Submitted:
20 December 2024
Posted:
24 December 2024
Read the latest preprint version here
Abstract
Keywords:
| Contents | ||
| 1 | Introduction | 2 |
| 2 | Mathematical Formulation of Creep Mechanisms | 3 |
| 2.1 Diffusion-Controlled Creep............................................................................................................................ | 3 | |
| 2.1.1 Nabarro-Herring Creep (Lattice Diffusion)............................................................................................... | 3 | |
| 2.1.2 Coble Creep (Grain Boundary Diffusion).................................................................................................. | 5 | |
| 2.2 Dislocation Creep..................................................................................................................................... | 7 | |
| 2.3 Grain Boundary Sliding................................................................................................................................ | 11 | |
| 2.3.1 Nanomaterials........................................................................................................................... | 15 | |
| 2.3.2 Application to High-Strength Steel and Tungsten filaments............................................................................... | 15 | |
| 2.3.3 Superplastic forming (SPF) technique.................................................................................................... | 17 | |
| 2.4 Solute Drag Creep..................................................................................................................................... | 19 | |
| 2.4.1 Portevin–Le Chatelier effect.......................................................................................................... | 21 | |
| 2.5 Dislocation Climb-Glide Creep......................................................................................................................... | 22 | |
| 2.6 Harper-Dorn Creep..................................................................................................................................... | 24 | |
| 2.7 Sequential and Parallel Process in Creep.............................................................................................................. | 26 | |
| 3 | Phenomenological Description of Creep | 27 |
| 3.1 Stages of Creep....................................................................................................................................... | 27 | |
| 3.1.1 Primary Creep........................................................................................................................... | 27 | |
| 3.1.2 Secondary Creep......................................................................................................................... | 30 | |
| 3.1.3 Tertiary Creep.......................................................................................................................... | 33 | |
| 3.2 Constitutive Models................................................................................................................................... | 36 | |
| 3.2.1 Norton-Bailey Law....................................................................................................................... | 36 | |
| 3.2.2 Time-Hardening Model.................................................................................................................... | 38 | |
| 3.2.3 Strain-Hardening Model.................................................................................................................. | 40 | |
| 4 | Sintering | 46 |
| 5 | Examples of Creep Deformation | 48 |
| 5.1 Surface materials..................................................................................................................................... | 48 | |
| 5.2 Metals................................................................................................................................................ | 49 | |
| 5.3 Polymers.............................................................................................................................................. | 50 | |
| 5.4 Glass................................................................................................................................................. | 51 | |
| 5.5 Concrete.............................................................................................................................................. | 52 | |
| 5.6 Wood.................................................................................................................................................. | 52 | |
| 6 | Larson-Miller Parameter and Monkman Grant Rule | 53 |
| 7 | Multi-Axial Creep and Damage Mechanics | 55 |
| 8 | Creep Testing | 57 |
| 9 | Environmental Effects on Creep: Oxidation and Irradiation | 59 |
| 10 | Increasing Creep Resistance | 61 |
| 10.1 Working Conditions.................................................................................................................................... | 62 | |
| 10.1.1 Time.................................................................................................................................... | 64 | |
| 10.1.2 Temperature............................................................................................................................. | 64 | |
| 10.1.3 Applied stress.......................................................................................................................... | 65 | |
| 10.2 Material Selection.................................................................................................................................... | 66 | |
| 11 | Creep Prevention in Superalloys | 69 |
| 12 | Famous Accidents | 72 |
| 12.1 Big Dig tunnel ceiling collapse in Boston on July 2006................................................................................................ | 72 | |
| 12.2 Collapse of the World Trade Center on September 11, 2001.............................................................................................. | 72 | |
| 13 | Conclusions | 73 |
| 14 | Acknowledgments | 73 |
| References | 73 | |
1. Introduction
2. Mathematical Formulation of Creep Mechanisms
2.1. Diffusion-Controlled Creep
- Nabarro-Herring Creep (Lattice Diffusion)
- Coble Creep (Grain Boundary Diffusion)
2.1.1. Nabarro-Herring Creep (Lattice Diffusion)
| Mechanism | Diffusion Path | Strain Rate Dependence | Grain Size Scaling |
|---|---|---|---|
| Nabarro-Herring | Lattice diffusion | ||
| Coble Creep | Grain boundary |
2.1.2. Coble Creep (Grain Boundary Diffusion)
2.2. Dislocation Creep
- Stress Sensitivity: The cubic dependence of on highlights the nonlinear impact of stress on dislocation creep. Stress sensitivity in dislocation creep refers to the mathematical and physical relationship between the applied differential stress, , and the resulting steady-state strain rate, , in materials undergoing deformation at high temperatures and low to moderate stresses, where dislocation motion is the dominant mechanism. This relationship is described by a power-law equation, where A encapsulates material-specific constants such as dislocation density and grain size, Q is the activation energy required for creep processes, R is the universal gas constant, T is the absolute temperature, and n is the stress exponent, which characterizes the stress sensitivity of the deformation process. The term signifies the nonlinear dependence of the strain rate on stress, with the stress exponent n reflecting the dominant dislocation mechanisms. A smaller n, typically near 3, indicates that dislocation creep is governed primarily by climb processes, where the rate-limiting step involves vacancy diffusion that enables dislocations to bypass obstacles. In contrast, larger values of n, often exceeding 5, suggest that glide-controlled mechanisms dominate, where dislocations overcome stronger obstacles such as precipitates or other dislocations, requiring higher stresses to sustain significant deformation. The stress exponent can be rigorously derived from experimental data by examining the logarithmic relationship between and , specifically by differentiating with respect to , yielding . This procedure highlights the sensitivity of the strain rate to variations in stress, with a higher n signifying a more pronounced influence of stress on deformation kinetics. The exponential term further incorporates the thermal activation of atomic processes, underscoring the role of temperature in facilitating dislocation motion through enhanced diffusion. Stress sensitivity thus provides profound insights into the interplay of stress, temperature, and microstructural factors, enabling predictive modeling of material behavior under high-temperature conditions and informing the design of alloys and geological models where dislocation creep mechanisms predominate.
-
Thermal Effects: The exponential temperature dependence underscores the dominance of thermally activated mechanisms in governing creep behavior. Thermal effects in dislocation creep refer to the profound influence of temperature on the mechanisms that govern the movement of dislocations and, consequently, the rate of creep deformation. Dislocation creep is a thermally activated process, meaning that an increase in temperature facilitates the atomic and microstructural processes required for dislocation motion. These processes are captured mathematically in the creep equation where is the steady-state strain rate, A is a material-specific constant, is the applied differential stress, n is the stress exponent, Q is the activation energy for the creep process, R is the universal gas constant, and T is the absolute temperature. The exponential term explicitly describes the thermal activation of the creep process. At low temperatures, the exponential factor significantly suppresses the strain rate because the energy available thermally is insufficient to overcome the barriers associated with dislocation motion, such as lattice resistance, pinning at obstacles, or vacancy formation. Conversely, as temperature increases, the thermal energy becomes comparable to or exceeds the activation energy Q, facilitating dislocation climb and glide.Dislocation climb, in particular, is a diffusion-controlled process, where atoms or vacancies must migrate to or away from dislocations for them to bypass obstacles. The rate of atomic diffusion is strongly temperature-dependent and typically follows an Arrhenius relationship, which is directly reflected in the term. The activation energy Q is a critical parameter that depends on the dominant diffusion mechanism in the material. For instance, in pure metals, Q is often associated with self-diffusion, whereas in alloys or complex materials, it may reflect vacancy diffusion through solute-rich regions. The temperature dependency is so significant that even slight changes in T can result in orders-of-magnitude variations in . Moreover, at very high temperatures approaching the melting point of a material, the mobility of dislocations becomes exceptionally high due to enhanced diffusion, leading to faster creep rates. However, if the temperature becomes excessively high, other deformation mechanisms, such as diffusional creep or grain boundary sliding, may become more dominant than dislocation creep. In summary, thermal effects in dislocation creep control the rate at which dislocations overcome obstacles via thermally activated mechanisms such as climb and glide. The temperature dependence, encapsulated in the Arrhenius-type term , highlights the critical role of thermal energy in facilitating the microstructural processes that drive dislocation motion. Understanding these effects is essential for predicting and mitigating high-temperature deformation in engineering materials.
