Home Temperature-porosity-dependent elastic modulus model for metallic materials
Article Open Access

Temperature-porosity-dependent elastic modulus model for metallic materials

  • Haibo Kou EMAIL logo , Yaowen Gao , Jiaxing Shao , Kaiyue Dou and Nan Wang
Published/Copyright: October 20, 2022
Become an author with De Gruyter Brill

Abstract

Elastic modulus plays a key role in the application of porous metallic materials. However, to the best of our knowledge, few attempts have been made to model the simultaneous dependence of elastic modulus on temperature and porosity for metallic materials. The present article contributes to a rational temperature-porosity-dependent elastic modulus model for metallic materials with all parameters having definite physical significance. The model can well predict the elastic moduli of porous metallic materials, from extremely low temperature to ultrahigh temperature, and from dense material to about 0.9 porosity, with reference to an easy-to-access elastic modulus. In a special case, when intrinsic elastic modulus [M] = 2 and critical porosity P C = 1, a phenomenological parameter-free predictive model can be obtained. The model can be applied when the matrix Poisson ratio is 0.1 < v < 0.4 for Young’s modulus and 0.17 < v < 0.27 for shear modulus, which covers most metallic porous materials.

1 Introduction

Porous metal has drawn a great deal of attention in the last decades, and has been extensively used in aerospace, shipping, building, automotive engineering, etc. This is of course owing to their outstanding thermal, acoustic, electromagnetic, and mechanical properties [1,2,3,4]. Elastic modulus is one of the most important mechanical properties and plays a key role in the abovementioned application fields. Therefore, immense literature, both in experiment and theory, is available on their porosity-dependent elastic modulus [5,6]. In addition, due to their advantages in heat insulation, packaging, and energy-absorption, porous metals are also widely used in high-temperature service environments, e.g., evaporative cooling rocket nozzle, steam or gas turbine combustors, and thermal protection system of the exhaust port on aircraft [1,7]. Therefore, investigation of the temperature-dependent elastic modulus of porous metals is of great importance in practical application. And numerous research works in this field can also be found. However, to the best of our knowledge, only occasional attempts have been made to confirm the combined effects of both porosity and temperature on the elastic modulus of metallic materials from the side of experiment, and even less attention was paid from the side of theory. Munro [8] proposed a temperature-porosity-dependent elastic modulus model for ceramic materials:

(1) E ( T , P ) = E 0 ( 1 α T ) ( 1 P ) n

where E(T, P) denotes the elastic modulus at different temperatures and porosities, E 0 denotes the elastic modulus of dense material at ambient temperature, P denotes the porosity, and α and n are the fitting parameters. In this model, the temperature-dependence of elastic modulus is considered by a linear relation. Although the variation in the elastic modulus in certain temperature range above Debye temperature is almost linear [9], the model is not rational at low temperatures, since the slope of elastic modulus vs temperature curve will approach zero as the temperature approaches absolute zero, according to the third law of thermodynamics and experimental results.

Zhang et al. [10] suggested another model for ceramic materials:

(2) E ( T , P ) = ( E 0 B T exp ( T D / T ) ) 1 P P C n

where B denotes the slope of Young’s modulus vs temperature curve above room temperature. T D is a parameter and suggested to correlate with Debye temperature [9]. P C denotes the critical porosity (which can be interpreted as the percolation threshold in percolation theory [11,12]), at which the elastic modulus of porous material reaches zero. And n is a fitting parameter. This model takes a big step forward compared with Munro’s model, which can be applied in wide temperature and porosity range. However, due to its intrinsic defect, this model inherits too many phenomenological fitting parameters.

As can be seen from the above actualities, studies on theoretical modeling of elastic modulus with simultaneous dependence on temperature and porosity are still not enough. The present article attempts to make up for the deficiency. In Section 2, we review the porosity-dependent and temperature-dependent elastic modulus model first, and then contribute to a more rational temperature-porosity-dependent elastic modulus model for metallic materials. In a special case, the model can be deduced to be a phenomenological parameter-free predictive one, and is validated by a systematic comparison between the predicted results and available experimental data in a wide range of temperature and porosity. Furthermore, determination of parameters and application criterion of the model are also discussed.

2 Temperature-porosity-dependent elastic modulus modeling

To deduce a reasonable temperature-porosity-dependent elastic modulus model, we carefully investigated previous studies on this theme. Pabst et al. [6] began his modeling process with an assumption that the effects of temperature and porosity on elastic modulus may be separated, and the validations support his assumption. Moreover, the experimental results by Armstrong [13] and Pabst et al. [14] also support that the relative effects of porosity on elastic modulus are independent of temperature. On the contrary, different statements also exist. Werner et al. [15] experimentally studied the influence of porosity on Young’s modulus of carbon-bonded alumina from room temperature up to 1,450°C, and found that Munro’s decoupling model can be proved valid up to 1,025°C for the composite material. At higher temperatures, the interaction of temperature and porosity should be noted. Following Werner et al. [15], Zhang et al. [10] began their modeling with a conservative consideration that the combined effects of temperature and porosity should be taken into account. Their verification results, however, voted that decoupling of temperature and porosity is suitable. To sum up, decoupling effects of temperature and porosity on elastic modulus may not be suitable for certain materials in high temperature range, but it can be acceptable for the other majority of conditions. Therefore, in this study, we take the following decoupling form for the temperature-porosity-dependent elastic modulus model:

(3) E ( T , P ) = E ( T ) E ( P )

where E(P) and E(T) denote the porosity-dependent and temperature-dependent elastic modulus, respectively. And next we need to find suitable model for E(P) and E(T).

