Next Article in Journal
Investigation of the Microcharacteristics of Asphalt Mastics under Dry–Wet and Freeze–Thaw Cycles in a Coastal Salt Environment
Next Article in Special Issue
Validation and Investigation on the Mechanical Behavior of Concrete Using a Novel 3D Mesoscale Method
Previous Article in Journal
Effect of Multi-Walled Carbon Nanotubes on Improving the Toughness of Reactive Powder Concrete
Previous Article in Special Issue
Generalized Softened Variable Angle Truss Model for RC Hollow Beams under Torsion
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nonlinear Stress-Strain Model for Confined Well Cement

1
College of Petroleum Engineering, China University of Petroleum, Beijing 102249, China
2
Mewbourne School of Petroleum and Geological Engineering, The University of Oklahoma, Norman, OK 73019-1003, USA
3
School of Civil Engineering, Dalian University of Technology, Dalian 116024, China
*
Author to whom correspondence should be addressed.
Materials 2019, 12(16), 2626; https://doi.org/10.3390/ma12162626
Submission received: 29 June 2019 / Revised: 5 August 2019 / Accepted: 15 August 2019 / Published: 17 August 2019
(This article belongs to the Special Issue Modeling of Cementitious Materials and Structures)

Abstract

:
The cement sheath is the key for providing the zonal isolation and integrity of the wellbore. Oil well cement works under confining pressure, so it exhibits strong nonlinear and ductile behavior which is very different from that without confining pressure. Therefore, for the accuracy of the simulation and the reliability of well construction design, a reliable compression stress–strain model is essential for confined well cement. In this paper, a new axial stress–strain model for confined well cement is developed based on uniaxial and triaxial test data, examinations of failure mechanisms, and the results of numerical analysis. A parametric study was conducted to evaluate and calibrate the model. The model is simple and suitable for direct use in simulation studies and well design. Results from this study show the nonlinear compressive behavior of confined well cement can be predicted using the traditional uniaxial compressive strength test measurements.

1. Introduction

It is increasingly important to develop safer, environmentally friendly, and cost-effective technologies to satisfy the growing oil demand while conventional resources are decreasing [1]. As the main barrier to ensure zonal isolation and maintain the integrity of the wellbore, cement sheath must have an acceptable mechanical strength to withstand severe loading conditions resulting from the application of these new technologies. Consequently, more accurate mechanical strength models and better assessment of cement integrity are necessary to reduce the risk of failure. Numerical simulation is commonly applied to investigate the mechanical integrity of cement [2], also by using some software such as ABAQUS [3,4], ANSYS [5], and ADINA [1]. In all types of applications, an accurate well cement stress–strain model is required.
The linear elastic stress–strain model of hardened cement paste is universally used as a material function to perform numerical simulations [6]. Well cement is a brittle material when it is unconfined. Thus, using a linear elastic model is appropriate because cement fails quickly once the stress reaches the ultimate strength limit. For example, a linear elastic model can be used to simulate uniaxial tests of cement specimens. However, when well cement is set in the annulus and confined between casing and formation, it exhibits mechanical properties which are greatly different from that of unconfined cement [7]. Under confining pressure, the ultimate strain and strength of cement increase significantly, and its behavior becomes more ductile and nonlinear [6,8,9,10]. The peak volumetric strain and its corresponding axial strain increase linearly with increasing confining pressures [11,12]. The cement shows more plasticity and can hold more pressure cycles as the confining pressure increases [13]. Obviously, strong nonlinear characteristics cannot be described by a linear elastic stress–strain model. Moreover, the nonlinearity becomes stronger with the use of strength-enhancing additives [14].
In fact, compressive strength and the corresponding strain, ultimate strain, and post-peak behavior required to describe the nonlinearity of confined cement are difficult to obtain from triaxial test data [4], particularly those monotonically increasing curves shown in Figure 1. In addition, the post-peak material behavior of well cement is less reported, even for simple stress states [4]. Some investigations on the material behavior of confined concrete (i.e., a cementitious material) can be referred to due to the similarities of mechanical properties between cement and concrete. They both show brittleness without confinement and nonlinearity and ductility under confining conditions, though they have different properties. Confined concretes are classified based on the types of confinement as: Steel-confined concrete, steel tube confined concrete, and fiber-reinforced plastic (FRP) confined concrete. All of these confining conditions are similar to the entrapment of cement between the casing and formation.
The stress–strain model of Popovics [15] and Mander et al. [16] was originally proposed for steel-confined concrete, and subsequently used in FRP confined concrete [17,18], and then adopted for steel tube confined concrete [19]. This model uses a single equation to relate the compressive strength of confined concrete to the unconfined concrete compressive strength and confining pressure. The corresponding strain is determined based on the ratio of confined compressive strength to unconfined compressive strength, and the strain at failure during the unconfined test. In addition, Mander’s model and ACI 318-08 models were also applied for confined geopolymer concretes [20,21,22,23,24].
Sakino et al. [25] proposed a similar stress–strain model for steel tube-confined concrete. The model is adopted from earlier studies [26,27,28,29]. A single equation is used to express the stress–strain model. The strength of confined concrete is determined by multiplying the strength of unconfined concrete by a confinement factor. The strain at failure is related to its mechanical properties and dimensions of the steel tube. The enhancement of the strain at failure due to confinement increases with the reduction of the diameter to thickness ratio of a steel tube.
Another model [30] is developed for FRP-confined concrete. In this model, the nearly linear portion of the stress–strain curve is characterized by its slope and intercept. The model is appropriate for sufficiently confined concrete because it represents a monotonously increasing stress–strain curve. Other existing models [16,25] represent a decreasing post-peak stress–strain curve. In summary, using these single-equation models can represent the behavior of confined concrete. Though the models seem simple, they lead to a more complex equation form and complicated stress analysis procedures [31].
The typical bilinear stress–strain model, which includes two straight lines was applied for confined concrete [27,28]. This approach is simple, but straight lines cannot represent the nonlinear characteristic of the stress–strain curve. Therefore, a nonlinear function, such as a parabola, is commonly used to describe the pre-peak portion of the stress–strain model. Popovics’ curve was applied [32,33] followed by exponential function [34] or equations established using experimental data. Parabolas are directly used [35] or applied with some modification such as Hognestad’s parabola [31,36]. Other studies [19,37] used a parabola followed by a single or multiple straight lines. The parabolic function is also incorporated into some codes of practice such as BS 8110 [38] and EN 1992-1-1 [39] for concrete design.
In this study, the stress–strain model of confined cement is proposed based on experimental results, failure mechanism analyses, and trial calculations. The model can reasonably predict the compressive strength of cement under confining pressure. It describes the nonlinear characteristics of the material, which is not captured using the linear model. The proposed model was used in commercial software to carry out simulation studies and the predicted results are compared with experimental data to calibrate the model and evaluate its performance.

