Next Article in Journal
Numerical Analysis on Static Performances of Graphene Platelet-Reinforced Ethylene-Tetrafluoroethylene (ETFE) Composite Membrane Under Wind Loading
Previous Article in Journal
Comparative Study on the Impact of Various Non-Metallic Fibres on High-Performance Concrete Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Fatigue Failure of Adhesive Joints in Fiber-Reinforced Composite Material Under Step/Variable Amplitude Loading—A Critical Literature Review

Department of Mechanical and Materials Engineering, College of Engineering and Applied Science, University of Cincinnati, 2901 Woodside Drive, Cincinnati, OH 45221, USA
*
Author to whom correspondence should be addressed.
J. Compos. Sci. 2024, 8(11), 477; https://doi.org/10.3390/jcs8110477
Submission received: 18 September 2024 / Revised: 21 October 2024 / Accepted: 29 October 2024 / Published: 18 November 2024
(This article belongs to the Section Fiber Composites)

Abstract

:
Most fatigue-loading research has concentrated on constant-amplitude tests, which seldom represent actual service conditions. Because of the significant time and expense associated with variable-amplitude experiments, researchers often employ block/step-loading tests to evaluate the effects of variable-amplitude loading. These tests utilize various sequences of low-to-high and high-to-low loads to simulate real-world scenarios. Empirical investigations have shown inconsistencies in the damage accumulation under different load sequences. Although literature reviews exist for simulation and experimental methods, there is limited research examining the impact of step/variable-amplitude loading on adhesive joints in composite materials. This review aims to address this gap by comprehensively analyzing the effects of load sequence and block loading on fatigue damage progression in fiber-reinforced polymer composites. Additionally, the applicability of various step-loading fatigue damage accumulation models to adhesive materials is evaluated through numerical simulation to study its suitability in predicting fatigue failure. This review also explores recent theoretical advancements in this field over the past few years, examining more than 100 fatigue damage accumulation models categorized into seven subcategories: (i) linear damage rules, (ii) nonlinear damage curve and two-stage linearization models, (iii) life curve modification models, (iv) models based on crack growth concepts, (v) continuum damage mechanics-based models, (vi) material degradation models, and (vii) energy-based models. Finally, numerical simulations using the most common nonlinear cumulative fatigue damage accumulation models were conducted to predict fatigue failure in adhesively bonded joints under four step-loading tests, and the results were compared with the experimental data. Numerical simulations revealed the need and scope of further development of a fatigue failure model under step/variable loading. This comprehensive review offers valuable insights into the complex nature of fatigue failure in adhesive joints under variable loading conditions and highlights current state-of-the-art nonlinear fatigue damage accumulation models for adhesive materials.

1. Introduction

To accurately assess the fatigue characteristics of adhesive materials, it is crucial to replicate real-world fatigue loading conditions as closely as possible during testing. Adhesive joints under fatigue in real-life service usually encounter variable/step loading and rarely encounter a constant amplitude. However, fatigue testing is often performed at a constant amplitude owing to high cost and variable-amplitude experimental time, fatigue testing facility limitations, and the unclear nature of the in-service loading spectrum.
Despite its long history, cumulative fatigue damage remains an unresolved issue. In 1945, Palmgren and Miner introduced a mathematical model to address cumulative fatigue damage under variable loading, which can be expressed by Equation (1) [1], as follows:
D = n i N i
where D is the total damage, n i is the number of fatigue cycles under the ith constant-amplitude loading level, and N i is the number of fatigue cycles until failure under the ith constant-amplitude loading level. Building on Palmgren and Miner’s damage accumulation theory, numerous researchers have created models to describe how fatigue damage accumulates under varying fatigue load conditions. First, fatigue damage caused by a specific stress level must be determined. Second, the cumulative fatigue damage incurred at multiple stress levels must be accounted for accurately. Finally, the critical damage threshold that triggers failure must be identified when the material or structure reaches its limit.
Several experimental investigations were conducted on the fatigue life of adhesively bonded joints under block/variable loading conditions [2,3,4,5]. These investigations have uncovered notable inconsistencies between empirical results and conventional models that assume uniform fatigue damage accumulation. The observed discrepancies can be attributed to several factors.
  • Neglecting to factor in the load history;
  • Inability to consider the loading sequence when determining fatigue failure;
  • Disregarding the impact of load interaction effects.
Miner’s rule typically yields conservative fatigue life predictions for loading sequences that progress from low load to high load ( l o a d 1 < l o a d 2 ). Conversely, when the loading sequence transitions from high load to low load ( l o a d 1 > l o a d 2 ), the results tend to be nonconservative, as illustrated in Figure 1. The observed material behavior may be attributed to various factors, some of which are as follows:
  • Substantial strain often leads to early development of initial microscopic cracks. These tiny fissures then expand under less intense cycles, causing accelerated deterioration compared with conditions with consistent stress levels.
  • An increased number of load cycles on a specimen can lead to surface roughening during repeated plastic deformation. This roughening process generates additional locations where cracks could be initiated under lower load cycles.
  • It should be emphasized that the accumulation of damage is not a linear process.
To address these problems, researchers have developed several theories regarding nonlinear damage accumulation. However, these theories fail to account for one or more of the factors listed above [6,7,8].
The second section of the study examines various nonlinear cumulative fatigue damage accumulation models created to address the nonlinear nature of damage accumulation during step/variable loading these models, which focus on cumulative fatigue damage accumulation, can be classified into seven categories: (i) linear damage rules [1], (ii) nonlinear damage curve and two-stage linearization models [9,10,11,12,13], (iii) life curve modification model [14,15,16,17,18,19], (iv) models based on crack growth concepts [20,21,22,23,24,25,26,27], (v) continuum damage mechanics-based models [28,29,30,31,32,33,34], (vi) material degradation models [7,35,36,37,38], and (vii) energy-based models [39,40,41,42,43,44].
In the third section, several nonlinear cumulative fatigue damage accumulation models (discussed in Section 2) are used to predict the fatigue failure of adhesively bonded fiber-reinforced composite (FRP) joints under variable fatigue loading. The authors used a double-lap joint with four different step-loading cases (two high-load to low-load transition cases and two low-load to high-load transition cases) and compared the results with experimental data. This study also provides a comprehensive discussion of the limitations and applicability of fatigue damage accumulation models.

2. Fatigue Damage Accumulation Models

Expanding on the work done by Fatemi and Yang [3], the fatigue damage accumulation rules can be categorized into seven groups:
  • Linear damage rules (LDRs);
  • Nonlinear damage curve and two-stage linearization models;
  • Fatigue damage accumulation theories are based on life curve modification methods;
  • Fatigue damage accumulation approaches are based on crack growth concepts;
  • Fatigue damage accumulation theories are based on continuum damage mechanics;
  • Material degradation-based model;
  • Fatigue damage accumulation theory based on energy.
However, these categories are not clearly distinguishable from one another [9]. Some of the theories cannot be placed into these categories; these damage rules are placed in the miscellaneous category.

2.1. Linear Damage Rule (LDR) Models

The origins of fatigue damage modeling can be traced back to 1924, when Palmgren introduced the concept of linear summation of fatigue damage. In 1933, French conducted a notable study on the impact of overstress on endurance limits [45]. In 1938, Kommers suggested using changes in endurance limits as an indicator of damage [46]. In 1937, Langer suggested dividing fatigue damage into two phases, crack initiation and propagation [10], proposing a linear rule for each phase. These three early ideas—linear summation, endurance limit changes, and two-stage damage process–formed the basis for phenomenological cumulative fatigue damage models. In 1939, Gassner et al. 49 introduced the concept of linear damage accumulation in adhesive materials [47].
Miner introduced the first rule for fatigue damage accumulation under variable loading in 1945 [1]. Miner’s rule quantifies fatigue damage as a cycle ratio of applied number of cycles vs. total number of cycles to failure at any given load, which can be expressed as
D = n i N i
where D is the total damage, n i is the number of applied fatigue cycles under the ith constant amplitude loading level, and N i is the number of fatigue cycles until failure under the ith constant amplitude loading level.
The linear damage rule (LDR) postulates a consistent amount of work performed per cycle and the properties of the work absorbed at the point of failure. Material degradation occurs because of the accumulation of energy, resulting in a linear addition of the cycle ratio, which is expressed as D = n i N i = 1 . The progression of damage according to the LDR is illustrated in Figure 2.
Machlin introduced an alternative form of LDR in 1949 and proposed a metallurgically based cumulative damage theory [48]. During the 1950s, Coffin et al. expressed LDR in terms of the plastic strain range, linking it to the fatigue life through the Coffin–Manson relationship [49,50]. Later, Topper et al. utilized strain-based LDR to analyze experimental results [51,52,53,54,55]. Miller reviewed LDR applications in strain-controlled fatigue damage analyses in 1970 [56].
Jeans et al. first investigated the spectrum loading effects on adhesively bonded composite materials in 1983 [57]. Subsequently, Yang et al. explored the impact of service-loading spectra on fatigue damage, focusing on the reduction in fatigue life and residual strength in composites [58]. Jones et al. studied fatigue failure under adhesive bonding with variable loading by using LDR in 1989 [59]. In 2001, Erpolat improved the LDR by using a cycle mix fracture for adhesively bonded composite materials [60].

2.2. Non-Linear Damage Curve and Two-Stage Linearization Models

Richart and Newmark introduced the damage curve concept in 1948 to address the shortcomings of LDR by linking the damage to the cycle ratio [61]. They hypothesized that the stress levels influenced the D-r curves, where r denotes the cycle ratio n/N. Building on this foundation, Marco and Starkey proposed the nonlinear damage rule (NLDR) in 1954. This rule posits that the damage progression follows a power relationship with the number of cycles [9], as follows:
D = n i N i x σ i
where D is the total damage, n i is the number of applied fatigue cycles under the ith constant-amplitude loading level, N i is the number of fatigue cycles to failure under the ith constant-amplitude loading level, and x σ i is a variable related to the ith load level. For LDR, the exponent x σ i = 1 .
A comparison of the linear damage rule, Freudenthal–Heller rule (also known as the two-stage linearization rule), and damage curve approach (DCA) rule for variable fatigue loading is illustrated in Figure 3.
Langer proposed that the accumulation of damage be evaluated using two LDRs for each load level [10]. In 1960, Grover introduced a two-stage linear damage rule based on observations of fracture mechanics [62]. The initial stage represents crack initiation, whereas the subsequent stage represents crack propagation. Although Grover’s approach was qualitative, it lacked a quantitative formula for separating the total life into the initiation and propagation phases. Manson proposed a double linear damage rule (DLDR) in 1966 based on Grover’s work [62,63]. In the following year, Manson et al. modified the DLDR to represent two effective stages of the fatigue process rather than literal crack initiation and propagation [64]. This modification made the DLDR dependent on the material and load, addressing some of the limitations of their original proposal. Experimental comparisons between DLDR and Miner’s rule by Manson and Halford revealed that Miner’s rule led to an average overestimation of 37%, while DLDR resulted in an average overestimation of only 12% [11,65].
Following the work of Marco and Starkey, Manson and Halford built upon the findings of DLDR to propose a damage curve approach (DCA) [9,11]. The damage exponent for multi-level loading sequences in NDLR was empirically formulated by Manson and Halford, which is represented by the equation [11]
D = n 1 N 1 α 1 , 2 + n 2 N 2 α 2 , 3 + n 3 N 3 α 3 , 4 + + n i 1 N i 1 α i 1 , i + n i N i = 1
where
α i 1 , i = N i 1 i 0.4
where D is the total damage, n i is the number of applied fatigue cycles under the ith constant-amplitude loading level, N i is the number of fatigue cycles to failure under the ith constant-amplitude loading level, and α i 1 , i is a variable related to the ith load level.
In the case of multilevel loading sequences, a minor discrepancy was observed between the DLDR and DCA approaches, despite the DLDR being considered a linear damage concept. It is crucial to recognize that, while DLDR does not consider the loading sequence, DCA incorporates its effects. Numerous studies have performed experimental comparisons between the NLDR, DLDR, and DCA methods. These investigations consistently demonstrated that DLDR yields the most precise results [11,34,66,67,68].
Although the NLDR, DLDR, and DCA models consider load-level dependency, they fail to account for the effects of load interactions and cycles with amplitudes below the fatigue threshold [6,7,8]. Corten and Dolan tackled this limitation by introducing a novel theory for fatigue damage accumulation. They hypothesized that fatigue damage results from the formation of microscopic voids, leading to crack initiation and growth [13]. This phenomenon can be mathematically expressed using the following Equation (6):
D = m r n a
where n is the number of cycles applied; m is the number of damaged nuclei; r is the damage propagation rate coefficient, which is a function of the stress condition; and a is the material constant related to material failure. In the context of two-level block-loading experiments, Corten and Dolan proposed a stress-dependent ratio denoted as R, which is defined as R = σ 1 / σ 2 . This concept led to the development of a novel empirical equation to determine the number of cycles until failure occurred, as follows:
D = n i N i σ i σ i 1 d
where D represents the total damage and n i denotes the number of applied fatigue cycles under the ith constant-amplitude loading level. N i signifies the number of fatigue cycles leading to failure under the same ith constant-amplitude loading level. The load level at the ith load block is represented by σ i , and σ i 1 indicates the load level at the preceding load block. The variable D corresponds to the gradient of a linear curve in the l o g R 1 / a l o g σ 2 / σ 1 diagram.
Chen et al. examined the fatigue life of 304 stainless steel under various loading sequences, including axial/torsional, torsional/axial, in-phase/90° out-of-phase, and 90° out-of-phase/in-phase [69,70]. Their research also evaluated NDLR, DLDR, and DCA, which yielded non-conservative outcomes compared with experimental tests. To address the impact of the loading path under variable loading conditions, researchers incorporated a non-proportionality function J into the DCA model exponent α based on experimental findings. For two-block loading, the exponent α was defined as
α i 1 , i = 1 1 + β J N i 1 N i 0.4
where
J = 1.57 T ε 1 , m a x T 0 s i n ξ t ε t d t
where ε 1 t and ξ t are the absolute value and the angle of the maximum principal strain at time t, respectively. T and ε 1 , m a x are the time for a cycle and the maximum value of ε t in a cycle.
Based on Corten and Dolan’s work, Xu et al. acknowledged that DCA failed to account for the loading sequence [13,71]. They proposed modifying the parameter α i 1 , i to incorporate the load interaction and effective stress related to loading. Inspired by Corten and Dolan’s DCA research and Freudenthal and Heller’s studies on load interaction fatigue models, Gao et al. refined the parameter α i 1 , i by incorporating the minimum ratio of the applied stress amplitude [13,72,73,74], which can be expressed as
α i 1 , i = N i 1 N i 0.4 m i n σ i 1 σ i , σ i σ i 1
where α i 1 , i is the exponent factor in the DCA equation; N i and N i 1 are the numbers of cycles to failure in the current load and the previous load, respectively; and σ i and σ i 1 are the current load and the previous load, respectively.
Drawing on earlier research by Gao et al., Yuan et al. proposed that the assessment of fatigue damage should also consider the deterioration of the material strength [74,75]. Consequently, the formula for total fatigue damage D was enhanced to
D = γ n i N i N i 1 N i 0.4 m i n σ i 1 σ i , σ i σ i 1
where
γ = e x p α 1 A 1
where A is the residual strength of the degradation coefficient that reflects the relationship between the strength degradation and fatigue damage accumulation, and α is a material coefficient obtained from the experimental data.
In 2019, Zhou et al. introduced a novel fatigue damage accumulation model, drawing inspiration from the DCA and Corten and Dolan models [13,76], as follows:
F = n 1 N 1 N 1 N 2 α σ m a x , 1 σ f + n 2 N 2 N 2 N 3 α σ m a x , 2 σ f + + n i 1 N i 1 N i 1 N i α σ m a x , i 1 σ f + n i N i
where σ m a x , i is the maximum stress at the ith stress level, σ f is the fatigue limit, and α i is the damage exponent from the DCA equation.