2.1 Porosity-dependent elastic modulus model

The porosity dependence of elastic modulus is an important and old theme. There has been much effort dedicating to the understanding of it since 1960s [15]. A great many theoretical and experimental research works have been conducted, and numerous models have been put forward [5,6,16]. However, this does not mean that this old theme has been fully understood. Generally speaking, existing porosity-dependent elastic modulus models can be roughly classified into linear relation, polynomial relation, power-law relation, and exponential relation, and almost all of them are empirical or semi-empirical relations. Although the linear relation, E(P) = E(P 0)(1 − aP) (where a is a fitting parameter), can obtain a good fit with experiment results at certain porosity ranges [17,18,19], researchers familiar with this field know it for sure that it cannot describe the elastic modulus in a large porosity range, since it is obviously not linear. Other researchers also proposed power-law relation, exponential relation, and other relation forms after observing the variation in elastic modulus with porosity [11,20,21,22,23], and different relation forms can well describe the elastic modulus at different porosity ranges [6,8]. However, strictly speaking, there is not much physical significance left in the abovementioned relations, and the studied elastic moduli are also not zero when P = 100%. In order to satisfy this basic boundary condition and assign physical significance to parameters in the model, Pabst et al. [6] and Pabst and Gregorová [24,25] studied the classic research by Coble and Kingery [26], Mackenzie [27], and Dewey [28], then summarized and improved the expression of Coble-Kingery relation.

(4) E ( P ) = E ( P 0 ) ( 1 [ M ] P + ( [ M ] 1 ) P 2 )

where E(P 0) denotes the elastic modulus of dense material, [M] denotes the intrinsic elastic modulus, [M] = 3(1 − ν)(9 + 5ν)/[2(7 − 5ν)] for Young’s modulus, [M] = 15(1 − ν)/(7 − 5ν) for shear modulus, and [M] = 3(1 − ν)/[2(1 − 2ν)] for bulk modulus, ν denotes the matrix Poisson ratio. In a special case, when [M] = 2 (Roberts and Garboczi [11] and Pabst and Gregorová [24] analyzed this special case, which corresponds to porous material with spherical pores and the matrix Poisson ratio is 0.2), the equation is deduced to a simple power-law relation, E(P) = E(P 0)(1 − P)2.

The more frequently used power-law relation is a semi-empirical equation proposed by Phani and Niyogi [16,29].

(5) E ( P ) = E ( P 0 ) 1 P P C n

This equation is also the one that Zhang et al. [10] chose to consider the porosity-dependence of elastic modulus. The equation shows good agreement with the experimental results over a wide range of porosity and is approved by many researchers [10,11,30]. In addition, the equation allows for the possibility that elastic modulus reaches zero at a porosity lower than 100%, which conforms to most cases of practical interest that the elastic modulus reduces to zero already at such porosity [6,24]. Pabst et al. [6] and Pabst and Gregorová [12] summarized the existing elastic modulus-porosity relations, and retained important features of equations (4) and (5), then proposed a combined relation.

(6) E ( P ) = E ( P 0 ) ( 1 [ M ] P + ( [ M ] 1 ) P 2 ) 1 P / P C 1 P

This relation inherits the advantages of the aforementioned equations. It can fully satisfy the boundary conditions, i.e., E(P) = E(P 0), when P = 0, and E(P) = 0, when P approaches 1. And also allows the possibility of E(P) = 0, when P = P C, inheriting from Phani–Niyogi equation. More importantly, it is also self-consistent and can deduce to many of the classical relations, e.g., Coble and Kingery relation [26], Phani and Niyogi relation [16,29], Hasselman relation [31], and Gibson and Ashby relation [32]. With those advantages, the model can well describe the porosity-dependent elastic modulus of lots of materials. Although this is also a semi-empirical relation, it is probably the most rational porosity-dependent elastic modulus model at ambient temperature for now.

2.2 Temperature-dependent elastic modulus model

The elastic modulus of porous metals with different porosities is extensively studied at ambient temperature. However, their temperature dependence does not gain as much attention as their porosity dependence, although the temperature-dependent elastic modulus is also a basic mechanical parameter for their high-temperature application [14,33]. Few researchers systematically studied the temperature-elastic modulus of porous metals in theory. Fortunately, there are some temperature-elastic modulus models that do not restrict their applications to specific materials. And the decoupling of temperature and porosity is reasonably based on the discussion in Section 2.1. Thus, these models can be used on porous metals.

One of the most frequently used temperature-dependent Young’s modulus models is Wachtman et al.’s relation [34].

(7) E ( T ) = E ( T 0 ) B T exp ( T D / T )

where E(T 0) denotes the elastic modulus at room temperature. This model is also the one that Zhang et al. [10] chose in their model for considering the temperature-dependence of elastic modulus, and shows a good description of experimental data. To compensate for the deficiency that Wachtman’s equation cannot describe more complicated changes in Young’s modulus at a much higher temperature range, Li et al. [35] made a contribution to improving the Wachtman’s equation.

