situ electrical resistivity measurements and correlative X-ray diffraction data (a) Ex-situ measured X-ray diffractograms of Cr2AlC samples annealed to various temperatures up to 800 °C indicated by dotted lines; (b) resistivity measured in-situ while heating (black) compared to DSC results by Abdulkadhim et al. (dashed)25. Red and green highlighted areas indicate average values and standard deviation of the onset of characteristic changes of resistivity for all measured samples; (c) lattice parameters determined from ex-situ XRD measurements. For better comparability of the c parameter the height of 3 stacked unit cells of (Cr,Al)2Cx is compared to a single unit cell of the MAX phase.

The average film thickness was measured on 4 samples to be 3.8 ± 0.2 μm. Resistivity measurements at room temperature yielded an average of 2.15 ± 0.13 μΩ m for as deposited X-ray amorphous samples.

Figure 1b shows the result of one in-situ resistivity measurement during annealing from 500 °C to 800 °C. Up to a temperature of 550 °C, a negative temperature coefficient of resistance (TCR) αas dep, 300K of −5.4 × 10−5 ± 0.7 × 10−5 K−1 was observed. This observation is in accordance with Mooij’s correlation predicting negative TCRs for amorphous transition metals with ρ(0) > 1.50 μΩ m and can be rationalized based on the mean free paths of electrons in the order of interatomic distances32,33,34.

At a temperature of 552 ± 4 °C a pronounced change in resistivity of 5.0 ± 0.7% within 20 °C was observed. As the temperature is increased to 585 ± 13 °C a second significant resistivity change of 3.0 ± 1.7% within 20 °C (17.0 ± 2.1% within 60 °C) is measured. The above given temperatures are averaged values and the corresponding temperature ranges represent the standard deviations stemming from 14 and 9 individual measurements for the first and second pronounced change of resistivity, respectively. For this purpose, the inflection point of the resistivity curve is employed as an indicator for the onset of the second pronounced change in resistivity. Above 700 °C the sample exhibited a positive TCR indicating a metal-like behavior of the temperature dependency as expected for MAX phases2,10,35.

A similar trend was observed for the ex-situ resistivity measured at room temperature after annealing at temperatures of 500, 540, 560, 580, 600 and 800 °C as shown in Fig. 2. For an annealing temperature of ≤540 °C the measured resistivity at room temperature was reduced by less than 2% compared to the initial resistivity at room temperature. However, annealing to 560 °C resulted in a permanent change of resistivity by −18.5%, while heating a sample to 640 to 800 °C resulted in a resistivity of less than one third of the initial resistivity. Very good agreement between the temperature induced in-situ and ex-situ resistivity changes was obtained.

Figure 2
figure 2

Microstructure and ex-situ SAED measurements compared to temperature dependence of resistivity (af): Bright field images and SAED patterns of FIB lamellae from Cr2AlC thin films after annealing at 500, 540, 560, 580, 600 and 800 °C in vacuum; Center: In-situ measured resistivity (blue) and ex-situ resistivity at room temperature (black) after annealing to various temperatures. The ex-situ resistivity data is normalized with respect to the initial resistivity at room temperature.

Structural changes due to annealing at different temperatures were probed by ex-situ XRD and are shown in Fig. 1a. A comparison of diffractograms of an as deposited thin film and one annealed to 540 °C does not reveal structural changes due to the heat treatment. Both diffractograms indicate the presence of X-ray amorphous thin films. Crystallization was observed for an annealing temperature of 560 °C which is consistent with Grieseler et al. reporting crystalline thin films after annealing Cr-Al-C multilayer systems at 550 °C36. At temperatures of 600 °C and higher an additional peak at 13.78° appears. At 560 to 580 °C a disordered solid solution (Cr,Al)2Cx is formed, which is consistent with Abdulkadhim et al.25,26. (Cr,Al)2Cx is structurally similar to the Cr2AlC MAX phase, which was suggested to consist of three perfectly ordered (Cr,Al)2Cx unit cells25. Due to the structural similarity between (Cr,Al)2Cx and Cr2AlC a positive phase identification of the MAX phase can be challenging. However, the (002) peak of Cr2AlC at 13.8° as well as the (101) peak at 36.90° are distinct indicators for MAX phase formation. Upon phase transformation the (006) peak of Cr2AlC MAX phase, which is structurally related to the (002) peak of (Cr,Al)2Cx at 41.35°, shifts to higher angles past the (103) MAX phase peak at 42.16°. Thus, samples annealed at 600 °C to 800 °C were identified as Cr2AlC MAX phase based on the presence of the (002), (006) and (101) peaks. Due to the peak overlap for both phases an unambiguous appraisal of the phase purity of the obtained MAX phase thin films based on diffraction data alone is not feasible.

