Skip to main content

ORIGINAL RESEARCH article

Front. Mater., 07 February 2024
Sec. Structural Materials

An improved approach for the continuous retardation spectra of concrete creep and applications

Xinzhu ZhouXinzhu Zhou1Linhong BaiLinhong Bai1Hua RongHua Rong2Xinglang Fan
Xinglang Fan2*Jianjun ZhengJianjun Zheng1Yan GengYan Geng2
  • 1College of Civil Engineering, Zhejiang University of Technology, Hangzhou, China
  • 2Central Research Institute of Building and Construction, Metallurgical Group Corporation of China, Beijing, China

Creep is an important physical property of concrete and can lead to additional displacement, stress redistribution, and even cracking in concrete structures, inducing prestress loss of large-scale prestressed concrete structures. When an exponential algorithm is used to calculate the long-term creep of concrete, it is usually necessary to apply the continuous retardation spectra of the material. In the improved approach proposed here, the continuous retardation spectra can be obtained by the Weeks inverse Laplace transform. The CEB MC90 creep model is taken as an example to analyze the computational process, efficiency, and error of the approach. The improved approach is further applied to the ACI 209R-92, JSCE, and GL2000 concrete creep models. Through comparison with other methods, the advantages of the improved approach are illustrated, and some useful conclusions are drawn.

1 Introduction

Concrete creep is defined as the time-dependent deformation of a concrete specimen under sustained load. Its magnitude is closely related to the stress applied, time, cement type, mix proportion of concrete, and environmental conditions (ACI Committee 209, 2005). In practice, concrete creep can lead to additional displacements, stress redistribution, and even cracking in concrete structures during their service life (Hubert Rüsch and Hilsdorf, 1983; Bažant et al., 1997). As a result, a prestress loss is often observed in many large-scale prestressed concrete structures, such as long-span bridges and nuclear containments, which could significantly affect their safety and durability.

Under low stress, concrete can be considered an aging viscoelastic material, with concrete creep following the Boltzmann superposition principle. Thus, its strain rate can be expressed as (Bažant and Jirásek, 2018)

ε˙t=σ˙tEt+0tJ˙t,t0dσt0,(1)

where t is the age of concrete, t0 is the time when the load is applied, J˙t,t0 is the first derivative of the compliance function Jt,t0 with respect to t, Et is the instantaneous elastic modulus, and ε˙t and σ˙t are the strain and stress rates, respectively. Equation (1) can be solved by the finite difference method. However, this method has some limitations. First, the entire stress history is required to obtain the strain increment at the current time step. As a result, in the process of solving creep by the finite element method, the entire stress history needs to be stored for the integration points of each element. For a large-scale problem with many time steps, the evaluation of these history variables is quite time-consuming. Second, the effects of some variable factors, such as temperature, humidity, and concrete cracks, cannot be considered.

It should be noted that the calculation method of Jt,t0 comes from different creep models, reflecting the ratio of strain value to stress of the material at time t (loaded at t0), which is usually related to conditions such as material mix ratio, specimen shape, and environmental factors.

To overcome the difficulties of the finite difference method, Eq. (1) can be transformed into a differential rate-type equation, and an efficient exponential algorithm can be employed to calculate the concrete creep which exhibits a quadratic convergence rate and is unconditionally stable. Zienkiewicz et al. (1968) first applied this method to nonaging viscoelastic materials, and Bažant extended it to aging viscoelastic materials. This method can effectively improve computational efficiency with fixed internal variables. When using the efficient exponential algorithm, it is crucial to select a proper rheological model, such as the Kelvin chain model, to describe the viscoelastic behavior of the material. From a mathematical point of view, the creep compliance function of such a viscoelastic material can be approximated by the Dirichlet series. This can be achieved by curve fitting—the so-called retardation spectrum method.

The curve fitting method is usually based on the least squares method to obtain the coefficients of the Dirichlet series from test data. However, this method lacks actual physical meanings and does not follow the second law of thermodynamics, which sometimes leads to negative coefficients when the test data is not statistically ideal (Schapery, 1962). Many efforts have been made to solve the issue (Baumgaertel and Winter 1989; Elster and Honerkamp, 1991; Kaschta and Schwarzl, 1994; Mead, 1994; Emri and Tschoegl, 1995; Ramkumar et al., 1997; Park and Kim, 2001). Furthermore, for aging viscoelastic materials such as concrete, the coefficients of the Dirichlet series are time-dependent, and the curve fitting method becomes inefficient as the computational process needs additional optimization techniques.

With the continuous retardation spectrum method, the coefficients of the Dirichlet series can be determined by discretizing the continuous retardation spectrum, avoiding the issues encountered in the curve fitting method. Bažant and Xi (1995) studied the continuous retardation spectrum for concrete solidification theory and used the Post–Widder method to approximate the spectrum. In practice, however, a high-order Post–Widder formula is often needed to meet the precision requirement, which significantly increases the analytical complexity. Fortunately, Jirásek and Havlásek (2014) solved this issue using a low-order Post–Widder formula with time adjustment factors of retardation times. A high convergence speed and good accuracy are demonstrated for various concrete creep models. However, the method is heuristically based on empirical analyses, and the determination of time adjustment factors is highly dependent on personal experience and numerical experiments for different creep models.

The purpose of this paper is to develop an improved approach for efficiently approximating the continuous retardation spectra of various concrete creep models. The continuous retardation spectrum is first introduced, then the process of calculating the continuous retardation spectrum by the Post–Widder method and its corresponding shortcomings are analyzed, and an improved approach for solving the continuous retardation spectra based on the Weeks inverse Laplace transform method is proposed. By taking the CEB MC90 creep model as an example, the numerical solution of the continuous retardation spectra solved by the improved approach is analyzed. The proposed approach is then applied to the ACI 209R-92, JSCE, and GL2000 concrete creep models. Finally, the numerical results are compared with the other methods and some conclusions are drawn.