(8) E ( T ) = E ( T 0 ) B T exp ( T m / T ) + B 1 ( T B 2 T m + T B 2 T m ) exp ( T m / T )

where T m denotes the melting point of materials, B 1 and B 2 are the material constants. Good agreement between the calculated results of the model and experimental data of HfB2, HfC0.67, and HfC0.98 has been obtained at ultrahigh temperatures. After that, Zhang et al. [10] improved Wachtman’s equation in another way. Considering that grain boundary sliding will have effects on Young’s modulus at ultrahigh temperatures (Wachtman’s equation ignored this), Zhang et al. also added an additional term based on Wachtman’s equation, B 1(TT S + |TT S|)exp(−T 0/(TT S)), where T S denotes the temperature at which grain boundary sliding begins to have an influence on Young’s modulus [10]. The model also shows good applicability at ultrahigh temperatures. Although the aforementioned works make great contributions to the characterization of temperature dependence of elastic modulus, their defects are also obvious, i.e., all of them are semi-empirical models. In spite of the good phenomenological description of the experimental data, their prediction values are very limited. In addition, determination of the many fitting parameters in the models is another trouble one has to tackle with when using them.

Recently, Li et al. [36] reviewed the temperature-dependent elastic modulus models and derived a new model without phenomenological fitting parameters:

(9) E ( T ) = 1 + 0 T 0 α ( t ) d t 3 1 + 0 T α ( t ) d t 3 × 1 T 0 T C V ( t ) d t T 0 T m C V ( t ) d t + Δ fus H 1 / 2 E ( T 0 )

where α denotes the linear expansion coefficient, C v denotes the heat capacity at constant volume, T 0 denotes an arbitrary temperature, and Δfus H denotes the molar enthalpy of fusion at the melting point. The model is capable of predicting Young’s modulus, elastic constant, and shear modulus at different temperatures, and has been validated by many kinds of dense metals. It can predict the elastic modulus well from extremely low temperature to almost the melting point of metallic materials, which are difficult to obtain through experiments, with reference to an easy-to-access elastic modulus. Compared with various other temperature-dependent elastic modulus models for metallic bulk materials, this model shows many advantages [36,37].

Given the above investigations, in this study, we will take Pabst’s model and Li’s model for E(P) and E(T) in equation (3), respectively, and deduce the following model:

(10) E ( T , P ) = E ( T 0 , P 0 ) 1 + 0 T 0 α ( t ) d t 3 1 + 0 T α ( t ) d t 3 × 1 T 0 T C V ( t ) d t T 0 T m C V ( t ) d t + Δ fus H 1 / 2 × ( 1 [ M ] P + ( [ M ] 1 ) P 2 ) 1 P / P C 1 P

Hence, a new temperature-porosity-dependent elastic modulus model is obtained. And all parameters in the relation have definite physical significance; thus, the model has predictive value on elastic modulus at different temperatures and porosities with reference to an easy-to-access elastic modulus.

3 Validation and discussion

In this part, the newly developed temperature-porosity-dependent elastic modulus model is validated by comparison between the experimental data and calculated results of the model. Unfortunately, as mentioned in Section 1, although the individual effects of temperature and porosity on elastic modulus are extensively studied, their simultaneous effects are rarely reported and immethodical. We made the greatest efforts, however, only the experimental data of elastic modulus of porous nickel (Ni) [38,39] and aluminum (Al) [40] were found, and comparison with them is made in this section.

In calculation, the temperature-dependent molar heat capacity at constant pressure (for solid materials, the difference between molar heat capacity at constant pressure and at constant volume is very small, thus, the molar heat capacity at constant pressure was used in the calculation) above room temperature of Ni and Al was obtained from the literature [41], with a multinomial expression of C P = a + b × 10−3 T + c × 105 T −2 + d × 10−6 T 2. The value of parameters in the expression in different temperature ranges is shown in Table 1. Their molar heat capacity at constant pressure at low temperatures was obtained from literature [42]. The melting point of Ni and Al are 1,726 and 933 K [43], respectively. And the enthalpy of fusion at the melting point of them are 17,472 and 10,719 J·mol−1 [43], respectively. The temperature-dependent linear thermal expansion coefficient of them was obtained from literature [42,44].

Table 1

Parameters in the multinomial expression of temperature dependent molar heat capacity at constant pressure of Ni and Al

Porous metal a b c d Temperature range (K)
Ni 19.083 23.497 298.15–500
−251.166 356.439 259.454 500–631
467.194 −678.737 631–640
−385.698 404.225 654.532 640–700
−10.874 54.668 56.476 −16.489 700–1,400
36.192 1,400–1,726
Al 31.376 −16.393 −3.607 20.753 298.15–933

Parameters used in the calculation, i.e., the reference Young’s modulus or shear modulus, intrinsic elastic modulus [M], and critical porosity P C, are shown in Table 2. As can be seen from Figures 16, the comparisons between experimental and calculated Young’s modulus or shear modulus of open cell porous Ni and open/closed cell porous Al at different porosities and different temperatures, indicate that very good agreement has been obtained. It means the temperature-porosity-dependent elastic modulus model for metallic materials developed in this study has been validated.

Table 2

Parameters used in the calculation

Figure Reference parameter [M] P C
Temperature Porosity Young’s/Shear modulus (GPa)
1 T/T m = 0.1 P = 0.961 0.34 2 1
2 T = 20°C 0 213.12
3 T/T m = 0.1 P = 0.961 0.13
4 T = 330°C 0 86.42
5 T = 21°C P = 0.42 45.78 1
6 T = 21°C P = 0.42 9.67
Figure 1 
               Comparison between experimental and calculated Young’s modulus of open cell porous Ni at different temperatures with porosity P = 0.961, P = 0.969, P = 0.971, and P = 0.973, respectively.
Figure 1

Comparison between experimental and calculated Young’s modulus of open cell porous Ni at different temperatures with porosity P = 0.961, P = 0.969, P = 0.971, and P = 0.973, respectively.

Figure 2 
               Comparison between experimental and calculated Young’s modulus of open cell porous Ni at different porosities with temperatures T = 20°C, T = 330°C, and T = 700°C, respectively.
Figure 2

Comparison between experimental and calculated Young’s modulus of open cell porous Ni at different porosities with temperatures T = 20°C, T = 330°C, and T = 700°C, respectively.

Figure 3 
               Comparison between experimental and calculated shear modulus of open cell porous Ni at different temperatures with porosity P = 0.961, P = 0.969, P = 0.971, and P = 0.973, respectively.
Figure 3

Comparison between experimental and calculated shear modulus of open cell porous Ni at different temperatures with porosity P = 0.961, P = 0.969, P = 0.971, and P = 0.973, respectively.

Figure 4 
               Comparison between experimental and calculated shear modulus of open cell porous Ni at different porosities with temperatures T = 330°C and T = 700°C.
Figure 4

Comparison between experimental and calculated shear modulus of open cell porous Ni at different porosities with temperatures T = 330°C and T = 700°C.

Figure 5 
               Comparison between experimental and calculated Young’s modulus of closed cell porous Al at different temperatures with porosity P = 0.42, P = 0.54, P = 0.65, P = 0.77, and P = 0.88, respectively.
Figure 5

Comparison between experimental and calculated Young’s modulus of closed cell porous Al at different temperatures with porosity P = 0.42, P = 0.54, P = 0.65, P = 0.77, and P = 0.88, respectively.

Figure 6 
               Comparison between experimental and calculated Young’s modulus of open cell porous Al at different temperatures with porosity P = 0.42, P = 0.54, P = 0.65, P = 0.77, and P = 0.88, respectively.
Figure 6

Comparison between experimental and calculated Young’s modulus of open cell porous Al at different temperatures with porosity P = 0.42, P = 0.54, P = 0.65, P = 0.77, and P = 0.88, respectively.

Inheriting advantages from the temperature-dependent elastic modulus model established by Li et al. [36], the model developed in this study can predict elastic moduli at extremely high and low temperatures based on a reference elastic modulus. Since experiments at room temperature are easier to conduct than those at high and low temperatures, the elastic modulus at near room temperature can be chosen as the reference parameter preferentially (reference parameters used in the calculation in this article are shown in Table 2). Moreover, as can be seen in Figures 1 and 3, the model well predicts the elastic moduli of porous metallic materials from extremely low temperature to near melting point of Ni, which also indicates that the model is suitable for a very wide temperature range.

As for porosity dependence, as can be seen in Figures 2 and 4–6, comparisons between experimental and calculated Young’s modulus and shear modulus of Ni and Al at different porosities with different temperatures indicate that good agreement has been obtained. More significantly, as shown in Figures 2 and 4, the good agreement is maintained from dense material to about 0.9 porosity, which is certainly a very wide porosity range. In calculation, the intrinsic elastic modulus [M] and critical porosity P C for Al are both 1, and for Ni, they are chosen as 2 and 1, respectively. It should be noted that the coefficient in the classic Coble-Kingery relation have clear physical significance. Although the first order in the relation is determined only for dilute system of non-interacting spherical pores [24,26], the main purpose of the second order is tackling with the strong non-linear porosity dependence of elastic modulus. Thus, the applicability of the relation has been extended to high porosity range. Moreover, the derivation of the coefficient is based on the premise of isotropy materials. Therefore, this should be adopted as the basic limitation for its application.

In addition, in a special case of [M] = 2 and P C = 1 (when the spatial connectivity of studied porous metal is supposed to be sufficient to sustain an applied stress at super high porosity [8]), i.e., E(P) = E(P 0)(1 − P)2, the same results of the relation and Gibson-Ashby model [32] can be found when fitting a large number of experimental results. And the latter is derived without the assumptions of spherical pore shape or non-interacting system. Therefore, in this case, it is not the spherical pores shape or connectedness of pores, but the approximate isometry is the decisive criterion of application conditions of the relation [6]. In this case, relation (10) has the form of

(11) E ( T , P ) = E ( T 0 ) 1 + 0 T 0 α ( t ) d t 3 1 + 0 T α ( t ) d t 3 × 1 T 0 T C V ( t ) d t T 0 T m C V ( t ) d t + Δ fus H 1 / 2 ( 1 P ) 2

which is a very simple and elegant relation considering the simultaneous dependence of elastic modulus on the variables of temperature T and porosity P. All parameters in the relation have definite physical significance and can be determined easily. Thus, the model has a predictive value on the elastic modulus at different temperatures and porosities, with reference to an easy-to-access elastic modulus. Since the elastic modulus of dense material at room temperature is easier to obtain, it is suggested to be chosen as the reference parameter preferentially. It is worth noting that in the validation of relation (10), Young’s/shear modulus at different temperatures and porosities is predicted with [M] = 2 and P C = 1, as can be seen from Figures 14. Thus, it is actually a validation and classic application of model (11). Furthermore, according to the relationship between intrinsic elastic modulus and the Poisson ratio of matrix metal [6,25], the intrinsic Young’s modulus of 2 or approximately 2 can be obtained when the matrix Poisson ratio is 0.1 < v < 0.4. And the same results for intrinsic shear modulus can be obtained when 0.17 < v < 0.27. Thus, the authors suggest that the application conditions of relation (11) are restricted to 0.1 < v < 0.4 for Young’s modulus and 0.17 < v < 0.27 for shear modulus, which covers most of the metallic materials.

As aforementioned, all parameters in the relation (10) have definite physical significance. However, determination of the intrinsic elastic modulus [M] and critical porosity P C is sometimes a challenge in practice, especially for critical porosity, since it is a posteriori parameter for real materials due to the difficulties in assessing and quantifying the pore size distribution and connectivity. Considering this, some researchers, even including Pabst et al. themselves [6,12], tend to regard critical porosity P C, and sometimes including the intrinsic elastic modulus [M], as fitting parameter. In this way, relation (10) can be quite conveniently and universally used to provide highly effective descriptions of variation in elastic modulus with mean porosity and temperature. The calculated results in Figures 5 and 6 are typical applications of this condition. However, this advantage is obtained at the sacrifice of physical significance and predictive value of the relation. Since the intrinsic elastic modulus is a priori parameter, the authors suggest that one can retain the physical significance of it if Poisson ratio is obtained. However, as aforementioned, derivation of the intrinsic elastic modulus is based on the premise of isotropy materials, one has to pay attention to this application restriction in this condition. This seems to be a contradiction, and future efforts can be made to solve this problem.

4 Conclusion

In this study, the simultaneous and individual dependence of elastic modulus on temperature and porosity are analyzed, and relative theoretical models are reviewed. In this way, a new temperature-porosity-dependent elastic modulus model is developed, and all parameters in the relation have definite physical significance. Thereafter, the model is validated by comparisons between the experimental and calculated Young’s modulus and shear modulus of open cell porous Ni and open/closed cell porous Al. And very good agreements, from extremely low to ultrahigh temperature, from dense material to about 0.9 porosity, have been obtained with reference to an easy-to-access elastic modulus. Since experiments on dense material at room temperature are easier to conduct, the elastic modulus at room temperature of dense material is suggested to be chosen as the reference parameter preferentially. The original model (relation (10)) is restrained to isotropy materials and the critical porosity is difficult to assess. However, when [M] = 2 and P C = 1, a very simple and elegant relation can be obtained. This relation is absolutely a phenomenological parameter-free predictive one. It has the decisive criterion of pores approximate isometry, and 0.1 < v < 0.4 for Young’s modulus and 0.17 < v < 0.27 for shear modulus, which covers most of the metallic porous materials. In addition, if the intrinsic elastic modulus [M] and critical porosity P C are regarded as fitting parameters, the relation can be quite conveniently and universally used to provide highly effective descriptions of variation in the elastic modulus with mean porosity and temperature.

  1. Funding information: This work was supported by the Scientific Research Plan Projects of Shaanxi Education Department (CN) [grant number 20JK0762]; and the National Natural Science Foundation of China [grant numbers 12002268 and 12002288].

  2. Author contributions: H.K.: methodology, investigation, data analysis, and writing-review and editing, and funding acquisition; Y.G.: methodology, investigation, and writing-review and editing; J.S.: investigation, data analysis, and writing-review and editing; K.D.: writing-review and editing; N.W.: writing-review and editing. The authors applied the SDC approach for the sequence of authors.

  3. Conflict of interest: Authors state no conflict of interest.

  4. Informed consent: Informed consent has been obtained from all individuals included in this study.

  5. Data availability statement: The data used to support the findings of this study have not been made available because they also form part of an ongoing study.

References

[1] Aly, M. S. Behavior of closed cell aluminium foams upon compressive testing at elevated temperatures: Experimental results. Materials Letters, Vol. 61, 2007, pp. 3138–3141.10.1016/j.matlet.2006.11.046Search in Google Scholar

[2] Ashby, M. F., A. Evans, N. A. Fleck, L. J. Gibson, J. W. Hutchinson, and H. N. G. Wadley. Metal foams: a design guide, Elsevier Science, Oxford, UK, 2000.10.1115/1.1421119Search in Google Scholar

[3] Banhart, J. and D. Weaire. On the road again: metal foams find favor. Physics Today, Vol. 55, No. 7, 2002, pp. 37–42.10.1063/1.1506749Search in Google Scholar

[4] Movahedi, N., E. Linul, and L. Marsavina. The temperature effect on the compressive behavior of closed-cell aluminum-alloy foams. Journal of Materials Engineering and Performance, Vol. 27, No. 1, 2018, pp. 99–108.10.1007/s11665-017-3098-4Search in Google Scholar

[5] Boccaccini, A. R. and Z. Fan. A new approach for the Young’s modulus-porosity correlation of ceramic materials. Ceramics International, Vol. 23, No. 3, 1997, pp. 239–245.10.1016/S0272-8842(96)00033-8Search in Google Scholar

[6] Pabst, W., E. Gregorová, and G. Tichá. Elasticity of porous ceramics-A critical study of modulus-porosity relations. Journal of the European Ceramic Society, Vol. 26, No. 7, 2006, pp. 1085–1097.10.1016/j.jeurceramsoc.2005.01.041Search in Google Scholar

[7] Li, Z. Mechanical behaviors of closed-cell aluminum foams and sandwich panels at elevated temperatures. Engineering mechanics, University of Science and Technology of China, Hefei, 2013.Search in Google Scholar

[8] Munro, R. G. Analytical representations of elastic moduli data with simultaneous dependence on temperature and porosity. Journal of Research of the National Institute of Standards Technology, Vol. 109, No. 5, 2004, pp. 497–503.10.6028/jres.109.036Search in Google Scholar PubMed PubMed Central

[9] Anderson, O. L. Derivation of wachtman’s equation for the temperature dependence of elastic moduli of oxide compounds. Physical Review, Vol. 144, No. 2, 1966, pp. 553–557.10.1029/SP026p0166Search in Google Scholar

[10] Zhang, X., W. Li, H. Kou, J. Shao, Y. Deng, and J. Ma. Modeling the temperature and porosity dependent Young’s modulus of porous ceramics. Materials Research Express, Vol. 5, No. 1, 2018, id. 015204.10.1088/2053-1591/aaa7cbSearch in Google Scholar

[11] Roberts, A. P. and E. J. Garboczi. Elastic properties of model porous ceramics. Journal of the American Ceramic Society, Vol. 83, No. 12, 2000, pp. 3041–3048.10.1111/j.1151-2916.2000.tb01680.xSearch in Google Scholar

[12] Pabst, W. and E. Gregorová. New relation for the porosity dependence of the effective tensile modulus of brittle materials. Journal of Materials Science, Vol. 39, No. 10, 2004, pp. 3501–3503.10.1023/B:JMSC.0000026961.12735.2aSearch in Google Scholar

[13] Armstrong, P. E. The dependence of Young’s modulus on porosity in a series of isotropic graphites at elevated temperature. Carbon, Vol. 10, No. 3, 1972, pp. 350–350.10.1016/0008-6223(72)90483-6Search in Google Scholar

[14] Pabst, W., E. Gregorová, and M. Černý. Isothermal and adiabatic Young’s moduli of alumina and zirconia ceramics at elevated temperatures. Journal of the European Ceramic Society, Vol. 33, No. 15–16, 2013, pp. 3085–3093.10.1016/j.jeurceramsoc.2013.06.012Search in Google Scholar

[15] Werner, J., C. G. Aneziris, and S. Schafföner. Influence of porosity on Young׳s modulus of carbon-bonded alumina from room temperature up to 1,450°C. Ceramics International, Vol. 40, No. 9, 2014, pp. 14439–14445.10.1016/j.ceramint.2014.07.013Search in Google Scholar

[16] Phani, K. K. and S. K. Niyogi. Young’s modulus of porous brittle solids. Journal of Materials Science, Vol. 22, No. 1, 1987, pp. 257–263.10.1007/BF01160581Search in Google Scholar

[17] Janowski, K. R. and R. C. Rossi. Elastic behaviour of MgO matrix composites. Journal of the American Ceramic Society, Vol. 5011, 1967, pp. 599–603.10.1111/j.1151-2916.1967.tb15007.xSearch in Google Scholar

[18] Warren, W. E. and A. M. Kraynik. The linear elastic properties of open-cell foams. Journal of Applied Mechanics, Vol. 55, No. 2, 1988, pp. 341–346.10.1115/1.3173680Search in Google Scholar

[19] Christensen, R. M. Mechanics of cellular and other low-density materials. International Journal of Solids Structures, Vol. 37, 2000, pp. 93–104.10.1016/S0020-7683(99)00080-3Search in Google Scholar

[20] Spriggs, R. M. Expression for effect of porosity on elastic modulus of polycrystalline refractory materials, particularly aluminum oxide. Journal of the American Ceramic Society, Vol. 4412, 1961, pp. 628–629.10.1111/j.1151-2916.1961.tb11671.xSearch in Google Scholar

[21] Pabst, W. and E. Gregorová. Mooney-type relation for the porosity dependence of the tensile modulus of ceramics. Journal of Materials Science, Vol. 39, 2004, pp. 3213–3215.10.1023/B:JMSC.0000025863.55408.c9Search in Google Scholar

[22] Wang, J. C. Young’s modulus of porous materials. Journal of materials science, Vol. 19, No. 3, 1984, pp. 801–808.10.1007/BF00540451Search in Google Scholar

[23] Wagh, A., R. Poeppel, and J. Singh. Open pore description of mechanical properties of ceramics. Journal of materials science, Vol. 26, No. 14, 1991, pp. 3862–3868.10.1007/BF01184983Search in Google Scholar

[24] Pabst, W. and E. Gregorová. Note on the so-called Coble-Kingery formula for the effective tensile modulus of porous ceramics. Journal of Materials Science Letters, Vol. 22, No. 13, 2003, pp. 959–962.10.1023/A:1024600627210Search in Google Scholar

[25] Pabst, W. and E. Gregorová. Effective elastic properties of alumina–zirconia composite ceramics-Part 2: Micromechanical modeling. Ceramics-Silikáty, Vol. 48, 2004, pp. 14–23.Search in Google Scholar

[26] Coble, R. L. and W. D. Kingery. Effect of porosity on physical properties of sintered alumina. Journal of the American Ceramic Society, Vol. 3911, 1956, pp. 377–385.10.1111/j.1151-2916.1956.tb15608.xSearch in Google Scholar

[27] Mackenzie, J. The elastic constants of a solid containing spherical holes. Proceedings of the Physical Society. Section B, Vol. 63, No. 1, 1950, id. 2.10.1088/0370-1301/63/1/302Search in Google Scholar

[28] Dewey, J. M. The elastic constants of materials loaded with non-rigid fillers. Journal of Applied Physics, Vol. 18, No. 6, 1947, pp. 578–581.10.1063/1.1697691Search in Google Scholar

[29] Phani, K. K. and S. K. Niyogi. Elastic modulus-porosity relation in polycrystalline rare-earth oxides. Journal of the American Ceramic Society, Vol. 70, 1987, pp. 362–366.10.1111/j.1151-2916.1987.tb04920.xSearch in Google Scholar

[30] Ostrowski, T., A. Ziegler, R. K. Bordia, and J. Rödel. Evolution of Young’s modulus, strength, and microstructure during liquid-phase sintering. Journal of the American Ceramic Society, Vol. 81, 1998, pp. 1852–1860.10.1111/j.1151-2916.1998.tb02557.xSearch in Google Scholar