With increasing annealing temperature, the lattice parameters change, see Fig. 1c. For a better comparability in-between MAX phase and disordered solid solution three stacked unit cells of (Cr,Al)2Cx are considered here. Thus, the actual c parameter of (Cr,Al)2Cx is one third of the here employed value. While the a lattice parameter increases with higher annealing temperatures from 2.831 to 2.865 Å, the c lattice parameter decreases from 13.105 to 12.826 Å, for temperatures of 560 °C and 800 °C, respectively. For the phase transition from disordered solid solution to the MAX phase between 580 and 600 °C relative changes in the lattice parameters of 0.90 and −1.66% are observed for a and c, respectively. Lattice parameters of both phases are in very good agreement with previously reported values25. However, both lattice parameters vary within the temperature region of the MAX phase, which may indicate higher crystal quality.

The XRD results are in agreement with ex-situ SAED measurements on lamellae extracted from annealed thin films by FIB as shown in Fig. 2. For the thin film which was annealed to 500 °C the SAED pattern exhibits only a diffuse ring indicating the presence of an amorphous structure and bright field images do not exhibit any features. However, small crystallites are already visible in the bright field image for the sample annealed to 540 °C. Due to the crystallite size and the small volume fraction of the crystalline material XRD is insensitive to this phase formation. Annealing to 560 °C induced a fully crystalline sample with elongated grains. However, only after annealing to 600 °C the (002) basal plane of the MAX phase structure was detected.

Both ex-situ diffraction techniques were applied to samples after heating and cooling. Therefore, compared to the in-situ experiment the annealing process was prolonged by the cool down procedure. This procedure may lead to further crystallization of the samples prior to analysis shifting the necessary temperature for observations to lower values. In an effort to narrow down the phase transition temperature range, in-situ heating TEM measurements were conducted. Representative SAED patterns are shown in Fig. 3. Up to a temperature of 564 °C (Fig. 3a) the SAED data is consistent with XRD measurements indicating the presence of amorphous Cr2AlC. At 567 °C a diffraction signal belonging to the (101) plane of (Cr,Al)2Cx was detected (Fig. 3b), whereas the first diffraction signal stemming from the (002) plane of the MAX phase was identified at 594 °C (Fig. 3d). With increasing temperature, the intensity of the diffraction signals stemming from crystalline material increased as the fraction of amorphous material was reduced (Fig. 3e). The phase transformation from (Cr,Al)2Cx to Cr2AlC MAX phase at 591 to 594 °C narrows down the transition temperature range of 580 to 600 °C obtained by ex-situ XRD and SAED analysis. However, the reported temperatures for the amorphous to (Cr,Al)2Cx transition at 564 to 567 °C are slightly above the temperature range from 540 to 560 °C identified by ex-situ analysis. This difference may on the one hand be caused by uncertainties associated with the temperature calibration as our appraisal is based on the assumption that the chip employed in this study shows identical behavior as the chip employed by Niekiel et al.29. On the other hand, the extended annealing time by the cooling procedure for the ex-situ analyzed samples and the smaller analyzed sample volume may also have influenced the observed transition temperatures. These results are supported by in-situ SAED measurements performed on a lamella extracted by FIB milling from an as deposited amorphous Cr-Al-C thin film which yield comparable results (not shown here).

Figure 3
figure 3

In-situ SAED measurements compared to temperature dependence of resistivity (ae) SAED patterns obtained from in-situ heating TEM measurements at indicated temperatures; Bottom: In-situ measured resistivity. Red and green highlighted areas indicate average values and standard deviation of the onset of characteristic changes of resistivity for all measured samples. Highlighted diffraction signals are initial diffraction signals linked to (Cr,Al)2Cx disordered solid solution (red) and Cr2AlC MAX phase (green).

Abdulkadhim et al. performed DSC measurements on amorphous Cr-Al-C powder at a heating rate of 10 K min−1 and observed an endothermic reaction corresponding to the presence of (Cr,Al)2Cx at 560 °C25, see Fig. 1b. They identified a second peak in the DSC signal at 610 °C and linked it to the phase formation of Cr2AlC MAX phase25. Further analysis of these DSC data reveals that the onset temperatures of the phase transformations from amorphous to the disordered solid solution (Cr,Al)2Cx as well as from disordered solid solution (Cr,Al)2Cx to Cr2AlC MAX phase occur at 560 °C and 585 °C, respectively.