-
Implications for Material Design: The implications of dislocation creep for material design are fundamentally rooted in the quantitative relationship between strain rate, applied stress, and temperature, encapsulated in the constitutive equation . This equation reveals that the steady-state strain rate, , is exponentially sensitive to the activation energy Q, inversely proportional to temperature T, and nonlinearly dependent on the applied stress through the stress exponent n. The precise interplay of these parameters highlights critical pathways for designing materials that resist dislocation creep under high-temperature and high-stress conditions. One of the most significant factors is the activation energy Q, which represents the energy barrier for thermally activated dislocation mechanisms such as climb and glide. Increasing Q through microstructural engineering—such as the introduction of solute atoms, precipitates, or dispersoids—creates barriers that hinder dislocation mobility, thereby reducing the strain rate exponentially. These obstacles can also influence the stress dependence characterized by n, with materials designed to promote a higher effective stress exponent being more resistant to small increments in applied stress.The pre-exponential factor A and its relation to the material’s microstructure further illustrate the intricacy of design strategies. This parameter reflects the density and mobility of dislocations, which are heavily influenced by grain size, dislocation substructures, and precipitate distribution. Grain size, in particular, governs the Hall-Petch effect, where finer grains reduce dislocation mobility through grain boundary strengthening. However, at high temperatures, fine grains may encourage grain boundary sliding or diffusional creep mechanisms, necessitating an optimal grain size tailored to the specific temperature and stress regime. Precipitates, often introduced through precipitation hardening, provide long-range elastic fields that impede dislocation motion. The coherency, size, and spacing of these precipitates are critical; too small or incoherent precipitates may lose effectiveness at elevated temperatures, while excessive precipitate coarsening can degrade mechanical properties over time.Another central consideration in material design is the effect of solid solution strengthening, where alloying elements create local lattice distortions that interact with dislocations, increasing the stress required for motion. The solute-dislocation interaction becomes particularly significant for dislocation climb, a diffusion-controlled process essential for bypassing obstacles. This interaction underscores the importance of the diffusivity term implicit in the exponential factor , where Q often correlates with the self-diffusion or vacancy diffusion energy of the host lattice or secondary phases. Furthermore, the temperature-dependent exponential term underscores the dominance of thermal stability. Materials that retain their microstructural integrity, such as stable precipitates or low-diffusivity phases, exhibit superior performance at elevated temperatures. For example, nickel-based superalloys, employed in turbine blades, achieve remarkable creep resistance through the optimization of Q and A via careful alloying and processing to stabilize the -phase precipitates and suppress diffusional degradation.Additionally, environmental considerations, such as oxidation or corrosion, interact synergistically with creep mechanisms, exacerbating deformation through surface or grain boundary degradation. The application of protective coatings or the incorporation of oxidation-resistant alloying elements can mitigate these effects. In totality, the design of creep-resistant materials involves a multi-faceted optimization of the parameters A, Q, n, and T through advanced microstructural engineering, alloy development, and environmental control. This rigorous approach, grounded in the mathematical formalism of dislocation dynamics and thermal activation theory, enables the creation of materials capable of withstanding extreme service conditions, ensuring structural integrity and prolonged lifespans in critical applications. This framework provides predictive tools for designing materials with enhanced creep resistance by optimizing parameters such as and Q.
2.3. Grain Boundary Sliding
- Dislocation Movement [68]: Dislocations are line defects within the crystal lattice that move under stress, contributing significantly to plastic deformation. Dislocation movement occurs through processes such as glide and climb. Glide refers to the motion of dislocations along specific crystallographic planes, driven by shear stress, allowing for deformation by sliding of atomic planes relative to each other. This sliding is characterized by a decrease in elastic energy, making the material more ductile. Climb, on the other hand, is the process where dislocations move out of their glide planes by the absorption or emission of vacancies. This movement can relieve local stress fields more directly, allowing dislocations to reconfigure and reduce their energy, thereby accommodating deformation more effectively. The combination of glide and climb mechanisms allows for continuous strain accommodation, maintaining structural integrity under stress.
- Elastic Distortion [63]: Elastic deformation is the reversible change in shape of the grains within a material when subjected to small deformations. In this mechanism, the grains themselves can deform elastically, meaning the atoms within the grains move under stress but return to their original positions once the stress is removed. This elastic response is crucial for accommodating small strains without permanent deformation. When a material undergoes small deformations, the grains can distort elastically, which not only absorbs the applied stress but also permits the recovery of energy without the accumulation of dislocations. This mechanism helps in distributing the stress more evenly across the material, preventing localized plastic deformation and crack initiation, thereby maintaining compatibility and preventing failure.
- Diffusional Accommodation [63]: Diffusional creep is a critical mechanism for accommodating plastic deformation, especially at elevated temperatures. In diffusional accommodation, atoms or vacancies migrate through the grain boundaries or within the grains themselves. This diffusion process allows for the redistribution of material, helping to relieve stress by moving atoms to areas where the material is more compressed. The movement can occur along grain boundaries where atomic mobility is enhanced due to the reduced energy barrier compared to intra-grain diffusion. This redistribution enables the material to deform more uniformly and adapt to stress by adjusting its microstructure without permanent change in the lattice. This diffusional process can effectively increase the material’s ductility, enabling it to handle higher strains before failure.
- Role of Grain Boundaries: Grain boundary sliding occurs along grain boundaries, and the density of grain boundaries increases as the grain size decreases. The number of grain boundaries per unit volume scales inversely with the grain size, . A higher grain boundary density means there are more locations where sliding can occur, which increases the contribution of GBS to the overall strain rate.
- Diffusion Mechanisms: Grain boundary sliding is often accommodated by diffusion, either along grain boundaries or through the bulk. Smaller grains shorten the diffusion path between boundaries, facilitating faster accommodation of sliding. The diffusion time scales inversely with d, further enhancing the rate of grain boundary sliding for smaller grains.
- Stress Distribution and Sliding: For a given applied stress, smaller grains experience higher stress concentrations at the grain boundaries due to reduced grain size. This enhanced stress at the boundaries promotes sliding, leading to faster strain accumulation in smaller grains.
2.3.1. Nanomaterials
2.3.2. Application to High-Strength Steel and Tungsten Filaments
-
High-Temperature Structural Stability and Atomic Mobility:
- -
- Tungsten filaments operate at temperatures around 2000°C, where the material’s mechanical properties are governed by grain boundary sliding.
- -
- At these temperatures, the atomic interactions at grain boundaries are critically important for accommodating deformation.
- -
- Grain boundary sliding involves the migration of vacancies and interstitial atoms through the grain boundary, facilitated by increased thermal energy promoting atomic diffusion.
- -
- The atomic-scale mechanism maintains the structural integrity of tungsten filaments under prolonged thermal exposure.
-
Micromechanical Mechanisms of Grain Boundary Sliding:
- -
- The sliding of grains across their boundaries involves the activation of localized stress fields within the grain boundary regions.
- -
- These stress fields are a consequence of the anisotropic distribution of atomic interactions along the boundary.
- -
- The process follows an Arrhenius-type relationship for the sliding velocity v dependent on temperature:where Q is the activation energy for sliding, R is the universal gas constant, and T is the absolute temperature.
- -
- The rate of grain boundary sliding in tungsten is influenced by impurity content and their segregation to grain boundaries, which alters the activation energy for sliding.