2 General solution for the continuous retardation spectrum

To describe viscoelastic materials, their constitutive properties can be represented by the Kelvin chain model. In the Kelvin chain model, the deformation of a material can be characterized by a number of Kelvin units and an additional spring unit assembled in series (Figure 1). Each Kelvin unit is composed of a spring and a dashpot assembled in parallel. All these units bear the same stress, and the total strain ε is equal to the sum of the deformations of each Kelvin unit and the spring unit.

ε=ε0+ε1+ε2++εM.(2)

FIGURE 1
www.frontiersin.org

FIGURE 1. Nonaging Kelvin chain model.

For the nonaging Kelvin unit μ shown in Figure 2, the total stress is equal to the sum of the elastic stress and the viscous stress.

σt=Eμεμt+ημε˙μt,(3)

where both the spring elastic modulus Eμ and the dashpot viscosity ημ do not change with age.

FIGURE 2
www.frontiersin.org

FIGURE 2. Nonaging Kelvin unit.

For aging viscoelastic materials such as concrete, the aging Kelvin chain model is needed. For the aging Kelvin unit μ, the stress–strain relationship can be expressed in rate form as

σ˙t=Dμtε˙μt+ημε¨μt,(4)

where the modified age-dependent modulus Dμt is equal to

Dμt=Eμt+η˙μt.(5)

Since σ˙t is equal to zero for creep tests, the boundary condition of εμt0=0 is satisfied and the strain of the aging Kelvin unit μ can be obtained as

εμt=σDμt01exptt0/τμ,(6)

where τμ=ημt/Dμt is the retardation time of the Kelvin unit μ and t0 is the time when the load is applied.

When the aging Kelvin chain model is subjected to a unit stress, the strains of all the Kelvin units (μ=0,1,2,,M) and the spring unit are superimposed; the compliance function is given by (Bažant, 1988)

Jt,t0=1E0t0+μ=1M1Dμt01exptt0/τμ,(7)

which can be considered a finite Dirichlet series. In the practical application process, once the compliance function is constructed in the form of Eq. (7), the long-term creep value of materials can be numerically calculated by exponential algorithm. For a given time t0, when there are infinite number of Kelvin units in the chain and the retardation time distributes continuously, Eq. (7) can be converted into an integral form (Bažant and Xi, 1995):

Jt,t0=1E0t0+01τLτ1exptt0/τdτ,(8)

where Lτ is the continuous retardation spectrum. For a specific creep model, once the continuous retardation spectrum is given, the direct discrete method can be used to construct the compliance function in the form of finite Dirichlet series. For convenience sake, the compliance function Jt,t0 is rewritten as

Jt,t0=1E0t0+φξ,(9)

where ξ=tt0 is the loading duration, and φξ is defined as

φξ=01τLτ1expξ/τdτ.(10)

By using the inverse Laplace transform method, Tschoegl (Nicholas, 1989) approximated Lτ from the Post–Widder method as

Lkτ=kτkk1!φkkτ,k=1,2,,(11)

where φkkτ is the kth order derivative of φkτ. When k, Lkτ converges to the exact solution:

Lτ=limkLkτ.(12)

In this method, a sufficiently smooth function φξ and a higher order k are required for an acceptable approximation of the continuous retardation spectrum. This method is straightforward and efficient for some problems with simple compliance functions. In practice, however, the computation of the high-order derivatives is usually quite complicated, and the process of solving them will become complicated. To improve this method, Jirásek and Havlásek (2014) analyzed the difference between the low-order Post–Widder method and the exact solution and proposed an approach for adjusting the retardation time by multiplying correction factor ατ to improve the accuracy of calculating continuous retardation spectrum. Thus, the continuous retardation spectrum is defined as

Lαττ=Lkτ.(13)

When the correction factor ατ is applied, the accuracy of the numerical solution of the compliance function is effectively improved, and the approximation order is well controlled within low ranges. In this method, however, the value of the correction factor is determined empirically and is different for different creep models.

In Section 3, in view of the shortcomings of the Post–Widder method, an improved method for solving continuous retardation spectra based on the Weeks inverse Laplace transform is proposed. Here, Eq. (10) is processed in advance.

The differentiation of Eq. (10) with respect to ξ yields

φξ=01τ2Lτexpξ/τdτ.(14)

By setting δ=1/τ, we have

φξ=0L1/δexpξδdδ.(15)

It can be seen from Eq. (15) that φξ is the Laplace transform of L1/δ. Therefore, once the inverse Laplace transform of φξ is determined, L1/δ can be obtained.

3 Weeks method for the continuous retardation spectrum

In the method by Weeks (1966), the Laguerre polynomials are used to numerically calculate the inverse Laplace transform. The main advantage is that an explicit solution can be obtained. In applying this method, the following two conditions should be fulfilled: the Laplace space function is a smooth function with bounded exponential growth, and it can be expressed as a Laguerre series. The above two conditions are fulfilled for commonly used creep models, including the CEB MC90, ACI 209R-92, JSCE, and GL2000 models.

For Laplace space function Fx, the analytical expression of the time-domain function ft can be obtained by the Weeks method as

ft=expσtn=0anexpbtLgn2bt.(16)

In Eq. (16), σ and b are two parameters that fulfill the conditions of b>0 and σ>σ0, where σ0 is the Laplace convergence abscissa. an contains the information of the Laplace space function. It could be a scalar, vector, or matrix but does not change with time. With the Maclaurin series, the analytical expression of an is given by

2b1ωFσ+2b1ωb=n=0anωn,ω<R,(17)

where R is the radius of convergence of the Maclaurin series.

Lgnx in Eq. (16) is the Laguerre polynomial of degree n defined as

Lgnx=expxn!dndxnexpxxn.(18)

For practical numerical calculation, Eq. (16) can approximately be expressed as

fNt=expσtn=0N1anexpbtLgn2bt,(19)

where N is an integer. The exact solution of ft can be obtained as N approaches infinity

ft=limNfNt.(20)

By using the Bromwich integral and the fast Fourier transform, an approximate expression of an in Eq. (17) can be obtained as

anexpinπ2N2Nm=NN12bexpinθm1expiθm+1/2Fσbexpiθm+1/2+1expiθm+1/21,(21)

where θm=mπ/N.

In the Laplace transform formula of Eq. (15), the Laplace space function is φξ, the complex variable is ξ, the time-domain function is L1/δ, and the time-domain variable is δ. From Eq. (19), L1/δ can be obtained through the inverse Laplace transform as

LN1/δ=expσδn=0N1anexpbδLgn2bδ.(22)

Substitution of τ=1/δ into Eq. (22) leads to

LNτ=expσ/τn=0N1anexpb/τLgn2b/τ,(23)
anexpinπ2N2Nm=NN12bexpinθm1expiθm+1/2φσbexpiθm+1/2+1expiθm+1/21.(24)

In applying the Weeks method, it is obvious that the truncation error can be reduced by using a larger value of N. In addition, a proper choice of σ and b can lead to a higher convergence speed. Therefore, it is very important to choose the values of σ and b reasonably. For Eq. (19), according to the method provided by Weeks (1966), when solving fNt in the range of 0<t<tmax, the value can be simply determined as

σ=σ0+1tmaxuσ0+1tmax,b=N2tmax,(25)
ux=0  if  x<0,1  if  x0.(26)

For Eq. (23), let tmax=1/τ, and, considering that for all creep models σ0=0, then σ and b can be taken as

σ=τ,b=Nτ2.(27)

Equation (27) is simple and suitable for all creep models and can directly improve the computational accuracy by increasing N, which is recommended to take a value between 10 and 50. In order to fully improve the computational efficiency, N is taken as 10 in this paper, and, to further determine the values of σ and b for different models, the analysis process is shown as follows.

To determine the parameters σ and b, Weideman (1999) proposed a method by minimizing the theoretical error E

E=expσtn=N2N1an+ϵn=0N1an,(28)

where ϵ is the machine round-off error.

Further analysis according to the method of Weideman (Weeks, 1966), taking the CEB MC90 creep model (CEB-FIP, 1993) as an example, the compliance function is

Jt,t0=ρt0+ϕ0βctt0Ecm28,(29)

where Ecm28 is the average elastic modulus of concrete at 28 days, ρt0 is the correction factor related to the loading time t0, ϕ0 is the nominal creep coefficient related to the material strength, loading time, and the relative humidity of the environment, and the creep development coefficient βctt0 is equal to

βctt0=tt0βh+tt00.3,(30)

with βh being a coefficient related to the volume/surface ratio, the relative humidity, and the material strength, generally ranging from 250 to 1,500. By setting φξ=βctt0, the first-order derivative of φξ is given by

φξ=0.3βhξ0.7βh+ξ1.3.(31)

When βh, N, and τ are given, the error E as a function of σ and b can be obtained by substituting Eqs (24) and (31) into Eq. (28). N should be selected by considering computational accuracy and efficiency. When a larger N is used, the choice of σ and b is more flexible.

By setting βh=500 and N=10, logE is obtained as a function of σ and b for different retardation times τ (Figure 3). The figure shows that the error E can be significantly reduced by a proper choice of the two parameters. In some zones where σ and b follow a certain relationship, the error E is kept within relatively low levels as the dark part shown in Figure 3.

FIGURE 3
www.frontiersin.org

FIGURE 3. Distributions of logE (CEB MC90 model) for (A) τ=10; (B) τ=100; (C) τ=1000.

Combine the three graphs in Figure 3 and represent the horizontal and vertical coordinates in exponential growth. Select several points with the smallest error E within different parameter ranges and use a few crosses to represent them, forming Figure 4.

FIGURE 4
www.frontiersin.org

FIGURE 4. Relationship between σ and b (CEB MC90 model).

If the relationship between σ and b can be determined under the condition of minimum error E, the method of parameter selection can be further determined. It was found that the relationship is related to the poles and branch points of the Laplace function (Weideman, 1999). If s1=α1+β1i and s2=α2+β2i are two farthest points to the origin of the Laplace function on the complex plane, we have

b2σ2+s22s12α2α1σ+α2s12α1s22α2α1=0.(32)

Equation (31) has two branch points: ξ=βh and ξ=0. By using Eq. (32), the relationship between σ and b for the CEB MC90 model can be obtained as

b=βhσ+σ2.(33)

By taking βh=500, Eq. (33) is plotted as the solid line shown in Figure 4. It is apparent from Figure 4 that the numerical results from error minimization are very close to Eq. (33) for the CEB MC90 model. Therefore, for the CEB MC90 creep model, Eq. (33) can be determined as the relationship between σ and b.

As seen from Eq. (23), the relationship between σ (or b) and τ is also required to perform the inverse Laplace transform and will be determined empirically through error analysis. Thus, a reference solution Ldτ is obtained by a combination of the Durbin inverse Laplace transformation (Durbin, 1974) and an adaptive numerical integration. It should be pointed out that, although the reference solution Ldτ has higher accuracy, it is very time-consuming and needs to be calculated by long-term iterative calculation and is thus not suitable for practical applications. The relative error Er is defined as

Er=LdτLNτLdτ.(34)

When βh=500, N=10, under the premise of using Eq. (33) for parameter values, the relative errors of the Weeks inverse Laplace transform for τ=0.01,1,100,1000 are shown in Figure 5.

FIGURE 5
www.frontiersin.org

FIGURE 5. Relative errors of the Weeks method (CEB MC90 model) for (A) τ=0.01; (B) τ=1; (C) τ=100; (D) τ=1000.

Figure 5 shows that, when τ is smaller than 1, the relative error exhibits larger changes from 10−4 to 1 (Figures 5A,B). When τ is larger than 100, the relative error is smaller than 10−1 and has a downward tendency as σ/τ increases (Figures 5C,D).

Based on the numerical results in Figure 5, an empirical value of σ for the CEB MC90 model is suggested as follows:

σ=0.038βhτ2,forτ<0.02βh,σ=0.038τ2,for0.02βhτ<100,σ=τ,forτ100.(35)

For different values of τ, the relative error calculated from Eq. (35) is also obtained as the red cross shown in Figure 5. It is seen from Figure 5 that, when using Eq. (35), the relative error is smaller than 10−1. Particularly when τ is equal to 100 or 1,000, the relative error is smaller than 10−3. Therefore, computational accuracy is guaranteed.

For N=10 and different values of βh, the continuous retardation spectra obtained from different methods are compared in Table 1, where Laτ is calculated from Eqs (33) and (35) while Lbτ is calculated from Eq. (27), Era and Erb are the relative errors between the calculated results obtained by the two calculation methods and the reference solution Ldτ, respectively. It is seen from Table 1 that, compared with the Weeks parameter value method—Eq. (27)—the relative errors of the continuous retardation spectra given by Eqs (33) and (35) are smaller in most cases, and the computational accuracy is obviously improved. For the Weeks method, however, the computational accuracy can also be improved by increasing the value of N.

TABLE 1
www.frontiersin.org

TABLE 1. Comparison of Lτ calculated from two different methods (CEB MC90 model).

4 Application of the Weeks method to different creep models

In this section, the Weeks method is applied to the CEB MC90, ACI 209R-92, JSCE, and GL2000 creep models. φξ is calculated and analyzed based on the aging Kelvin chain and different methods for the continuous retardation spectrum. N=10 is used in all cases. The results obtained by a combination of the Durbin inverse Laplace transformation (Durbin, 1974) and an adaptive numerical integration are used as a reference solution.

4.1 CEB MC90 model

As discussed in the previous section, the continuous retardation spectrum of the CEB MC90 model can be obtained by the Weeks method. When βh=500, the continuous retardation spectra calculated by the Weeks method, the Post–Widder method with different orders, and the reference solution are shown in Figure 6.

FIGURE 6
www.frontiersin.org

FIGURE 6. Continuous retardation spectra calculated by different methods (CEB MC90 model).

Once Lτ is known, φξ can be obtained from Eq. (10). However, the integral involved should be approximated by the Dirichlet series for an aging material. Therefore, the discrete retardation times τμ are selected based on accuracy and efficiency. If τm is assumed to be a geometric series with an initial value of τ0=104 and a base of 10, that is,

τm=104,103,,106m=0,1,2,,10,(36)

then Eq. (10) is changed to

φξ=τ=0τ=10τ01τLτ1-expξ/τdτ+m=110τ=10τm1τ=10τm1τLτ1-expξ/τdτ+τ=10τ10τ=+1τLτ1-expξ/τdτ.(37)

When τ<10τ0, 1-expξ/τ is close to unity, and the first term in Eq. (37) can be assumed to be 1/E0, which is not affected by the duration ξ. When τ>10τ10, Lτ is close to zero, and the third term in Eq. (37) tends to be zero. Thus, Eq. (37) reduces to

φξ=1E0+m=110τ=10τm1τ=10τm1τLτ1-expξ/τdτ.(38)

The integral in Eq. (38) can be approximated by the two-point Gaussian quadrature rule:

φξ1E0+ln102m=110L103/6τm1expξ103/6τm+L103/6τm1expξ103/6τm.(39)

The following definitions are used:

τμ=103/6τμ+1/2,for μ is odd,τμ=103/6τμ/2,for μ is even.(40)

Equation (39) then becomes

φξ1E0+μ=120ln102Lτμ1expξ/τμ.(41)

If 1/Dτμ is set to be ln10Lτμ/2, φξ of the aging Kelvin chain model is changed to

φξ=1E0+μ=1201Dτμ1expξ/τμ.(42)

Equation (42) is consistent with the form of Eq. (7), where the number of Kelvin units is set to M=20. 1/E0 can be obtained by numerical integration algorithms (Jirásek and Havlásek, 2014), but this will cause some errors. In this paper, in order to reduce the effect of numerical integration at low retardation times, 1/E0 is solved by subtracting the difference between the numerical solution and the exact solution of φξ, which is βctt0 in Eq. (30) for the CEB MC90 model:

1E0=βcξμ=1201Dτμ1expξ/τμ,(43)

where ξ is the loading duration when βcξ reaches the maximum value (ξ=105 for the CEB MC90 model). The analysis process is shown in Figure 7.

FIGURE 7
www.frontiersin.org

FIGURE 7. Analysis process of 1/E0.

By taking βh as 500, 800, and 1,200, the values of φξ are obtained by the Weeks and Post–Widder methods with different orders (Figure 8).

FIGURE 8
www.frontiersin.org

FIGURE 8. φξ calculated by different methods (CEB MC90 model) for (A) βh=500; (B) βh=800; (C) βh=1,200.

The error is defined as φξβcξ for the CEB MC90 model. Thus, the errors of different methods are shown in Figure 9. At the same time, the errors obtained from the method of Jirásek and Havlásek (2014) are also compared in Figure 9.

FIGURE 9
www.frontiersin.org

FIGURE 9. Errors of φξ calculated by different methods (CEB MC90 model) for (A) βh=500; (B) βh=800; (C) βh=1,200.

Equation (42) shows that φξ is related to τμ. For the two-point Gaussian quadrature rule, Lτμ corresponding to different τμ is able to describe the retardation spectrum curve as accurately as possible. If Lτμ at two adjacent points τμ are connected by line segments, the approximation of Lτμ to the retardation spectrum curve can be observed more intuitively. As shown in Figure 10, the retardation spectrum curve can be described well when τμ takes values from Eq. (40).

FIGURE 10
www.frontiersin.org

FIGURE 10. Approximation of Lτμ to the retardation spectrum (CEB MC90 model and βh=500).

In Eq. (42), the number of Kelvin units is taken as 20. If higher computational accuracy is required, the number of Kelvin units can be increased: M=20k,(k=1,2,3,). Thus, Eq. (42) becomes

φξ=1E0+μ=1M1Dτμ1expξ/τμ,(44)

where

τm,k=104,104+1/k,104+2/k,,106m=0,1,2,,10k,(45)
τμ=103/6kτμ+1/2,k,forμ is odd,τμ=103/6kτμ/2,k,forμ is even,(46)
1E0=βcξμ=1M1Dτμ1expξ/τμ,(47)
1Dτ=ln101kLτ2.(48)

It should be noted that, for most concrete creep models (CEB MC90, ACI 209R-92, and GL2000), the retardation spectra are relatively smooth, and taking k=1 in Eq. (45) usually satisfies the precision requirement. However, if the retardation spectrum changes sharply in terms of retardation time, it is necessary to increase the value of k, which will be discussed in Section 4.3.

4.2 ACI 209R-92 model

The creep compliance function of the ACI 209R-92 model (ACI committee 209, 2008) is

Jt,t0=1+ϕt,t0Et0,(49)

where ϕt,t0 is defined as

ϕt,t0=tt0ψd+tt0ψϕμ.(50)

In Eq. (50), ϕμ is the ultimate creep coefficient related to the curing conditions, slump, and ambient humidity of concrete structures and ψ and d are two parameters—usually taken as 0.6 and 10 according to the recommendations in ACI 209R-92, respectively.

To simplify the analysis process, φξ is defined as

φξ=ξψd+ξψ.(51)

Derivation of Eq. (51) with respect to ξ gives

φξ=dψξψ1d+ξψ2.(52)

Since Eq. (52) has a branch point of ξ=0, Eq. (32) cannot be directly used. For the ACI 209R-92 model, when N=10, the computational accuracy of the creep compliance function obtained from Eq. (27) is not very satisfactory (Figure 13). With reference to the analysis of the CEB MC90 model, Eq. (28) is used to calculate E and analyze the relationship between σ and b. Thus b is given by

b=46σ+σ2.(53)

When N=10, the empirical expression of σ is expressed as

σ=0.05τ2,forτ<20,σ=τ,forτ20.(54)

The continuous retardation spectra calculated by the Weeks method and the Post–Widder method with different orders are shown in Figure 11.

FIGURE 11
www.frontiersin.org

FIGURE 11. Continuous retardation spectra calculated by different methods (ACI 209R-92 model).

From Eq. (42), the values of φξ of the ACI 209R-92 model for different durations ξ are obtained as shown in Figure 12; the errors are shown in Figure 13. For comparison, the errors of φξ calculated from Eq. (27) are also shown in Figure 13 for N=10 and N=50. Figure 13 shows that, when Eq. (53) and (54) are adopted, the computational accuracy is obviously improved under the same N value.

FIGURE 12
www.frontiersin.org

FIGURE 12. φξ calculated by different methods (ACI 209R-92 model).

FIGURE 13
www.frontiersin.org

FIGURE 13. Errors of φξ calculated by different methods (ACI 209R-92 model).

4.3 JSCE model

The creep compliance function of the JSCE model (Uomoto et al., 2008) is defined as

Jt,t0=1exp0.09tt00.6εcr,(55)

where εcr is the ultimate creep strain under unit stress and related to the ambient relative humidity, temperature, and the volume/surface ratio. φξ is defined as

φξ=1exp0.09ξ0.6.(56)

Derivation of φξ with respect to ξ gives

φξ=0.054ξ0.4exp0.09ξ0.6.(57)

Since Eq. (57) has only a branch point ξ=0, Eq. (32) cannot be directly used to evaluate the values of σ and b. Instead, Eq. (27) is adopted to determine the values of the two parameters. For N=10, the continuous spectra calculated by the Weeks and Post–Widder methods with different orders are shown in Figure 14.

FIGURE 14
www.frontiersin.org

FIGURE 14. Continuous retardation spectra calculated by different methods (JSCE model).

It is seen from Figure 14 that, compared with the CEB MC90 and ACI 209R-92 models, the continuous retardation spectra calculated by the Weeks method are steeper at the peak. The continuous retardation spectra of the CEB and ACI models show significant changes in the range from τ=102 to τ=105, while that of the JSCE model mainly shows changes in the range from τ=101 to τ=103. If k=1 is set in Eq. (45), Lτμ cannot effectively describe the retardation spectrum curve. As shown in Figure 15A, the curve around τ=102 is not well reproduced. Therefore, it is necessary to increase the value of k. By setting the number of Kelvin units to M=40k=2 in Eq. (45)—it is seen from Figure 15B that the computational accuracy is greatly improved. Therefore, for the JSCE model, it is necessary to adopt this retardation time value method. The numerical results and the errors for different methods are shown in Figures 16, 17, respectively.

FIGURE 15
www.frontiersin.org

FIGURE 15. Approximation of Lτμ to the retardation spectrum (JSCE model) for (A) k=1; (B) k=2.

FIGURE 16
www.frontiersin.org

FIGURE 16. φξ calculated by different methods (JSCE model).

FIGURE 17
www.frontiersin.org

FIGURE 17. Errors of φξ calculated by different methods (JSCE model).

4.4 GL2000 model

The creep compliance function of the GL2000 model (Gardner and Lockman, 2002) is defined as

Jt,t0=1Ecmt0+ϕ28t,t0Ecm28,(58)

where Ecmt0 is the elastic modulus at loading time and ϕ28t,t0 is the creep coefficient at 28 days. The expression of ϕ28t,t0 is

ϕ28t,t0=ϕc2tt00.3tt00.3+14+7t00.5tt0tt0+70.5+2.511.086h2tt0tt0+0.12V/S20.5.(59)

In Eq. (59), ϕc is the correction term for the drying effect before loading, h is the relative humidity, and V/S is the volume/surface ratio of specimens. To simplify the calculation, φξ is expressed as

φξ=2ξ0.3ξ0.3+14+7t00.5ξξ+70.5+2.511.086h2ξξ+0.12V/S20.5.(60)

If φ1ξ and φ2ξ,γ are defined as

φ1ξ=ξ0.3ξ0.3+14,φ2ξ,γ=ξξ+γ0.5,(61)

the derivation of φξ with respect to ξ yields

φξ=2φ1ξ+7t00.5φ2ξ,7+2.511.086h2φ2ξ,0.12V/S2.(62)

Because the function form of φ2 is similar to Eq. (30), and the function form of φ1 is similar to Eq. (51), the parameter value forms of the CEB MC90 and ACI 209R-92 models can be referred to respectively. For N=10, the values of σ and b recommended for φ1ξ and φ2ξ,γ are

b1=σ1,σ1=0.6τ,(63)
b2=σ22+γσ2,σ2=15/γτ2,forτ<0.04γ,σ2=0.6τ,for0.04γτ<100,σ2=τ,forτ100.(64)

Two cases are considered, and the parameters are listed in Table 2. The continuous retardation spectra calculated by the Weeks and Post–Widder methods for different orders are shown in Figure 18.

TABLE 2
www.frontiersin.org

TABLE 2. Parameters for the two cases.

FIGURE 18
www.frontiersin.org

FIGURE 18. Continuous retardation spectra calculated by different methods (GL2000 model) for (A) case 1; (B) case 2.

It is apparent from Figure 18 that the continuous retardation spectra begin to grow significantly from τ=104 and decrease gradually at a long retardation time. Therefore, more Kelvin units are needed to cover the range of the continuous retardation spectra. Consequently, τ0 is set to 105, the maximum value of m in Eq. (36) is taken as 18, τ18=1013 for this model, and the number of Kelvin units in Eq. (42) changes from 20 to 36. The values of φξ of the GL2000 model for different duration ξ are obtained as per Figure 19, and the errors are shown in Figure 20. Since the GL2000 model was not discussed by Jirásek and Havlásek (2014), the error obtained from their method is not compared in Figure 20.

FIGURE 19
www.frontiersin.org

FIGURE 19. φξ calculated by different methods (GL2000 model) for (A) case 1; (B) case 2.

FIGURE 20
www.frontiersin.org

FIGURE 20. Errors of φξ calculated by different methods (GL2000 model) for (A) case 1; (B) case 2.

4.5 Discussion

For all the concrete creep models discussed in this section, the proposed approach based on the Weeks method improves the computational accuracy and efficiency of the continuous retardation spectra and the creep compliance. It should be noted that, for the case of long duration, the errors of φξ decrease with the increase of ξ and approach zero. There are two main reasons for this phenomenon. First, the creep compliances used have negligible variation for a long duration and have a clear upper limit. Thus, the approximation through the Dirichlet series will maintain the same characteristics as long as the retardation time range selected is wide enough. Second, by calculating 1/E0 based on the most accurate compliance value—Eq. (43)—the effect of numerical integration for a short duration on the final results is reduced.

For the creep models considered, when ξ1, the absolute error of φξ given by the improved approach (using the Weeks method) is smallest and maintains a lower level (usually smaller than 0.01). For the ACI 209R-92 and JSCE models, the results from the proposed approach achieve higher accuracy than the Post–Widder method. When calculated by the improved approach, φξ of the ACI 209R-92 model is closest to the reference solution, while φξ of the JSCE model exhibits the most significant improvement on the Post–Widder method.

The errors of φξ obtained from the improved approach and the Jirásek–Havlásek method are both controlled at a relatively low level. For the CEB MC90 model, the results obtained from Jirásek–Havlásek are more accurate for low retardation times, while those obtained from Weeks are more accurate for high retardation times. As shown in Figure 9, the errors corresponding to the Weeks method are smaller compared to the Jirásek–Havlásek method when ξ>10. For the ACI 209R-9 and JSCE models, the results obtained from Weeks are more accurate and the errors of φξ are smaller.

For the other creep models which are not discussed in this paper, the proposed approach is still applicable when they fulfill the requirement of the Weeks inverse transform. In general, a large value of N can be used to achieve the required computational accuracy and a proper choice of σ and b can significantly improve computational efficiency. In general, Eq. (27) can be used to determine the parameters, and 10N50 is recommended. The larger the value of N, the more accurate the calculation results, but computational efficiency will be reduced. If φξ has two or more poles and branch points, Eq. (32) can also be used to determine the parameters, and precision requirements are usually satisfied when N=10.

5 Comparison with experimental results

To further verify the validity of the improved approach, the finite element method is used to compute the long-term creep of concrete by combining the retardation spectrum obtained by the Weeks method with the second-order exponential algorithm (Bažant and Jirásek, 2018). For this purpose, two sets of experimental data, OPC and SF10, were selected from Mazloom et al. (2004). They had different mix proportions, and a pressure of 10 MPa was applied on the OPC and SF10 cylinders on the 28th and 7th days, respectively. Two UMAT user subroutines for material behavior—ACI UMAT and CEB UMAT—of the commercial finite element software ABAQUS were coded. The concrete strain was calculated using the CEB MC90 and ACI 209R-92 creep models. The results are shown in Figure 21, which shows that the finite element results are in good agreement with the experimental results. For the OPC group, the ACI 209R-92 model has higher accuracy, while, for the SF10 group, the CEB MC90 has higher accuracy.

FIGURE 21
www.frontiersin.org

FIGURE 21. Comparison between experimental results and numerical results for (A) OPC; (B) SF10.

It is noted that one of the purposes of this paper is to ensure that the numerical results agree well with the corresponding expressions of Jt,t0. As a result, the accuracy of the final numerical results is mainly dependent on whether the analytical expression of Jt,t0 given by design codes is close to the practical situation. Therefore, before the concrete creep is calculated, it is necessary to select a reasonable model according to the practical situation, including the mix proportion, the component shape, loading, and environmental conditions, to ensure the computational accuracy of the numerical results. It should also be noted that, since this paper is mainly concerned with improving on the continuous retardation method, finding a viable, stable strategy to identify the optimal dimension of the approximation in the presence of an error (uncertainty) on data is not addressed. This limitation is expected to be removed in our future work.

6 Conclusion

Based on the Weeks method, an efficient approximation approach has been developed for the continuous retardation spectra of aging viscoelastic materials. Compared with the existing methods, the approach has several advantages.

(1) It can calculate the continuous retardation spectrum more accurately by only using the first-order derivative of the creep compliance function. The difficulty of calculating the high-order derivatives in the Post–Widder method is avoided.

(2) Unlike the method proposed by Jirásek and Havlásek (2014), in which the correction factor is empirically determined for each concrete creep model at a given derivative order, the proposed approach is based on a solid theoretical foundation and can be conveniently applied to various concrete creep models.

(3) Better computational accuracy can be achieved for a long loading duration. As illustrated by different concrete creep models, the error of the creep compliance function obtained by the proposed approach is controlled within 0.02 for a loading duration of 101<ξ<106. Therefore, the approach is applicable to long-term creep analyses of aging viscoelastic materials, such as concrete.

It should be noted that the proposed approach is only applicable to concrete creep models when the first-order derivative of the compliance function fulfills the requirement of the Weeks method. For concrete creep models with logarithmic compliance functions, such as the fib model (CEB-FIP, 2010), the inversion formula of the Laplace transform has an analytical solution and does not require the Weeks method for the continuous retardation spectrum. In this research, to achieve high computational efficiency and accuracy, the values of σ and b for the CEB MC90, ACI 209R-92, and GL2000 models are determined empirically for N=10. As for the JSCE model, a good precision requirement can be achieved when the parameters are taken directly through Eq. (27) for N=10. If the proposed approach is applied to other creep models, the parameters can be determined by referring to the analytical process of this paper or directly by increasing the value of N to meet the precision requirements.

Data availability statement

The raw data supporting the conclusion of this article will be made available by the authors, without undue reservation.

Author contributions

XZ: Conceptualization, Methodology, Writing–review and editing. LB: Formal Analysis, Validation, Writing–original draft. HR: Data curation, Software, Writing–original draft. XF: Funding acquisition, Project administration, Supervision, Writing–original draft. JZ: Funding acquisition, Resources, Writing–review and editing. YG: Investigation, Visualization, Writing–original draft.

Funding

The author(s) declare financial support was received for the research, authorship, and/or publication of this article. This research was funded by the National Science Foundation for Excellent Young Scholars (Grant No. 52222808), the Zhejiang Provincial Natural Science Foundation (Grant No. LY20E080027), and the National Natural Science Foundation of the People’s Republic of China (Grant Nos 52008413, 52078509, 52178255, and 52278279).

Conflict of interest

Authors HR, XF, and YG were employed by Metallurgical Group Corporation of China.

The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

The handling editor J-GD declared a past co-authorship with the author JZ.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors, and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Supplementary material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fmats.2024.1340883/full#supplementary-material

References

ACI Committee 209 (2005). Report on factors affecting shrinkage and creep of hardened concrete. Farmington Hills: American Concrete Institute.

Google Scholar

ACI Committee 209 (2008). Guide for modeling and calculating shrinkage and creep in hardened concrete. Farmington Hills: American Concrete Institute.

Google Scholar

Baumgaertel, M., and Winter, H. H. (1989). Determination of discrete relaxation and retardation time spectra from dynamic mechanical data. Rheol. Acta 28, 511–519. doi:10.1007/BF01332922

CrossRef Full Text | Google Scholar

Bažant, Z. P. (1988). Mathematical modeling of creep and shrinkage of concrete. New York: John Wiley and Sons.

Google Scholar

Bažant, Z. P., Hauggaard, A. B., Baweja, S., and Ulm, F. J. (1997). Microprestress-solidification theory for concrete creep. I: aging and drying effects. J. Eng. Mech. 123, 1188–1194. doi:10.1061/(asce)0733-9399(1997)123:11(1188)(1997)123:11(1188)

CrossRef Full Text | Google Scholar

Bažant, Z. P., and Jirásek, M. (2018). Creep and hygrothermal effects in concrete structures. Netherland: Springer Dordrecht.

Google Scholar

Bažant, Z. P., and Xi, Y. (1995). Continuous retardation spectrum for solidification theory of concrete creep. J. Eng. Mech. 121, 281–288. doi:10.1061/(asce)0733-9399(1995)121:2(281)(1995)121:2(281)

CrossRef Full Text | Google Scholar

CEB-FIP (1993). Design of concrete structures, CEB-FIP model-code 1990. London: British Standard Institution.

Google Scholar

CEB-FIP (2010). Fib Model Code for concrete structures 2010. London: British Standard Institution.

Google Scholar

Durbin, F. (1974). Numerical inversion of laplace transforms: an efficient improvement to dubner and abate's method. Comput. J. 17, 371–376. doi:10.1093/comjnl/17.4.371

CrossRef Full Text | Google Scholar

Elster, C., and Honerkamp, J. (1991). Modified maximum entropy method and its application to creep data. Macromolecules 24, 310–314. doi:10.1021/ma00001a047

CrossRef Full Text | Google Scholar

Emri, I., and Tschoegl, N. W. (1995). Determination of mechanical spectra from experimental responses. Int. J. Solids Struct. 32, 817–826. doi:10.1016/0020-7683(94)00162-P

CrossRef Full Text | Google Scholar

Gardner, N. J., and Lockman, M. J. (2002). Design provisions for drying shrinkage and creep of normal-strength concrete. ACI Mater. J. 98, 159–167. doi:10.14359/10199

CrossRef Full Text | Google Scholar

Hubert Rüsch, D. J., and Hilsdorf, H. K. (1983). Creep and shrinkage. New York: Springer.

Google Scholar

Jirásek, M., and Havlásek, P. (2014). Accurate approximations of concrete creep compliance functions based on continuous retardation spectra. Comput. Struct. 135, 155–168. doi:10.1016/j.compstruc.2014.01.024

CrossRef Full Text | Google Scholar

Kaschta, J., and Schwarzl, R. R. (1994). Calculation of discrete retardation spectra from creep data-Ⅰ. Method. Rheol. Acta 33, 517–529. doi:10.1007/BF00366336

CrossRef Full Text | Google Scholar

Mazloom, M., Ramezanianpour, A. A., and Brooks, J. J. (2004). Effect of silica fume on mechanical properties of high-strength concrete. Cem. Concr. Compos. 26, 347–357. doi:10.1016/s0958-9465(03)00017-9

CrossRef Full Text | Google Scholar

Mead, D. W. (1994). Numerical interconversion of linear viscoelastic material functions. J. Rheology 38, 1769–1795. doi:10.1122/1.550526

CrossRef Full Text | Google Scholar

Nicholas, W. T. (1989). The phenomenological theory of linear viscoelastic behavior. Heidelberg: Springer.

Google Scholar

Park, S. W., and Kim, Y. R. (2001). Fitting prony-series viscoelastic models with power-law presmoothing. J. Mater. Civ. Eng. 13, 26–32. doi:10.1061/(asce)0899-1561(2001)13:1(26)(2001)13:1(26)

CrossRef Full Text | Google Scholar

Ramkumar, D. H. S., Caruthers, J. M., Mavridis, H., and Shroff, R. (1997). Computation of the linear viscoelastic relaxation spectrum from experimental data. J. Appl. Polym. Sci. 64, 2177–2189. doi:10.1002/(SICI)1097-4628(19970613)64:11<2177::AID-APP14>3.0.CO;2-1

CrossRef Full Text | Google Scholar

Schapery, R. (1962). "A simple collocation method for fitting viscoelastic models to experimental data", in: Graduate Aeronautical Laboratory. (California: California Institute of Technology).

Google Scholar

Uomoto, T., Ishibashi, T., Nobuta, Y., Satoh, T., Kawano, H., Takewaka, K., et al. (2008). Standard specifications for concrete structures-2007 by Japan society of Civil engineers. Concr. J. 46, 3–14. doi:10.3151/coj1975.46.7_3

CrossRef Full Text | Google Scholar

Weeks, W. T. (1966). Numerical inversion of laplace transforms using Laguerre functions. J. ACM 13, 419–429. doi:10.1145/321341.321351

CrossRef Full Text | Google Scholar

Weideman, J. a.C. (1999). Algorithms for parameter selection in the weeks method for inverting the laplace transform. Siam J. Sci. Comput. 21, 111–128. doi:10.1137/s1064827596312432

CrossRef Full Text | Google Scholar

Zienkiewicz, O. C., Watson, M., and King, I. P. (1968). A numerical method of visco-elastic stress analysis. Int. J. Mech. Sci. 10, 807–827. doi:10.1016/0020-7403(68)90022-2

CrossRef Full Text | Google Scholar

Keywords: improved approach, concrete creep, continuous retardation spectra, Weeks inverse Laplace transform, Dirichlet series

Citation: Zhou X, Bai L, Rong H, Fan X, Zheng J and Geng Y (2024) An improved approach for the continuous retardation spectra of concrete creep and applications. Front. Mater. 11:1340883. doi: 10.3389/fmats.2024.1340883

Received: 19 November 2023; Accepted: 02 January 2024;
Published: 07 February 2024.

Edited by:

Jian-Guo Dai, Hong Kong Polytechnic University, Hong Kong SAR, China

Reviewed by:

Roberto Fedele, Polytechnic University of Milan, Italy
Xiaohua Li, Chongqing University, China

Copyright © 2024 Zhou, Bai, Rong, Fan, Zheng and Geng. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Xinglang Fan, fanxinglang@cribc.com

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.