[31] Hasselman, D. On the porosity dependence of the elastic moduli of polycrystalline refractory materials. Journal of the American Ceramic Society, Vol. 45, 1962, pp. 452–453.10.1111/j.1151-2916.1962.tb11191.xSearch in Google Scholar

[32] Gibson, I. and M. F. Ashby. The mechanics of three-dimensional cellular materials. Proceedings of the royal society A. Mathematical, physical, and engineering sciences, Vol. 382, No. 1782, 1982, pp. 43–59.10.1098/rspa.1982.0088Search in Google Scholar

[33] Zhang, X., W. Li, Z. Zhao, Y. He, P. Dong, Y. Ma, et al. A theoretical model for the tensile modulus of polymer/CNT nanocomposites over a wide temperature range. Composites Communications, Vol. 28, 2021, id. 100971.10.1016/j.coco.2021.100971Search in Google Scholar

[34] Wachtman, J. B., W. J. Tefft, D. G. Lam, and C. S. Apstein. Exponential temperature dependence of Young’s modulus for several oxides. Physical Review, Vol. 122, No. 6, 1961, pp. 1754–1759.10.1103/PhysRev.122.1754Search in Google Scholar

[35] Li, W., R. Wang, D. Li, and D. Fang. A model of temperature-dependent Young’s modulus for ultrahigh temperature ceramics. Physics Research International, Vol. 2011, 2011, id. 791545.10.1155/2011/791545Search in Google Scholar

[36] Li, W., H. Kou, X. Zhang, J. Ma, Y. Li, P. Geng, et al. Temperature-dependent elastic modulus model for metallic bulk materials. Mechanics of Materials, Vol. 139, 2019, id. 103194.10.1016/j.mechmat.2019.103194Search in Google Scholar

[37] Li, W., J. Shao, H. Kou, and X. Zhang. Research progress on the theoretical characterization methods for the high-temperature mechanical properties of materials. Chinese Journal of Solid Mechanics, Vol. 38, No. 2, 2017, pp. 93–123.Search in Google Scholar

[38] Zhang, J. Y., P. Zhang, Q. L. Gan, and X. G. Guo. Temperature-dependence of open-cell nickel foams properties. Journal of Materials Science Letters, Vol. 22, No. 23, 2003, pp. 1701–1703.10.1023/B:JMSL.0000004653.15071.a0Search in Google Scholar

[39] Zhang, J. Mechanical properties and breakage mechanism of cellular materials, Xiangtan University, Hunan, 2003.Search in Google Scholar

[40] Hosseini, H. S. M., A. Kharaghani, C. Kirsch, and A. Öchsner. Temperature dependence of elastic properties of aluminum foam Structures, Defect and Diffusion Forum, 2012, pp. 233–237.10.4028/www.scientific.net/DDF.326-328.233Search in Google Scholar

[41] Liang, Y., M. Che, X. Liu, and N. Li. Handbook of thermodynamic data of inorganic compounds, Northeastern University Press, Shenyang, 1994, p. 270.Search in Google Scholar

[42] Gray, D. E. American institute of physics handbook, McGraw-Hill book Campany, New York, Section 4, 1982, pp. 66–67.Search in Google Scholar

[43] Haynes, W. M. CRC handbook of chemistry and physics, CRC Press, New York, Chapter 12, 2013, pp. 217–218.10.1201/b17118Search in Google Scholar

[44] Abdullaev, R. N., Y. M. Kozlovskii, R. A. Khairulin, and S. V. Stankus. Density and thermal expansion of high purity nickel over the temperature range from 150 K to 2,030 K. International Journal of Thermophysics, Vol. 36, No. 4, 2015, pp. 603–619.10.1007/s10765-015-1839-xSearch in Google Scholar

Received: 2022-07-15
Revised: 2022-08-15
Accepted: 2022-08-29
Published Online: 2022-10-20