2.3. Non-Linear Damage Accumulation Models Based on Wöhler (S-N) Curve

The Wöhler curve or S-N curve-based damage models modify damage evolution based on the evolution of the S-N curve. In 2011, Kwofie and Rahbar introduced a fatigue damage model based on the fatigue driving stress (FDS) based on the Basquin equation [14,77]. The authors stated that FDS is the primary factor in fatigue damage and can be utilized to estimate the remaining fatigue life under fluctuating loads. The FDS value calculated from the prior loads was used to forecast the remaining lifespan of the material. The FDS resulting from the applied cyclic stress σ i can be represented as follows:
S D i = σ i N b n i N i
where n i is the number of cycles applied at ith load, N i is the number of cycles to failure at ith load, and b is the fatigue strength exponent. Based on this, Kwofie and Rahbar described the damage due to fatigue as
D = n i N i l n N i l n N 1
where n i is the number of cycles applied at the ith load; N i is the number of cycles to failure at the ith load, and N 1 is the number of cycles to failure in first load. The ratio l n N i / l n N 1 accounts for load sequence and interaction effects.
Based on this Kwofie and Rahbar damage model, Zuo et al. and Zhu et al. developed a novel fatigue damage accumulation rule that incorporates load interaction effects using the FDS method [14,15,77,78]. This newly proposed rule enables the estimation of the remaining fatigue life and can be expressed as
n i N i = 1 n 1 N 1 n 2 N 2 n i 1 N i 1 l n N 1 l n N i σ i 1 σ i
where n i is the number of cycles applied at the ith load, N i is the number of cycles to failure at the ith load, and σ i is the load in the ith cycle.
Similar to the Kwofie and Rahbar damage model, Si-Jian et al. introduced a damage model grounded in the Basquin equation, which only necessitates the Wöhler curve or S-N curve [17]. The damage accumulation parameter D is characterized as follows:
D = 1 n 1 N 1 σ 2 σ 1 + n 2 N 2 σ 3 σ 2 + + n i 1 N i 1 σ i σ i 1 + n i N i
where n i is the number of cycles applied at the ith load, N i is the number of cycles to failure at the ith load, and σ i is the load in the ith cycle.
A damage model based on the Wöhler (S-N) curve was introduced by Aeran et al. without the inclusion of any additional parameters [16,79]. The model depends exclusively on the number of cycles to failure ( N i ). The absolute value of the proposed damage index D i represents the fatigue damage D, expressed as D = D i ,
D i = 1 1 n i N i δ i
where
δ i = 1.25 l n N i
Additionally, Aeran et al. introduced an alternative damage model incorporating a novel factor µ, which considers the effects of load sequence and load interaction [16,79]. This interaction factor is characterized as follows:
μ i + 1 = σ i σ i + 1 2
where σ i and σ i + 1 represent consecutive stress levels and μ i + 1 denotes the interaction of the loads between these levels. The authors used parameter μ i + 1 to redefine the damage parameter D, which can be expressed through the following set of equations:
D i = 1 1 n i + 1 , e f f N i + 1 δ i + 1 μ i + 1
where
n i + 1 , e f f = 1 1 D i μ i + 1 δ i + 1 . N i + 1
This concept can be expanded by defining n i + 1 as the number of cycles for the stress state σ i + 1 ; the total number of cycles of the loading step i+1 is defined as
n i + 1 , t o t a l = n i + 1 , e f f + n i + 1
D = D i + 1 = 1 n i + 1 , t o t a l N i + 1 δ i + 1
A novel approach to predicting fatigue life was introduced by Theil, which utilizes linearized damage growth curves on a double linear Wöhler (S-N) curve, with the S-N curve indicating failure [67]. This technique aims to estimate overloads that are approximately 0.2% of the yield strength and slightly exceed it by approximately 10%. The damage-calculation process involves iterative computations.

2.4. Life Curve or Iso-Damage Line Modification Models

Life curve modification models were developed to adjust the Wöhler (S-N) curve by considering various factors, including the load sequence and interaction effects. Modification of life curves or iso-damage lines can be divided into two primary categories.
  • Models based on the Corten–Dolan and Freudenthal–Heller models;
  • Models based on the iso-damage line.

2.4.1. Models Based on the Corten–Dolan and Freudenthal–Heller Models

Both the Corten–Dolan and Freudenthal–Heller models exhibit a common feature: the S-N curve rotates in a clockwise direction around a specific reference point [72]. Based on the Corten–Dolan model, Kaechele discovered that a larger number of damaged nuclei caused by high-stress amplitudes resulted in increased growth. This suggests that the nuclei were damaged at the lower stress levels [13,80]. Freudenthal and Heller set the low load cycle to high load cycle transition as a benchmark, which is between 103 and 104 cycles [72,73].
Owing to the intricate interaction effects described by Freudenthal and Heller and Corten and Dolan, the Freudenthal–Heller method is more challenging to align with experimental data than the Corten–Dolan model. Consequently, researchers tend to prefer the Corten–Dolan model. Spitzer and Corten determined the gradient of a modified N curve by averaging the slopes of several S-N curves derived from two-level block loading and tests [81]. Numerous studies have showcased the wide range of applications and high level of accuracy of the Corten–Dolan model [37,82,83,84,85].
In contrast to the exponent d in the Corten–Dolan equation as a material constant, Zhu et al. proposed expressing it as a function of the applied stresses [7,86,87]. They introduced a revised formulation for the parameter d, which is presented as follows:
d σ i = μ σ 1 λ δ f 1 λ σ i
where μ represents a material constant and λ is a coefficient that considers the effects of the load sequence, ranging from 0 to 1 (0 <  λ  < 1). For constant-amplitude loading, λ equals 0 because of the absence of load interaction. To simplify the calculations, λ was approximated as n1/N1. The variable δ f denotes the initial strength of the specimen, which was determined experimentally.
Expanding on a research by Zhu et al., Gao et al. enhanced the definition of the damage exponent d by incorporating a material parameter γ . This parameter was obtained from the experimental results and selected failure criterion, such as n/N = 1 [7,86,87,88].
d = e x p n i N i σ i / σ m a x + γ
where n i is the number of cycles applied at the ith load, N i is the number of cycles to failure at the ith load, σ m a x is the maximum load applied, and σ i is the load in the ith cycle.
The model of Gao et al., when compared with Zhu’s model, reduced errors by 50% [88]. Xue et al. further introduced the stress ratios of consecutive blocks [89], as follows:
d = σ i σ i 1 e x p n i N i σ i / σ m a x + γ
where n i is the number of cycles applied at the ith load, N i is the number of cycles to failure at the ith load, σ m a x the maximum load applied, and σ i is the load in the ith cycle.
Liu et al. introduced nonlinearity to the effects in the stress ratios of consecutive blocks by introducing the parameter k, which is expressed as
d = σ i σ i 1 k e x p n i N i σ i / σ m a x + γ
where n i is the number of cycles applied at the ith load, N i is the number of cycles to failure at the ith load, σ m a x is the maximum load applied, σ i is the load in the ith cycle, and k is a user-defined constant.
A novel approach was developed by Leipholz et al. by replacing the conventional S-N curve with an S- N ^ model. This new model accounts for life-reducing load interaction effects that occur between cycles of varying amplitudes [90,91,92,93]. The researchers formulated their model as follows:
N = i = 1 k β i N ^ i 1
where
β i = n i N i
where N is the total accumulated fatigue life and N ^ i is the modified life at the ith load. N ^ i is obtained by comparing the virgin S-N data and the experimental data.

2.4.2. Life Modification Models Based on Iso-Damage Lines

The S-N curve is commonly employed to assess the fatigue life of a structure and determine its remaining lifespan after a certain number of fatigue cycles. The iso-damage line concept, derived from the S-N curve, posits that the curve represents 100% fatigue damage, whereas stress-cycle combinations ( σ i , n i ) with equal damage values form smooth curves when graphed. In 1976, Subramanyan introduced a set of straight iso-damage lines on an S-log(N) diagram, converging near the knee point of the S-N curve, which is demonstrated in Figure 4 [18].
Damage was quantified as the ratio between the iso-damage line slope inclination and the limiting slope of the S−N curve. Expressing the damage as a logarithmic function of load cycles implies that the cycle count required for a specific damage amount increases with decreasing stress amplitude and decreases with increasing stress amplitude. This model accounts for the extended lifespan observed in high-to-low block-loading sequences and the shortened lifespan observed in low-to-high block-loading sequences.
D i = t a n θ i t a n θ f = l o g N f l o g N i l o g N f l o g n i
where N f represents the cycle count to failure at the knee point of the S-N curve, N i is the cycle count to failure at the ith load, n i denotes the cycle count at the ith load, and σ i is the load in the ith cycle. The iso-damage line concept can be utilized to calculate the equivalent cycle count n i , i + 1 in the i+1-th load σ i + 1 cycle. The equivalent cycle count n i , i + 1 can be determined as follows:
n i , i + 1 = N i + 1 n i N i α i
where
α i = l o g N f l o g N i + 1 l o g N f l o g N i = σ i + 1 σ f σ i σ f
Form the above equations, the residual fatigue life N i + 1 R at the (i+1)-th load cycle can be defined as
N ( i + 1 ) R = N i n i + 1
Based on iso-damage lines, Hashin and Rotem proposed a novel method for predicting damage. According to the authors, the iso-damage line converges at a point on the S-axis instead of the N-axis, which is in contrast with the Subramanyan model [94]. Figure 5 demonstrates Hashin and Rotem life modification models.
When both models are compared, the Subramanyan model becomes more favorable since the Hashin–Rotem model becomes invalid at low-stress levels [18,94,95]. However, the Subramanyan model does not consider cycles within the fatigue limit [18]. Multiple studies have been carried out in the Subramanyan and Hashin–Rotem models, and it was found that although both models outperformed Miner’s rule model, the results were still non-conservative in nature [96,97,98,99,100].
Research conducted by Hu et al. examined the use of 2D and 3D Hashin models to forecast the damage behavior and failure mechanisms in carbon fiber-reinforced aluminum laminates (CARALLs) [101]. Researchers found that both 2D and 3D Hashin models closely predicted the tensile and bending properties. In contrast, the 3D Hashin model considers out-of-plane stress components and employs an additional element deletion technique to calculate the damaged area. This approach leads to enhanced precision and prevents substantial mesh distortion.
Rege and Pavlou enhanced the non-conservative predictions of the Subramanyan model by introducing nonlinearity through the parameter q ( σ i ) , which varies based on stress amplitude [18,102]. This modification aims to address the limitations of the original Subramanyan model.
D i = t a n θ i t a n θ f q ( σ i ) = l o g N f l o g N i l o g N f l o g n i 1 , i + n i q ( σ i )
where
l o g n i 1 , i = l o g N f l o g N f l o g N i D i 1 1 / q ( σ i )
where N i represents the number of cycles until failure at the ith load level, N f denotes the number of cycles at the fatigue limit, n i indicates the number of cycles at the ith load level, and q ( σ i ) is a function of the stress amplitude, which can be expressed as follows:
q σ i = a σ i b = 2 σ i σ s b
where σ s represents the stress amplitude value at the point where the S-N curve intersects the stress axis. The variable b is specific to the material, with steel having a b value of −0.75. There was no significant improvement in Rege and Pavlou’s fatigue model for steel compared with Subramanyan’s model [18,102].
To address these inconsistencies, a new formula for q ( σ i ) was introduced by Zhu et al. [103], as follows:
q σ i = l l o g n i + s l o g ε a , i
where ε a , i represents the amplitude of the strain for the ith load, whereas l and s denote the coefficients for the load sequence and load weighting, respectively.
A novel fatigue damage model termed the S-N fatigue damage envelope was introduced by Pavlou and incorporated alterations to the iso-damage line [104]. The authors suggest that the area bounded by the S-axis, N-axis, and S-N curves represents the macroscopic effects of the damage mechanism for any given S-N curve pair. In this model, dimensionless S-N axes are established by defining σ i * and n i * , as follows:
σ i * = σ i σ f σ u σ f
n i * = n i N i
where σ i represents the amplitude of the stress for the ith block of loading, σ f denotes the material’s fatigue threshold, and σ u indicates the maximum tensile stress that the material can withstand before failure.
A significant limitation of life curve modification models is their difficulty in calculating the damage during frequent load amplitude fluctuations. Furthermore, these models employ bilinear S-N curves, which are not suitable for all scenarios. Several alternative approaches have been developed to overcome these challenges.
In 2013, El and Khaled introduced the virtual target life curve (VTLC) model, which proposes that materials possess a virtual expected life exceeding their actual life under a constant load amplitude [105]. The model suggests that as the number of cycles increases, the expected life of the material decreases, and the load amplitude influences the damage life. In addition, the VTLC model accounts for the overloading effects on the materials. Dan et al. compared the fatigue life of asphalt under two-stage loading with theoretical fatigue models, including VTLC [106]. Liu et al. developed a novel fatigue damage model by integrating VTLC and Manson models to predict material fatigue life [63,105,107]. However, this new model failed to demonstrate significant improvements in the fatigue life estimation for 300 CVM steel compared with the previously discussed models.
To address these problems, Batsoulas introduced a hyperbolic iso-damage curve damage model [108]. The hyperbolic curves can be defined as
l o g σ σ f = c l o g N N e 1
where σ f represents the fatigue strength coefficient, N e denotes the cycle count necessary for damage initiation, and c is a constant value.
For each individual curve, the damage accumulation can be defined by
D i = l o g σ i σ f l o g n i N e l o g σ i σ f l o g N i N e
Based on the above equation, the rule for accumulating damage under multiple levels of loading can be formulated as
D i = n 1 N 1 φ 2 , 3 + n 2 N 2 φ 2 , 3 + + n i N i 1 φ i 1 , i + n i N i
φ i 1 , i = l o g σ i 1 / σ f l o g σ i / σ f
Batsoulas and Giannopoulos used this model to compute the fatigue life of composite material under step/variable loading [109].
Xie et al. modified the damage accumulation exponent φ i 1 , i [110]. They integrated the Batsoulas model with the Ye and Wang model, drawing from the material degradation model [35,108,109]. The resulting modified exponent φ i 1 , i can be described as follows:
φ i 1 , i = l o g σ i 1 / σ f l o g σ i / σ f l o g N f l o g N i l o g N f l o g N i 1

