This section presents the experimental results and their analysis, focusing on the creep behaviors and microstructural evolution of CZ1 Zr alloys under distinct stress levels (140 MPa and 260 MPa). The findings are organized into three key areas: pre-creep microstructure, creep curve analysis, and dynamic microstructural changes during creep.
3.2. Creep Behavior
Figure 2a,b illustrate the creep curves for the CZ1-1 and CZ1-2 alloys under low (140 MPa) and high (260 MPa) stress conditions at 380 °C, respectively. Under low-stress conditions (140 MPa), CZ1-1 accumulated a total creep strain of 5.69% after 1000 h, whereas CZ1-2 exhibited a significantly lower strain of 1.56%. Steady-state creep rates were determined from narrow linear segments of the ε–t curves (small steady-state windows), selected to minimize edge fluctuations associated with transient and tertiary creep stages, with linear fits yielding R
2 > 0.995. For CZ1-1, the steady-state window spans 200–500 h, during which strain increases from 1.85% to 4.47%, corresponding to a rate of 8.73 × 10
−5 h
−1 (~2.42 × 10
−8 s
−1). For CZ1-2, the steady-state window spans 250–550 h, with strain increasing from 0.45% to 1.13%, yielding a rate of 2.28 × 10
−5 h
−1 (~6.33 × 10
−9 s
−1). Both rates have an estimated uncertainty of ±10% due to specimen dimension measurements. These results indicate that CZ1-2, with its higher degree of recrystallization, exhibits superior resistance to low-stress creep, likely resulting from a refined microstructure and lower dislocation density. As shown in
Figure 2a, the creep curve for CZ1-2 increases more gradually, indicating its enhanced resistance to deformation under these conditions. In contrast, the CZ1-1 curve displays a steeper rise in strain, reflecting its higher sensitivity to deformation at low stress levels.
Under high-stress conditions (260 MPa), the trend reverses. CZ1-2 reaches a creep strain of 8% after just 9 h, whereas CZ1-1 requires 48.5 h to reach 9.53%. Steady-state creep rates were determined from narrow linear segments of the ε–t curves (small steady-state windows, R
2 > 0.995) to ensure precise rate evaluation. For CZ1-1, the steady-state window spans 5–25 h, during which strain increases from 2.15% to 7.63%, yielding a rate of 2.74 × 10
−3 h
−1 (~7.61 × 10
−7 s
−1). For CZ1-2, the window spans 1–4 h, with strain increasing from 1.52% to 5.39%, corresponding to a rate of 1.29 × 10
−2 h
−1 (~3.58 × 10
−6 s
−1). Both rates carry an estimated uncertainty of ±10%. These results indicate that CZ1-1 exhibits superior creep resistance under high stress, likely due to dislocation entanglement, Hall–Petch strengthening in non-recrystallized regions, and recovery-driven subgrain formation with localized low-angle boundary migration—consistent with its grain refinement trend (
Table 2: average grain size reduced to 0.53 ± 0.26 μm post-260 MPa creep). As depicted in
Figure 2b, the creep curve for CZ1-1 shows a faster initial increase in creep strain, indicating the onset of Stage III (accelerated creep). In contrast, CZ1-2 shows a sharp rise in strain early in the test, suggesting a rapid transition into the accelerated creep phase under high stress.
High-temperature creep in metallic materials is a thermally activated phenomenon. Assuming independent temperature and time effects on the steady-state creep rate (
), the relationship between steady-state creep rate, temperature (
), and stress (
) generally follows the power-law equation [
32]:
where
is a material constant,
is the apparent creep activation energy,
is the gas constant,
is the absolute temperature, and
n is the stress exponent. Taking the natural logarithm of both sides yields:
At constant temperature, let
(a constant). Equation (2) simplifies to:
This linear relationship between
and
at a fixed temperature allows for the determination of the stress exponent,
n, from the slope. Higher
n values indicate a greater stress sensitivity of the steady-state creep rate. The magnitude of
provides critical insights into the underlying creep mechanisms [
33,
34]: a value of approximately 1 is characteristic of diffusion-controlled mechanisms, while a value around 2 suggests grain boundary sliding. Dislocation glide is often associated with
n ≈ 3, and dislocation climb typically falls within the
n ≈ 4–6 range. A stress exponent exceeding 7 generally signifies power-law breakdown.