2. Experimental Study

To calibrate and evaluate the new model, uniaxial, and triaxial compressive tests were conducted on conventional oil well cement specimens. Class G cement, silica flour, and additives were used to prepare test specimens. The compositions of all materials in the slurry formulation are presented in Table 1. Sample series 2 and 3 are conventional neat cements, which were obtained from the China National Petroleum Corporation (CNPC) and a previous study [40], respectively. All samples presented in Table 1 were tested using a triaxial rock testing system (GCTS Co. Model RTR-1000).
For specimens in Sample 1 series, one was tested in uniaxial (Sample 1-0), second (Sample 1-10), and third (Sample 1-20) in triaxial compression with 10 and 20 MPa confining pressure. These specimens were prepared with conventional cement and toughness-modifier was added during the slurry preparation. Therefore, the samples represent toughness-enhanced cement. The cement slurries were prepared according to API RP 10B-2 [41] standards. For Samples 1, the slurries were poured into cylindrical molds and cured for 7 days in 70°C water under atmospheric pressure. The size of the cylindrical samples was 2.54 cm in diameter and 5.08 cm in height (i.e., the height-diameter ratio was 2). This size is commonly used in standards and cement studies [14,42,43]. The height to diameter ratio of 2 meets applicable ISRM, ASTM, and cement testing standards [43], which is considered to be the accepted specimen geometry to test the compressive strength of cement [42,44,45,46,47,48]. Mechanical properties of cement Samples 1 in the uniaxial and triaxial tests were measured using an RTR-1000 triaxial rock testing system (GCTS Co., Tempe, AZ, USA). During the triaxial tests, confining pressure was first applied with the ramp rate of 50 N/S until 10 MPa for Sample 1-10 and 20 MPa for Sample 1-20. Then, an axial strain ramp rate of 0.05% per minute was maintained while measuring the deviator stress.
In Sample 2 series, two specimens were tested in uniaxial (Sample 2-0A and Sample 2-0B) and the other two specimens (Sample 2-10A and Sample 2-10B) in triaxial compression with 10 MPa confining pressure. For Sample 3 series, Sample 3-0 was used in uniaxial and Sample 3-20 in a triaxial test with 20 MPa confinement. The preparation of cement specimens and compressive strength test procedures were both according to API RP 10B-2 [41] and SY/T 6466 [43] standards. They were also 2.54 cm in diameter and 5.08 cm in height.
Figure 1a shows the schematics of stress loads applied on specimens during uniaxial and triaxial compressive tests. The surrounding temperature was approximately 22°C for Samples 1 and 2, and 60°C for Sample 3. Figure 1b–d presents the stress responses in the compression experiments, i.e., deviator stress (y-axis) as a function of axial strain (x-axis). Deviator stress is equal to σ1-σ3. σ1 is the axial stress and σ3 is the confining stress. It can be seen that both the deviator stress and the ultimate strain are significantly enhanced under confining conditions. In addition, the enhancement increased with the level of confinement in Figure 1b.