2.5. Crack-Growth-Based Fatigue Damage Accumulation Models

In 1921, A. A. Griffith first described the concept of linear elastic fracture mechanics (LEFM) leading to material failure [111]. During the 1950s and the 1960s, several fatigue crack growth theories based on LEFM were introduced and gained widespread acceptance owing to the direct relationship between cracks and damage [112,113,114,115,116,117]. According to Makkonen, the total fatigue life of a material can be divided into three stages [118].
  • Crack initiation;
  • Stable crack growth;
  • Unstable crack growth.
Crack initiation typically accounts for 40–90% of the total fatigue life of a material [118].
In 1963, Paris and Erdogan developed a formulation for stable fatigue crack growth [115], which is expressed as
d a d N = C Δ K m
where a represents the depth of the crack, K denotes the material’s stress intensity, and C and m are constants specific to the material properties.
Wheeler introduced a modification to the Paris law’s constant amplitude crack growth rate in 1972 by adding a retardation factor R i [115]. The authors proposed that crack propagation occurs owing to the interaction of plastic zones near the crack tip, which develops as a result of residual compressive stresses caused by overloading. This modified law expresses the crack growth rate as follows:
d a d N = R i C Δ K m
R i = r p i r m a x p
where r p i represents the size of the plastic zone corresponding to the ith loading cycle, and r m a x denotes the distance between the current crack tip and the largest previous elastic-plastic zone generated by the overload. The variable p is an empirical parameter that shapes the equation based on the material characteristics and spectrum of the applied loading.
The Willenborg–Wood model, an extension of the Wheeler model, incorporates a retardation mechanism that employs the effective stress intensity factor at the crack tip ( Δ K e f f ) [20,21]. This factor diminishes the applied crack-tip stress intensity factor ( Δ K ) owing to the increased compressive stress at the crack tip resulting from the overloads. The reduction in the applied Δ K is determined by comparing the current plastic zone size at the ith load cycle with the maximum plastic zone size generated by the overload. Unlike its predecessor, the Willenborg model eliminates the need for an empirical shaping parameter.
Elber proposed that fatigue-induced cracks can close because of the formation of a compressive residual stress zone at the crack tip [22,23]. This phenomenon reduces the driving force for crack propagation, resulting in decreased stress intensity at the crack tip, which is influenced by the effective stress range.
Δ S e f f = S m a x S o p
where Δ S e f f represents the effective stress, S m a x denotes the maximum stress, and S o p signifies the crack-tip opening stress.
Newman et al. formulated a model for fatigue crack growth under variable loading conditions, focusing on short-crack closures [119,120,121,122,123,124,125]. Their approach, reminiscent of Wheeler’s model, involved redefining the effective stress intensity to characterize crack propagation [126,127]. Newman proposed that stress intensity could be expressed as
Δ K e f f = S m a x S o p × π c × F c / w
where Δ K e f f represents the effective stress intensity factor and S m a x denotes the maximum stress. The term S o p indicates the stress at which the crack tip opens, and F is a factor that corrects for the boundary conditions.
Using the finite element method (FEM), Newman et al. developed an iterative approach to estimate the crack opening stress through cycle-by-cycle closure analysis [126,127]. Ritchie and Suresh accounted for environmental factors such as corrosion and fracture surface roughness during crack fracture growth [128,129]. Furthermore, several variable loading fatigue models based on crack-closure fatigue have been developed [130,131,132,133]. However, these models have limitations, including the exclusion of plasticity or nonlinear behavior, difficulty in application to complex variable amplitude loading, and a limited scope for small crack growth theories.
To address these concerns, Miller and Zachariah introduced a double exponential crack growth fatigue law that established an exponential relationship between the crack length and elapsed life for each phase [134,135]. Later, Miller and Ibrahim enhanced this model by identifying stages I (crack initiation) and II (stable crack growth) within the context of the linear elastic fracture mechanics (LEFM) theory [135,136].
d a d N = ϕ Δ γ p α a
where ϕ and α are constants specific to the material, Δ γ p denotes the range of the plastic shear, and a indicates the current length of the crack. A power function was utilized to establish the relationship between the number of cycles N i and instantaneous crack length a i with their corresponding values in the plastic strain range Δ γ p . The equation for damage is expressed as follows:
D = a a f = a 1 a f 1 r 1 r I
where a f is the final crack length, r is the cycle ratio, and r I is the cycle ratio in stage I.
However, the double exponential crack growth model, which was developed within the framework of LEFM, does not consider plasticity in fatigue crack growth.
To address this issue, Miller et al. introduced a novel fatigue damage accumulation model focused on crack growth [137,138,139,140,141,142,143]. Their research highlighted that, in metal fatigue, crack initiation occurs at the onset, and the entire fatigue lifespan consists of crack propagation starting from an initial flaw size a 0 .
Miller et al. combined linear elastic fracture mechanics (LEFM) and elastic-plastic fracture mechanics (EPFM) theories to develop crack growth fatigue failure theory [137,138,139,140,141,142,143]. The authors defined two different elastic-plastic fracture mechanics phases: (i) microstructurally short cracks (MSCs) and (ii) physically short cracks (PSCs). The crack growth model for MSCs is defined as follows:
d a d N = A ( Δ γ p ) α ( d a )
where, a 0 < a < a t .
Similarly, the crack growth model for MSCs is defined as
d a d N = B Δ γ p β a C
where, a t < a < a f ,
where A, B, α , and β are determined by fitting experimental data to the model. Δ γ p represents the range of shear strain, while a t denotes the crack length at which the transition from microstructurally small cracks (MSCs) to physically small cracks (PSCs) occurs. The variable a f is defined as the crack length at failure, and C indicates the rate of crack growth under threshold conditions.
Dugdale and Barenblatt introduced a traction–separation law to model the crack initiation and propagation along ductile material interfaces, drawing from the short-crack regime and Paris’s law [144,145]. Turon derived the overall damage accumulation rate d by examining the relationship between the total damage parameter d = d s + d f and the area encompassed by the cyclic loading curve in the traction–separation law [146]. Harper established the fatigue damage accumulation rate d f by suggesting the concept of fatigue crack length across an element, and Landry utilized the same idea to derive the fatigue damage accumulation rate d f [147]. The research conducted in these investigations offers a significant understanding of how FCGR and the bilinear cohesive law are interconnected. Tabiei and Zhang presented a new bilinear law that connects FCGR to the damage parameter, and after obtaining the damage parameter, it is used to reduce the critical energy release rate in the cohesive zone model [148,149]. The damage accumulation under bilinear traction separation law regime can be defined as follows:
For mode I:
d f N = b A c z , I a N = 32 T 2 9 π E G I C C Δ G m
For mode II:
d f N = b A c z , I I a N = S 2 E G I I C C Δ G m
where b is the mix ratio of mode I and mode II loads; A c z , I and A c z , I I are the cohesive zone size ( A c z ) for mode I and mode II; Tand S are tractions in the normal (mode I) and shear (mode II) direction; E n and E t are the elastic modulus in normal (mode I) and shear (mode II) directions; Δ G is the energy release rate; and C and m are Paris law parameters, which are experimentally obtained for mode I and mode II. Figure 6 demonstrates the damage accumulation in Bilinear Traction Separation Law.
Ma and Laird expanded upon the short-crack regime theory by introducing the concept of crack population, P, which is similar to MSC, and is linearly related to the strain damage amplitude and used life [25,150]. This concept can serve as a damage indicator, as the authors defined cumulative damage accumulation as
D = P i P c r i t = K n i Δ γ p / 2 i α i Δ γ p / 2 l i m i t
where Δ γ p / 2 l i m i t represents the strain at which fatigue limit occurs. K = C / P c r i t , where C is a constant found in the strain-life equation, and P c r i t denotes the critical crack population when failure happens. Additionally, α i signifies the factor related to loading history at the ith level of load, and is defined as follows:
α i = Δ γ p / 2 i / Δ γ p / 2 m a x i n p r e l o a d i n g h i s t o r y
This model considers the loading sequence, that is, the high-to-low load transition and low-to-high load transition.
Valek and Polak proposed a new two-stage damage model based on experimental observations and analysis [151]. The crack initiation regime is described as follows:
d a d N = ν i
where a 0 < a < a c .
Meanwhile, Stable crack growth regime was defined by
d a d N = ν i + k a a 0
where a c < a < a f ,
where ν i represents the rate at which cracks grow, independent of the applied cycle. The constant k is utilized, while a 0 , a c , and a f denote the initial, critical, and final crack lengths, respectively.
Based on the above equations, the damage accumulation for stages I and II were defined through
D = 2 D c r
For initiation: D r 1 / 2
D = D c + D c m e m 2 r 1 1
For propagation: 1 / 2 r 1
where r is the n/N, D c is the critical damage ( a c / a f ), and m is a constant that can be defined as follows:
m = k N f 2
In 2005, Noroozi et al. introduced the UniGrow fatigue crack growth rate model based on the EPFM crack growth theory for cumulative fatigue damage accumulation. This model builds on the work of Walker and Kujawski [27,46,152,153,154,155]. The UniGrow model presents a novel method for calculating the stress intensity factor, resulting in a new definition of crack propagation. The research concluded that crack propagation can be characterized as
d a d N = ρ * / N f
where
ρ * = ψ y , 1 2 2 π Δ K t h Δ σ ˜ t h a 2
The UniGrow fatigue crack growth rate model was developed by Mikheevskiy et al., who analyzed the elastic-plastic stress/strain behavior near the crack tip [26,156]. This model underwent modifications to accommodate various variable loading spectra and later incorporated the environmental effects on cumulative fatigue damage under variable loading. Abduallah et al. enhanced the UniGrow model developed by Noroozi et al. by integrating additional crack growth rate models [27,46,152,153,154,155,157,158,159]. The impact of plasticity-induced crack closure using the UniGrow model was investigated by Pedrosa et al. [160]. Bang et al. employed the UniGrow model to explore fatigue damage accumulation under variable loading for different aluminum alloys in both short and long fatigue crack regimes [161,162,163].
Other models have been created that employ statistical or probabilistic techniques to determine the cumulative fatigue damage under crack growth conditions. One such model was developed by Manjunatha, who proposed using the root mean square of the stress intensity factor ( Δ K r m s ) in the fatigue crack growth rate regime to account for cumulative fatigue damage under spectrum loading [164]. Several authors have developed probabilistic models for cumulative fatigue damage under a crack growth regime. Nonetheless, these models have several limitations, including the need for experimental data, sensitivity to assumptions, limited applicability across various problems, and unreliable estimation parameters [165,166,167,168,169,170,171,172,173,174].

2.6. Continuum Damage Mechanics-Based Models

Continuum damage mechanics (CDM) is a branch of mechanics that investigates the initiation and evolution of material damage. In CDM, damage is considered a measurable and quantifiable physical phenomenon. This theory emerged from the studies of Kachanov and Rabotnov on creep damage issues [175,176,177]. Chaboche pioneered the application of CDM to fatigue life prediction by introducing a nonlinear continuum damage (NLCD) model that describes progressive deterioration prior to macroscopic crack formation [28,29,178,179]. This model extends the nonlinear damage mechanics models of Marco and Starkey and the DCA model of Manson and Halford, all of which are underpinned by the CDM principles [9,11,65,180]. Subsequently, Chaboche and Lense examined the CDM model for uniaxial loading [181], and damage was expressed as
D = 1 1 n N 1 1 α 1 1 + β
where
α = 1 a σ a σ l 0 1 b σ ¯ σ u σ m a x
and
N F = 1 1 + β 1 α σ m a x σ ¯ M 0 1 b σ ¯ β
where α , β , and M 0 represent material constants. The variable σ l 0 denotes the fatigue limit under fully reversed conditions, whereas σ u indicates the ultimate tensile strength. The applied stress is represented by σ a , where σ m a x signifies the maximum stress experienced.
Chaboche’s NLCD model presents multiple benefits compared with alternative theories regarding cumulative fatigue damage [28,29,178,179]. It considers damage growth beneath the initial fatigue threshold in damaged materials, incorporates interaction effects by including a variable for strain hardening, directly connects damage to mean stress, and allows for the easy determination of material parameters from standard S-N curves. Sun et al. [182] demonstrated these advantages.
Numerous fatigue damage accumulation models have been developed based on Chaboche’s work that encompass various variables, damage rate equations, and boundary conditions [183,184,185,186,187,188,189]. Bhattacharaya and Ellingwood created a continuum damage mechanics (CDM)-based model for predicting fatigue life based on fundamental thermodynamic principles and a stochastic ductile damage growth model [31,190]. This approach differs from other CDM-based methods by deriving damage accumulation equations from the first principles of mechanics and thermodynamics rather than starting with a dissipation potential function or a kinetic equation of damage growth. Damage accumulation is defined as follows:
d D ε p = A ε p 1 D ε p d ε p + B ε p d W ε p
where A and B are coefficients that depend on ε p , σ f , M, K, and ε 0 . ε p is the plastic strain, σ f is the failure stress, M and K are the hardening coefficients.
A thermo-elastoplastic damage model was created by Oller et al., who linked the remaining strength of materials to the evolution of damage thresholds [32,191,192]. To represent the nonlinear damage behavior observed during fatigue, they introduced a novel fatigue state variable, called the reduction function f r e d N , σ m a x , R , θ . This model differs from the approach adopted by Chaboche and Lense, as it does not explicitly define the number of cycles within the damage model [181].
Dattoma et al. introduced an innovative nonlinear fatigue damage model for uniaxial structures subjected to multilevel sequence loading, expanding on Chaboche’s initial framework [33]. The researchers proposed a new approach to damage accumulation for multi-level sequence loading, which is described as
D i = 1 1 n i + N i N f i 1 1 α i 1 1 + β
where N f i represents the number of cycles until failure at the ith load level, and N i denotes the equivalent number of cycles applied at the same load level. The latter produces an identical amount of damage to n i 1 at the previous load level (i−1). Furthermore, the authors redefined α as follows:
α = 1 1 H σ a σ f σ u σ a a
where a and H are experimentally obtained parameters.
According to Dattoma et al., the quantity N f i , representing the number of cycles until failure for the ith load, was established as follows [193]:
N f i = 1 1 α i 1 1 + β σ a M 0 β
This approach provided an accurate estimate of the linear correlation between log S and Log N31. Nevertheless, the primary limitation of Dattoma’s model is its requirement for S-N curve data specific to each geometry and various loading conditions [193]. Zhang and colleagues developed an enhanced version of the Chaboche and Lense model by integrating a strengthening function, f s , for cycles with low amplitude [194]. This function accounts for the inconsistencies found in low-amplitude cycles where the Chaboche model failed prematurely.
Further modification to the Chaboche and Lesne model was introduced by Zhang et al., who incorporated a strengthening function ( f s ) to account for low-amplitude cycles [195]. This function was designed to address the shortcomings identified in the Chaboche model for low-amplitude cycles, where it would prematurely fail [196,197,198,199]. The strengthening function ( f s ) is defined as follows:
f s = 1 , σ i 0 , σ f e m σ i , σ i σ f , σ 0
where m represents the coefficient of strengthening, which depends on the characteristics of the material and can be determined through experimental methods. Additionally, σ f denotes the minimum stress required for strengthening.
Li et al. developed a novel fatigue model to evaluate damage accumulation in bridges subjected to traffic loads, building on the Lemaitre CDM fatigue model [184,200]. This enhanced model incorporates the concepts of microplastic strain, strain energy density release rate, and current damage state. Within this framework, damage is characterized as
D i + 1 = 1 1 D i α 1 + 1 α 1 + 1 N b l i B ( β + 3 ) j = 1 m r b σ j + 2 σ m j σ j β + 3 2 1 D i α 1 α j 1 1 + α
where
α i = k α σ i + α 0
where the material parameters B, β , k α , and α 0 can be derived from the S-N curve. m r b represents the cycle count where the peak stress in the representative block exceeds the fatigue threshold. N b l denotes the total number of loading blocks. For the jth cycle, σ j indicates the stress level, whereas σ m j signifies the mean stress.
Several studies have been conducted using this method to assess the damage in long-span bridges [200,201,202,203,204,205]. However, the primary drawback of the model is that it was specifically designed to address high-cycle fatigue and is not reliable under low-cycle conditions.
A novel damage stress model (DSM) based on cumulative fatigue damage accumulation was introduced by Mesmacque et al. This model postulates that the fatigue life of a material is solely determined by the loading conditions when the physical damage state remains constant [34], which is demonstrated in Figure 7. The authors defined the damage accumulation using the residual life ( N i R ) associated with the damage stress ( σ e d , i ) following n i loading cycles, which was derived from the Wöhler (S-N) curve.
The remaining lifespan ( N i R ) can be calculated as N i R = N i n i , where n i represents the number of cycles applied to the ith load. For multiple levels of loading, damage accumulation is defined as
D i = σ e d , i σ i σ u σ i = σ e q u i v σ i + 1 σ u σ i + 1
where σ u is the ultimate stress and σ e q u i v is the equivalent damage stress.
Several studies have examined the use of DSM to address multistep loading fatigue problems [100,206,207,208,209]. These studies have shown that the results are considerably more consistent than those obtained using Miner’s rule. The application of DSM to predict the fatigue life of riveted railway bridge components and corroded bridge elements was investigated by Siriwardane et al. [210,211]. Researchers have advised the use of DSM when stress histories are available. Nevertheless, a study by Rege and Pavlou on fatigue life estimations using DSM revealed poor performance when the S-N curve exhibited linearity in a semi-log plot [102].
Researchers have expanded DSM through various models. Pitoiset and Preumont introduced a method utilizing von Mises stress for multiaxial loading scenarios, and they shifted from the time domain to the frequency domain, drawing on the Crossland criterion [212,213]. Subsequently, Aid et al. formulated a novel fatigue damage model that integrated the material’s S-N curve, DSM, and Miner’s rules [214]. This model also incorporates the von Mises stress approach for multiaxial loading cases, as proposed by Pitoiset and Preumont, along with Sines and Crossland’s criterion [212,213,215]. Multiple iterations of the DSM have been developed by Socie et al., Shen et al., Benkabouche et al., Wang-Brown, and Lagoda-Macha, each employing distinct criteria for assessing cumulative fatigue damage accumulation [100,216,217,218,219,220,221].

2.7. Material Degradation Based Models

Numerous theoretical and practical investigations have shown that materials display varying static mechanical characteristics, such as Young’s modulus, yield stress, and ultimate tensile stress, under both their initial and post-fatigue conditions [222,223,224,225,226]. Consequently, researchers have formulated various fatigue damage accumulation models to explain the deterioration behavior of materials.
Through experimentation, Ye and Wang discovered that the internal energy of a material increases with each subsequent fatigue cycle, leading to an increase in dislocations and internal flaws [225]. Researchers introduced a novel fatigue damage model that considers damage as the irreversible dissipation of cyclic plastic strain energy, ultimately resulting in fatigue fracture. The newly proposed damage model is expressed as follows [35]:
D = D N f 1 l n N i l n 1 n i N i
where
D N f 1 = 1 σ a 2 2 E U T 0
where E represents Young’s modulus and U T 0 denotes the static toughness of the material before damage occurs. The applied stress amplitude is denoted by σ a , where N i denotes the number of cycles until failure at the ith load level. Additionally, n i represents the number of cycles applied at the ith load and D N f 1 represents the critical value of the damage variable.
Lv et al. further refined the Ye and Wang model by incorporating a load interaction factor, as proposed by Corten and Dolan and Morrow, to consider the load sequence effects [7,13,227]. The damage accumulation parameter becomes
D = D N f 1 l n N i l n 1 n i N i σ i σ m a x
Researchers have employed this model to predict the fatigue life of C45 and 16Mn steels under two distinct block-loading sequences. Unlike the Ye and Wang model, the Lv et al. model exhibited superior accuracy in predicting fatigue life, underscoring the importance of considering load sequence effects [7].
Böhm et al. introduced an alternative approach to material degradation. They formulated a fatigue damage accumulation model based on the psychological theory, specifically utilizing the Ebbinghaus forgetting curve. This exponential function describes how memory retention decreases over time when it is not actively maintained [36]. To represent the material degradation due to fatigue, the authors adapted the time function into a cycle-based function. The concept of material memory is defined as follows:
m = a c e + N f d + c
where m represents the material’s memory performance, a denotes the rate of memorization, d stands for the inverse of the forgetting rate expressed as the number of cycles, and c indicates the function’s horizontal asymptote. Based on this equation, researchers have developed models for damage assessment.
D = t = 1 f N f e n t d d 1 e n t d N i
In 2018, Peng et al. introduced a novel method that merges the material memory concept with a residual S-N curve [36]. Traditionally, the residual S-N curve illustrates the remaining lifespan of a material or component after exposure to multiple stress cycles. This curve typically maintains the same slope as the S-N curve of the virgin material but with a different S-intercept. However, the researchers proposed that the slope of the residual S-N curve, Δ b, is influenced by the loading history. At the outset, the slope of the residual S-N curve matches that of the virgin material (b); however, as fatigue damage progresses, the slope Δ b increases. The ratio of the slopes, b/ Δ b, is theorized to indicate the accumulation of fatigue damage. Drawing from the material memory degradation concept, a decay coefficient α is established that correlates with the slope ratio b/ Δ b. This memory degradation concept is employed to model the alterations in the residual S-N curve, yielding a more precise representation of the fatigue life of the material for the ith loading block.
b Δ b = α i = e n i N i e 1 1 e 1
Using this relation, Peng et al. redefined the accumulative fatigue damage accumulation rule [36], as follows:
D = 1 i n i N i × j = 1 i 1 N i N i + 1 k = 1 j e n k N k e 1 1 e 1 1
An evaluation of fatigue life prediction outcomes from two-block and multi-block models, using Miner’s rule and Kwofie and Rhabar’s rule, indicated that the model by Peng et al. yielded more precise results [1,36,77].
Zhou et al. introduced a novel fatigue damage accumulation rule, building upon the material memory degradation rule proposed by Bohm et al. [36,38]. The damage model is expressed as follows:
D i = 1 e n i N i 1 e 1 ζ σ m a x , i σ a , i δ
where ζ and δ are the fitting parameters of Zhou’s fatigue model. The authors used this model for fatigue life prediction of 30NiCrMoV12 steel; however, the residual life ζ was omitted without proper reasoning. Moreover, because this model uses fitting parameters, it cannot be generalized.
In 2020, echoing the work of Zhou et al., Chapman et al. introduced a model for the cumulative fatigue damage in anisotropic composite materials based on material degradation [38,228]. Their research postulates that the material strength ( R n ) decreases as a function of the number of cycles to failure (n), with the rate of this decline approximated using a power-law relationship.
d R n d n = A σ / m R n m 1
where A σ is a function dependent on the peak cyclic stress, and m is a constant. Furthermore, this study introduces elastic property degradation through the degradation parameter ( η ).
E n = η E s
where
η = 1 l o g n l o g 0.25 l o g N f l o g 0.25 λ 1 γ 1 σ E s ε f + σ E s ε f
where λ and γ are the experimental curve fitting parameters, E s is the static stiffness, and ε f is the average strain to failure.