Figure 3 shows the relationship between ln
and ln
σ for the CZ1-1 and CZ1-2 alloys. As shown in the figure, CZ1-1 exhibits a stress exponent of
n ≈ 5 at 380 °C, which is consistent with a dislocation climb–dominated creep mechanism. In contrast, the fully recrystallized CZ1-2 alloy shows a distinct stress sensitivity: it maintains a power-law creep behavior (
n ≈ 5) under 140 MPa but transitions to a much higher stress exponent (
n ≈ 10) when the stress increases to 260 MPa. This apparent shift implies a change in the rate-controlling mechanism at high stress levels.
This interpretation is well supported by the canonical α-Zr creep mechanism map proposed by Hayes and Kassner [
14], which delineates the regime boundaries for homologous temperatures (T/Tm) around 0.3—closely matching our test condition of 380 °C (T/Tm ≈ 0.29 for Zr, where Tm = 1855 °C). According to their framework, the boundary between power-law creep (dominated by dislocation climb,
n = 4–6) and power-law breakdown (where conventional dislocation mechanisms lose rate control,
n > 7) occurs near ~200 MPa at this temperature range. The high-stress condition applied in this study (260 MPa) lies well above this boundary, providing a mechanistically consistent explanation for the observed
n ≈ 10.
Importantly, this mechanistic conclusion is not solely inferred from the stress exponent value. It is further corroborated by the microstructural evidence obtained for CZ1-2 under 260 MPa, including the rapid onset of tertiary creep (
Figure 2b), the absence of dynamic recovery, and the formation of localized dislocation networks. These features are characteristic of the power-law breakdown regime, where steady-state dislocation climb can no longer accommodate plastic strain. The integration of stress exponent analysis, established creep mechanism frameworks, and microstructural observations thus provides a robust basis for identifying power-law breakdown at high stress.
3.3. Microstructural Evolution During Creep
To quantitatively compare grain size variations before and after creep, statistical measurements were conducted on both CZ1-1 and CZ1-2 alloys. For each condition, at least 200 grains were measured to determine the average grain size, as summarized in
Table 2. For CZ1-1, the pre-creep sample exhibited a low degree of recrystallization, as shown in
Figure 1a, with only a few grains nucleating and growing into equiaxed grains. Therefore, the average grain size was not statistically evaluated for the pre-creep condition. The results show distinct trends between low-stress (140 MPa) and high-stress (260 MPa) creep conditions. Under low stress, both alloys exhibit grain coarsening after creep exposure, whereas high-stress creep leads to grain refinement.
Grain size quantification was performed using ImageJ software (version 1.53e; National Institutes of Health, Bethesda, MD, USA), a widely used image analysis tool for microstructural statistics. For TEM bright-field images (e.g.,
Figure 1), thresholding and segmentation were based on grayscale contrast: grain boundaries were identified by darker grayscale values (relative to grain interiors) and manually corrected to avoid missegmentation of overlapping grains or subgrain boundaries. At least 200 grains were sampled from 3 to 5 non-overlapping TEM regions per condition to ensure statistical representativeness.
Regarding 2D-to-3D bias: Due to the use of thin TEM foils (~30–50 μm thick,
Section 2.3), the measured grain sizes reflect 2D cross-sections rather than true 3D grain morphology. This may introduce minor underestimation of large, irregularly shaped grains or overestimation of small equiaxed grains—an inherent stereological limitation of thin-foil TEM analysis. To mitigate this bias, we ensured consistent foil thickness across all samples and avoided sampling regions with obvious foil bending, thereby minimizing variability in 2D cross-section representation.