3. Stress–Strain Model of Confined Well Cement

Based on experimental results and extensive modeling studies on well cement, a stress–strain model is developed to represent the deformation behavior of cement under confined condition (Figure 2). The model considers three parts (segments) in the stress–strain curve with various parameters that are dependent on cement properties and confinement characteristics. Segment OA is represented by a second-degree parabola with point A at εc0 and fc0. The fc0 represents the ultimate compressive strength of confined well cement. Peak stress segment (line AB) is a horizontal line, in which εc0 and εc1 are the minimum and maximum strains that correspond to the ultimate strength, fc0. The descending segment (BD) represents the post-peak (softening) region.

3.1. Modeling the Ascending Segment

Since many researchers and standards [19,35,37,38,39] consider that the appropriate approach for describing the first portion of the stress–strain curve is to use a modified parabola with various parameters. Hence, the ascending portion is modeled as a quadratic function. Thus:
σ = α 1 f c 0 ε c 0 2 ε 2 + β 1 f c 0 ε c 0 ε , ε ε c o ,
where σ and ε are the axial stress and axial strain, respectively. α1 and β1 are model parameters which control the slope and curvature of segment OA. The parameters are dependent on elastic modulus (E) and the stress at the linear limit (σp0). The parameters E, σp0, fc0, εc0, α1, and β1 need to be obtained from a uniaxial test to fully describe stress–strain characteristics of cement in this segment.

3.1.1. Determination of Elastic Modulus E and Stress at the Linear Limit σp0

Elastic modulus affects mechanical properties of cement during the elastic deformation phase. The comparisons of the uniaxial and triaxial test data shown in Figure 3 demonstrate that the linear elastic phase of confined cement is basically close to that of unconfined cement. Particularly, the elastic modulus of confined cement is very close to that of unconfined one. For the Sample 2 series, the strain trends of uniaxial and triaxial tests are very similar when the axial strain is less than 0.25% (Figure 3d). This indicates that within the elastic limit, mechanical properties are little affected by confinement and the confinement effect is triggered once micro-cracks develop in the cement due to excessive loading. Therefore, in this study, elastic modulus E and the stress at the linear limit σp0 of confined cement are considered to be the same as that of unconfined cement.
It can be seen from the uniaxial test results (Figure 4) that the stress–strain curve is basically linear when the stress is below 70% of fc0 and begins to deviate from the straight line when it is between 70% to 90% of fc0. Therefore, the stress at the linear limit (σp0) of confined cement is expected to be between 0.7 fc0 and 0.9 fc0. If σp0 = φfc0, then φ can be taken as the range between 0.7 and 0.9.

3.1.2. Determination of the Ultimate Strength fc0 and the Corresponding Strain εc0

After the linear elastic phase, as the cement approaches its unconfined ultimate strength, micro-cracks develop and grow to the extent that the Poisson ratio can no longer describe the relationship between the lateral and axial strains. In addition, the confinement, which is a result of the formation and casing becomes the sole restraining mechanism against crack development and catastrophic failure. In fact, it is difficult to obtain the ultimate strength of confined cement from the triaxial test results because as the load increases the cement continues to maintain its strength and resist the loading due to the confinement (Figure 3). The monotonous increase in the stress occurs even though parts of the material have reached the ultimate strength and can maintain it. Based on the above analyses of failure mechanism, fc0 and εc0 in the stress–strain model of confined cement are considered to be equal to the ultimate strength and corresponding strain of unconfined cement, which can be obtained from uniaxial test data.

3.1.3. Determination of α1 and β1

According to Equation (1), parameters α1 and β1 control the slope and curvature of the ascending segment (OA). The peak point A (εc0, fc0) and the point at the linear limit (φfc0/E, φfc0) are both in the curve. Hence, substituting the coordinates of point A (εc0, fc0) into Equation (1), the following equation can be obtained.
α 1 + β 1 = 1
Then, substituting point (φfc0/E, φfc0) into Equation (1), the following equation can be obtained.
α 1 φ ( f c 0 E ε c 0 ) 2 + β 1 f c 0 E ε c 0 = 1
Therefore, the expressions for α1 and β1 can be obtained combining Equations (2) and (3). Thus:
α 1 = E ε c 0 f c 0 1 1 φ f c 0 E ε c 0 ,
β 1 = E ε c 0 f c 0 φ f c 0 E ε c 0 1 φ f c 0 E ε c 0

3.2. Modeling Peak Stress Segment

The strain increases rapidly after the peak point A, but the cement continues to resist the loading rather than immediately failing. A horizontal straight line or slightly ascending straight line is used in confined cementitious materials to describe the peak stress portion [31,38,39]. Besides this, some studies reported strain doubling in this phase [37,49]. Hence, this study assumes the peak stress portion (AB) to be a horizontal line, which extends from εc0 to 2εc0 (Figure 2). The equation for the peak stress portion is described as follows:
σ = f c 0 ,   ε c 0     ε   ε c 1
where εc0 and εc1 are the minimum and the maximum strain of the peak stress portion and εc1 is equal to 2εc0.