2.8. Energy-Based Damage Rule

In the 1960s, Morrow and Halford initially established a failure criterion based on hysteresis energy and fatigue behavior, which Inglis introduced [11,39,229]. Subsequently, Zuchowski, Budiansky-O’Connell, and Glinka developed preliminary failure models that incorporated the strain energy [41,230,231]. Researchers have developed strain-energy-density-based fatigue damage rules for high-cycle fatigue, drawing on the observation that plastic energy dissipation ( Δ W p ) during cyclic loading can serve as a criterion for low-cycle fatigue failure [42,232]. However, this approach is restricted to low-cycle fatigue, because the plastic component Δ ε diminishes. Consequently, the plastic strain energy density Δ W p also approaches zero, rendering this method less effective for high-cycle fatigue.
To address this limitation, Ellyin et al. proposed a novel failure rule that combines high- and low-cycle fatigue by introducing the concept of total energy ( Δ W t ), which is the sum of elastic energy ( Δ W e ) and plastic energy ( Δ W p ) [232,233,234,235]. The researchers introduced a double logarithmic curve of SED versus fatigue life cycle ( l o g Δ W l o g N ), analogous to the Wöhler (S-N) curve. This curve is founded on a power law, which can be expressed as
Δ W t = Δ W e + Δ W p = κ N α + C
where κ , α , and C are material constants. The constant C denotes the proportion of elastic energy input in the tensile mode and is associated with the material’s fatigue limit strain energy density Δ W l i m . The elastic component of a material under tensile stress promotes crack propagation, whereas the plastic component ultimately leads to material degradation. This theoretical framework is applicable to both the Masing and non-Masing materials. For either type, a master curve can be generated by shifting the loop along its linear response segment (in the case of non-Masing materials). The cyclic plastic strain for the non-Masing behavior can be represented as
Δ W p = 1 n * 1 + n * Δ σ δ σ 0 Δ ε p + δ σ 0 Δ ε p
and the for the ideal Masing behavior can be represented as
Δ W p = 1 n 1 + n Δ σ Δ ε p
where n * and n are the cyclic exponents of the mater curve and the Masing material, respectively. Further developments have been made based on this theory, accounting for both the onset of crack formation and crack propagation.
Therefore, Ellyin et al. proposed the following fatigue damage law for a multi-block loading sequence [232,233,234,235]:
D = n 1 N 1 l o g N 2 N * l o g N 1 N * + n 2 N 2 l o g N 3 N * l o g N 2 N * + + n i 1 N i 1 l o g N i N * l o g N i 1 N * + n i N i
where N * corresponds to W * , which is extrapolated from the l o g Δ W l o g N curve and is referred to as the reduced fatigue limit.
Kreiser et al. introduced a novel energy-based nonlinear cumulative fatigue damage model. This model is particularly effective for materials and structures subjected to low-cycle, high-amplitude loading conditions where significant plastic strain is observed [236]. The authors employed Ellyin-Golos’s model, which considers loading sequence history, and integrated it with Coffin–Mason relations, which account for plastic strain in both stable and unstable hysteresis [234]. The proposed damage model is described as follows:
φ = f ψ , p m = 1 l o g 10 Δ W p / Δ W e
where φ is the cumulative fatigue damage, ψ is the cumulative damage parameter, and p m is the material parameter. Using the above equation, Manson derived the fatigue life of a material using the following equation:
Δ ε = Δ ε e + Δ ε p = 3.5 σ B E N f 0.12 + ε f 0.6 N f 0.6
where ε f is the fracture ductility, and σ B is the tensile strength of the material. Various studies have used this model to predict the fatigue life of materials under various loading conditions.
Leis introduced a nonlinear damage model based on energy that assumes the rate of damage is influenced by cycle-dependent alterations in the overall microstructure [43]. The damage rule he put forward is expressed as follows:
D = 4 σ f E 2 N f 2 b 1 + 4 σ f ε f 2 N f 2 b 1 + c 1
where
c 1 = 1 1 + 5 n 1
b 1 = n 1 + 5 n 1
where σ f is the fatigue strength of the material, ε f is the ductility coefficient of the material, and n 1 is defined as a function of the accumulation of the plastic strain. Xiaode et al. after conducting fatigue tests under a constant strain amplitude and discovering a new cyclic stress–strain relationship, researchers further developed Leis’s model [43,237].
Niu et al. observed that the strain hardening coefficient changed during cyclic loading, whereas the strain hardening exponent changed only negligibly [238,239]. The authors introduced a new cyclic stress-strain relation as follows:
Δ σ 2 = K Δ ε p 2 n * r β
where K * and n * are the cyclic strain hardening coefficients, and β is the cyclic hardening rate. β is found through
β = a Δ σ 2 Δ ε p 2 b
where a and b are constants.
Based on this, the damage is defined as
D = Φ 1 / n + α 1 + β = r 1 / n + α
where
α = Δ σ Δ ε p 4 2 b a
where n is the strain hardening parameter.
Radhakrishnan proposed that the crack growth rate is proportional to the plastic strain energy density, which accumulates linearly until failure [44,240]. For m-level load variation, an expression for predicting the remaining life fraction at the last load step is formulated as follows:
D m = 1 i = 1 m 1 W f i W f m D i
where W f i is the total plastic strain at failure for the ith load level, and W f m is the total plastic strain at failure for the mth load level. Subsequently, a similar approach was adopted by Kliman for the harmonic loading cycles.
Lagoda et al. showed in 2001 that the normal strain energy density on the critical plane effectively predicts high-cycle fatigue life under random and non-proportional cyclic loading [241,242]. They also introduced a novel strain energy density parameter that distinguishes between positive strain energy density in tension and negative strain energy density in compression. Researchers have applied this model to assess the fatigue life of steel and cast iron, yielding a strong experimental correlation.
Park et al. modified the study by Ellyin et al. by introducing the plastic strain energy consumed by mean stress [232,233,234,235,243]. This formulation for the total work conducted is expressed as
Δ W t = Δ W e + Δ W t + f ε m N f γ = κ N α + C
where f ε m represents the plastic strain energy associated with the mean strain and γ is the material constant. This model was employed to estimate the fatigue life of AZ31 alloy with anisotropic properties owing to rolling, and the predictions were in good agreement with the experimental findings.
A novel approach for predicting fatigue life based on energy was introduced by Jahed and Varvani-Farahani [244,245]. Their method utilizes the upper and lower limits of the experimentally determined lifespan of a material. The equation for calculating the number of cycles until failure ( N F ) is as follows:
N F = Δ E A Δ E N A + Δ E T Δ E N T
where Δ E is the total elastic-plastic energy, and Δ E A and Δ E T are the energies due to purely tensile loading and purely torsion loading, respectively. Furthermore, the authors extended this to non-proportional loading. Gu et al. used this model to predict the fatigue life of welded frames for mining trucks with a good experimental correlation [246].
Ozaltun et al. proposed a new energy-based fatigue damage accumulation model that considered the aging effects caused by cyclic loading on the fatigue strength of structural components in gas turbine engines [247]. The model also accounts for plastic strain energy dissipation. Fatigue life (N) was estimated as follows:
N = C σ n ε n σ 0 2 E + ε 0 σ 0 c o s h σ n σ 0 1 2 σ c σ a σ c s i n h 2 σ a σ c c o s h 2 σ a σ c 1 + β 1 2 ε f 2 ε n 2 + β 0 ε f ε n 2 σ c σ a σ c s i n h 2 σ a σ c c o s h 2 σ a σ c 1
where σ n is the necking stress, ε n is the necking strain, ε f is the fracture strain, σ a is the generalized cyclic stress, ε a is the generalized strain, β 0 and β 1 are the d intercept of the stress–strain relationship in the necking region, σ 0 , σ c , and C are the curve fitting parameters.
An energy-based variant of DSM by Mesmacque et al. [34,248,249] was introduced by Djebli et al. In this new approach, the cumulative damage accumulation parameter ( D i ) is established as
D i = W e d , i W i W u W i
where W e d , i represents the energy associated with the damage stress, W i denotes the energy related to the applied stress, and W u denotes the energy corresponding to the ultimate stress of the material. Researchers have employed this model to estimate the fatigue lives of metals and polymers [248,249].
Peng et al. introduced a novel method called the fatigue driving energy (FDE) approach, which integrates Kwofie and Rhabar’s fatigue driving stress model with strain energy density (SED) [14,250]. This model for fatigue damage accumulation is expressed as
D = W D W D 0 W D C W D 0 = N 2 b n N 1 N 2 b 1
where,
W D = 1 2 E σ 2 N 2 b n N
W D 0 = σ 2 2 E , n N = 0
W D c = A 2 2 E , n N = 1
where W is the strain energy density, W D is the fatigue driving energy, W D 0 is the stress-energy density for the initial stage without damage, W D c is the critical stress-energy density at failure, A is the material constant, and b is the exponent of the Basquin equation. For multi-block loading, the residual fatigue life of the material can be described as
n i N i = 1 1 2 b l n N i × l n N i 2 b 1 N i 1 2 b n i 1 N i 1 + 1 n i 1 N i 1 m p 1 N i 1 2 b 1 σ i 2 σ i 1 × σ i σ i 1 + 1
Peng’s and Miner’s models provided a more accurate fatigue life estimation for five different experimental datasets with different geometries.
Gan et al. proposed a new energy-based cumulative fatigue damage accumulation model based on a series of energy-life curves as an alternative to incremental plasticity theories [251]. The authors modified the multiaxial Wöhler (S-N) curve to redefine a new multiaxial S-N curve based on axial and torsion S-N curves, called the modified Wöhler curve method (MWCM).
N f = N A Δ τ A ρ Δ τ k ρ
where Δ τ is the shear stress range; N A is a reference fatigue life usually taken in the interval 106–107 cycles; Δ τ A ρ and k ρ are the reference shear stress amplitude and inverse slope related to the multiaxial S-N curve, respectively, which can be defined by
Δ τ A ρ = Δ τ A 1 Δ τ A 0 ρ + Δ τ A 0 , k ρ = k 1 k 0 ρ + k 0
where ρ is the stress ratio associated with the multiaxiality and non-proportionality of the local stress field, which can be expressed as
ρ = σ m a x 2 Δ τ
Lu et al. proposed a new energy-based cumulative fatigue damage accumulation model for low-cycle fatigue life prediction under multiaxial irregular loading [252]. The model combines both cyclic plastic work and the non-proportional (NP) hardening effect. The model defines the damage accumulation parameters (D) by defining the effective plastic work ( W e f f ), as follows:
D = N f i = 1 m W i e f f W i f
where the effective energy parameter W e f f is defined as
W e f f = 1 + α w F N P W e q
where W e q is the equivalent uniaxial plastic work, W i f is the fatigue toughness under the ith loading path in a cycle, α w is a material constant reflecting the material sensitivity to an NP path, F N P accounts for the path dependency of NP cyclic hardening, and N f is the number of cycles to failure.
This study utilized this model to predict the fatigue life of composite materials under multiaxial loading with multi-block loading sequences, and it showed an improved correlation with experimental data compared with Miner’s rule.

3. Fatigue Life Prediction of Adhesive Joint in FRP Composite Under Step/Variable Fatigue Loading

3.1. Numerical Model

A double-lap joint testing model is used to test the applicability of the aforementioned theories. Khabbaz et al. have carried out experimental studies to determine the fatigue life of adhesively bonded FRP structural joints under constant and variable loading [2].
A double-lap joint model shown in Figure 8 was used for the numerical simulation. The total length of each specimen is 350 mm. The outer and inner GFRP laminates were 200 mm long. The blue section of Figure 8 of the outer GFRP laminates was fixed and 50 mm long, whereas the load was applied to the yellow section (50 mm) at the end of the inner GFRP laminate.
The inner and outer tabs were made of pultruded GFRP laminates consisting of E-glass fibers, and an isophthalic polyester resin was used to fabricate the bonded joints. The table below (i.e., Table 1) shows the material properties of the GFRP laminates.
Sikadur 330 (Sika AG, Zugerstrasse 50, CH-6340 Baar (ZG), Switzerland), a two-component epoxy adhesive system, was used for the adhesive joint [253]. The material properties for the Sikadur 330 are listed in Table 2.
The fatigue test was performed under load control at a constant frequency of 10 Hz for constant loading. The load ratio, which was the ratio of the maximum applied cyclic load to the minimum applied cyclic load, was 0.1.
Four different two-block loading experiments were performed. The experiments consisted of two low-high-cycle loadings and two high-low-cycle loadings. A constant loading frequency of 10 Hz was used for the block loading, and the load ratio was maintained at 0.1. Table 3 lists the two-block loading experiments.
where Fmax1 is the max load applied in first loading block, Fmax2 is the max load applied in second loading block, n1 is the number of cycles with the first loading block, n2 is the number of cycles with the second loading block until failure, N1 is the number of cycles to failure in the first loading block, N2 is the number of cycle to failure in the second loading block, D1 is the damage caused in the first loading block, D2 is the damage caused in the second loading block, D is the total damage according to Miner’s rule (D1 + D2).

3.2. Simulation Results

To assess the validity of the theoretical models for fatigue damage accumulation under step/variable loading, as discussed in the previous section, 12 models were selected based on four criteria: popularity, relevance to adhesive materials, recency, and uniqueness. A comparison of the selected models with experimental data for fatigue damage accumulation under step/variable loading is presented in Table 4. Table 5 shows a comparison through relative error of prediction (REP), which can be defined as
R E P = N e x p N s i m N e x p × 100
The majority of the models demonstrated lower error percentages than Miner’s rule predictions for the number of cycles to failure, particularly for low-to-high load cycle scenarios (Cases 1 and 2), with most models exhibiting REP of approximately 1%. However, the Morrow’s damage accumulation model deviated from this trend, showing REP of approximately 2.5% for both cases. However, Kwofie’s fatigue damage accumulation model achieved the most favorable results for these scenarios, with REP of approximately 0.15% in both cases.
For the high-to-low load cycle cases (Cases 3 and 4), most models also showed lower error percentages than Miner’s rule predictions. Nevertheless, all theoretical damage accumulation models had a substantial degree of error when compared with the experimental data. Radhakrishnan’s damage accumulation model had the least REP, with 13.5% for Case 3 and 52% for Case 4 when compared to the experimental data.
The reason for the low percentage error in the low-to-high load cycle scenarios (i.e., Cases 1 and 2) was that the number of cycles to failure post-load transition was much smaller than the number of pre-load transition cycles.
The impact of the loading sequence on the fatigue life of the examined joints was related to the crack growth rate during the applied loading blocks. The high-to-low load cycle cases (i.e., Cases 3 and 4) resulted in a retardation of the crack growth rate, whereas acceleration was observed under the low-to-high load cycle cases (i.e., Cases 1 and 2).
An analysis of the damage accumulation rate per cycle revealed that the experimental results showed a faster rate for low-to-high loading cycle scenarios (Cases 1 and 2), and then predicted the damage accumulation rate. By contrast, high-to-low loading situations demonstrated a reduced damage accumulation rate per cycle, which was aligned with the experimentally observed trends [2].
A primary factor contributing to this discrepancy is the fatigue crack propagation, which is affected by the energy release rate. Many existing models do not account for the energy release rate when calculating failure, resulting in a linear and steady failure progression [144,145]. However, empirical evidence has demonstrated that the energy release rate increases exponentially as fractures begin to form in a material, which leads to material instability and rapid failure [7,13,227]. Even models that consider the energy release rate and crack propagation as the primary mechanisms of fatigue failure do not account for loading history. This limitation applies to fatigue damage accumulation models that are based on crack growth.
These findings suggest that most existing fatigue failure models are insufficient for accurately predicting fatigue failure in adhesive joints. These models often overlook crucial aspects, such as load history factors, the importance of loading sequences in determining fatigue failure, and the effects of load interaction. Additionally, many models fail to account for the swift and progressive nature of crack propagation during fatigue failure cycles. Consequently, there is a need for further research to develop an improved failure model that incorporates all of these critical factors when predicting cumulative fatigue failure in adhesive joints. Such a model would provide a more comprehensive and reliable approach to understanding and predicting the fatigue failure in these materials.

4. Conclusions

In this paper, a critical review of cumulative fatigue damage accumulation under step/variable fatigue loading experiments on adhesive joints in fiber-reinforced polymer composites is presented. This study investigated experimental observations of fatigue failure under step/variable loading and attempted to explore the theoretical reasoning behind the experimental observations. Cumulative fatigue damage has been extensively studied, and many approaches have been developed to treat it. This study investigated both linear and nonlinear accumulations, which can be grouped into eight categories, exploring over 100 cumulative fatigue damage accumulation models. A double-lap joint test model with four different step-load cases (two high-to-low cases and two how-to-high cases) was used to examine the applicability of the cumulative fatigue damage accumulation models. However, the cumulative fatigue damage accumulation model could not provide accurate predictions of failure in the aforementioned experimental cases.
This study emphasizes the significance of accounting for the loading history, load interaction effects, and material behavior when predicting fatigue life. Additionally, numerical simulations have demonstrated the limitations of conventional cumulative fatigue damage models, as they frequently yield conservative results under both low-to-high and high-to-low loading conditions. This reveals the necessity for more comprehensive models that integrate all the relevant factors to better predict the fatigue life of adhesive joints under variable loading scenarios. Therefore, additional research is necessary to fully understand the consequences of step/variable fatigue loading on the accumulation of fatigue damage in adhesive joints in fiber-reinforced polymer composites.

Author Contributions

Conceptualization, A.P. and A.T.; methodology, A.P. and A.T.; software, A.T.; validation, A.P. and A.T.; formal analysis, A.P.; investigation, A.P.; resources, A.T.; data curation, A.P.; writing—original draft preparation, A.P.; writing—review and editing, A.P. and A.T.; visualization, A.P.; supervision, A.T.; project administration, A.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Miner, M.A. Cumulative Damage in Fatigue. J. Appl. Mech. 1945, 12, A159–A164. [Google Scholar] [CrossRef]
  2. Sarfaraz Khabbaz, R. Fatigue Life Prediction of Adhesively-Bonded Fiber-Reinforced Polymer Structural Joints Under Spectrum Loading Patterns; EPFL: Lausanne, Switzerland, 2012; Volume 5573, p. 219. [Google Scholar]
  3. Vassilopoulos, A.P.; Shahverdi, M.; Keller, T. Mode I fatigue and fracture behavior of adhesively-bonded pultruded glass fiber-reinforced polymer (GFRP) composite joints. In Fatigue and Fracture of Adhesively-Bonded Composite Joints; Woodhead Publishing: Sawston, UK, 2015; pp. 149–186. [Google Scholar] [CrossRef]
  4. Vassilopoulos, A.P. Fatigue and Fracture of Adhesively-Bonded Composite Joints; Elsevier: Amsterdam, The Netherlands, 2014; Volume 52. [Google Scholar]
  5. Vassilopoulos, A.P.; Nijssen, R.P. Fatigue life prediction under realistic loading conditions. In Fatigue Life Prediction of Composites and Composite Structures; Woodhead Publishing: Sawston, UK, 2020; pp. 387–423. [Google Scholar] [CrossRef]
  6. Fatemi, A.; Yang, L. Cumulative fatigue damage and life prediction theories. Int. J. Fatigue 1998, 20, 9–34. [Google Scholar] [CrossRef]
  7. Lv, Z.; Huang, H.Z.; Zhu, S.P.; Gao, H.; Zuo, F. A modified nonlinear fatigue damage accumulation model. Int. J. Damage Mech. 2014, 24, 168–181. [Google Scholar] [CrossRef]
  8. Hectors, K.; De Waele, W. Cumulative damage and life prediction models for high-cycle fatigue of metals: A review. Metals 2021, 11, 204. [Google Scholar] [CrossRef]
  9. Marco, S.M.; Starkey, W.L. A Concept of Fatigue Damage. J. Fluids Eng. 1954, 76, 627–632. [Google Scholar] [CrossRef]
  10. Langer, B.F. Fatigue failure from stress cycles of varying amplitude. Asme J. Appl. Mech. 1937, 59, A160–A162. [Google Scholar] [CrossRef]
  11. Manson, S.S.; Halford, G.R. Practical implementation of the double linear damage rule and damage curve approach for treating cumulative fatigue damage. Int. J. Fract. 1981, 17, 169–192. [Google Scholar] [CrossRef]
  12. Miller, K.J. Metal Fatigue—A New Perspective. In Topics in Fracture and Fatigue; Springer: New York, NY, USA, 1992; pp. 309–330. [Google Scholar] [CrossRef]
  13. Corten, H.T.; Dolan, T. Cumulative fatigue damage. In Proceedings of the International Conference on Fatigue of Metals, London, UK, 10–14 September 1956; Institution of Mechanical Engineering and American Society of Mechanical Engineers: London, UK, 1956; Volume 1, pp. 235–242. [Google Scholar]
  14. Kwofie, S.; Rahbar, N. A fatigue driving stress approach to damage and life prediction under variable amplitude loading. Int. J. Damage Mech. 2013, 22, 393–404. [Google Scholar] [CrossRef]
  15. Zuo, F.J.; Huang, H.Z.; Zhu, S.P.; Lv, Z.; Gao, H. Fatigue life prediction under variable amplitude loading using a non-linear damage accumulation model. Int. J. Damage Mech. 2015, 24, 767–784. [Google Scholar] [CrossRef]
  16. Aeran, A.; Siriwardane, S.; Mikkelsen, O.; Langen, I. A new nonlinear fatigue damage model based only on S-N curve parameters. Int. J. Fatigue 2017, 103, 327–341. [Google Scholar] [CrossRef]
  17. Si-Jian, L.; Wei, L.; Da-Qing, T.; Jun-Bi, L. A new fatigue damage accumulation model considering loading history and loading sequence based on damage equivalence. Int. J. Damage Mech. 2017, 27, 707–728. [Google Scholar] [CrossRef]
  18. Subramanyan, S. A Cumulative Damage Rule Based on the Knee Point of the S-N Curve. J. Eng. Mater. Technol. 1976, 98, 316–321. [Google Scholar] [CrossRef]
  19. Epaarachchi, J.A.; Clausen, P.D. A new cumulative fatigue damage model for glass fibre reinforced plastic composites under step/discrete loading. Compos. Part Appl. Sci. Manuf. 2005, 36, 1236–1245. [Google Scholar] [CrossRef]
  20. Wheeler, O.E. Spectrum Loading and Crack Growth. J. Basic Eng. 1972, 94, 181–186. [Google Scholar] [CrossRef]
  21. Willenborg, J.D.; Engle, R.M.; Wood, H.A. A Crack Growth Retardation Model Using an Effective Stress Concept; Air Force Flight Dynamics Laboratory: Greenbrier, OH, USA, 1971. [Google Scholar] [CrossRef]
  22. Wolf, E. Fatigue crack closure under cyclic tension. Eng. Fract. Mech. 1970, 2, 37–45. [Google Scholar] [CrossRef]
  23. Elber, W. The Significance of Fatigue Crack Closure. In Damage Tolerance in Aircraft Structures; ASTM International: West Conshohocken, PA, USA, 1971; Volume 486, pp. 230–242. [Google Scholar] [CrossRef]
  24. Miller, K. The behavior of short fatigue cracks and their initiation. Mech. Behav. -Mater.- Proc. Fifth Int. Conf. 1987, 1, 1357–1381. [Google Scholar]
  25. Bao-Tong, M.; Laird, C. Overview of fatigue behavior in copper single crystals—II. Population, size distribution and growth kinetics of Stage I cracks for tests at constant strain amplitude. Acta Metall. 1989, 37, 337–348. [Google Scholar] [CrossRef]
  26. Mikheevskiy, S.; Glinka, G. The UniGrow Fatigue Crack Growth Model for Spectrum Loading. Theor. Appl. Fract. Mech. 2015, 79, 25–33. [Google Scholar] [CrossRef]
  27. Noroozi, A. Development of a Two-Parameter Model (Kmax, ΔK) for Fatigue Crack Growth Analysis. Ph.D. Thesis, University of Waterloo, Waterloo, ON, Canada, 2007. [Google Scholar]
  28. Chaboche, J.L. Continuum Damage Mechanics: Part II—Damage Growth, Crack Initiation, and Crack Growth. J. Appl. Mech. 1988, 55, 65–72. [Google Scholar] [CrossRef]
  29. Chaboche, J.L. Continuum Damage Mechanics: Part I—General Concepts. J. Appl. Mech. 1988, 55, 59–64. [Google Scholar] [CrossRef]
  30. Xiao, Y.C.; Li, S.; Gao, Z. A continuum damage mechanics model for high cycle fatigue. Int. J. Fatigue 1998, 20, 503–508. [Google Scholar] [CrossRef]
  31. Bhattacharya, B.; Ellingwood, B. A new CDM-based approach to structural deterioration. Int. J. Solids Struct. 1999, 36, 1757–1779. [Google Scholar] [CrossRef]
  32. Oller, S.; Salomón, O.; Oñate, E. A continuum mechanics model for mechanical fatigue analysis. Comput. Mater. Sci. 2005, 32, 175–195. [Google Scholar] [CrossRef]
  33. Dattoma, V.; Giancane, S.; Nobile, R.; Panella, F. Fatigue life prediction under variable loading based on a new non-linear continuum damage mechanics model. Int. J. Fatigue 2006, 28, 89–95. [Google Scholar] [CrossRef]
  34. Mesmacque, G.; Garcia, S.; Amrouche, A.; Rubiogonzalez, C. Sequential law in multiaxial fatigue, a new damage indicator. Int. J. Fatigue 2005, 27, 461–467. [Google Scholar] [CrossRef]
  35. Duyi, Y.; Zhenlin, W. A new approach to low-cycle fatigue damage based on exhaustion of static toughness and dissipation of cyclic plastic strain energy during fatigue. Int. J. Fatigue 2001, 23, 679–687. [Google Scholar] [CrossRef]
  36. Böhm, E.; Kurek, M.; Junak, G.; Cieśla, M.; Łagoda, T. Accumulation of Fatigue Damage Using Memory of the Material. Procedia Mater. Sci. 2014, 3, 2–7. [Google Scholar] [CrossRef]
  37. Peng, Z.; Huang, H.Z.; Zhou, J.; Li, Y.F. A New Cumulative Fatigue Damage Rule Based on Dynamic Residual S-N Curve and Material Memory Concept. Metals 2018, 8, 456. [Google Scholar] [CrossRef]
  38. Zhou, J.; Huang, H.Z.; Barnhart, M.V.; Huang, G.; Li, Y.F. A novel non-linear cumulative fatigue damage model based on the degradation of material memory. Int. J. Damage Mech. 2019, 29, 610–625. [Google Scholar] [CrossRef]
  39. Feltner Jo Dean, C.E.M. Microplastic Strain Hysteresis Energy as a Criterion for Fatigue Fracture. J. Basic Eng. 1961, 83, 15–22. [Google Scholar] [CrossRef]
  40. Noroozi, A.; Glinka, G.; Lambert, S. Prediction of fatigue crack growth under constant amplitude loading and a single overload based on elasto-plastic crack tip stresses and strains. Eng. Fract. Mech. 2008, 75, 188–206. [Google Scholar] [CrossRef]
  41. Zuchowski, R. Specific strain work as both failure criterion and material damage measure. Res. Mech. 1989, 27, 309–322. [Google Scholar]
  42. Lefebvre, D.; Neale, K.W.; Ellyin, F. A Criterion for Low-Cycle Fatigue Failure Under Biaxial States of Stress. J. Eng. Mater. Technol. 1981, 103, 1–6. [Google Scholar] [CrossRef]
  43. Leis, B.N. A Nonlinear History-Dependent Damage Model for Low Cycle Fatigue. Low Cycle Fatigue 1988, 143–159. [Google Scholar] [CrossRef]
  44. Radhakrishnan, V.M. An analysis of low cycle fatigue based on hysteresis energy. Fatigue Fract. Eng. Mater. Struct. 1980, 3, 75–84. [Google Scholar] [CrossRef]
  45. French, H. Fatigue and hardening of Steels. Trans. Am. Soc. Steel Treat. 1933, 21, 899. [Google Scholar]
  46. Kommers, J. The effect of overstressing and understressing in fatigue. Am. Soc. Test. Mater. 1938, 38, 249–268. [Google Scholar]
  47. Gaßner, E. Festigkeitsversuche mit wiederholter Beanspruchung im Flugzeugbau. Luftwissen 1939, 6, 61–64. [Google Scholar]
  48. Machlin, E.S. Dislocation Theory of the Fatigue of Metals; NASA Technical Reports: Washington, DC, USA, 1949. [Google Scholar]
  49. Coffin, L.F. Design Aspects of High-Temperature Fatigue with Particular Reference to Thermal Stresses. J. Fluids Eng. 1956, 78, 527–531. [Google Scholar] [CrossRef]
  50. Baldwin, E.E.; Sokol, G.J.; Coffin, L.F. Cyclic Strain Fatigue Studies on AISI Type 347 Stainless Steel; Knolls Atomic Power Lab.: Schenectady, NY, USA, 1957. [Google Scholar]
  51. Havard, D.G.; Topper, T.H. New equipment for cyclic biaxial testing—The behavior of the equipment and the modes of failure of the specimens are described and some test data are presented. Exp. Mech. 1969, 9, 550–557. [Google Scholar] [CrossRef]
  52. Gowda, C.V.; Topper, T.H. Performance of miniature resistance strain gages in low-cycle fatigue—Paper reports investigation of the performance characteristics of miniature resistance strain gages subjected to cyclic strains of high amplitude. Exp. Mech. 1970, 10, 27N–38N. [Google Scholar] [CrossRef]
  53. Gowda, C.V.; Topper, T.H. Cyclic-deformation behavior of notched mild-steel plates in plane stress—Stress and strain responses of thin mild-steel plates with a central circular hole due to cyclic loading into the plastic-deformation range. Exp. Mech. 1972, 12, 359–367. [Google Scholar] [CrossRef]
  54. Watson, P.; Topper, T.H. Fatigue-damage evaluation for mild steel incorporating mean stress and overload effects—The effects of mean stress and overstrain on the fatigue life of mild steel are demonstrated. These effects are then incorporated in a fatigue-damage evaluation technique. Exp. Mech. 1972, 12, 11–17. [Google Scholar] [CrossRef]
  55. Abdel-Raouf, H.; Plumtree, A.; Topper, T.H. Temperature and strain rate dependence of cyclic deformation response and damage accumulation in ofhc copper and 304 stainless steel. Metall. Trans. 1974, 5, 267–277. [Google Scholar] [CrossRef]
  56. Miller, K. An experimental linear cumulative-damage law. J. Strain Anal. 1970, 5, 177–184. [Google Scholar] [CrossRef]
  57. Jeans, L.; Grimes, G.; Kan, H. Fatigue sensitivity of composite structure for fighter aircraft. J. Aircr. 1983, 20, 102–110. [Google Scholar] [CrossRef]
  58. Yang, J.; Du, S. An exploratory study into the fatigue of composites under spectrum loading. J. Compos. Mater. 1983, 17, 511–526. [Google Scholar] [CrossRef]
  59. Jones, T.; Williams, N. Influence of section design and loading regime on the fatigue properties of spot welded/adhesive bonded box sections. In Proceedings of the IAVD Congress on Vehicle Design and Components, Geneva, Switzerland, 3–6 March 1986; pp. 262–273. [Google Scholar]
  60. Erpolat, S.; Ashcroft, I.A.; Crocombe, A.D.; Abdel-Wahab, M.M. A study of adhesively bonded joints subjected to constant and variable amplitude fatigue. Int. J. Fatigue 2004, 26, 1189–1196. [Google Scholar] [CrossRef]
  61. Richart, F.; Newmark, N. An hypothesis for the determination of cumulative damage in fatigue. In Selected Papers By Nathan M. Newmark: Civil Engineering Classics; ASCE: Reston, VA, USA, 1948; pp. 279–312. [Google Scholar]
  62. Grover, H.J. An Observation Concerning the Cycle Ratio in Cumulative Damage. In Symposium on Fatigue of Aircraft Structures; ASTM International: Reston, VA, USA, 1960. [Google Scholar] [CrossRef]
  63. Manson, S.S. Interfaces Between Fatigue, Creep, and Fracture. Int. J. Fract. Mech. 1966, 2, 327–363. [Google Scholar] [CrossRef]
  64. Manson, S.S.; Freche, J.C.; Ensign, C.R. Application of a double linear damage rule to cumulative fatigue. In Fatigue Crack Propagation; National Aeronautics and Space Administration: Washington, DC, USA, 1967; pp. 329–384. [Google Scholar] [CrossRef]
  65. Manson, S.; Halford, G. Complexities of High Temperature Metal Fatigue: Some Steps Toward Understanding; NASA Technical Memorandum, National Aeronautics and Space Administration: Washington, DC, USA, 1983. Available online: https://ntrs.nasa.gov/citations/19840006473 (accessed on 17 September 2024).
  66. Schönborn, S.; Kaufmann, H.; Sonsino, C.M.; Heim, R. Cumulative Damage of High-strength Cast Iron Alloys for Automotive Applications. Procedia Eng. 2015, 101, 440–449. [Google Scholar] [CrossRef]
  67. Theil, N. Fatigue life prediction method for the practical engineering use taking in account the effect of the overload blocks. Int. J. Fatigue 2016, 90, 23–35. [Google Scholar] [CrossRef]
  68. Inoma, J.E.; Pavlou, D.G.; Zec, J. Implementation of linear, double-linear, and nonlinear fatigue damage accumulation rules for fatigue life prediction of offshore drilling top-drive tie-rods. In IOP Conference Series: Materials Science and Engineering; IOP Publishing: Bristol, UK, 2019; Volume 700, p. 012025. [Google Scholar] [CrossRef]
  69. Chen, X.; Jin, D.; Kim, K.S. Fatigue life prediction of type 304 stainless steel under sequential biaxial loading. Int. J. Fatigue 2006, 28, 289–299. [Google Scholar] [CrossRef]
  70. Itoh, T.; Chen, X.; Nakagawa, T.; Sakane, M. A Simple Model for Stable Cyclic Stress-Strain Relationship of Type 304 Stainless Steel Under Nonproportional Loading. J. Eng. Mater. Technol. 1999, 122, 1–9. [Google Scholar] [CrossRef]
  71. Xu, J.; Shang, D.G.; Sun, G.Q.; Chen, H.; Liu, E.T. Fatigue life prediction for GH4169 superalloy under multiaxial variable amplitude loading. Beijing Gongye Daxue Xuebao/Journal Beijing Univ. Technol. 2012, 38, 1462–1466. [Google Scholar]
  72. Freudenthal, A.M.; Heller, R.A. On Stress Interaction in Fatigue and a Cumulative Damage Rule. J. Aerosp. Sci. 1959, 26, 431–442. [Google Scholar] [CrossRef]
  73. Freudenthal, A.M. Physical and statistical aspects of cumulative damage. In Colloquium on Fatigue/Colloque de Fatigue/Kolloquium über Ermüdungsfestigkeit; Springer: Berlin/Heidelberg, Germany, 1956; pp. 53–62. [Google Scholar] [CrossRef]
  74. Gao, H.; Huang, H.Z.; Zhu, S.P.; Li, Y.F.; Yuan, R. A Modified Nonlinear Damage Accumulation Model for Fatigue Life Prediction Considering Load Interaction Effects. Sci. World J. 2014, 2014, 164378. [Google Scholar] [CrossRef]
  75. Yuan, R.; Li, H.; Huang, H.Z.; Zhu, S.P.; Gao, H. A nonlinear fatigue damage accumulation model considering strength degradation and its applications to fatigue reliability analysis. Int. J. Damage Mech. 2014, 24, 646–662. [Google Scholar] [CrossRef]
  76. Zhou, Y.; Mu, D.S.; Han, Y.B.; Huang, X.W.; Zhang, C.C. Application of the Nonlinear Fatigue Damage Cumulative on the Prediction for Rail Head Checks Initiation and Wear Growth. In Structural Integrity; Springer: Cham, Germany, 2019; pp. 239–244. [Google Scholar] [CrossRef]
  77. Kwofie, S.; Rahbar, N. An equivalent driving force model for crack growth prediction under different stress ratios. Int. J. Fatigue 2011, 33, 1199–1204. [Google Scholar] [CrossRef]
  78. Zhu, S.P.; Hao, Y.Z.; de Oliveira Correia, J.A.; Lesiuk, G.; De Jesus, A.M. Nonlinear fatigue damage accumulation and life prediction of metals: A comparative study. Fatigue Fract. Eng. Mater. Struct. 2018, 42, 1271–1282. [Google Scholar] [CrossRef]
  79. Aeran, A.; Siriwardane, S.C.; Mikkelsen, O.; Langen, I. An accurate fatigue damage model for welded joints subjected to variable amplitude loading. Iop Conf. Ser. Mater. Sci. Eng. 2017, 276, 012038. [Google Scholar] [CrossRef]
  80. Kaechele, L. Review and Analysis of Cumulative-Fatigue-Damage Theories; Rand Corporation: Santa Monica, CA, USA, 1963; Available online: https://www.rand.org/content/dam/rand/pubs/research_memoranda/2008/RM3650.pdf (accessed on 17 September 2024).
  81. Spitzer, R.; Corten, H.T. Effect of Loading Sequence on Cumulative Fatigue Damage of 7075-T6 Aluminum Alloy; Department of Theoretical and Applied Mechanics, University of Illinois: Champaign, IL, USA, 1961; Volume 193, Available online: https://books.google.com/books?id=kCgvkGMZawUC (accessed on 17 September 2024).
  82. Rao, J.S.; Pathak, A.; Chawla, A. Blade Life: A Comparison by Cumulative Damage Theories. J. Eng. Gas Turbines Power 1999, 123, 886–892. [Google Scholar] [CrossRef]
  83. Jiao, J.; Lei, H.; Chen, Y.F. Numerical simulation and experimental study on constant amplitude fatigue behavior of welded cross plate-hollow sphere joints. J. Southeast Univ. 2018, 34, 62–70. [Google Scholar] [CrossRef]
  84. Marsh, K.J.; Mackinnon, J.A. Random-Loading and Block-Loading Fatigue Tests on Sharply Notched Mild Steel Specimens. J. Mech. Eng. Sci. 1968, 10, 48–58. [Google Scholar] [CrossRef]
  85. Zhao, S. Study on the accuracy of fatigue life predictions by the generally used damage accumulation theory. J. Mech. Strength 2000, 22, 206–209. [Google Scholar]
  86. Peng, Z.; Huang, H.Z.; Zhu, S.P.; Gao, H.; Lv, Z. A fatigue driving energy approach to high-cycle fatigue life estimation under variable amplitude loading. Fatigue Fract. Eng. Mater. Struct. 2015, 39, 180–193. [Google Scholar] [CrossRef]
  87. Zhu, S.P.; Huang, H.Z.; Liu, Y.; He, L.P.; Liao, Q. A Practical Method for Determining the Corten-Dolan Exponent and Its Application to Fatigue Life Prediction. Int. J. Turbo Jet-Engines 2012, 29, 79–87. [Google Scholar] [CrossRef]
  88. Gao, H.; Huang, H.Z.; Lv, Z.; Zuo, F.J.; Wang, H.K. An improved Corten-Dolan’s model based on damage and stress state effects. J. Mech. Sci. Technol. 2015, 29, 3215–3223. [Google Scholar] [CrossRef]
  89. Xue, Q.; Du, X.; Wang, S. An Improved Fatigue Life Prediction Model Based on Loading Sequence. Zhongguo Tiedao Kexue/China Railw. Sci. 2019, 40, 88–93. [Google Scholar] [CrossRef]
  90. Leipholz, H.H.E. Lifetime Prediction for Metallic Specimens Subjected to Loading with Varying Intensity. Comput. Struct. 1985, 20, 239–246. [Google Scholar] [CrossRef]
  91. Leipholz, H.H.E. On the modified S-N curve for metal fatigue prediction and its experimental verification. Eng. Fract. Mech. 1986, 23, 495–505. [Google Scholar] [CrossRef]
  92. Leipholz, H.H.E.; Topper, T.; El Menoufy, M. Lifetime prediction for metallic components subjected to stochastic loading. Comput. Struct. 1983, 16, 499–507. [Google Scholar] [CrossRef]
  93. Dowdell, D.J.; Leipholz, H.H.E.; Topper, T.H. The modified life law applied to SAE-1045 steel. Int. J. Fract. 1986, 31, 29–36. [Google Scholar] [CrossRef]