-
Enhancement of Fatigue Resistance:
- -
- Grain boundary sliding significantly enhances the fatigue resistance of tungsten filaments.
- -
- The process allows for efficient stress redistribution during cyclic loading, mitigating the formation of dislocation pile-ups and microcracks.
- -
- The redistribution of stresses through grain boundary sliding reduces the likelihood of fatigue failure under thermal cycling conditions.
-
Thermal Management and Stress Relaxation:
- -
- Grain boundary sliding plays a pivotal role in thermal management within tungsten filaments.
- -
- The sliding mechanism accommodates thermal expansion mismatches between adjacent grains, preventing the development of thermal gradients that could lead to stress concentrations and microcracking.
- -
- Stress relaxation provided by grain boundary sliding is crucial for preventing the accumulation of stress-induced defects such as dislocation tangles and microcracks.
-
Microstructural Optimization:
- -
- The effectiveness of grain boundary sliding in tungsten filaments is closely tied to the material’s microstructural characteristics.
- -
- Key factors include grain size, grain boundary misorientation, and impurity content.
- -
- Smaller grain sizes enhance grain boundary sliding by increasing the volume fraction of grain boundaries relative to the total volume.
- -
- Low-angle grain boundaries, characterized by high dislocation densities, are particularly conducive to sliding due to the lower activation energy for atomic diffusion.
- -
- The presence of impurities such as oxygen, carbon, or metal alloying elements can significantly affect the grain boundary properties, influencing the sliding behavior.
2.3.3. Superplastic Forming (SPF) Technique
- Unusually Weak Grain Boundary Sliding (GBS): During the initial stage of superplastic deformation in commercial fine-grained Al-Mg alloys, grain boundary sliding (GBS) is observed to be unusually weak. This phenomenon is due to the presence of a complex microstructure, where the grain boundaries are not easily activated for sliding. The resistance to GBS is primarily due to the strengthening precipitates within the grains and along the grain boundaries, which hinder dislocation movement and grain boundary mobility. These precipitates serve as barriers to dislocation glide and grain boundary motion, creating a high-energy configuration that requires significant strain energy for activation [74].
- Tensile Test and Grain Elongation: A tensile test was conducted to evaluate the superplastic properties, where grains elongated significantly along the tensile direction to a range of 50% to 70%. This substantial elongation indicates effective plastic strain accommodation through GBS, which is crucial for the material’s superplasticity. The tensile test quantifies the material’s ability to elongate without fracturing, which is a hallmark of superplasticity. The deformation mechanism in this context involves the cooperative action of grain boundary sliding, dislocation movement, and the formation of subgrains, all of which contribute to the observed strain [74].
- Increased Precipitation Depletion Zone Fractions: Increased Mg content leads to the formation of larger precipitation depletion zones within the grains. These zones are regions devoid of precipitates, effectively weakening the grain boundaries by reducing the pinning force on dislocations. As Mg content increases, the volume fraction of these depletion zones becomes more significant, which facilitates grain boundary sliding by lowering the energy barrier for sliding at these boundaries. The depletion zones act as initiation sites for GBS, making the grain boundaries more compliant to sliding during deformation. The quantitative relationship between Mg content and the size of these depletion zones can be described by the empirical increase in their volume fraction, leading to an enhanced ability of the material to undergo GBS [74].
- Particle Segregation on Longitudinal Grain Boundaries: The segregation of particles along the longitudinal grain boundaries is a crucial factor enhancing GBS. With higher Mg content, these particles, which are typically Mg-rich phases, segregate at the grain boundaries. The presence of these particles reduces the grain boundary strength, creating lubricated interfaces that are easier to slide. The particle segregation provides a lower friction interface along the grain boundaries, significantly lowering the activation energy required for GBS. The quantitative description of this phenomenon can be derived from the increased particle volume fraction, which correlates directly with the increased Mg content [74].
- Dislocation Activity and Subgrains: Dislocation activity within the grains plays a pivotal role in the deformation mechanism. As Mg content increases, there is an increased density of dislocations and subgrains. These subgrains are small, misoriented regions formed by dislocations within the grains, which reduce the overall resistance to deformation. The dislocation activity within the grains leads to the formation of these subgrains, which in turn facilitate GBS by reducing the grain boundary friction. The formation of subgrains is a quantitative indicator of the enhanced plasticity of the alloy, as they reduce the internal stress and energy associated with dislocation interactions. The density of subgrains and the mobility of dislocations are critical for achieving higher elongations during superplastic deformation [74].
- Mg Content and Grain Size Stability: Increasing the Mg content from 4.8% to 6.5–7.6% significantly aids grain size stability during increased temperature processes. The higher Mg content stabilizes the grain boundaries, preventing excessive grain coarsening at elevated temperatures. This stabilization is crucial for maintaining superplastic properties, as finer grains are more conducive to GBS. The Mg content helps in pinning the grain boundaries by forming Mg-rich particles within the matrix, which limit grain boundary mobility and prevent grain growth. The relationship between Mg content and grain size stability can be quantitatively described by the reduction in grain growth rate with increasing Mg content [74].
- Simplified Grain Boundary Sliding (GBS) and Reduced Diffusion Creep: The increased Mg content simplifies GBS by providing a more lubricated grain boundary interface. This lubrication effect reduces the resistance to grain boundary sliding, which is a primary mechanism for superplastic deformation. Additionally, the higher Mg content reduces the contribution of diffusion creep to deformation. Diffusion creep is a mechanism where atomic diffusion occurs along the grain boundaries to accommodate deformation. As Mg content increases, the alloy becomes less reliant on diffusion creep, as the grain boundaries are already weakened by Mg segregation and depletion zones, reducing the need for atomic diffusion. The decreased diffusion creep contribution can be quantitatively described by the reduced activation energy for deformation and the corresponding decrease in the diffusion coefficients at high temperatures [74].
- Increased Failure Strain: The modification of Mg content leads to a significant increase in the material’s failure strain from 300% to 430%. This increase is a direct result of the enhanced GBS, reduced diffusion creep contribution, and the improved microstructural stability due to higher Mg content. The increased failure strain quantifies the alloy’s ability to sustain higher levels of deformation without fracture. The quantitative measure of failure strain can be linked to the increased volume fractions of precipitation depletion zones, the enhanced particle segregation, and the higher density of dislocations and subgrains. These factors collectively contribute to a more ductile and superplastic material, capable of large elongations during deformation [74].
2.4. Solute Drag Creep
2.4.1. Portevin–Le Chatelier Effect
- Dislocation-Solute Interaction: When the applied stress becomes sufficiently large, dislocations—imperfections within the crystal structure—are able to break away from solute atoms. This happens because the dislocation velocity increases with the applied stress, making it easier for dislocations to move past solute atoms.
- Stress Dynamics: After dislocations break away from solute atoms, the stress initially decreases, and the dislocation velocity also decreases. During this period, solute atoms can approach and reattach to the previously departed dislocations, leading to an increase in local stress. This reattachment happens because the solute atoms have a tendency to obstruct dislocation movement.
- Repetitive Stress Behavior: The process of dislocations moving away and then being obstructed by solute atoms can repeat, creating a cyclic pattern of local stress maxima and minima. These repetitions of stress maxima and minima are characteristic of solute drag creep. The cyclic nature indicates a dynamic interaction between dislocations and solute atoms under stress, which can be observed experimentally in the form of fluctuating stress levels in the material.