3.3. Modeling Descending Segment

The descending segment BD exhibits post-peak behavior. A number of studies [19,37,50] used straight lines to model the descending segment, which is simple and suitable for designing and performing numerical simulations. Therefore, in this study, the straight-line approach is adopted to describe the descending portion of the stress–strain model. Point C is necessary because experiments often do not prolong beyond this point due to the sudden failure of specimens. It would be better not to set the stress at point C too low. In this study, it is considered to be 0.65fc0. The strain at point C, εc2, and the slope of BC are required to formulate an equation for this portion of the curve. The equation of this portion (BC) can be expressed as:
σ f c 0 ε ε c 1 = 0.35 f c 0 ε c 1 ε c 2 ,   ε c 1 ε ε c 2 .
Let the slope of portion BC be -CZZfc0, and the following equation can be obtained:
C Z Z = 0.35 ε c 2 ε c 1
where both Cz and Z need to be determined. Based on trial calculations, a reasonable value of Z is found to be 30 and that of Cz is found to be between 0.85 and 2.5. Considering εc1 = 2εc0 and applying Equation (8), εc2 can be determined using the following equation:
ε c 2 = 0.35 C Z Z + 2 ε c 0 .
Beyond point C, the straight line can be expected to continue with the same slope until the cement completely fails. fr represents the residual strength, and εr represents the corresponding strain. If the confinement is large enough, it can be assumed that the cement is completely crushed at this stage; thus, fr is close to zero. However, to avoid numerical instability in the model, a value close to zero is recommended for fr, instead of zero.

4. Results and Discussions

4.1. Model Evaluation

To evaluate and calibrate the proposed stress–strain model, simulation studies were conducted using commercial software (ABAQUS) and results are compared with experimental measurements. First, the uniaxial test of Sample 1-0 was simulated using ABAQUS to calibrate the calculation process. The software is used to reproduce the test data shown in Figure 5. It can be seen that the simulation results predominantly match the experimental measurements, indicating the applicability of the calculation process with ABAQUS.
Arjomand et al. [4] used a uniaxial test stress–strain curve as a stress–strain model of cement to simulate its confined stress–strain behavior using ABAQUS. This procedure was applied in this study to reproduce triaxial test results of Samples 1-10 and 1-20 (Figure 6). It can be observed that numerically obtained curves do not fit the triaxial test data, though they have reasonably reproduced uniaxial test measurements (Figure 5). The results demonstrate that the stress–strain model of confined cement is significantly different from that of unconfined cement. Therefore, it is essential to propose an accurate and suitable stress–strain model for confined well cement.
The new stress–strain model is used to simulate the triaxial test results of all cement samples. The related parameters of the stress–strain model during simulations are presented in Table 2. These parameters were obtained according to the procedures proposed in Section 3. Figure 7 shows the corresponding models, in which Samples 1 represent toughness-enhanced cement and Samples 2 and 3 plain cement. Compared to the plain cement samples, the toughness-enhanced ones have higher strength and ultimate strain, thereby exhibiting significantly increased toughness and ductility.
It can be seen from Figure 8, using the proposed models, the numerical simulation results closely match the experimental data for both types of cement samples, while using Mander’s model [16] does not provide good predictions. When the new model is used, the predicted and experimental curves at the initial stage are in good agreement indicating that the values of parameters E, α1, β1, and φ are reasonable. The curves in the plastic stage also match well, which demonstrates the suggestions about the peak stress portion and descending portion are also appropriate for confined cement. Therefore, the proposed stress–strain model, including the equations and parameters are reasonable and suitable for not only plain cement but also toughness-enhanced cement under confining conditions.

4.2. Effects of Model Parameters

4.2.1. Effects of Strength fc0

As discussed in Section 3, the strength of confined cement in the model, fc0, is equal to that of unconfined cement obtained from uniaxial test data. The peak stresses in the uniaxial tests were 60 MPa for Sample 1, 47.5 MPa for Sample 2, and 37.8 MPa for Sample 3. To investigate the effects of the fc0 on the accuracy of the model, a parametric study has been conducted. The trial values of fc0 are 57, 60, and 65 MPa for Sample 1-10 and 1-20, 45–50 MPa for Sample 2, and 35–40 MPa for Sample 3. Other parameters of the stress–strain model are presented in Table 2. Figure 9 shows the numerical and experimental results for all samples. The deviator stresses increase slightly with the value of fc0. In fact, all numerical results are close to the experimental values, as long as fc0 in the model is taken as a value near the strength of unconfined cement obtained from uniaxial tests. It indicates that the suggested value of fc0 provides stable and accurate simulation results.