  94. Hashin, A.Z.R. A cumulative damage theory of fatigue failure. Mater. Sci. Eng. 1978, 34, 147–160. [Google Scholar] [CrossRef]
  95. Hashin, Z.; Rotem, A. A Fatigue Failure Criterion for Fiber Reinforced Materials. J. Compos. Mater. 1973, 7, 448–464. [Google Scholar] [CrossRef]
  96. Pavlou, D.G. A phenomenological fatigue damage accumulation rule based on hardness increasing, for the 2024-T42 aluminum. Eng. Struct. 2002, 24, 1363–1368. [Google Scholar] [CrossRef]
  97. Pavlou, D.G. Fatigue design challenges: Recent linear and nonlinear models. Iop Conf. Ser. Mater. Sci. Eng. 2019, 700, 012028. [Google Scholar] [CrossRef]
  98. Bjørheim, F.; Pavlou, D.G.; Siriwardane, S.C. Nonlinear fatigue life prediction model based on the theory of the S-N fatigue damage envelope. Fatigue Fract. Eng. Mater. Struct. 2022, 45, 1480–1493. [Google Scholar] [CrossRef]
  99. Lee, Y.L.; Pan, J.; Hathaway, R.B.; Barkey, M.E. Fatigue testing and analysis: Theory and practice. In Fatigue Testing and Analysis Theory and Practice; Elsevier: Amsterdam, The Netherlands, 2004; pp. 1–402. [Google Scholar] [CrossRef]
  100. Shen, C.; Talha, A.; Hamdi, A.; Benseddiq, N. A new cumulative fatigue damage model under biaxial loading. Proc. Inst. Mech. Eng. Part J. Mater. Des. Appl. 2020, 234, 962–973. [Google Scholar] [CrossRef]
  101. Hu, H.; Wei, Q.; Liu, B.; Liu, Y.; Hu, N.; Ma, Q.; Wang, C. Progressive Damage Behaviour Analysis and Comparison with 2D/3D Hashin Failure Models on Carbon Fibre–Reinforced Aluminium Laminates. Polymers 2022, 14, 2946. [Google Scholar] [CrossRef]
  102. Rege, K.; Pavlou, D.G. A one-parameter nonlinear fatigue damage accumulation model. Int. J. Fatigue 2017, 98, 234–246. [Google Scholar] [CrossRef]
  103. Zhu, S.P.; Liao, D.; Liu, Q.; Correia, J.A.; De Jesus, A.M. Nonlinear fatigue damage accumulation: Isodamage curve-based model and life prediction aspects. Int. J. Fatigue 2019, 128, 105185. [Google Scholar] [CrossRef]
  104. Pavlou, D.G. The theory of the S-N fatigue damage envelope: Generalization of linear, double-linear, and non-linear fatigue damage models. Int. J. Fatigue 2018, 110, 204–214. [Google Scholar] [CrossRef]
  105. El Aghoury, I.; Galal, K. A fatigue stress-life damage accumulation model for variable amplitude fatigue loading based on virtual target life. Eng. Struct. 2013, 52, 621–628. [Google Scholar] [CrossRef]
  106. Dan, H.C.; Wang, Z.; Cao, W.; Liu, J. Fatigue characterization of porous asphalt mixture complicated with moisture damage. Constr. Build. Mater. 2021, 303, 124525. [Google Scholar] [CrossRef]
  107. Liu, Y.; Ye, N.; Hu, X.; Zhao, G. An improved fatigue damage model based on the virtual load spectrum of golden section method. Fatigue Fract. Eng. Mater. Struct. 2021, 44, 2101–2118. [Google Scholar] [CrossRef]
  108. Batsoulas, N.D. Cumulative Fatigue Damage: CDM-Based Engineering Rule and Life Prediction Aspect. Steel Res. Int. 2016, 87, 1670–1677. [Google Scholar] [CrossRef]
  109. Batsoulas, N.D.; Giannopoulos, G.I. Cumulative Fatigue Damage of Composite Laminates: Engineering Rule and Life Prediction Aspect. Materials 2023, 16, 3271. [Google Scholar] [CrossRef]
  110. Xia, F.L.; Zhu, S.P.; Liao, D.; Dantas, R.; Correia, J.A.; De Jesus, A.M. Isodamage curve-based fatigue damage accumulation model considering the exhaustion of static toughness. Eng. Fail. Anal. 2020, 115, 104575. [Google Scholar] [CrossRef]
  111. Griffith, A.A. The phenomena of rupture and flow in solids. Philos. Trans. R. Soc. London. Ser. Contain. Pap. Math. Phys. Character 1921, 221, 163–198. [Google Scholar] [CrossRef]
  112. Murray, W.M. (Ed.) Fatigue and Fracture of Metals; Cambridge, Mass., Technology Press of the Massachusetts Institute of Technology Wiley: New York, NY, USA, 1952; 313p. [Google Scholar] [CrossRef]
  113. Williams, M.L. On the Stress Distribution at the Base of a Stationary Crack. J. Appl. Mech. 1957, 24, 109–114. [Google Scholar] [CrossRef]
  114. Irwin, G.R. Analysis of Stresses and Strains Near the End of a Crack Traversing a Plate. J. Appl. Mech. 1957, 24, 361–364. [Google Scholar] [CrossRef]
  115. Paris, P.; Erdogan, F. A Critical Analysis of Crack Propagation Laws. J. Basic Eng. 1963, 85, 528–533. [Google Scholar] [CrossRef]
  116. Erdogan, F.; Sih, G.C. On the Crack Extension in Plates Under Plane Loading and Transverse Shear. J. Basic Eng. 1963, 85, 519–525. [Google Scholar] [CrossRef]
  117. Rice, J.R. Mathematical analysis in the mechanics of fracture. Fract. Adv. Treatise 1968, 2, 191–311. [Google Scholar]
  118. Makkonen, M. Predicting the total fatigue life in metals. Int. J. Fatigue 2009, 31, 1163–1175. [Google Scholar] [CrossRef]
  119. Newman, J., Jr.; Wu, X.; Venneri, S.; Li, C. Small-crack effects in high-strength aluminum alloys. In NASA/CAE Cooperative Program, NASA RP-1309; NASA: Washington, DC, USA, 1994. Available online: https://ntrs.nasa.gov/api/citations/19940029793/downloads/19940029793.pdf (accessed on 17 September 2024).
  120. Newman, J.C. Fatigue-Life Prediction Methodology Using a Crack-Closure Model. J. Eng. Mater. -Technol.-Trans. Asme 1995, 117, 433–439. [Google Scholar] [CrossRef]
  121. Newman, J., Jr.; Phillips, E.P.; Everett, R.A., Jr. Fatigue life and crack growth prediction methodology. In AGARD Structures and Materials Panel Workshop on an Assessment of Fatigue Damage and Crack Growth Prediction Techniques; NASA: Washington, DC, USA, 1993; number NAS 1.15: 109044. Available online: https://ntrs.nasa.gov/citations/19940030077 (accessed on 17 September 2024).
  122. Newman, J. Advances in fatigue life prediction methodology for metallic materials. Int. J. Fatigue 1992, 16, 156. [Google Scholar] [CrossRef]
  123. Newman, J. The Merging of Fatigue and Fracture Mechanics Concepts: A Historical Perspective. Prog. Aerosp. Sci. 1998, 34, 347–390. [Google Scholar] [CrossRef]
  124. Newman, J.; Phillips, E.; Swain, M. Fatigue-life prediction methodology using small-crack theory. Int. J. Fatigue 1999, 21, 109–119. [Google Scholar] [CrossRef]
  125. Newman, J., Jr. Advances in fatigue and fracture mechanics analyses for metallic aircraft structures. Aeronaut. Fatigue 2000. Available online: https://ntrs.nasa.gov/api/citations/20000040431/downloads/20000040431.pdf (accessed on 17 September 2024).
  126. Newman, J.; Phillips, E.; Everett, R. Fatigue Analyses Under Constant- and Variable-Amplitude Loading Using Small-Crack Theory; NASA: Washington, DC, USA, 1999. Available online: https://ntrs.nasa.gov/citations/19990046065 (accessed on 17 September 2024).
  127. Newman, J. Prediction of Crack Growth Under Variable-Amplitude Loading in Thin-Sheet 2024-T3 Aluminum Alloys; NASA: Washington, DC, USA, 1997. Available online: https://ntrs.nasa.gov/citations/20040110410 (accessed on 17 September 2024).
  128. Suresh, S.; Ritchie, R.O. Some Considerations on Fatigue Crack Closure at Near-Threshold Stress Intensities Due to Fracture Surface Morphology. Metall. Trans. 1982, 13, 937–940. [Google Scholar] [CrossRef]
  129. Suresh, S.; Ritchie, R.O. Near-Threshold Fatigue Crack Propagation: A Perspective on the Role of Crack Closure; Lawrence Berkeley Laboratory University of California: Berkeley, CA, USA, 1983; Volume 84, p. 30202. Available online: https://www.osti.gov/biblio/6924455 (accessed on 17 September 2024).
  130. Dill, H.; Saff, C. Spectrum Crack Growth Prediction Method Based on Crack Surface Displacement and Contact Analyses. In Fatigue Crack Growth Under Spectrum Loads; ASTM International: West Conshohocken, PA, USA, 2009; pp. 306–319. [Google Scholar] [CrossRef]
  131. Dill, H.; Saff, C.; Potter, J. Effects of Fighter Attack Spectrum on Crack Growth. In Effect of Load Spectrum Variables on Fatigue Crack Initiation and Propagation; ASTM International: West Conshohocken, PA, USA, 2009; pp. 205–217. [Google Scholar] [CrossRef]
  132. Führing, H.; Seeger, T. Dugdale crack closure analysis of fatigue cracks under constant amplitude loading. Eng. Fract. Mech. 1979, 11, 99–122. [Google Scholar] [CrossRef]
  133. De Koning, A.U. A Simple Crack Closure Model for Prediction of Fatigue Crack Growth Rates Under Variable-Amplitude Loading. In Fracture Mechanics; ASTM International: West Conshohocken, PA, USA, 1981; pp. 63–85. [Google Scholar] [CrossRef]
  134. Miller, K.J.; Zachariah, K.P. Cumulative damage laws for fatigue crack initiation and stage i propagation. J. Strain Anal. Eng. Des. 1977, 12, 262–270. [Google Scholar] [CrossRef]
  135. Ibrahim, M.F.E.; Miller, K.J. Determination of Fatigue Crack Initiation Life. Fatigue Fract. Eng. Mater. Struct. 1979, 2, 351–360. [Google Scholar] [CrossRef]
  136. Miller, K.J.; Ibrahim, M.F.E. Damage Accumulation During Initiation and Short Crack Growth Regimes. Fatigue Fract. Eng. Mater. Struct. 1981, 4, 263–277. [Google Scholar] [CrossRef]
  137. Miller, K.J.; Mohamed, H.J.; De Los Rios, E.R. Fatigue damage accumulation above and below the fatigue limit. In The Behavior of Short Fatigue Cracks; Mechanical Engineering Publications: West Conshohocken, PA, USA, 1986; pp. 491–511. Available online: https://www.gruppofrattura.it/ocs/index.php/esis/EGF1/paper/viewFile/9666/6377 (accessed on 17 September 2024).
  138. Miller, K.; de los Rios, E. The Behaviour of Short Fatigue Cracks. In Fatigue & Fracture of Engineering Materials & Structures; Wiley: Hoboken, NJ, USA, 1987. [Google Scholar] [CrossRef]
  139. Miller, K. The Three Thresholds for Fatigue Crack Propagation. In Fatigue and Fracture Mechanics: 27th Volume; ASTM International: West Conshohocken, PA, USA, 1997; pp. 477–500. [Google Scholar] [CrossRef]
  140. Miller T, K.J.G. High-Temperature cumulative damage for stage I crack growth. J. Strain Anal. Eng. Des. 1977, 12, 253–261. [Google Scholar] [CrossRef]
  141. Miller, K.J. The behaviour of short fatigue cracks and their initiation part I—A review of two recent books. Fatigue Fract. Eng. Mater. Struct. 1987, 10, 75–91. [Google Scholar] [CrossRef]
  142. Miller, K.J. Short Fatigue Cracks. In Advances in Fatigue Science and Technology; Springer: Dordrecht, The Netherlands, 1989; pp. 3–22. [Google Scholar] [CrossRef]
  143. Hobson, P.D. The Growth of Short Fatigue Cracks in a Medium Carbon Steel. Ph.D. Thesis, University of Sheffield, Sheffield, UK, 1985. Available online: https://core.ac.uk/download/pdf/42605957.pdf (accessed on 17 September 2024).
  144. Dugdale, D.S. Yielding of steel sheets containing slits. J. Mech. Phys. Solids 1960, 8, 100–104. [Google Scholar] [CrossRef]
  145. Barenblatt, G.I. The Mathematical Theory of Equilibrium Cracks in Brittle Fracture. Adv. Appl. Mech. 1962, 7, 55–129. [Google Scholar] [CrossRef]
  146. Turon, A.; Camanho, P.P.; Costa, J.; Dávila, C.G. A damage model for the simulation of delamination in advanced composites under variable-mode loading. Mech. Mater. 2006, 38, 1072–1089. [Google Scholar] [CrossRef]
  147. Harper, P.W.; Hallett, S.R. Cohesive zone length in numerical simulations of composite delamination. Eng. Fract. Mech. 2008, 75, 4774–4792. [Google Scholar] [CrossRef]
  148. Zhang, W.; Tabiei, A. Fatigue life prediction of composite material’s adhesive joints in automotive applications. Int. J. Automot. Compos. 2017, 3, 61–79. [Google Scholar] [CrossRef]
  149. Tabiei, A.; Zhang, W. Composite laminate delamination simulation and experiment: A review of recent development. Appl. Mech. Rev. 2018, 70, 030801. [Google Scholar] [CrossRef]
  150. Bao-Tong Campbell, M.L. Overview of fatigue behavior in copper single crystals—V. Short crack growth behavior and a new approach to summing cumulative damage and predicting fatigue life under variable amplitudes. Acta Metall. 1989, 37, 369–379. [Google Scholar] [CrossRef]
  151. Polák, J.; Obrtlík, K.; Hájek, M. Cyclic Plasticity in Type 316L Austenitic Stainless Steel. Fatigue Fract. Eng. Mater. Struct. 1994, 17, 773–782. [Google Scholar] [CrossRef]
  152. Noroozi, A.; Glinka, G.; Lambert, S. A two parameter driving force for fatigue crack growth analysis. Int. J. Fatigue 2005, 27, 1277–1296. [Google Scholar] [CrossRef]
  153. Noroozi, A.; Glinka, G.; Lambert, S. A study of the stress ratio effects on fatigue crack growth using the unified two-parameter fatigue crack growth driving force. Int. J. Fatigue 2007, 29, 1616–1633. [Google Scholar] [CrossRef]
  154. Walker, K. The effect of stress ratio during crack propagation and fatigue for 2024-T3 and 7075-T6 aluminum. In Effects of Environment and Complex Load History on Fatigue Life: A Symposium; ASTM International: West Conshohocken, PA, USA, 1970; pp. 1–14. [Google Scholar] [CrossRef]
  155. Dinda, S.; Kujawski, D. Correlation and prediction of fatigue crack growth for different R-ratios using Kmax and ΔK+ parameters. Eng. Fract. Mech. 2004, 71, 1779–1790. [Google Scholar] [CrossRef]
  156. Mikheevskiy, S.; Glinka, G. Elastic-plastic fatigue crack growth analysis under variable amplitude loading spectra. Int. J. Fatigue 2009, 31, 1828–1836. [Google Scholar] [CrossRef]
  157. Abdullah, N.; Correia, J.; Jesus, A.; Hafezi, M.H.; Abdullah, S. Assessment of fatigue crack growth data available for materials from Portuguese bridges based on UniGrow model. Procedia Eng. 2011, 10, 971–976. [Google Scholar] [CrossRef]
  158. Abdullah, N.N.; Hafezi, M.H.; Abdullah, S. Analytical Concepts for Recent Development in Fatigue Crack Growth under Variable Amplitude Loading. Part I: Qualitative Interpretation. Key Eng. Mater. 2011, 462–463, 59–64. [Google Scholar] [CrossRef]
  159. Hafezi, M.H.; Abdullah, N.; Correia, J.; Jesus, A. An assessment of a strain-life approach for fatigue crack growth. Int. J. Struct. Integr. 2012, 3, 344–376. [Google Scholar] [CrossRef]
  160. Pedrosa, B.; Correia, J.; Lesiuk, G.; Rebelo, C.; Veljković, M. Fatigue crack growth modelling for S355 structural steel considering plasticity-induced crack-closure by means of UniGrow model. Int. J. Fatigue 2022, 164, 107120. [Google Scholar] [CrossRef]
  161. Bang, D.; Ince, A.; Tang, L. A modification of UniGrow 2-parameter driving force model for short fatigue crack growth. Fatigue Fract. Eng. Mater. Struct. 2018, 42, 45–60. [Google Scholar] [CrossRef]
  162. Bang, D.; Ince, A.; Noban, M. Modeling approach for a unified crack growth model in short and long fatigue crack regimes. Int. J. Fatigue 2019, 128, 105182. [Google Scholar] [CrossRef]
  163. Bang, D.; Ince, A. A short and long crack growth model based on 2-parameter driving force and crack growth thresholds. Int. J. Fatigue 2020, 141, 105870. [Google Scholar] [CrossRef]
  164. Manjunatha, C. Fatigue Crack Growth Prediction under Spectrum Load Sequence in an Aluminum Alloy by K*-RMS Approach. Int. J. Damage Mech. 2008, 17, 477–492. [Google Scholar] [CrossRef]
  165. Correia, J.; Jesus, A.; Fernández-Canteli, A. Local unified probabilistic model for fatigue crack initiation and propagation: Application to a notched geometry. Eng. Struct. 2013, 52, 394–407. [Google Scholar] [CrossRef]
  166. Correia, J.; Jesus, A.; Fernández-Canteli, A. Modelling probabilistic fatigue crack propagation rates for a mild structural steel. Fract. Struct. Integr. 2014, 9, 80–96. [Google Scholar] [CrossRef]
  167. Correia, J.; de Jesús, P.; Calçada, R.; Fernández-Canteli, A. Fatigue Crack Propagation Rates Prediction Using Probabilistic Strain-Based Models. In Fracture Mechanics; IntechOpen: London, UK, 2016. [Google Scholar] [CrossRef]
  168. Raposo, P.; Correia, J.; Jesus, A.; Calçada, R. Probabilistic fatigue S-N curves derivation for notched components. Frat. Integrita Strutt. 2017, 11, 2017. [Google Scholar] [CrossRef]
  169. Lopes, B.; Arruda, M.R.T.; Almeida-Fernandes, L.; Castro, L.; Silvestre, N.; Correia, J.R. Assessment of mesh dependency in the numerical simulation of compact tension tests for orthotropic materials. Compos. Part Open Access 2020, 1, 100006. [Google Scholar] [CrossRef]
  170. Nagarajappa, N.; Malipatil, S.G.; Majila, A.N.; Fernando, D.; Manjuprasad, M.; Manjunatha, C. Fatigue Crack Growth Prediction in a Nickel-Base Superalloy Under Spectrum Loads Using FRANC3D. Trans. Indian Natl. Acad. Eng. 2021, 7, 533–540. [Google Scholar] [CrossRef]
  171. Malipatil, S.G.; Nagarajappa, N.; Majila, A.N.; Fernando, D.C.; Bojja, R.; Jagannathan, N.; Manjuprasad, M.; Manjunatha, C. A Study on the Fatigue Crack Growth Behaviour of GTM718 Nickel Based Super Alloy Under cold-TURBISTAN Spectrum Loads. Theor. Appl. Fract. Mech. 2022, 120, 103386. [Google Scholar] [CrossRef]
  172. Ince, A.; Glinka, G. Approximation modeling framework for elastic-plastic stress-strain fields near cracks with a small finite crack tip radius. Theor. Appl. Fract. Mech. 2022, 121, 103452. [Google Scholar] [CrossRef]
  173. Gates, N.R.; Fatemi, A. Experimental fatigue crack growth behavior and predictions under multiaxial variable amplitude service loading histories. Eng. Fract. Mech. 2017, 174, 80–103. [Google Scholar] [CrossRef]
  174. Ding, Z.; Wang, X.; Gao, Z.; Bao, S. An experimental investigation and prediction of fatigue crack growth under overload/underload in Q345R steel. Int. J. Fatigue 2017, 98, 155–166. [Google Scholar] [CrossRef]
  175. Kachanov, L.M. Rupture Time Under Creep Conditions. Int. J. Fract. 1999, 97, 11–18. [Google Scholar] [CrossRef]
  176. Kachanov, L. Introduction to Continuum Damage Mechanics; Springer: Berlin/Heidelberg, Germany, 1986. [Google Scholar] [CrossRef]
  177. Rabotnov, Y.N.; Leckie, F.A.; Prager, W. Creep Problems in Structural Members. J. Appl. Mech. 1970, 37, 249. [Google Scholar] [CrossRef]
  178. Chaboche, J.L. Continuous damage mechanics—A tool to describe phenomena before crack initiation. Nucl. Eng. Des. 1981, 64, 233–247. [Google Scholar] [CrossRef]
  179. Chaboche, J.L. Continuum damage mechanics: Present state and future trends. Nucl. Eng. Des. 1987, 105, 19–33. [Google Scholar] [CrossRef]
  180. Halford, G.R.; Manson, S.S. Reexamination of Cumulative Fatigue Damage Laws. In Structure Integrity and Durability of Reusable Space Propulsion Systems; NASA: Washington, DC, USA, 1985. Available online: https://ntrs.nasa.gov/citations/19850019650 (accessed on 17 September 2024).
  181. Chaboche, J.L.; Lesne, P.A. A Non-Linear Continuous Fatigue Damage Model. Fatigue Fract. Eng. Mater. Struct. 1988, 11, 1–17. [Google Scholar] [CrossRef]
  182. Sun, Y.; Lu, S.; Meng, Q.; Wang, X. Online damage model of steam turbine based on continuum damage mechanics. In Proceedings of the 2017 36th Chinese Control Conference (CCC), Dalian, China, 26–28 July 2017; IEEE: Piscataway, NJ, USA, 2017; pp. 10392–10397. [Google Scholar] [CrossRef]
  183. Wang, W.Z.; Liu, Y.Z. Continuum Damage Mechanics Theory and Application-Continuum Damage Mechanics Theory and Applications; Springer: Berlin/Heidelberg, Germany, 1987. [Google Scholar] [CrossRef]
  184. Lemaitre, J.; Plumtree, A. Application of Damage Concepts to Predict Creep-Fatigue Failures. J. Eng. Mater. Technol. 1979, 101, 284–292. [Google Scholar] [CrossRef]
  185. June, W. A continuum damage mechanics model for low-cycle fatigue failure of metals. Eng. Fract. Mech. 1992, 41, 437–441. [Google Scholar] [CrossRef]
  186. Wang, T.; Lou, Z. A continuum damage model for weld heat affected zone under low cycle fatigue loading. Eng. Fract. Mech. 1990, 37, 825–829. [Google Scholar] [CrossRef]