This contrast is particularly evident in CZ1-1, where the average grain size decreases to 0.53 ± 0.26 μm after creep at 260 MPa (
Table 2)—attributed to recovery-driven subgrain formation and localized low-angle boundary migration, which is consistent with the observed dislocation walls—in contrast to grain growth observed under 140 MPa. A similar but less pronounced trend is observed in CZ1-2. These differences reflect the influence of applied stress on substructural evolution during creep. While both alloys exhibit stress-dependent grain size changes, the magnitude of refinement and coarsening varies between the two, with CZ1-1 showing greater sensitivity to stress-induced structural modification.
It is well established that fine, homogeneously distributed second-phase particles impede dislocation motion, thereby improving creep resistance [
35]. However, these particles tend to coarsen under prolonged creep, diminishing their strengthening effect and increasing the creep rate. To investigate the distribution and composition of second phases in the CZ1 alloys, energy-dispersive X-ray spectroscopy mapping was performed on CZ1-1 and CZ1-2 samples under both pre- and post-creep conditions as shown in
Figure 4,
Figure 5 and
Figure 6.
The pre-creep specimen (
Figure 4) exhibits a finely dispersed distribution of second-phase particles within the Zr alloy grains. In contrast, the post-creep specimens show significant coarsening of these particles with increasing creep exposure, indicating particle growth and redistribution during deformation.
As shown in
Figure 5, the EDS mapping of the CZ1-1 specimen after creep at 380 °C and 260 MPa reveals that the second-phase particles are substantially coarsened but remain primarily composed of Zr, Fe, Cr, and Nb, without any evident Cu enrichment. For comparison,
Figure 6 presents the microstructure of the CZ1-2 specimen after creep at 140 MPa, where localized Cu enrichment becomes evident in certain precipitates alongside Zr-Fe-Cr-Nb phases.
The detailed EDS quantification results for representative second-phase particles (
Figure 6) are summarized in
Table 3. The data indicate that, after creep at 140 MPa, CZ1-2 primarily contains Zr-Fe-Cr-Nb phases (~90%) with a minor fraction of Zr-Cu phases (~10%), reflecting the stress-dependent evolution of the precipitate chemistry during creep.
Zr alloys have a hexagonal close-packed (HCP) crystal structure, where the axial ratio (c/a) is less than 1.732. This favors dislocation glide on prismatic planes. <a> dislocations typically move on prismatic planes but may cross-slip to basal or pyramidal planes at elevated temperatures or under high stress. Pile-ups of <a> dislocations can block slip on prismatic planes, leading to the activation of <c + a> dislocations, which operate on pyramidal planes and contribute to c-axis strain [
36,
37].
To experimentally identify the types of dislocations present in the CZ1 alloy, TEM analyses were performed. The CZ1-1 alloy before creep was selected as an example. Under the [1
1
] zone axis, three two-beam conditions with diffraction vectors g = (1
01), g = (10
0), and g = (0
11) were applied, as shown in
Figure 7. Three types of dislocations, marked in orange, white, and blue, were observed under these conditions. Based on the comparison between the possible dislocation types in zirconium alloys and their invisibility behavior under different g vectors, the Burgers vectors of these dislocations could not be unambiguously determined from the above observations. Therefore, additional analyses were performed under the [01
2] zone axis using three distinct two-beam conditions with g = (2
0), g = (20
), and g = (01
), as shown in
Figure 8. According to the invisibility criteria, the blue dislocations were identified as b = <a>1/3[2
0], the orange dislocations as b = <a>1/3[
2
0], and the white dislocations as b = <c + a> type.