4.2.2. Effects of CZ

CZ controls the slope of the descending portion of the stress–strain model. The appropriate value of this parameter is between 0.85 and 2.5. The simulations were carried out to reproduce the test data of Samples 1-10 and 1-20. It can be seen from Figure 10a,b that CZ has some effects in predicting the last stage of the experiment. The larger the value of CZ is, the faster the end of the curve drops. Figure 10c,d also illustrates this trend. The best fit can be obtained with CZ = 1 for Sample 1-10, and CZ = 0.85 for Sample 1-20. Even if the formulation is the same for the two samples, but the confinement is different, slightly different model parameters are needed. In this case, their models are different in the descending portion. It demonstrates again that, for a given cement type, the model predictions are not very sensitive to the model parameters.

5. Conclusions

A new three-segment stress–strain model is proposed for confined well cement based on experimental data and simulation results. The model consists of a parabolic segment ascending to the peak stress, horizontal segment with constant stress, and linear descending segment. By applying the model in commercial software (ABAQUS), nonlinear characteristics of cement under confining pressure are accurately simulated. The results show good agreement between predictions and measurements obtained from neat cement under confining conditions. In addition, the model is used to simulate the triaxial test data of toughness-enhanced cement. The results show good agreement between simulation results and experimental data, indicating the accuracy and suitability of the proposed model. The new model is simple and can be directly implemented in simulation software used in designing. It requires model parameters that are obtained from uniaxial test data and reasonable estimation of CZ from available triaxial test measurements obtained from the same cement formulation. The sensitivity analysis conducted varying model parameters demonstrates that, for a given cement formulation, the model predictions are not very sensitive to the model parameters.

Author Contributions

Conceptualization, Y.L. and Y.L.; methodology, Y.L.; software, Y.L., R.A. and B.H. ; validation, Y.L. , Y.L. and R.A.; formal analysis, Y.L.; investigation, Y.L., Y.L. and R.A. ; resources, Y.L.; data curation, Y.L. and Y.L. ; writing---original draft preparation, Y.L.; writing---review and editing, R.A.; visualization, Y.L.; supervision, R.A. and B.H.; project administration, B.H. and Y.J.; funding acquisition, Y.J.

Funding

This research received no external funding.