© 2022 Haibo Kou et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Review Articles
  2. State of the art, challenges, and emerging trends: Geopolymer composite reinforced by dispersed steel fibers
  3. A review on the properties of concrete reinforced with recycled steel fiber from waste tires
  4. Copper ternary oxides as photocathodes for solar-driven CO2 reduction
  5. Properties of fresh and hardened self-compacting concrete incorporating rice husk ash: A review
  6. Basic mechanical and fatigue properties of rubber materials and components for railway vehicles: A literature survey
  7. Research progress on durability of marine concrete under the combined action of Cl erosion, carbonation, and dry–wet cycles
  8. Delivery systems in nanocosmeceuticals
  9. Study on the preparation process and sintering performance of doped nano-silver paste
  10. Analysis of the interactions between nonoxide reinforcements and Al–Si–Cu–Mg matrices
  11. Research Articles
  12. Study on the influence of structural form and parameters on vibration characteristics of typical ship structures
  13. Deterioration characteristics of recycled aggregate concrete subjected to coupling effect with salt and frost
  14. Novel approach to improve shale stability using super-amphiphobic nanoscale materials in water-based drilling fluids and its field application
  15. Research on the low-frequency multiline spectrum vibration control of offshore platforms
  16. Multiple wide band gaps in a convex-like holey phononic crystal strip
  17. Response analysis and optimization of the air spring with epistemic uncertainties
  18. Molecular dynamics of C–S–H production in graphene oxide environment
  19. Residual stress relief mechanisms of 2219 Al–Cu alloy by thermal stress relief method
  20. Characteristics and microstructures of the GFRP waste powder/GGBS-based geopolymer paste and concrete
  21. Development and performance evaluation of a novel environmentally friendly adsorbent for waste water-based drilling fluids
  22. Determination of shear stresses in the measurement area of a modified wood sample
  23. Influence of ettringite on the crack self-repairing of cement-based materials in a hydraulic environment
  24. Multiple load recognition and fatigue assessment on longitudinal stop of railway freight car
  25. Synthesis and characterization of nano-SiO2@octadecylbisimidazoline quaternary ammonium salt used as acidizing corrosion inhibitor
  26. Perforated steel for realizing extraordinary ductility under compression: Testing and finite element modeling
  27. The influence of oiled fiber, freeze-thawing cycle, and sulfate attack on strain hardening cement-based composites
  28. Perforated steel block of realizing large ductility under compression: Parametric study and stress–strain modeling
  29. Study on dynamic viscoelastic constitutive model of nonwater reacted polyurethane grouting materials based on DMA
  30. Mechanical behavior and mechanism investigation on the optimized and novel bio-inspired nonpneumatic composite tires
  31. Effect of cooling rate on the microstructure and thermal expansion properties of Al–Mn–Fe alloy
  32. Research on process optimization and rapid prediction method of thermal vibration stress relief for 2219 aluminum alloy rings
  33. Failure prevention of seafloor composite pipelines using enhanced strain-based design
  34. Deterioration of concrete under the coupling action of freeze–thaw cycles and salt solution erosion
  35. Creep rupture behavior of 2.25Cr1Mo0.25V steel and weld for hydrogenation reactors under different stress levels
  36. Statistical damage constitutive model for the two-component foaming polymer grouting material
  37. Nano-structural and nano-constraint behavior of mortar containing silica aggregates
  38. Influence of recycled clay brick aggregate on the mechanical properties of concrete
  39. Effect of LDH on the dissolution and adsorption behaviors of sulfate in Portland cement early hydration process
  40. Comparison of properties of colorless and transparent polyimide films using various diamine monomers
  41. Study in the parameter influence on underwater acoustic radiation characteristics of cylindrical shells
  42. Experimental study on basic mechanical properties of recycled steel fiber reinforced concrete
  43. Dynamic characteristic analysis of acoustic black hole in typical raft structure
  44. A semi-analytical method for dynamic analysis of a rectangular plate with general boundary conditions based on FSDT
  45. Research on modification of mechanical properties of recycled aggregate concrete by replacing sand with graphite tailings
  46. Dynamic response of Voronoi structures with gradient perpendicular to the impact direction
  47. Deposition mechanisms and characteristics of nano-modified multimodal Cr3C2–NiCr coatings sprayed by HVOF
  48. Effect of excitation type on vibration characteristics of typical ship grillage structure
  49. Study on the nanoscale mechanical properties of graphene oxide–enhanced shear resisting cement
  50. Experimental investigation on static compressive toughness of steel fiber rubber concrete
  51. Study on the stress field concentration at the tip of elliptical cracks
  52. Corrosion resistance of 6061-T6 aluminium alloy and its feasibility of near-surface reinforcements in concrete structure
  53. Effect of the synthesis method on the MnCo2O4 towards the photocatalytic production of H2
  54. Experimental study of the shear strength criterion of rock structural plane based on three-dimensional surface description
  55. Evaluation of wear and corrosion properties of FSWed aluminum alloy plates of AA2020-T4 with heat treatment under different aging periods
  56. Thermal–mechanical coupling deformation difference analysis for the flexspline of a harmonic drive
  57. Frost resistance of fiber-reinforced self-compacting recycled concrete
  58. High-temperature treated TiO2 modified with 3-aminopropyltriethoxysilane as photoactive nanomaterials
  59. Effect of nano Al2O3 particles on the mechanical and wear properties of Al/Al2O3 composites manufactured via ARB
  60. Co3O4 nanoparticles embedded in electrospun carbon nanofibers as free-standing nanocomposite electrodes as highly sensitive enzyme-free glucose biosensors
  61. Effect of freeze–thaw cycles on deformation properties of deep foundation pit supported by pile-anchor in Harbin
  62. Temperature-porosity-dependent elastic modulus model for metallic materials
  63. Effect of diffusion on interfacial properties of polyurethane-modified asphalt–aggregate using molecular dynamic simulation
  64. Experimental study on comprehensive improvement of shear strength and erosion resistance of yellow mud in Qiang Village
  65. A novel method for low-cost and rapid preparation of nanoporous phenolic aerogels and its performance regulation mechanism
  66. In situ bow reduction during sublimation growth of cubic silicon carbide
  67. Adhesion behaviour of 3D printed polyamide–carbon fibre composite filament
  68. An experimental investigation and machine learning-based prediction for seismic performance of steel tubular column filled with recycled aggregate concrete
  69. Effects of rare earth metals on microstructure, mechanical properties, and pitting corrosion of 27% Cr hyper duplex stainless steel
  70. Application research of acoustic black hole in floating raft vibration isolation system
  71. Multi-objective parametric optimization on the EDM machining of hybrid SiCp/Grp/aluminum nanocomposites using Non-dominating Sorting Genetic Algorithm (NSGA-II): Fabrication and microstructural characterizations
  72. Estimating of cutting force and surface roughness in turning of GFRP composites with different orientation angles using artificial neural network
  73. Displacement recovery and energy dissipation of crimped NiTi SMA fibers during cyclic pullout tests
Downloaded on 11.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/rams-2022-0270/html
Scroll to top button