  187. Socie, D.F.; Morrow, J. Review of Contemporary Approaches to Fatigue Damage Analysis. In Risk and Failure Analysis for Improved Performance and Reliability; Springer: Berlin/Heidelberg, Germany, 1980; pp. 141–194. [Google Scholar] [CrossRef]
  188. Weinacht, D.J.; Socie, D.F. Fatigue damage accumulation in grey cast iron. Int. J. Fatigue 1987, 9, 79–86. [Google Scholar] [CrossRef]
  189. Hua, C.T.; Socie, D.F. Fatigue Damage in 1045 Steel Under Constant Amplitude Biaxial Loading. Fatigue Fract. Eng. Mater. Struct. 1984, 7, 165–179. [Google Scholar] [CrossRef]
  190. Bhattacharya, B.; Ellingwood, B. A CDM analysis of stochastic ductile damage growth and reliability. Probabilistic Eng. Mech. 1999, 14, 45–54. [Google Scholar] [CrossRef]
  191. Salomón, O.; Oller, S.; Oñate, E. Fatigue Analysis of Materials and Structures using a Continuum Damage Model. Int. J. Form. Process. 2002, 5, 493–503. [Google Scholar] [CrossRef]
  192. Martinez, X.; Oller, S.; Barbu, L.G.; Barbat, A.H.; De Jesus, A.M. Analysis of Ultra Low Cycle Fatigue problems with the Barcelona plastic damage model and a new isotropic hardening law. Int. J. Fatigue 2015, 73, 132–142. [Google Scholar] [CrossRef]
  193. Giancane, S.; Nobile, R.; Panella, F.; Dattoma, V. Fatigue life prediction of notched components based on a new nonlinear continuum damage mechanics model. Procedia Eng. 2010, 2, 1317–1325. [Google Scholar] [CrossRef]
  194. Zhang, A.; Bui-Quoc, T.; Gomuc, R. A Procedure for Low Cycle Fatigue Life Prediction for Various Temperatures and Strain Rates. J. Eng. Mater. Technol. 1990, 112, 422–428. [Google Scholar] [CrossRef]
  195. Zhang, J.; Fu, X.; Lin, J.; Liu, Z.; Liu, N.; Wu, B. Study on Damage Accumulation and Life Prediction with Loads below Fatigue Limit Based on a Modified Nonlinear Model. Materials 2018, 11, 2298. [Google Scholar] [CrossRef] [PubMed]
  196. Sinclair, G. Investigation of coaxing effect in fatigue of metals. Proc. Am. Soc. Test. Mater. TAM R 28 1952, 52, 743–751. [Google Scholar]
  197. Xi, L.; Songlin, Z. Strengthening of transmission gear under low-amplitude loads. Mater. Sci. Eng. A 2008, 488, 55–63. [Google Scholar] [CrossRef]
  198. Xi, L.; Songlin, Z. Changes in mechanical properties of vehicle components after strengthening under low-amplitude loads below the fatigue limit. Fatigue Fract. Eng. Mater. Struct. 2009, 32, 847–855. [Google Scholar] [CrossRef]
  199. Xi, L.; Songlin, Z. Strengthening and damaging under low-amplitude loads below the fatigue limit. Int. J. Fatigue 2009, 31, 341–345. [Google Scholar] [CrossRef]
  200. Li, Z.X.; Chan, T.H.; Ko, J.M. Fatigue damage model for bridge under traffic loading: Application made to Tsing Ma Bridge. Theor. Appl. Fract. Mech. 2001, 35, 81–91. [Google Scholar] [CrossRef]
  201. Li, Z.X.; Chan, T.H.; Ko, J.M. Fatigue analysis and life prediction of bridges with structural health monitoring data—Part II: Application. Int. J. Fatigue 2001, 23, 55–64. [Google Scholar] [CrossRef]
  202. Li, Z.X.; Chan, T.H.T.; Ko, J.M. Evaluation of typhoon induced fatigue damage for Tsing Ma Bridge. Eng. Struct. 2002, 24, 1035–1047. [Google Scholar] [CrossRef]
  203. Li, Z.X.; Ko, J.M.; Chan, T.H.T. Modelling of load interaction and overload effect on fatigue damage of steel bridges. Fatigue Fract. Eng. Mater. Struct. 2001, 24, 379–390. [Google Scholar] [CrossRef]
  204. Xu, Y.L.; Chen, Z.W.; Xia, Y. Fatigue assessment of multi-loading suspension bridges using continuum damage model. Int. J. Fatigue 2012, 40, 27–35. [Google Scholar] [CrossRef]
  205. Xu, Y.L.; Liu, T.T.; Zhang, W.S. Buffeting-induced fatigue damage assessment of a long suspension bridge. Int. J. Fatigue 2009, 31, 575–586. [Google Scholar] [CrossRef]
  206. Wahidi, M. Real-time Structural Health Monitoring of Engineering Structures. Master’s Thesis, UIS, Selangor, Malaysia, 2023. Available online: https://uis.brage.unit.no/uis-xmlui/handle/11250/3080196 (accessed on 17 September 2024).
  207. Zengah, S.; Aid, A.; Benguediab, M. Comparative Study of Fatigue Damage Models Using Different Number of Classes Combined with the Rainflow Method. Eng. Technol. Appl. Sci. Res. 2013, 3, 446–451. [Google Scholar] [CrossRef]
  208. Yue, P.; Lei, Q.; Zhang, C.L.; Zhu, S.P.; Huang, H.Z. A Modified Cumulative Damage Model for Fatigue Life Prediction under Variable Amplitude Loadings. Appl. Mech. Mater. 2017, 853, 62–66. [Google Scholar] [CrossRef]
  209. Hectors, K.; Vanspeybrouck, D.; Plets, J.; Bouckaert, Q.; De Waele, W. Open-Access Experiment Dataset for Fatigue Damage Accumulation and Life Prediction Models. Metals 2023, 13, 621. [Google Scholar] [CrossRef]
  210. Bandara Sudath, C.; Dissanayake, U.I.; Dissanayake, R.C.S.S. Fatigue failure predictions for steels in the very high cycle region—A review and recommendations. Eng. Fail. Anal. 2014, 45, 421–435. [Google Scholar] [CrossRef]
  211. Adasooriya, N.D.; Siriwardane, S.C. Remaining fatigue life estimation of corroded bridge members. Fatigue Fract. Eng. Mater. Struct. 2014, 37, 603–622. [Google Scholar] [CrossRef]
  212. Pitoiset André, X.P. Spectral methods for multiaxial random fatigue analysis of metallic structures. Int. J. Fatigue 2000, 22, 541–550. [Google Scholar] [CrossRef]
  213. Crossland, B. Effect of large hydrostatic pressures on the torsional fatigue strength of an alloy steel. J. Mech. Eng. Sci. 1964, 6, 293–310. [Google Scholar]
  214. Aid, A.; Bendouba, M.; Aminallah, L.; Amrouche, A.; Benseddiq, N.; Benguediab, M. An equivalent stress process for fatigue life estimation under multiaxial loadings based on a new non linear damage model. Mater. Sci. Eng. -Struct. Mater. Prop. Microstruct. Process. 2012, 538, 20–27. [Google Scholar] [CrossRef]
  215. Sines, G.; Ohgi, G. Fatigue Criteria Under Combined Stresses or Strains. J. Eng. Mater. Technol. 1981, 103, 82–90. [Google Scholar] [CrossRef]
  216. Bannantine, J.A. A variable amplitude multiaxial fatigue life prediction method. In Proceedings of the Thirth International Conference on Biaxial/Multiaxial Fatigue, Champaign, IL, USA, 18 April 1989; University of Illinois at Urbana-Champaign: Champaign, IL, USA, 1989; pp. 12.1–12.20. Available online: https://fcp.mechse.illinois.edu/fcp_report151/ (accessed on 17 September 2024).
  217. Fatemi, A.; Socie, D.F. A Critical Plane Approach to Multiaxial Fatigue Damage Including out-of-Phase Loading. Fatigue Fract. Eng. Mater. Struct. 1988, 11, 149–165. [Google Scholar] [CrossRef]
  218. Socie, D. Critical plane approaches for multiaxial fatigue damage assessment. In Advances in Multiaxial Fatigue; ASTM: Washington, DC, USA, 1993. [Google Scholar] [CrossRef]
  219. Benkabouche, S.; Guechichi, H.; Amrouche, A.; Benkhettab, M. A modified nonlinear fatigue damage accumulation model under multiaxial variable amplitude loading. Int. J. Mech. Sci. 2015, 100, 180–194. [Google Scholar] [CrossRef]
  220. Wang, C.H.; Brown, M.W. A Path-Independent Parameter for Fatigue Under Proportional and Non-Proportional Loading. Fatigue Fract. Eng. Mater. Struct. 1993, 16, 1285–1297. [Google Scholar] [CrossRef]
  221. Łagoda, T.; Macha, E.; Będkowski, W. A critical plane approach based on energy concepts: Application to biaxial random tension-compression high-cycle fatigue regime. Int. J. Fatigue 1999, 21, 431–443. [Google Scholar] [CrossRef]
  222. Ye, L. On fatigue damage accumulation and material degradation in composite materials. Compos. Sci. Technol. 1989, 36, 339–350. [Google Scholar] [CrossRef]
  223. Pattin, C.A.; Caler, W.E.; Carter, D.R. Cyclic mechanical property degradation during fatigue loading of cortical bone. J. Biomech. 1996, 29, 69–79. [Google Scholar] [CrossRef]
  224. Lohbauer, U.; Belli, R.; Ferracane, J. Factors Involved in Mechanical Fatigue Degradation of Dental Resin Composites. J. Dent. Res. 2013, 92, 584–591. [Google Scholar] [CrossRef]
  225. Duyi, Y.; Zhenlin, W. Change characteristics of static mechanical property parameters and dislocation structures of 45# medium carbon structural steel during fatigue failure process. Mater. Sci. Eng. A 2001, 297, 54–61. [Google Scholar] [CrossRef]
  226. Cadenas, P.; Amrouche, A.; Mesmacque, G.; Jozwiak, K. Effect of the residual fatigue damage on the static and toughness properties. In Damage and Fracture Mechanics: Failure Analysis of Engineering Materials and Structures; Springer: Berlin/Heidelberg, Germany, 2009; pp. 323–330. [Google Scholar] [CrossRef]
  227. Kurath, P.; Sehitoglu, H.; Morrow, J.; Deves, T. Effect of Selected Subcycle Sequences in Fatigue Loading Histories; American Society of Mechanical Engineers, Pressure Vessels and Piping Division (Publication) PVP: New York, NY, USA, 1983; Volume 72, pp. 43–60. [Google Scholar]
  228. Nakai-Chapman, J.; Park, Y.H.; Sakai, J. Progressive Fatigue Life Prediction of Composite Materials Based on Residual Material Property Degradation Model; American Society of Mechanical Engineers: New York, NY, USA, 2020; Volume 8. [Google Scholar] [CrossRef]
  229. Inglis, N.P. Hysteresis and fatigue of Wohler rotating cantilever specimen. In The Metallurgist; Springer: Berlin/Heidelberg, Germany, 1927; pp. 23–27. [Google Scholar]
  230. Budiansky Richard J., B.O. Elastic moduli of a cracked solid. Int. J. Solids Struct. 1976, 12, 81–97. [Google Scholar] [CrossRef]
  231. Glinka, G. Relations Between the Strain Energy Density Distribution and Elastic-Plastic Stress-Strain Fields Near Cracks and Notches and Fatigue Life Calculation. In Low Cycle Fatigue; ASTM: Washington, DC, USA, 1988; pp. 1022–1047. [Google Scholar] [CrossRef]
  232. Kujawski Fernand, D.E. A cumulative damage theory for fatigue crack initiation and propagation. Int. J. Fatigue 1984, 6, 83–88. [Google Scholar] [CrossRef]
  233. Lefebvre Fernand, D.E. Cyclic response and inelastic strain energy in low cycle fatigue. Int. J. Fatigue 1984, 6, 9–15. [Google Scholar] [CrossRef]
  234. Ellyin K., F.G. Multiaxial Fatigue Damage Criterion. J. Eng. Mater. Technol. 1988, 110, 63–68. [Google Scholar] [CrossRef]
  235. Kujawski, D.; Ellyin, F. On the concept of cumulative fatigue damage. Int. J. Fract. 1988, 37, 263–278. [Google Scholar] [CrossRef]
  236. Kreiser, D.; Jia, S.X.; Han, J.J.; Dhanasekar, M. A nonlinear damage accumulation model for shakedown failure. Int. J. Fatigue 2007, 29, 1523–1530. [Google Scholar] [CrossRef]
  237. Feng, D.C.; Ren, X. Analytical examination of mesh-dependency issue for uniaxial RC elements and new fracture energy-based regularization technique. Int. J. Damage Mech. 2023, 32, 321–339. [Google Scholar] [CrossRef]
  238. Niu, X.; Li, G.; Lee, H. Hardening law and fatigue damage of a cyclic hardening metal. Eng. Fract. Mech. 1987, 26, 163–170. [Google Scholar] [CrossRef]
  239. Niu, X.D. Memory Behavior of Stress Amplitude Responses and Fatigue Damage Model of A Hot-Rolled Low Carbon Steel. In Mechanical Behaviour of Materials V; Elsevier: Amsterdam, The Netherlands, 1988; pp. 685–690. [Google Scholar] [CrossRef]
  240. Radhakrishnan, V.M. Cumulative damage in low-cycle fatigue: An analysis is presented for cumulative damage on the basis of absorbed plastic energy. Exp. Mech. 1978, 18, 292–296. [Google Scholar] [CrossRef]
  241. Lagoda, T. Energy models for fatigue life estimation under uniaxial random loading. Part II: Verification of the model. Int. J. Fatigue 2001, 23, 481–489. [Google Scholar] [CrossRef]
  242. Pawliczek, R.; Lagoda, T. Investigation of changes in fatigue damage caused by mean load under block loading conditions. Materials 2021, 14, 2738. [Google Scholar] [CrossRef]
  243. Park, S.H.; Hong, S.G.; Lee, B.H.; Lee, C.S. Fatigue life prediction of rolled AZ31 magnesium alloy using an energy-based model. Int. J. Mod. Phys. B 2008, 22, 5503–5508. [Google Scholar] [CrossRef]
  244. Jahed, H.; Varvani-Farahani, A. Upper and lower fatigue life limits model using energy-based fatigue properties. Int. J. Fatigue 2006, 28, 467–473. [Google Scholar] [CrossRef]
  245. Jahed, A.; Noban, M.; Khalaji, I.H.V.F. An energy-based fatigue life assessment model for various metallic materials under proportional and non-proportional loading conditions. Int. J. Fatigue 2007, 29, 647–655. [Google Scholar] [CrossRef]
  246. Gu, Z.; Mi, C.; Ding, Z.; Zhang, Y.; Liu, S.; Nie, D. An energy-based fatigue life prediction of a mining truck welded frame. J. Mech. Sci. Technol. 2016, 30, 3615–3624. [Google Scholar] [CrossRef]
  247. Ozaltun, H.; Shen, M.H.H.; George, T.; Cross, C. An Energy Based Fatigue Life Prediction Framework for In-Service Structural Components. Exp. Mech. 2011, 51, 707–718. [Google Scholar] [CrossRef]
  248. Djebli, A.; Aid, A.; Bendouba, M.; Amrouche, A.; Benguediab, M.; Benseddiq, N. A non-linear energy model of fatigue damage accumulation and its verification for Al-2024 aluminum alloy. Int. J.-Non-Linear Mech. 2013, 51, 145–151. [Google Scholar] [CrossRef]
  249. Djebli, A.; Mostefa, B.; Abdelkarim, A. Fatigue Life Prediction under Variable Loading Based a Non-Linear Energy Model. Int. J. Eng. Res. Afr. 2016, 22, 14–21. [Google Scholar] [CrossRef]
  250. Peng, Z.; Huang, H.; Wang, H.K.; Zhu, S.; Lv, Z. A new approach to the investigation of load interaction effects and its application in residual fatigue life prediction. Int. J. Damage Mech. 2016, 25, 672–690. [Google Scholar] [CrossRef]
  251. Gan, L.; Wu, H.; Zhong, Z. Multiaxial fatigue life prediction based on a simplified energy-based model. Int. J. Fatigue 2021, 144, 106036. [Google Scholar] [CrossRef]
  252. Lu, Y.; Wu, H.; Zhong, Z. A modified energy-based model for low-cycle fatigue life prediction under multiaxial irregular loading. Int. J. Fatigue 2019, 128, 105187. [Google Scholar] [CrossRef]
  253. Sikadur-330, 0. Available online: https://gcc.sika.com/en/construction/refurbishment/structural-strengthening/sikadur-330.html (accessed on 22 May 2024).
Figure 1. Typical low-to-high and high-to-low loading cycles and the theoretical damage accumulation example. (a,b) show different types of load transition. (c,d) show the theoretical nonlinear characteristics of the damage accumulation based on the type of load transition. (a) Low to high load transition, (b) High to low load transition, (c) D Low High = n 1 N 1 + n 2 N 2 > 1 , and (d) D High Low = n 1 N 1 + n 2 N 2 < 1 .
Figure 1. Typical low-to-high and high-to-low loading cycles and the theoretical damage accumulation example. (a,b) show different types of load transition. (c,d) show the theoretical nonlinear characteristics of the damage accumulation based on the type of load transition. (a) Low to high load transition, (b) High to low load transition, (c) D Low High = n 1 N 1 + n 2 N 2 > 1 , and (d) D High Low = n 1 N 1 + n 2 N 2 < 1 .
Jcs 08 00477 g001
Figure 2. Damage accumulation under miner rule, where the damage accumulates linearly based on the number of cycles per load, irrespective of the nature of the load transition. (a) Low-to-high load transition; (b) high-to-low load transition.
Figure 2. Damage accumulation under miner rule, where the damage accumulates linearly based on the number of cycles per load, irrespective of the nature of the load transition. (a) Low-to-high load transition; (b) high-to-low load transition.
Jcs 08 00477 g002
Figure 3. Fatigue damage accumulation comparison between Miner’s rule, Freudenthal–Heller rule (two-stage linearization rule), and damage curve approach (DCA) rule for the low-to-high loading case.
Figure 3. Fatigue damage accumulation comparison between Miner’s rule, Freudenthal–Heller rule (two-stage linearization rule), and damage curve approach (DCA) rule for the low-to-high loading case.
Jcs 08 00477 g003
Figure 4. Life modification models based on iso−damage lines by Subramanyan [18].
Figure 4. Life modification models based on iso−damage lines by Subramanyan [18].
Jcs 08 00477 g004
Figure 5. Life modification models based on iso−damage lines by Hashin and Rotem [95].
Figure 5. Life modification models based on iso−damage lines by Hashin and Rotem [95].
Jcs 08 00477 g005
Figure 6. Stress vs. displacement plot in bilinear traction separation law, where S is the static damage, F denotes the fatigue damage caused due to separation, and R is the residual strength of the material.
Figure 6. Stress vs. displacement plot in bilinear traction separation law, where S is the static damage, F denotes the fatigue damage caused due to separation, and R is the residual strength of the material.
Jcs 08 00477 g006
Figure 7. S−N curve defining parameters used in the Mesmacque damage model [34].
Figure 7. S−N curve defining parameters used in the Mesmacque damage model [34].
Jcs 08 00477 g007
Figure 8. Specifications of double-lap joint used in experimental study.
Figure 8. Specifications of double-lap joint used in experimental study.
Jcs 08 00477 g008
Table 1. Mechanical properties of GFRP laminates used in this study.
Table 1. Mechanical properties of GFRP laminates used in this study.
PropertyValue
Density2560 kg/m3
Young’s Modulus25 GPa
Tensile Strength307.5 MPa
Poisson’s Ratio0.23
Table 2. Mechanical properties of the adhesive (Sikadur 330) used in this study.
Table 2. Mechanical properties of the adhesive (Sikadur 330) used in this study.
PropertyValue
Flexural Strength60.6 MPa
Modulus of Elasticity in Flexure3489 MPa
Tensile Strength33.8 MPa
Elongation at Break1.2%
Heat Deflection Temperature50 °C
Table 3. Experimental cases used in this study to check the theoretical damage accumulation models.
Table 3. Experimental cases used in this study to check the theoretical damage accumulation models.
Fmax1 [kN]n1N1D1 = n1/N1Fmax2 [kN]n2N2D2 = n2/N2D
Low-High Sequence14.448,649142,9780.340321.648310700.45140.7917
12456,0381,291,9930.353019.2259244340.58460.9376
High-Low Sequence21.628710700.268214.4207,559142,9781.45171.7199
19.2133044340.300012.04,832,6881,291,9933.74054.040
Table 4. Comparison of the experimental number of cycles to failure with selected theoretical models.
Table 4. Comparison of the experimental number of cycles to failure with selected theoretical models.
Case 1Case 2Case 3Case 4
Experiment49,132458,630207,8464,834,018
Miner49,355458,906104,915905,784
Corten–Dolan48,963457,158236,0002,317,476
Freudenthal–Heller49,558459,86057,277301,967
Damage Curve Approach49,516459,558127,9921,109,683
Kwofie49,064457,749178,5651,517,491
Subramanyan48,650456,038142,9781,291,993
Dattoma48,650456,038142,9781,291,993
Xie49,018457,465138,3381,242,244
Radhakrishnana48,963457,158236,0002,317,476
Carpenteri49,551459,84882,304682,574
Iso-Damage line modification50,688470,08122,16097,819
Morrow218,3772,107,6927222901
Bilinear Cohesive model51,400300,000107,180913,610
Table 5. Comparison of REP for selected theoretical models.
Table 5. Comparison of REP for selected theoretical models.
Case 1Case 2Case 3Case 4
Experiment0000
Miner0.45390.0601849.522781.2623
Corten Dolan0.34400.321013.545652.0590
Fruedenthal Heller0.86710.268272.442693.7533
Damage curve approach0.78160.202338.419877.0443
Kwofie0.13840.192114.087868.6081
Subramanyan0.98100.565131.209673.2729
Dattoma0.98100.565131.209673.2729
Xie−0.23200.254033.442174.3020
Radhakrishnan0.34400.321013.545652.0590
Carpenteri0.85280.265660.401585.880
Iso-damage line modification3.16702.496889.338397.9764
Morrow344.470359.562699.652699.940
Bilinear cohesive model4.61634.58848.43381.100
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.

Share and Cite

MDPI and ACS Style

Patro, A.; Tabiei, A. Fatigue Failure of Adhesive Joints in Fiber-Reinforced Composite Material Under Step/Variable Amplitude Loading—A Critical Literature Review. J. Compos. Sci. 2024, 8, 477. https://doi.org/10.3390/jcs8110477

AMA Style

Patro A, Tabiei A. Fatigue Failure of Adhesive Joints in Fiber-Reinforced Composite Material Under Step/Variable Amplitude Loading—A Critical Literature Review. Journal of Composites Science. 2024; 8(11):477. https://doi.org/10.3390/jcs8110477

Chicago/Turabian Style

Patro, Abinash, and Ala Tabiei. 2024. "Fatigue Failure of Adhesive Joints in Fiber-Reinforced Composite Material Under Step/Variable Amplitude Loading—A Critical Literature Review" Journal of Composites Science 8, no. 11: 477. https://doi.org/10.3390/jcs8110477

APA Style

Patro, A., & Tabiei, A. (2024). Fatigue Failure of Adhesive Joints in Fiber-Reinforced Composite Material Under Step/Variable Amplitude Loading—A Critical Literature Review. Journal of Composites Science, 8(11), 477. https://doi.org/10.3390/jcs8110477

Article Metrics

Back to TopTop