Acknowledgments

The research is supported by the Project of the State Key Laboratory of Petroleum Resources and Prospecting of the China University of Petroleum and the Project of Key Laboratory of the China University of Geosciences and the University of Oklahoma. We thank the China National Petroleum Corporation (CNPC) for providing the test data of Sample 2 presented in this paper. We would also like to show our gratitude to the China Scholarship Council for sponsoring Yan Li for her one-year study at the University of Oklahoma.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. De Souza, W.R.M.; Bouaanani, N.; Martinelli, A.E.; Bezerra, U.T. Numerical simulation of the thermomechanical behavior of cement sheath in wells subjected to steam injection. J. Pet. Sci. Eng. 2018, 167, 664–673. [Google Scholar] [CrossRef]
  2. Gray, K.E.; Podnos, E.; Becker, E. Finite-element studies of near-wellbore region during cementing operations: Part I, SPE-106998. SPE Drill. Complet. 2009, 24, 127. [Google Scholar] [CrossRef]
  3. Wang, W.; Taleghani, A.D. Three-dimensional analysis of cement sheath integrity around Wellbores. J. Petrol. Sci. Eng. 2014, 121, 38–51. [Google Scholar] [CrossRef]
  4. Arjomand, E.; Bennett, T.; Nguyen, G.D. Evaluation of cement sheath integrity subject to enhanced pressure. J. Pet. Sci. Eng. 2018, 170, 1–13. [Google Scholar] [CrossRef]
  5. Andrade, J.; De Torsaeter, M.; Todorovic, J.; Opedal, N.; Stroisz, A.; Vralstad, T. Influence of Casing Centralization on Cement Sheath Integrity during Thermal Cycling. In Proceedings of the IADC/SPE Drilling Conference and Exhibition, Fort Worth, TX, USA, 4–6 March 2014. [Google Scholar]
  6. Thiercelin, M.; Baumgarte, C.; Guillot, D. A soil mechanics approach to predict cement sheath behavior. In Rock Mechanics in Petroleum Engineering; Society of Petroleum Engineers: Trondheim, Norway, 1998. [Google Scholar]
  7. Philippacopoulos, A.; Berndt, M. Mechanical property issues for geothermal well cements. Trans. Geoth. Resour. Counc. 2001, 25, 119–124. [Google Scholar]
  8. Lavrov, A.; Torsæter, M. Physics and Mechanics of Primary Well Cementing. 2016; 9–23. [Google Scholar]
  9. Noori, A.; Shekarchi, M.; Moradian, M.; Moosavi, M. Behavior of Steel Fiber-Reinforced Cementitious Mortar and High-Performance Concrete in Triaxial Loading. ACI Mater. J. 2015, 112, 95–104. [Google Scholar] [CrossRef]
  10. Folino, P.; Xargay, H. Recycled aggregate concrete–Mechanical behavior under uniaxial and triaxial compression. Constr. Build. Mater. 2014, 56, 21–31. [Google Scholar] [CrossRef]
  11. Yang, J.; Cai, X.; Guo, X.; Zhao, J. Effect of Cement Content on the Deformation Properties of Cemented Sand and Gravel Material. Appl. Sci. 2019, 9, 2369. [Google Scholar] [CrossRef]
  12. Lu, W.; Miao, L.; Zhang, J.; Zhang, Y.; Li, J. Characteristics of Deformation and Damping of Cement Treated and Expanded Polystyrene Mixed Lightweight Subgrade Fill under Cyclic Load. Appl. Sci. 2019, 9, 167. [Google Scholar] [CrossRef]
  13. Yuan, Z.; Teodoriu, C.; Schubert, J. Jerome Schubert. Low cycle cement fatigue experimental study and the effect on HPHT well integrity. J. Pet. Sci. Eng. 2013, 105, 84–90. [Google Scholar] [CrossRef]
  14. Li, M.; Yang, Y.; Liu, M.; Guo, X.; Zhou, S. Hybrid effect of calcium carbonate whisker and carbon fiber on the mechanical properties and microstructure of oil well cement. Constr. Build. Mater. 2015, 93, 995–1002. [Google Scholar] [CrossRef]
  15. Popovics, S. Numerical approach to the complete stress–strain relation for concrete. Cem. Concr. Res. 1973, 3, 583–599. [Google Scholar] [CrossRef]
  16. Mander, J.B.; Priestley, M.J.N.; Park, R. Theoretical Stress-Strain Model for Confined Concrete. J. Struct. Eng. 1988, 114, 1804–1826. [Google Scholar] [CrossRef]
  17. Saadatmanesh, H.; Ehsani, M.R.; Li, M.W. Strength and ductility of concrete columns externally reinforced with fiber composite straps. ACI Struct. J. 1994, 10, 434–447. [Google Scholar]
  18. Teng, J.G.; Hu, Y.M.; Yu, T. Stress-strain model for concrete in FRP-confined steel tubular columns. Eng. Struct. 2013, 49, 156–167. [Google Scholar] [CrossRef]
  19. Susantha, K.A.S.; Ge, H.; Usami, T. Uniaxial stress-strain relationship of concrete confined by various shipped steel tubes. Eng. Struct. 2001, 23, 1331–1347. [Google Scholar] [CrossRef]
  20. Noushini, A.; Aslani, F.; Castel, A.; Gilbert, R.I.; Uy, B.; Foster, S. Compressive stress-strain model for low-calcium fly ash-based geopolymer and heat-cured Portland cement concrete. Cem. Concr. Compos. 2016, 73, 136–146. [Google Scholar] [CrossRef]