The morphology of samples annealed to 500, 540, 560, 580, 600 and 800 °C were analyzed by ex-situ bright field images of prepared FIB lamellas depicted in Fig. 2. All three thin films exhibit dense microstructures. As shown above, the sample annealed to 500 °C is X-ray amorphous which agrees with the featureless homogeneous cross section with no indication of crystallization. For the samples annealed to 560 °C and above crystallization was observed. The sample heated up to 560 and 580 °C, both identified by SAED as (Cr,Al)2Cx, exhibited randomly oriented elongated grains (Fig. 2c,d). After the transition to the MAX phase by annealing to 600 °C a trend towards larger and equiaxed grains was observed. Further annealing to 800 °C leads to further grain growth. This observation may explain the decrease in measured ex-situ resistivity for the MAX phase samples annealed to temperatures ≥600 °C (Fig. 2). The electrical resistivity at 300 K after annealing to 680, 720 and 800 °C was measured to be 0.78, 0.57 and 0.56 μΩ m, respectively. Reported values for the Cr2AlC MAX phase are ρ(300 K) = 0.60 μΩ m by Ying et al.10, ρ(300 K) = 0.74 μΩ m by Hettinger et al.2,5, ρ(RT) = 0.63 μΩ m by Zhou et al.9 and ρ(RT) = 0.71 μΩ m by Tian et al.8. Thus, the values measured within this work range from the upper to the lower end of the known resistivity range with lower values at higher annealing temperatures. As the temperature is increased further after MAX phase formation the crystal quality improves by coarsening as indicated by diffraction peaks with larger intensity and decreased full width at half-maximum (Fig. 1a) as well as the larger grains observable in the bright field images (Fig. 2). The improvement in crystal quality enables lower resistivity values due to a reduced defect density. Thus, it is reasonable to assume that the above discussed resistivity range reported in literature2,5,8,9,10 is caused by a synthesis induced variation in crystal quality.

Comparing the phase transition temperature ranges obtained by XRD, SAED and DSC25 with the measured temperature dependent resistivity signal reveals that the resistivity changes are characteristic for the here observed phase transformations. Ex-situ SAED data indicate that the onset of crystallization of (Cr,Al)2Cx occurs below 540 °C while at 560 °C a fully crystalline structure is observed. Thus, the formation of a fully crystalline sample correlates with the first pronounced resistivity decrease at 552 ± 4 °C. The second pronounced decrease in resistivity measured at 585 ± 13 °C indicates the phase transition from (Cr,Al)2Cx to Cr2AlC MAX phase, which was shown to take place in the temperature range from 580 to 600 °C by both ex-situ diffraction techniques, XRD and SAED. Results on powder samples analyzed by in-situ SAED experiments indicate phase transitions at temperatures ≤15 °C above the average temperatures determined by in-situ resistivity measurements. DSC measurements on powder samples25 are in very good agreement with transition temperature ranges obtained by in-situ resistivity measurements for both phase formations. Hence, the correlation of measured structural changes with the measured resistivity changes reveals that changes in resistivity are characteristic for the here observed phase transformations.

Based on the correlation of measured structural changes with the measured resistivity changes it is evident that changes in resistivity are characteristic for the here observed phase transformations in Cr2AlC. Thus, measuring resistivity is proposed as a powerful yet technically comparatively simple tool to track the onset and progress of phase transitions without destructive material characterization. It is conceivable that this method can be employed to monitor structural changes during application. E.g. it would allow the estimation of amorphization due to irradiation in materials employed in nuclear applications. The here communicated research strategy for tracking phase changes may be utilized in fundamental research as well as in technological applications where phase changes are expected due to exposure to harsh environments, such as nuclear reactors.

Conclusions

Annealing experiments of magnetron sputtered amorphous Cr2AlC samples were performed in vacuum at temperatures up to 800 °C. In-situ resistivity measurements revealed two characteristic changes of resistivity which were observed at 552 ± 4 °C and 585 ± 13 °C. These are in excellent agreement with DSC measurement by Abdulkadhim et al.25 and correlate with phase changes from amorphous to hexagonal (Cr,Al)2CX as well as (Cr,Al)2CX to Cr2AlC MAX phase observed by XRD and ex-situ as well as in-situ SAED. The results clearly reveal that phase changes in Cr2AlC thin films can be tracked by non-destructive resistivity measurements. These findings are relevant for other materials systems provided that the different phases exhibit differences in resistivity.