2.5. Dislocation Climb-Glide Creep
2.6. Harper-Dorn Creep
2.7. Sequential and Parallel Process in Creep
3. Phenomenological Description of Creep
3.1. Stages of Creep
- Primary Creep
- Secondary Creep
- Tertiary Creep
3.1.1. Primary Creep
- Stress-induced deformation (via dislocation glide, climb, and diffusion): Stress-induced deformation in primary creep involves the movement of dislocations within the crystal lattice, facilitated by mechanisms such as glide, climb, and diffusion, in response to an applied stress. During this initial stage of creep, the strain rate decreases over time as the material adjusts to the applied load. Dislocation glide occurs when dislocations move along specific crystallographic planes under shear stress, allowing plastic deformation. However, at elevated temperatures, obstacles such as other dislocations, impurities, or grain boundaries can impede glide, requiring additional mechanisms to continue deformation. Dislocation climb becomes active under these conditions, where atoms diffuse around obstacles, enabling dislocations to move out of their glide planes. Diffusion also plays a role in redistributing atoms to accommodate stress concentrations, further facilitating deformation. These combined mechanisms allow the material to deform plastically while redistributing stress, leading to a reduction in strain rate as the material undergoes work hardening. Stress-induced deformation during primary creep is critical in establishing the microstructural changes that influence the material’s subsequent behavior in the steady-state (secondary) creep phase.
- Strain hardening (increased material resistance due to evolving dislocation interactions): Strain hardening in primary creep refers to the progressive increase in a material’s resistance to deformation due to the accumulation and interaction of dislocations as the material is subjected to sustained stress. During primary creep, the material initially deforms at a relatively high strain rate, but as dislocations are generated and move through the crystal lattice, they interact with each other and with other obstacles such as grain boundaries or second-phase particles. These interactions create a tangled network of dislocations that impede further dislocation motion, increasing the material’s strength and reducing the rate of deformation over time. This process of strain hardening counteracts the applied stress, leading to a gradual decrease in the creep strain rate characteristic of the primary creep stage. The extent of strain hardening depends on factors such as the material’s composition, temperature, and microstructure. Strain hardening is essential for establishing the microstructural framework that influences the material’s transition to the steady-state deformation observed in secondary creep, where a balance between hardening and recovery mechanisms is achieved.
- Microstructural recovery (dynamic reorganization of internal defects): Microstructural recovery in primary creep is the process by which the internal defects within a material, such as dislocations, dynamically reorganize and reduce their energy state under sustained stress and elevated temperatures. As the material deforms during the initial stages of creep, dislocations are generated and accumulate, leading to localized stress concentrations. However, at elevated temperatures, the enhanced atomic mobility allows for mechanisms such as dislocation annihilation, rearrangement, and the formation of lower-energy configurations. These processes help to alleviate the internal stresses caused by dislocation interactions, partially offsetting the effects of strain hardening. Microstructural recovery contributes to the gradual decrease in strain rate observed in primary creep, as it counterbalances the increasing resistance to deformation caused by dislocation tangling and interaction. The degree of recovery depends on factors such as the applied stress, temperature, and material properties, and it plays a crucial role in determining how the material transitions from the primary to the secondary creep stage, where a steady-state strain rate is achieved through a balance between recovery and hardening mechanisms.
3.1.2. Secondary Creep
- Work Hardening: Work hardening in secondary creep refers to the process by which a material becomes progressively stronger and more resistant to deformation as dislocations accumulate and interact within the crystal structure during plastic deformation. In the secondary creep stage, the strain rate reaches a steady state due to a dynamic balance between work hardening and recovery mechanisms. As the material deforms under sustained stress, dislocations are generated and move through the crystal lattice. These dislocations interact with each other, forming tangles and networks that create barriers to further dislocation motion, thereby increasing the material’s strength. This phenomenon is known as work hardening and contributes to the deceleration of the strain rate that characterizes the transition from primary to secondary creep. However, at the elevated temperatures typical of creep conditions, recovery processes, such as dislocation annihilation and rearrangement, also occur, mitigating the effects of work hardening. The balance between these opposing mechanisms maintains the steady strain rate observed in secondary creep. Work hardening plays a crucial role in the material’s ability to sustain long-term deformation under high-temperature and high-stress conditions, influencing its overall creep resistance and structural integrity.
- Dynamic Recovery: Dynamic recovery in secondary creep is the process by which a material undergoing sustained deformation at high temperatures continuously relieves internal stress through the rearrangement and annihilation of dislocations. During secondary creep, the material reaches a steady-state strain rate due to a balance between work hardening, which increases the material’s resistance to deformation, and dynamic recovery, which counteracts it. As dislocations are generated and move through the crystal lattice under stress, interactions between them create localized regions of high stress and energy. At elevated temperatures, atomic mobility increases, allowing dislocations to rearrange into lower-energy configurations or to annihilate through mechanisms such as climb or cross-slip. This recovery process reduces the overall dislocation density and alleviates the internal stress, enabling the material to continue deforming without a significant increase in resistance. Dynamic recovery is particularly important in materials with high-temperature ductility, as it helps to stabilize the creep rate during the secondary stage, preventing rapid strain rate escalation. It is a key factor in determining a material’s long-term performance under high-stress, high-temperature conditions, as it influences the material’s ability to accommodate plastic deformation without accumulating excessive damage.
- Through the lattice (Nabarro-Herring creep): Atoms migrate through the bulk of the grains.
- Along grain boundaries (Coble creep): Atomic motion is concentrated at grain boundaries, particularly in fine-grained materials.
- : Diffusion creep (linear relationship with stress).
- : Dislocation creep (nonlinear relationship with stress).
- Dislocation Networks: Dislocation networks in secondary creep represent the stabilized structure of dislocations that form within a material under sustained stress and temperature during the steady-state phase of creep. In secondary creep, the material experiences a constant strain rate, which arises from a balance between the processes of dislocation generation, motion, and annihilation. Dislocations are line defects in the crystal lattice that serve as carriers of plastic deformation, and their density increases as deformation progresses. As the dislocations move and interact, they form tangles and networks, which act as barriers to further dislocation motion. These networks stabilize the creep rate by hindering the free movement of dislocations, requiring additional stress to drive further deformation. The formation and evolution of dislocation networks depend on the material’s microstructure, stress level, and temperature, with higher temperatures allowing dislocations to rearrange into more stable configurations. These networks are crucial in controlling the mechanical behavior of materials during secondary creep, as they dictate the material’s resistance to deformation and its ability to sustain steady-state operation under prolonged service conditions.
- Grain Boundary Sliding: Grain boundary sliding in secondary creep is a deformation mechanism that occurs when grains within a polycrystalline material slide relative to each other along their boundaries under the influence of sustained stress and elevated temperatures. This sliding accommodates plastic deformation and contributes to the steady-state strain rate characteristic of secondary creep. Unlike primary creep, where strain rate decreases over time, secondary creep achieves a balance between deformation mechanisms such as dislocation movement and grain boundary sliding. Grain boundary sliding in this phase typically occurs in a more controlled manner, facilitated by diffusion processes and the mobility of atoms at elevated temperatures. It is particularly prominent in materials with fine grains, where the increased number of grain boundaries provides more pathways for sliding. However, to maintain continuity at the boundaries, deformation is often accompanied by localized adjustments, such as the formation of voids or dislocation activity, which can evolve into more significant damage over time. Grain boundary sliding is a key factor in determining the creep resistance of materials, especially in high-temperature applications, as it plays a role in controlling the overall strain rate and contributes to the material’s ability to sustain loads during long-term service.
- Aerospace: Turbine blades must resist deformation over thousands of hours of operation.
- Power Plants: Steam generators and pressure vessels must maintain structural integrity under sustained loads.
- Nuclear Reactors: Materials must withstand radiation damage in addition to creep stresses.
3.1.3. Tertiary Creep
- Void Nucleation and Growth: Void nucleation and growth in tertiary creep is a critical phenomenon that plays a significant role in the eventual failure of materials subjected to prolonged high stress and elevated temperatures. During the tertiary stage of creep, the strain rate accelerates rapidly as the material undergoes microstructural degradation. This degradation is driven by the formation and evolution of microscopic cavities, or voids, which nucleate at stress concentration sites. These sites often include grain boundaries, second-phase particles, inclusions, or areas with high dislocation density. The initiation of voids is influenced by factors such as the stress distribution, material microstructure, and temperature. Once nucleated, the voids grow due to mechanisms like atomic diffusion away from the void surfaces and localized plastic deformation around the cavities. Elevated temperatures enhance atomic mobility, facilitating the diffusion process, while high stresses accelerate plastic deformation, further enlarging the voids. As the voids expand, they begin to coalesce, linking together to form microcracks. This coalescence significantly reduces the effective cross-sectional area of the material, causing an exponential increase in strain rate during tertiary creep. Ultimately, the material fails through mechanisms such as ductile fracture or intergranular cracking, depending on the temperature and stress conditions. Understanding the interplay of void nucleation, growth, and coalescence is crucial for predicting material failure and designing components capable of withstanding the demands of high-temperature, high-stress environments.
- Microcrack Formation: Microcrack formation in tertiary creep is a critical process that contributes to the accelerated degradation and eventual failure of materials under prolonged stress and elevated temperatures. During tertiary creep, the material experiences an increased strain rate driven by the accumulation of damage at the microstructural level. This damage begins with the nucleation and growth of voids, which eventually coalesce to form microcracks. These microcracks typically originate at regions of stress concentration, such as grain boundaries, inclusions, or second-phase particles, where the local stress exceeds the material’s capacity to deform plastically. The growth of microcracks is facilitated by mechanisms such as localized plastic deformation and atomic diffusion, both of which are enhanced at high temperatures. As these cracks propagate, they reduce the effective load-bearing cross-sectional area of the material, leading to a further increase in strain rate. The interaction and linkage of microcracks create a network of weakened zones that accelerate material failure. Ultimately, this progression leads to macroscopic cracking and fracture, marking the final stage of tertiary creep. Understanding microcrack formation is essential for predicting the lifespan of materials and preventing catastrophic failure in applications where sustained high-stress and temperature conditions are prevalent.
- Grain Boundary Sliding: Grain boundary sliding in tertiary creep is a deformation mechanism that occurs as grains within a material slide past one another along their boundaries under the influence of sustained stress and elevated temperatures. This process is a significant contributor to the accelerated strain rate observed during tertiary creep, as it facilitates the redistribution of stress and accommodates plastic deformation. Grain boundaries, being regions of atomic misalignment, serve as preferred pathways for sliding due to their relatively lower resistance compared to the grain interiors. At high temperatures, the atomic mobility increases, enabling the grains to slide more easily while maintaining the overall structural integrity of the material. However, this sliding is not uniform and often leads to the nucleation of voids or microcracks at triple junctions or regions where the boundaries are irregular. These defects grow and coalesce, accelerating the onset of failure. Grain boundary sliding is particularly prominent in materials with fine grains and weak boundary strength, and its extent is influenced by factors such as temperature, stress level, and the presence of impurities or secondary phases. This mechanism is critical in determining the creep resistance and long-term reliability of materials exposed to high-temperature environments, such as in power plants, turbines, and aerospace applications.
- Necking and Localization: Necking and localization in tertiary creep are phenomena that signify the final stages of deformation leading to material failure under sustained stress and high temperatures. Necking refers to the progressive reduction in cross-sectional area at a specific region of a material, which concentrates stress and accelerates the deformation in that localized area. This stress concentration creates a feedback loop where the localized thinning of the material increases the strain rate, further intensifying the deformation in that region. As the neck forms, the material’s ability to carry the applied load diminishes, and the deformation becomes highly localized, often accompanied by microstructural damage such as void formation, microcrack development, and grain boundary sliding. These processes collectively exacerbate the localization of strain, leading to catastrophic failure through ductile fracture or intergranular cracking. Necking and localization are influenced by factors such as the material’s ductility, the applied stress, and the operating temperature. These phenomena are critical to understanding the mechanical behavior of materials during tertiary creep, as they dictate the ultimate failure mechanism and limit the material’s service life in high-stress, high-temperature applications.
3.2. Constitutive Models
3.2.1. Norton-Bailey Law
- The creep strain rate is constant over time ().
- Work hardening and recovery mechanisms reach equilibrium, balancing the opposing effects of dislocation generation and annihilation.
- Indicates diffusional creep, where strain is governed by atomic diffusion under a stress gradient.
- Suggests dislocation creep, where the rate-controlling step involves thermally activated dislocation climb or glide.
- It does not account for transient or tertiary creep.
- Its parameters must be experimentally calibrated.
- Microstructural degradation is not explicitly included.
3.2.2. Time-Hardening Model
- Material Homogeneity and Isotropy: The material is considered homogeneous and isotropic at the macroscopic scale, ensuring that the creep behavior depends only on stress magnitude and time t, not on directionality.
- Time-Driven Evolution: The primary driver of inelastic strain is time elapsed under stress, decoupled from the total strain magnitude.
- Stress-Dependent Strain Rate: Creep strain rate scales with the applied stress according to a power-law relationship, reflecting dominant dislocation-based mechanisms.
- Dislocation Creep: Dominated by dislocation climb and glide. Characterized by n in the range 3–10 and .
- Diffusional Creep: Includes Nabarro-Herring (bulk diffusion) and Coble (grain boundary diffusion) mechanisms. Typically exhibits and .
- Grain Boundary Sliding: Common in polycrystalline materials. Results in moderate n and m, dependent on grain size.
3.2.3. Strain-Hardening Model
- 1.
-
Return-Mapping Algorithms: Enforcing plastic consistency conditions numerically. Return-mapping algorithms represent the backbone of incremental-iterative schemes for solving nonlinear problems in computational inelasticity. Their goal is to project the trial stress state back onto the admissible stress manifold (e.g., yield surface) while ensuring thermodynamic consistency, stability, and accuracy. This requires robust formulations, iterative solution strategies, and advanced integration schemes. The foundation of computational mechanics begins with the principle of virtual work. For a deformable body, the internal and external virtual work must balance:where is the Cauchy stress tensor, is the virtual strain tensor, is the body force, is the traction on boundary , is the displacement vector. Plasticity models govern the constitutive behavior . Plasticity models must satisfy thermodynamic laws, including:
- (a)
- First Law (Energy Balance):where U is internal energy, is external power, and is internal dissipation.
- (b)
- Second Law (Clausius-Duhem Inequality): The dissipation potential :where is the Helmholtz free energy, and is the thermodynamic force conjugate to .
We now have to analyze the Constitutive Relations. The general structure of the constitutive relations for strain-hardening plasticity includes Elastic Stress-Strain Law which states thatwhere is the elastic stiffness tensor. The Yield Function states thatwhere is the stress measure, and is the hardening parameter. The Flow Rule states thatwith g as the plastic potential. The Hardening Law states thatwhere H is the hardening modulus. The return-mapping algorithm resolves the incremental evolution of stresses and internal variables within the framework of strain-hardening plasticity. It is composed of the elastic predictor and plastic corrector steps. The elastic predictor step assumes no plastic deformation occurs within a load increment:The trial stress is evaluated against the yield function:- If , the material remains elastic.
- If , plastic correction is required.
We now state the Plastic Corrector step. The plastic corrector step enforces the yield condition:Regarding the Consistency Condition, The yield function is linearized using Taylor expansion:We now need to do Stress Correction. The stress update is obtained by projecting onto the yield surface:where is the plastic tangent operator. We have to do Numerical Integration of Hardening Variables. The hardening parameter evolves according to:where is determined from the consistency condition. Regarding Radial Return-Mapping for von Mises Plasticity, For von Mises plasticity with isotropic hardening, we first compute the equivalent stress:where . We then solve for the plastic multiplier :We then update the stress and internal variables:For Thermodynamic Consistency, The return-mapping algorithm must satisfy the dissipation inequality:For Algorithmic Tangent Moduli, The consistent tangent operator is derived from the linearization of the constitutive equations to ensure quadratic convergence:Regarding Viscoplastic Extensions, Viscoplasticity incorporates rate dependence:requiring regularization of the yield surface. In Large Deformation Framework, for large strains, multiplicative decomposition and logarithmic strain measures replace additive frameworks. - 2.
-
Incremental Formulation: Computing stress and strain increments iteratively. The incremental formulation in the strain hardening model is a mathematically rigorous framework employed to describe the evolution of stresses and strains in a material undergoing plastic deformation, where the material’s resistance to further deformation increases due to strain hardening. This formulation is rooted in the decomposition of the total strain tensor, , into its elastic and plastic components, such that , where represents the recoverable elastic strain and signifies the irrecoverable plastic strain. The governing equations are derived incrementally, acknowledging the fact that material deformation is typically path-dependent in the plastic regime. The incremental stress-strain relationship is derived from the generalized Hooke’s law for elastic behavior, modified to account for plastic deformation. This relationship is expressed aswhere is the increment in the Cauchy stress tensor, is the fourth-order elastic stiffness tensor, is the total strain increment, and is the plastic strain increment. The plastic strain increment, , is determined using an associated flow rule derived from the yield function, f, which characterizes the boundary between elastic and plastic behavior in stress space. Mathematically, the plastic strain increment is expressed aswhere is a scalar plastic multiplier determined by enforcing the consistency condition, and defines the direction of plastic flow. The strain hardening behavior of the material is incorporated through the evolution of the yield surface, which depends on internal state variables such as the accumulated plastic strain. A typical hardening law relates the current yield stress, k, to the effective plastic strain, , via a hardening modulus, H. For isotropic hardening, this relationship can be expressed aswhere is the initial yield stress andquantifies the accumulated plastic deformation. This formulation ensures that the yield surface expands uniformly in stress space as plastic deformation progresses, reflecting the material’s increased resistance to plastic flow. The yield function, , is updated incrementally to account for the evolving yield stress. In the von Mises yield criterion, for example, the function takes the formwhere is the deviatoric component of the stress tensor, and k is the current yield stress. When plasticity occurs, the stress state must satisfy the consistency condition , which is enforced using numerical methods. If the stress state violates this condition, a return mapping algorithm is employed to project the stress back onto the updated yield surface. This ensures that the solution remains consistent with the yield criterion while incorporating the effects of strain hardening.The numerical implementation of the incremental formulation involves solving a set of coupled nonlinear equations iteratively, often using methods such as Newton-Raphson. The computational framework requires updating the stress tensor, plastic strain tensor, and internal state variables incrementally within each load or time step. This approach ensures a rigorous treatment of the path-dependent nature of plastic deformation, accounting for the interplay between elastic recovery, plastic flow, and strain hardening. Through this mathematically rigorous and systematic formulation, the incremental strain hardening model enables accurate simulations of material behavior under complex loading conditions.
4. Sintering
5. Examples of Creep Deformation
5.1. Surface Materials
5.2. Metals
5.3. Polymers
5.4. Glass
5.5. Concrete
5.6. Wood
6. Larson-Miller Parameter and Monkman Grant Rule
7. Multi-Axial Creep and Damage Mechanics
8. Creep Testing
9. Environmental Effects on Creep: Oxidation and Irradiation
10. Increasing Creep Resistance
10.1. Working Conditions
10.1.1. Time
10.1.2. Temperature
10.1.3. Applied Stress
- Material Selection: Choosing materials with high Q and low A values for the creep-resistant phase.
- Microstructural Design: Optimizing grain size and precipitate distribution to strengthen the material.
- Stress Management: Minimizing stress through design to reduce the creep rate.
- Thermal Control: Controlling temperature to lower the activation energy for creep processes.
10.2. Material Selection
11. Creep Prevention in Superalloys
12. Famous Accidents
12.1. Big Dig Tunnel Ceiling Collapse in Boston on July 2006
12.2. Collapse of the World Trade Center on September 11, 2001
13. Conclusions
Acknowledgments
References
- Kassner, M. E. (2015). Fundamentals of creep in metals and alloys. Butterworth-Heinemann.
- Poirier, J. P. (1985). Creep of crystals: high-temperature deformation processes in metals, ceramics and minerals. Cambridge University Press.
- Penny, R. K. , and Marriott, D. L. (2012). Design for creep. Springer Science and Business Media.
- Naumenko, K. , and Altenbach, H. (2007). Modeling of creep for structural analysis. Springer Science and Business Media.
- Boyle, J. T. , and Spence, J. (2013). Stress analysis for creep. Elsevier.
- Ashby, M. F. , and Jones, D. R. (2012). Engineering materials 1: an introduction to properties, applications and design (Vol. 1). Elsevier.
- Frost, Harold J.; Ashby, Michael F. (1982). Deformation-Mechanism Maps: The Plasticity and Creep of Metals and Ceramics. Pergamon Press. ISBN 978-0-08-029337-0.
- Turner, S. (2001). Creep of polymeric materials. Encyclopedia of materials: science and technology, 1813-1817.
- Passmore, E. M., Duff, R. H., and Vasilos, T. Creep of dense, polycrystalline magnesium oxide. Journal of the American Ceramic Society 1966, 49(11), 594–600. [CrossRef]
- Tremper, R. T. , and Gordon, R. S. (1971). Effect of nonstoichiometry on the viscous creep of iron-doped polycrystalline magnesia (No. COO-1591-15; UTEC-MSE-71-102; CONF-711030-8). Utah Univ., Salt Lake City. Div. of Materials Science and Engineering.
- Vandervoort, R.R.; Barmore, W.L. Compressive creep of polycrystalline beryllium oxide. Journal of the American Ceramic Society 1963, 46(4), 180–184. [Google Scholar] [CrossRef]
- Hollenberg, G. W., and Gordon, R. S. Effect of oxygen partial pressure on the creep of polycrystalline Al2O3 doped with Cr, Fe, or Ti. Journal of the American Ceramic Society 1973, 56(3), 140–147. [CrossRef]
- Ahres, T. J. (1995). Rock Physics and Phase Relations a handbook of physical constants: Washington. American Geophysical Union, 190-191.
- Berry, J., Rottler, J., Sinclair, C. W., and Provatas, N. Atomistic study of diffusion-mediated plasticity and creep using phase field crystal methods. Physical Review B 2015, 92, 134103. [CrossRef]
- Goldsby, D. L. (2006). Superplastic flow of ice relevant to glacier and ice-sheet mechanics. Glacier science and environmental change, 308-314.
- Courtney, T. H. (2005). Mechanical behavior of materials. Waveland Press.
- Herring, C. Diffusional viscosity of a polycrystalline solid. Journal of applied physics 1950, 21, 437–445. [Google Scholar] [CrossRef]
- Neu, R.W.; Sehitoglu, H. Thermomechanical fatigue, oxidation, and creep: Part I. Damage mechanisms. Metallurgical transactions A 1989, 20, 1755–1767. [Google Scholar] [CrossRef]
- Neu, R. W.; Sehitoglu, H. Thermomechanical fatigue, oxidation, and creep: Part II. Life prediction. Metallurgical transactions A 1989, 20, 1769–1783. [Google Scholar] [CrossRef]
- Meyers, M. A., and Chawla, K. K. (2008). Mechanical behavior of materials. Cambridge university press.
- Coble, R. L. A model for boundary diffusion controlled creep in polycrystalline materials. Journal of applied physics 1963, 34, 1679–1682. [Google Scholar] [CrossRef]
- Cottrell, A. H., and Jaswon, M. A. Distribution of solute atoms round a slow dislocation. Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences 1949, 199, 104–114.
- Cottrell, A. H. LXXXVI. A note on the Portevin-Le Chatelier effect. The London, Edinburgh, and Dublin Philosophical Magazine and Journal of Science 1953, 44, 829–832. [Google Scholar] [CrossRef]
- Penning, P. Mathematics of the portevin-le chatelier effect. Acta Metallurgica 1972, 20, 1169–1175. [Google Scholar] [CrossRef]
- Yilmaz, A. The Portevin–Le Chatelier effect: a review of experimental findings. Science and technology of advanced materials 2011, 12, 063001. [Google Scholar] [CrossRef] [PubMed]
- Rizzi, E. Rizzi, E., and Hähner, P. On the Portevin–Le Chatelier effect: theoretical modeling and numerical results. 2004, 20, 121–165. [Google Scholar]
- Jiang, H., Zhang, Q., Chen, X., Chen, Z., Jiang, Z., Wu, X., and Fan, J. Three types of Portevin–Le Chatelier effects: Experiment and modelling. Acta materialia 2007, 55, 2219–2228. [CrossRef]
- Burton, B. The influence of solute drag on dislocation creep. Philosophical Magazine A 1982, 46, 607–616. [Google Scholar] [CrossRef]
- Heald, P.T.; Harbottle, J.E. Irradiation creep due to dislocation climb and glide. Journal of Nuclear Materials 1977, 67, 229–233. [Google Scholar] [CrossRef]
- Mansur, L.K. Irradiation creep by climb-enabled glide of dislocations resulting from preferred absorption of point defects. Philosophical Magazine A 1979, 39, 497–506. [Google Scholar] [CrossRef]
- Weertman, J. Steady-state creep of crystals. Journal of Applied Physics 1957, 28, 1185–1189. [Google Scholar] [CrossRef]
- Mott, N. F. Dislocations, work-hardening and creep. Nature. 1955, 175, 365–367. [Google Scholar]
- Harper, J., and Dorn, J. E. Viscous creep of aluminum near its melting temperature. Acta Metallurgica 1957, 5(11), 654–665. [CrossRef]
- Mohamed, F. A. , Murty, K. L., and Morris, J. W. (1973). Harper-dorn creep in al, pb, and sn. Metallurgical Transactions, 4, 935-940.
- Barrett, C. R. , Muehleisen, E. C., and Nix, W. D. (1972). High temperature-low stress creep of Al and Al+ 0.5 Fe. Materials Science and Engineering, 10, 33-42.
- Kassner, M. E. , and Pérez-Prado, M. T. (2000). Five-power-law creep in single phase metals and alloys. Progress in materials science, 45(1), 1-102.
- Poirier, J. P. (1985). Creep of crystals: high-temperature deformation processes in metals, ceramics and minerals. Cambridge University Press.
- Singh, S. P. , Kumar, P., and Kassner, M. E. (2020). The low-stress and high-temperature creep in LiF single crystals: An explanation for the So-called Harper-Dorn creep. Materialia, 13, 100864.
- Srivastava, V., McNee, K. R., Jones, H., and Greenwood, G. W. Tensile creep behaviour of coarse grained copper foils at high homologous temperatures and low stresses. Materials science and technology 2005, 21(6), 701–707. [CrossRef]
- Malakondaiah, G., and Rao, P. R. Grain size independent viscous creep in alpha titanium and beta cobalt. Scripta Metallurgica 1979, 13(12), 1187–1190. [CrossRef]
- Blum, W. , and Maier, W. (1999). Harper-Dorn Creep—a Myth?. physica status solidi (a), 171(2), 467-474.
- Ginter, T. J. , Chaudhury, P. K., and Mohamed, F. A. (2001). An investigation of Harper–Dorn creep at large strains. Acta materialia, 49(2), 263-272.
- Ginter, T. J. , and Mohamed, F. A. (2002). Evidence for dynamic recrystallization during Harper–Dorn creep. Materials Science and Engineering: A, 322(1-2), 148-152.
- McClay, K. R. (1977). Pressure solution and Coble creep in rocks and minerals: a review. Journal of the Geological Society, 134(1), 57-70.
- Wheeler, J. (1992). Importance of pressure solution and coble creep in the deformation of polymineralic rocks. Journal of Geophysical Research: Solid Earth, 97(B4), 4579-4586.
- Blum, W. , and Eisenlohr, P. (2009). Dislocation mechanics of creep. Materials Science and Engineering: A, 510, 7-13.
- Hirth, G., and Tullis, J. A. N. Dislocation creep regimes in quartz aggregates. Journal of structural geology 1992, 14(2), 145–159. [CrossRef]
- Somekawa, H. , Hirai, K., Watanabe, H., Takigawa, Y., and Higashi, K. (2005). Dislocation creep behavior in Mg–Al–Zn alloys. Materials Science and Engineering: A, 407(1-2), 53-61.
- Renner, J. , Evans, B., and Siddiqi, G. (2002). Dislocation creep of calcite. Journal of Geophysical Research: Solid Earth, 107(B12), ECV-6.
- Pollock, T. M. , and Argon, A. S. (1992). Creep resistance of CMSX-3 nickel base superalloy single crystals. Acta Metallurgica et Materialia, 40(1), 1-30.
- Kakehi, K. (2000). Effect of primary and secondary precipitates on creep strength of Ni-base superalloy single crystals. Materials Science and Engineering: A, 278(1-2), 135-141.
- Sun, F. , Gu, Y. F., Yan, J. B., Xu, Y. X., Zhong, Z. H., and Yuyama, M. (2016). Creep deformation and rupture mechanism of an advanced wrought NiFe-based superalloy for 700° C class A-USC steam turbine rotor application. Journal of Alloys and Compounds, 687, 389-401.
- Zhang, J. X. , Murakumo, T., Harada, H., Koizumi, Y., and Kobayashi, T. (2004). Creep deformation mechanisms in some modern single-crystal superalloys. Superalloys, 2004, 189-195.
- Zhang, J. X. , Murakumo, T., Harada, H., and Koizumi, Y. J. S. M. (2003). Dependence of creep strength on the interfacial dislocations in a fourth generation SC superalloy TMS-138. Scripta Materialia, 48(3), 287-293.
- Ennis, P. J. , and Czyrska-Filemonowicz, A. (2003). Recent advances in creep-resistant steels for power plant applications. Sadhana, 28, 709-730.
- Pekguleryuz, M., and Celikin, M. Creep resistance in magnesium alloys. International Materials Reviews 2010, 55(4), 197–217. [CrossRef]
- Mo, N. , Tan, Q., Bermingham, M., Huang, Y., Dieringa, H., Hort, N., and Zhang, M. X. (2018). Current development of creep-resistant magnesium cast alloys: A review. Materials and Design, 155, 422-442.
- Darling, K. A. , Rajagopalan, M., Komarasamy, M., Bhatia, M. A., Hornbuckle, B. C., Mishra, R. S., and Solanki, K. N. (2016). Extreme creep resistance in a microstructurally stable nanocrystalline alloy. Nature, 537(7620), 378-381.
- Yang, J., Zhang, Z., Friedrich, K., and Schlarb, A. K. Creep resistant polymer nanocomposites reinforced with multiwalled carbon nanotubes. Macromolecular rapid communications 2007, 28(8), 955–961. [CrossRef]
- Bell, R. L. , and Langdon, T. G. (1967). An investigation of grain-boundary sliding during creep. Journal of Materials Science, 2, 313-323.
- Rachinger, W. A. (1956). Glide in lead telluride. Acta Metallurgica, 4(6), 647-649.
- Lifshitz, I. M. (1963). On the theory of diffusion-viscous flow of polycrystalline bodies. Sov. Phys. Jetp-USSR, 17(4), 909.
- Raj, R., and Ashby, M. F. (). On grain boundary sliding and diffusional creep. Metallurgical transactions 1971, 2, 1113–1127. [CrossRef]
- Bhaduri, A. (2018). Mechanical properties and working of metals and alloys (Vol. 264). Singapore: Springer Singapore.
- Langdon, T. G. (2006). Grain boundary sliding revisited: Developments in sliding over four decades. Journal of Materials Science, 41, 597-609.
- Rösler, J. , Harders, H., and Bäker, M. (2007). Mechanical behaviour of engineering materials: metals, ceramics, polymers, and composites. Springer Science and Business Media.
- Ree, J. H. (1994). Grain boundary sliding and development of grain boundary openings in experimentally deformed octachloropropane. Journal of Structural Geology, 16(3), 403-418.
- Gifkins, R. (1976). Grain-boundary sliding and its accommodation during creep and superplasticity. Metallurgical transactions a, 7, 1225-1232.
- Yang, H. , Gavras, S., and Dieringa, H. (2021). Creep characteristics of metal matrix composites. In Encyclopedia of Materials: Composites (Vol. 1, pp. 375-388). Elsevier.
- Adams, M. A. , and Murray, G. T. (1962). Direct Observations of Grain-Boundary Sliding in Bi-Crystals of Sodium Chloride and Magnesia. Journal of Applied Physics, 33(6), 2126-2131.
- Naziri, H. , Pearce, R., Brown, M. H., and Hale, K. F. (1975). Microstructural-mechanism relationship in the zinc/aluminium eutectoid superplastic alloy. Acta Metallurgica, 23(4), 489-496.
- Sergueeva, A. V., Mara, N. A., and Mukherjee, A. K. (). Grain boundary sliding in nanomaterials at elevated temperatures. Journal of materials science 2007, 42, 1433–1438. [CrossRef]
- Zhang, X., Li, W., Ma, J., Li, Y., Zhang, X., and Zhang, X. (). Modeling the effects of grain boundary sliding and temperature on the yield strength of high strength steel. Journal of Alloys and Compounds 2021, 851, 156747. [CrossRef]
- Mikhaylovskaya, A. V. , Yakovtseva, O. A., and Irzhak, A. V. (2022). The role of grain boundary sliding and intragranular deformation mechanisms for a steady stage of superplastic flow for Al–Mg-based alloys. Materials Science and Engineering: A, 833, 142524.
- Wright, P. K. (1978). The high temperature creep behavior of doped tungsten wire. Metallurgical Transactions A, 9, 955-963.
- Raj, R., and King, G. W. (). Life prediction of tungsten filaments in incandescent lamps. Metallurgical Transactions A 1978, 9, 941–946. [CrossRef]
- Schade, P. (2010). 100 years of doped tungsten wire. International Journal of Refractory Metals and Hard Materials, 28(6), 648-660.
- Rosato, M. G. , and Rosato, D. V. (Eds.). (2013). Plastics design handbook. Springer Science and Business Media.
- Meyers, M. A. , and Chawla, K. K. (2008). Mechanical behavior of materials. Cambridge university press.
- McCrum, N. G. , Buckley, C. P., and Bucknall, C. B. (1997). Principles of polymer engineering. Oxford University Press.
- Ozyhar, T. , Hering, S., and Niemz, P. (2013). Viscoelastic characterization of wood: Time dependence of the orthotropic compliance in tension and compression. Journal of Rheology, 57(2), 699-717.
- Jiang, J. , Erik Valentine, B., Lu, J., and Niemz, P. (2016). Time dependence of the orthotropic compression Young’s moduli and Poisson’s ratios of Chinese fir wood. Holzforschung, 70(11), 1093-1101.
- McLean, D. (1966). The physics of high temperature creep in metals. Reports on Progress in Physics, 29(1), 1.
- Pineau, A. , McDowell, D. L., Busso, E. P., and Antolovich, S. D. (2016). Failure of metals II: Fatigue. Acta Materialia, 107, 484-507.
- Chen, Z., Yang, F., Liu, S., Hu, X., Liu, C., Zhou, Z., ... and Liu, L. (). Creep behavior of intermetallic compounds at elevated temperatures and its effect on fatigue life evaluation of Cu pillar bumps. Intermetallics 2022, 144, 107526. [CrossRef]
- Pollock, T. M. , and Tin, S. (2006). Nickel-based superalloys for advanced turbine engines: chemistry, microstructure and properties. Journal of propulsion and power, 22(2), 361-374.
- Clemens, H., and Mayer, S. (). Design, processing, microstructure, properties, and applications of advanced intermetallic TiAl alloys. Advanced engineering materials 2013, 15(4), 191–215. [CrossRef]
- "Ceiling Collapse in the Interstate 90 Connector Tunnel". National Transportation Safety Board. Washington, D.C. "Ceiling Collapse in the Interstate 90 Connector Tunnel". National Transportation Safety Board. Washington, D.C.: NTSB. July 10, 2007. Retrieved. 2 December.
- Bažant, Z. P. , and Zhou, Y. (2002). Why did the world trade center collapse?—Simple analysis. Journal of Engineering Mechanics, 128(1), 2-6.
- Lakes, R. S. (2017). Viscoelastic Solids (1998). CRC press.
- "Is glass liquid or solid?". University of California, Riverside. Retrieved 2008-10-15. Available online: https://math.ucr.edu/home/baez/physics/General/Glass/glass.
- Abuzeid, O. M. , Al-Rabadi, A. N., and Alkhaldi, H. S. (2011). Recent Advancements in Fractal Geometric-Based Nonlinear Time Series Solutions to the Micro-Quasistatic Thermoviscoelastic Creep for Rough Surfaces in Contact. Mathematical Problems in Engineering, 2011(1), 691270.
- Long, H. , Mao, S., Liu, Y., Zhang, Z., and Han, X. (2018). Microstructural and compositional design of Ni-based single crystalline superalloys―A review. Journal of Alloys and Compounds, 743, 203-220.
- Pollock, T. M. , and Tin, S. (2006). Nickel-based superalloys for advanced turbine engines: chemistry, microstructure and properties. Journal of propulsion and power, 22(2), 361-374.
- Suzuki, A., Inui, H., Pollock, T.M. L12-Strengthened Cobalt-Base Superalloys. Annual Review of Materials Research 2015, 45, 345–368. [CrossRef]
- Pelleg, J. , and Pelleg, J. (2017). Creep in ceramics (pp. 41-61). Springer International Publishing.
- Bakunov, V. S., and Belyakov, A. V. Creep and structure of ceramics. Inorganic materials 2000, 36, 1297–1301. [CrossRef]
- Duddu, R., and Waisman, H. A temperature dependent creep damage model for polycrystalline ice. Mechanics of Materials 2012, 46, 23–41. [CrossRef]
- Pandey, M. C. , Taplin, D. M. R., Ashby, M. F., and Dyson, B. F. (1986). The effect of prior exposure-time on air-environment/creep interactions. Acta Metallurgica, 34(11), 2225-2233.
- Neville, A. M. (1971). Creep of concrete: plain, reinforced, and prestressed.
- Neville, A. M. (1964). Creep of concrete as a function of its cement paste content. Magazine of concrete Research, 16(46), 21-30.
- Bažant, Z. P. , and Osman, E. (1976). Double power law for basic creep of concrete. Matériaux et Construction, 9, 3-11.
- Yu, Q. , Bazant, Z. P., and Wendner, R. (2012). Improved algorithm for efficient and realistic creep analysis of large creep-sensitive concrete structures. ACI Structural Journal, 109(5), 665.
- Ross, A. D. (1958, March). Creep of concrete under variable stress. In Journal Proceedings (Vol. 54, No. 3, pp. 739-758).
- Bissonnette, B. , Pigeon, M., and Vaysburd, A. M. (2007). Tensile creep of concrete: study of its sensitivity to basic parameters. ACI Materials journal, 104(4), 360.
- Ranaivomanana, N., Multon, S., and Turatsinze, A. (2013). Basic creep of concrete under compression, tension and bending. Construction and Building Materials 2013, 38, 173–180. [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. |
© 2024 by the author. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (http://creativecommons.org/licenses/by/4.0/).