  21. Haider, G.M.; Sanjayan, J.G.; Ranjith, P.G. Complete triaxial stress-strain curves for geopolymer. Constr. Build. Mater. 2014, 69, 196–202. [Google Scholar] [CrossRef]
  22. Ganesan, N.; Abraham, R.; Deepa Raj, S.; Sasi, D. Stress-strain behaviour of confined Geopolymer concrete. Constr. Build. Mater. 2014, 73, 326–331. [Google Scholar] [CrossRef]
  23. Diaz-Loya, E.I.; Allouche, E.N.; Vaidya, S. Mechanical properties of fly-ash-based geopolymer concrete. ACI Mater. J. 2011, 108, 300–306. [Google Scholar]
  24. Diaz-Loya, E.I.; Allouche, E.; Cahoy, D. Statistical-based approach for predicting the mechanical properties of geopolymer concretes. In Geopolymer Binder Systems; ASTM International: West Conshohocken, PA, USA, 2013; pp. 119–143. [Google Scholar]
  25. Sakino, K.; Nakahara, H.; Morino, S.; Nishiyama, I. Behavior of Centrally Loaded Concrete-Filled Steel-Tube Short Columns. J. Struct. Eng. 2004, 130, 180–188. [Google Scholar] [CrossRef]
  26. Mirmiran, A.; Shahawy, M. Dilation characteristics of confined concrete. Mech. Cohesive Frict. Mater. 1997, 2, 237–249. [Google Scholar] [CrossRef]
  27. Gao, Y.; Karbhari, V.M. Composite Jacketed Concrete under Uniaxial Compression—Verification of Simple Design Equations. J. Mater. Civ. Eng. 1997, 9, 185–193. [Google Scholar]
  28. Xiao, Y.; Wu, H. Compressive behavior of concrete confined by carbon fiber composite jackets. J. Mater. Civ. Eng. ASCE 2000, 12, 139–146. [Google Scholar] [CrossRef]
  29. Harries, K.A.; Kharel, G. Behavior and modeling of concrete subject to variable confining pressure. ACI Mater. J. 2002, 99, 180–189. [Google Scholar]
  30. Samaan, M.; Mirmiran, A. Model of concrete confined by fiber composites. J. Struct. Eng. 1998, 124, 1025–1031. [Google Scholar] [CrossRef]
  31. Lam, L.; Teng, J.G. Design-oriented stress-strain model for FRP confined concrete. Constr. Build. Mater. 2003, 17, 471–489. [Google Scholar] [CrossRef]
  32. Thai, H.T.; Uy, B.; Khan, M.; Tao, Z.; Mashiri, F. Numerical modelling of concrete-filled steel box columns incorporating high strength materials. J. Constr. Steel Res. 2014, 102, 256–265. [Google Scholar] [CrossRef]
  33. Lim, J.C.; Ozbakkaloglu, T. Stress–strain model for normal- and light-weight concretes under uniaxial and triaxial compression. Constr. Build. Mater. 2014, 71, 492–509. [Google Scholar] [CrossRef]
  34. Binici, B. An analytical model for stress–strain behavior of confined concrete. Eng. Struct. 2005, 27, 1040–1051. [Google Scholar] [CrossRef]
  35. Miyauchi, K.; Inoue, S.; Kuroda, T.; Kobayashi, A. Strengthening effects of concrete columns with carbon fiber sheet. Trans. Jpn. Concr. Inst. 1999, 21, 143–150. [Google Scholar]
  36. Hognestad, E. A Study of Combined Bending and Axial Load in Reinforced Concrete Members; Bulletin Series No. 399; University of Illinois at Urbana Champaign: Champaign, IL, USA, 1951. [Google Scholar]
  37. Sheikh, S.A.; Uzumeri, S.M. Analytical Model for Concrete Confinement in Tied Columns. ASCE J. Struct. Div. 1980, 108, 2703–2722. [Google Scholar]
  38. BS 8110. Structural Use of Concrete, Part 1, Code of Practice for Design and Construction; British Standards Institution: London, UK, 1997. [Google Scholar]
  39. EN 1992-1-1. Eurocode 2: Design of Concrete Structures-Part 1: General Rules and Rules for Buildings; European Committee for Standardization: Brussels, Belgium, 2004. [Google Scholar]
  40. Chang, P. Research on the Deformation Ability of Asphalt Base Cement. Master’s Dissertation, Southwest Petroleum University, Chengdu, China, 2015. [Google Scholar]
  41. API RP 10B-2. Recommended Practice for Testing Well Cements; American Petroleum Institute: Washington, DC, USA, 2013. [Google Scholar]
  42. Ulusay, R.; Hudson, J.A. The Complete ISRM Suggested Methods for Rock Characterization, Testing and Monitoring; Springer: Berlin/Heidelberg, Germany, 2007; pp. 1974–2006. [Google Scholar]
  43. SY/T 6466-2016. Testing of set oilwell cement. In Chinese Standard; National Energy Administration: Beijing, China, 2016. [Google Scholar]
  44. ASTM C192 Standard. Standard Practice for Making and Curing Concrete Test Specimens in the Laboratory; American Society for Testing and Materials: West Conshohocken, PA, USA, 2000. [Google Scholar]
  45. ASTM C39/C39M-12. Standard Test Method for Compressive Strength of Cylindrical Concrete Specimens; ASTM International: West Conshohocken, PA, USA, 2012. [Google Scholar]