TEM analysis was further conducted on both pre- and post-creep specimens using similar two-beam conditions to identify dislocation types via contrast extinction techniques. To ensure the clarity and consistency of dislocation characterization, Miller–Bravais notation for crystal features is uniformly defined throughout this section: crystal planes (including diffraction vectors) are denoted with parentheses (hkil) (e.g., diffraction vectors g = (101), g = (20)), and crystal directions (including Burgers vectors and zone axes) with square brackets [uvtw] (e.g., zone axis [11], Burgers vector b = 1/3[20]).
For the two-beam TEM analysis (
Figure 7 and
Figure 8), the invisibility criterion (g·b = 0) is explicitly correlated with observed dislocation contrast to confirm Burgers vectors. Taking pre-creep CZ1-1 as an example: under the [01
2] zone axis, when the diffraction vector was set to g = (01
) (a crystal plane), the blue-marked dislocations exhibited contrast extinction. Calculations show g·b = (0) × (2/3) + (1) × (−1/3) + (−1) × (−1/3) + (−1) × (0) = 0 (where b = 1/3[2
0]), which satisfies the invisibility criterion. In contrast, when g = (2
0) was applied, g·b = (2) × (2/3) + (−1) × (−1/3) + (−1) × (−1/3) + 0 × 0 = 2 ≠ 0, and the blue dislocations remained visible. This direct match between the g·b = 0 criterion and contrast changes validates the accuracy of Burgers vector identification. The summarized results of dislocation characterization under various conditions are presented in
Table 4.
As can be seen from
Table 4, both CZ1-1 and CZ1-2 show <a> and <c + a> dislocations before creep. Post-creep, dislocation density increases significantly, especially under high-stress creep, due to accelerated dislocation generation and pile-up in response to applied stress.
The pure <c> dislocations identified in CZ1-2 after creep at 380 °C/260 MPa (
Table 4) are rarely documented in α-Zr, so their characterization was validated through multiple complementary steps using existing TEM data to ensure reliability. First, dislocations were observed and analyzed under two non-parallel zone axes—[1
1
] and [01
2] (
Figure 7 and
Figure 8)—which avoided misjudgment caused by single-orientation limitations and confirmed the consistency of dislocation features across different viewing angles. Meanwhile, dual validation criteria were applied: the invisibility criterion (g·b = 0) was used to confirm the Burgers vector, and the slip plane confirmation criterion (g·(b × u) = 0, where u denotes the dislocation line direction) was employed to rule out other dislocation types. For instance, when assuming a Burgers vector of b = [0001] (pure <c>) under the [01
2] zone axis, setting the diffraction vector to g = (0002) resulted in contrast extinction of the dislocations (satisfying g·b = 0); at the same time, g·(b × u) = 0 confirmed the dislocations resided on the basal plane, excluding <c + a> dislocations that require pyramidal slip planes. Additionally, the literature support further corroborates this identification: high stress (>200 MPa) has been reported to activate non-conventional slip systems in α-Zr, such as basal <c> slip [
36], providing a mechanistic basis for the occurrence of pure <c> dislocations in this study. Collectively, these multi-faceted validations confirm that the identification of pure <c> dislocations in CZ1-2 under 260 MPa is not spurious, but rather a reliable reflection of the alloy’s deformation behavior under high stress.
In
Figure 9a,b, second-phase particles are predominantly observed along grain boundaries, where their spatial distribution appears non-uniform. These particles are frequently located at triple junctions and high-angle boundaries, consistent with typical precipitation behavior at high-energy sites. Additionally, some dislocations are seen to terminate at or interact with these particles, forming localized pinning points. In certain regions, dislocation lines are arrested or deflected near particle interfaces, suggesting a structural constraint on dislocation motion. The dislocation lines are relatively sparse and appear to align along specific crystallographic directions. No significant dislocation entanglement or wall formation is observed, and the dislocation motion appears to proceed in an orderly fashion. These features indicate that under low-stress conditions, dislocation activity remains relatively unobstructed within the grain interior, with limited evidence of complex interactions or storage.
Figure 10a,c reveals the presence of dislocation networks within the grains of CZ1-2 alloy after creep at 260 MPa. These networks are composed of intersecting and entangled dislocations, forming locally dense structures. Compared to the low-stress condition, where dislocation lines appeared more isolated and directional, the formation of such networks suggests an increase in dislocation interactions under elevated stress. In
Figure 10b,d, parallel planar features are observed within the matrix, which are identified as stacking faults. These features appear as local interruptions of the lattice contrast, extending across parts of grains or intersecting with dislocation lines. Their presence is more pronounced under higher-stress conditions and may indicate a shift in the deformation mechanism, such as increased dislocation activity or planar slip. Regarding these stacking faults (i.e., extended stacking faults, ESFs) observed in CZ1-2 (
Figure 10b,d), their formation can be rationalized by the stacking fault energy (SFE) of the CZ1 alloy, supported by literature data. For Zr-Sn-Cu-Nb alloys with compositions similar to CZ1 (
Section 2.1), Lin et al. [
38] reported an SFE of ~15–20 mJ/m
2—significantly lower than that of pure α-Zr (~40 mJ/m
2). Low SFE promotes the dissociation of full dislocations into partial dislocations, which in turn generates the planar ESF contrasts captured in the TEM images. While explicit identification of fault vectors (i.e., I1 and I2, the two typical partial dislocation vectors corresponding to intrinsic and extrinsic stacking faults in hexagonal metals) was not conducted, the presence of ESFs and their correlation with low SFE align with established deformation behavior in Sn/Cu-doped Zr alloys [
38], reinforcing the mechanistic interpretation of creep deformation. No evidence of twinning was observed in the present samples.
The observed microstructural features in
Figure 10 differ markedly from those presented under lower stress, highlighting the evolution of dislocation structures with increased loading. The presence of both dislocation networks and stacking faults implies a shift in the deformation behavior, the nature of which will be further examined in the following chapter.
As shown in
Figure 11b,c, the dislocation structures in the CZ1-1 alloy after creep at 140 MPa exhibit kinked and step-like segments deviating from their original glide traces. These non-planar dislocation configurations, particularly at the interaction zones with second-phase particles or local barriers, suggest that part of the dislocation motion occurred out of the primary slip plane. Similar non-planar, segmented dislocation morphologies have been interpreted as evidence of vacancy-assisted climb in high-temperature creep conditions [
39]. In the present case, such features likely reflect local climb-assisted bypass of obstacles, facilitating strain accommodation and stress relaxation within dislocation-dense regions (
Figure 11a).
Figure 11d shows dislocation walls near grain boundaries, where dislocations of similar character align into low-angle planar arrays. These walls interrupt otherwise homogeneous dislocation distributions and appear more frequently in CZ1-1 than in the corresponding CZ1-2 alloy under the same stress condition. The coexistence of climb-related dislocation features and organized dislocation walls indicates a distinct structural evolution pathway during low-stress creep, involving local climb-assisted rearrangement and the progressive formation of sub-grain boundaries.
As shown in
Figure 12a, the microstructure of CZ1-1 alloy after creep at 260 MPa is characterized by the formation of dense dislocation networks and stacking faults within the grain interiors. The dislocation networks appear more entangled and spatially interconnected compared to those observed in CZ1-2 under the same conditions. These networks often span large areas and exhibit frequent intersections and junctions, reflecting a higher degree of dislocation accumulation (
Figure 12c).
Stacking faults are also visible in
Figure 12b,d as narrow, planar contrast features extending across grains. These defects are more pronounced and frequent in CZ1-1 than in CZ1-2, where their appearance is comparatively sparse. The combined presence of intricate dislocation networks and well-defined stacking faults distinguishes CZ1-1 from CZ1-2 in terms of microstructural response to high-stress deformation. While both alloys develop similar types of defects under elevated stress, the internal dislocation configuration in CZ1-1 reveals a more complex spatial arrangement and higher structural heterogeneity.