  46. Kim, J.K.; Seong-Tae, Y. Application of size effect to compressive strength of concrete members. Acad. Proc. Eng. Sci. 2002, 27, 467–484. [Google Scholar] [CrossRef] [Green Version]
  47. Lesti, M.; Tiemeyer, C.; Plank, J. CO2 stability of portland cement based well cementing systems for use on carbon capture & storage (CCS) wells. Cem. Concr. Res. 2013, 45, 45–54. [Google Scholar]
  48. Reddy, B.R.; Santra, A.K.; McMechan, D.E.; Gray, D.W.; Brenneis, C.; Dunn, R. Cement mechanical property measurements under wellbore conditions. In Proceedings of the SPE Annual Technical Conference and Exhibition, Dallas, TX, USA, 9–12 October 2005. [Google Scholar]
  49. Pan, J.L.; Xu, T.; Hu, Z.J. Experimental investigation of load carrying capacity of the slender reinforced concrete columns wrapped with FRP. Constr. Build. Mater. 2007, 21, 1991–1996. [Google Scholar] [CrossRef]
  50. Schneider, S.P. Axially loaded concrete-filled steel tubes. J. Struct. Eng. 1998, 124, 1125–1138. [Google Scholar] [CrossRef]
Figure 1. (a-d) Experimental results used for calibration of the model.
Figure 1. (a-d) Experimental results used for calibration of the model.
Materials 12 02626 g001
Figure 2. Proposed stress–strain model of confined well cement.
Figure 2. Proposed stress–strain model of confined well cement.
Materials 12 02626 g002
Figure 3. Determination of elastic modulus from test measurements (a-e).
Figure 3. Determination of elastic modulus from test measurements (a-e).
Materials 12 02626 g003aMaterials 12 02626 g003b
Figure 4. Determination of σp0 based on uniaxial tests.
Figure 4. Determination of σp0 based on uniaxial tests.
Materials 12 02626 g004
Figure 5. Simulation result and uniaxial test data of Sample 1-0.
Figure 5. Simulation result and uniaxial test data of Sample 1-0.
Materials 12 02626 g005
Figure 6. Simulations and triaxial test data for Sample 1. (a) Sample 1-10; (b) Sample 1-20.
Figure 6. Simulations and triaxial test data for Sample 1. (a) Sample 1-10; (b) Sample 1-20.
Materials 12 02626 g006
Figure 7. Stress–strain models of samples.
Figure 7. Stress–strain models of samples.
Materials 12 02626 g007
Figure 8. Comparison of numerical calculation with experiment results (a-d).
Figure 8. Comparison of numerical calculation with experiment results (a-d).
Materials 12 02626 g008
Figure 9. Effects of strength. (a) Sample 1-10, (b) Sample 1-20, (c) Sample 2, (d) Sample 3.
Figure 9. Effects of strength. (a) Sample 1-10, (b) Sample 1-20, (c) Sample 2, (d) Sample 3.
Materials 12 02626 g009
Figure 10. Effects of CZ. (a) Sample 1-10, (b) Sample 1-20, (c) Sample 2, (d) Sample 3.
Figure 10. Effects of CZ. (a) Sample 1-10, (b) Sample 1-20, (c) Sample 2, (d) Sample 3.
Materials 12 02626 g010
Table 1. Compositions of all materials in the slurry formulation.
Table 1. Compositions of all materials in the slurry formulation.
SampleCompositionsCuringConfinement [MPa]
Sample 1-0Class G cement + water (w/c = 0.44) + silica flour + dispersant + antifoam agent + hydroxyethyl cellulose + latex modified mix1 week at 70 °C and 0.1 MPa0
Sample 1-10Same as Sample 1-01 week at 70 °C and 0.1 MPa10
Sample 1-20Same as Sample 1-01 week at 70 °C and 0.1 MPa20
Sample 2-0AClass G cement + water (w/c = 0.44) + dispersant + stabilizing agent1 week at 80 °C and 0. 1 MPa0
Sample 2-0BSame as Sample 2-0A1 week at 80 °C and 0. 1 MPa0
Sample 2-10ASame as Sample 2-0A1 week at 80 °C and 0. 1 MPa10
Sample 2-10BSame as Sample 2-0A1 week at 80 °C and 0. 1 MPa10
Sample 3-0 [40]Class G cement + water (w/c=0.44) + dispersant + antifoam agent + hydroxyethyl cellulose1 week at 60 °C and 20.7 MPa0
Sample 3-20 [40]Same as Sample 3-01 week at 60 °C and 20.7 MPa20
Table 2. Parameters of the model used in numerical simulation.
Table 2. Parameters of the model used in numerical simulation.
CementE [GPa]Poisson’s Ratioφfc0 [MPa]εc0 [%]CzConfinement [MPa]
Sample 1-1010.20.20.8560.00.8481.0010
Sample 1-2010.20.20.8560.00.8480.8520
Sample 29.650.240.8847.50.5802.2010
Sample 35.850.240.7037.80.7602.2020

Share and Cite

MDPI and ACS Style

Li, Y.; Lu, Y.; Ahmed, R.; Han, B.; Jin, Y. Nonlinear Stress-Strain Model for Confined Well Cement. Materials 2019, 12, 2626. https://doi.org/10.3390/ma12162626

AMA Style

Li Y, Lu Y, Ahmed R, Han B, Jin Y. Nonlinear Stress-Strain Model for Confined Well Cement. Materials. 2019; 12(16):2626. https://doi.org/10.3390/ma12162626

Chicago/Turabian Style

Li, Yan, Yunhu Lu, Ramadan Ahmed, Baoguo Han, and Yan Jin. 2019. "Nonlinear Stress-Strain Model for Confined Well Cement" Materials 12, no. 16: 2626. https://doi.org/10.3390/ma12162626

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop