1 Introduction

Water frequently comes into touch with concrete structures. Water may permeate through concrete because it is porous with many macropores, capillary pores, flaws, and cracks. This can have a significant impact on the static/dynamic mechanical properties of concrete because it affects the capillary tension, pore water pressure, inter-particle cohesion, and friction [1,2,3,4]. This results in the viscosity effect and meniscus stress of free water. Studying the impact of free water on the mechanical reaction of concrete is, therefore, crucial for the construction of concrete structures, e.g. hydraulics, marine engineering, bridges, and tunnels. Chemically bonded, physically bonded, and free water are the three forms of water that are generally present in concrete. It is a well-known fact that wet concrete materials perform substantially differently in laboratory trials under quasi-static loading conditions compared to dynamic loading ones [5,6,7,8,9,10]. Additionally, the greater the initial porosity of concrete, the higher the effect of water saturation on the strength [11, 12]. The mechanical properties of concrete with a high water–cement ratio are also more sensitive to water saturation than concrete with a low water–cement ratio. In general, the response of saturated concrete under loading is similar to that of mortars and rocks. The impact of the moisture variation in concrete during loading is typically ignored in practice.

Numerous quasi-static laboratory studies demonstrate that increasing water saturation causes a drop in compressive and tensile strength. In compression, the effect of water saturation is more pronounced than in tension. Concrete's fracture toughness, splitting tensile strength, and compressive strength all show an approximately linearly decreasing trend as water saturation rises [11, 13, 14]. The negative effect of water saturation on compressive strength is similar during both uniaxial and triaxial compression [15,16,17]. However, certain laboratory studies reveal that the strength and fracture toughness exhibit the characteristics of first decreasing and then increasing, and reflect a non-linear or bilinear relationship with the growth of the water saturation [18, 19]. In the case of elastic modulus of concrete, it usually increases [11, 20], does not much change [13], or decreases with a growing saturation ratio [21]. Water content is also one of the key elements influencing the material strain rate effect. Free water may even boost the dynamic compressive and tensile strength of cement-based materials when the strain rate rises to a specific point [11, 22,23,24]. There are many interpretations for the physical mechanism causing the reduced strength of wet concrete in quasi-static tests including e.g. the increase of an internal humidity gradient [11, 25] and the loosening of a molecular system in ITZs [13]. Some authors regarded the growth of pore pressure due to pore closure [26], and the water wedging in cracks [11] as the main reasons. The weakened cohesion and van der Waals forces between microscopic particles [13, 27] were also found to be the reasons for the strength drop. Finally, the effects of capillary suction [2, 19], viscosity [2, 28], viscous friction [28] and temperature-induced expansion [28, 29] were highlighted as the main reasons for the strength reduction. However, no explaining agreement has yet been achieved on this issue. It should be noted that there may be some inconsistency between the experimental results that are now available due to the various ways that dry and wet concrete are prepared [30].

Our research aims are to explain the behavior of wet concrete in compression and tension under quasi-static and dynamic conditions. The current study deals with quasi-static uniaxial compression of wet concrete only. It aims to (a) check the capability of a fully coupled DEM-CFD approach to realistically simulate the quasi-static mechanical behavior of partially saturated concrete specimens at the mesoscale, (b) explain the physical reason for the strength reduction in wet concrete specimens, and (c) demonstrate a quantitative impact of free fluid (water and gas) migration on both the concrete strength and a fracture process. The research examines how the mechanical effects of fluid on dry and wet concrete specimens modify the quasi-static mechanical properties of the material in compression. A fully coupled hydro-mechanical technique based on 2D DEM/CFD was adopted to investigate the solid–fluid interaction [31, 32]. The 3D DEM/CFD model is still in the testing phase, hence the 2D model was used. The model proposed takes into account the two-phase laminar flow of immiscible and compressible fluids. No external pressure is needed to move the fluid; it can be driven by local pressure changes resulting only from changes in pore volumes. Owing to this, the proposed model can realistically reproduce the interaction between flowing fluid and solids during compression tests.

The proposed improved DEM/CFD model employs a direct numerical simulation method. DEM was used to represent the mechanical behavior of concrete specimens by using discrete spherical elements interacting locally through normal and tangential contacts that generate cracks during their failure. There are several different discrete models for concrete mechanical behavior developed in the literature, such as classical lattice models, rigid-body-spring models, lattice discrete particle models and discrete element methods [33]. We chose DEM since it realistically captures the fracture processes in concrete [34,35,36,37,38]. Additionally, it correctly depicts mesostructure and contact forces, using simple dynamic equations. Using CFD, a flow network made up of predefined channels was used to represent the laminar viscous two-phase fluid flow (water and gas) through pores and cracks in a continuous domain between the discrete elements. Calculations in isothermal conditions were performed in this research stage on 2D small-size bonded granular specimens of a simplistic mesostructure resembling concrete under uniaxial compression. Concrete was treated as a one-phase material only, composed of bonded spheres with different diameters without distinguishing aggregate, mortar, ITZs, and macro-pores [35, 37]. Additionally, there was a relatively low number of spheres and a narrow range of particle diameter sizes. Therefore, the mechanical concrete behavior (stress–strain curve and fracture process) in laboratory tests was approximately reproduced. The main attention was focused in the paper on the effect of pore fluid migration and pressure on the specimen’s compressive strength. In a series of DEM/CFD simulations, the different fluid saturation and fluid viscosity were the variables that were explored.

The capillary pressure, defined as the difference between the partial pressures of two phases: the wetting phase (water) and non-wetting phase (water vapor or moist air), was neglected in the current DEM analyses for the sake of simplicity. According to this definition, when the concrete is fully saturated, there is no capillary flow. In partially saturated concrete, capillary flow can occur even when there is no external fluid source (due to the flow of viscous fluid during compression). The flow of the liquid phase can fill in some capillaries with water to the point where the snap-off mechanism [39, 40] is activated. Even if capillaries are not filled in with water, this mechanism immediately fills them with water. When the capillary is filled in with water, the typical capillary flow may begin (a piston-like mechanism). The capillary pressure can reach even 1.5 MPa (and more) and strongly depends upon the initial saturation and size and number of capillaries. However, its impact is solely local. Capillary pressure and capillary flow mechanisms were implemented in the proposed DEM-CFD model [41]. The influence of capillary pressure on the compressive strength of wet concrete will be investigated in the next phase of the research.

Several coupled hydro-mechanical DEM-CFD models have been suggested in the literature (e.g. [42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57]). The review of those DEM-based coupled models was given in [31]. As compared to those models, the DEM/CFD-based mesoscopic method for fluid flow in partially fluid-saturated porous cohesive-frictional materials with very low porosity that is presented in the current study has several advantages such as.

  1. (1)

    The direct numerical simulation approach is used to solve fluid flow equations in a two-dimensional continuous fluid domain between discrete elements.

  2. (2)

    The fluid and discrete domains of solids reside in one physical system. Both domains are discretized into one triangular mesh.

  3. (3)

    Due to the triangular mesh in the fluid domain, the variable geometry, size, position, and volume of pores/cracks are considered to correctly track the liquid/gas content. Hence, the effect of material skeleton deformation on the fluid pressure distribution can be studied.

  4. (4)

    An effective method is developed to automatically mesh and re-mesh these domains to account for changes in the geometry and topology of particle and fluid domains.

  5. (5)

    Coarse meshes of solid and liquid domains are utilized to build a virtual fluid flow network (VPN).

  6. (6)

    The two-phase immiscible fluid contains both a liquid (water) and a gas (water vapor or moist air).

There also exist hydro-mechanical models, based on lattice and lattice discrete particle models [58,59,60,61,62,63]. The models are simplified since the pore fluid pressure does not depend on volume changes which are ignored. The fluid flow is solely caused by external fluid pressure or relative humidity difference. However, during compression, the fluid flow is mainly driven by changes in pore pressures, which are affected by changes in pore volumes.

The impact of free water content on the response of concrete under various stress levels and static/dynamic loading situations was investigated using several numerical coupled models. Most of these simulations were done within continuum mechanics (e.g. [64, 65]). DEM solutions also exist for this problem (e.g. [66]); they are approximations since they do not take into consideration the true coupling between solid (DEM) and fluid mechanics (CFD).

This article is a follow-up to the authors' prior mesoscopic numerical investigations, which employed a fully coupled DEM-CFD model to numerically analyze hydraulic fracturing in rocks [31, 32] and hydraulic/capillary flow in unsaturated mortars and concretes [41]. With thorough 3D CFD simulations using the Reynolds-averaged Navier–Stokes equations in the continuous domain between particles, the coupled DEM/CFD technique was validated [67, 68]. It was demonstrated that the turbulent kinetic energy and turbulent dissipation energy are relatively low and may be disregarded in the simplified fluid flow model. Heat transfer has recently been added to the authors’ coupled DEM/CFD model [69, 70].

The structure of the current paper is as follows. After the introduction in Sect. 1, Sect. 2 provides a mathematical model of the DEM/CFD-based coupled hydro-mechanical technique. Section 3 describes the input data for DEM-CFD simulations. Pure DEM simulation results are gathered in Sect. 4. Section 5 displays several numerical simulation results with DEM-CFD on the behavior of fully and partially saturated concrete under uniaxial compression. Section 6 contains some concluding remarks. The YADE open-source software program [71, 72] was enhanced by the authors to include the fully coupled DEM-CFD approach for partially saturated bonded granular materials.

2 Two-dimensional DEM/CFD-based model

In [31] and [32], the coupled DEM-CFD model was thoroughly explained. For the sake of clarity, Sect. 2 only contains the model's most crucial details. The direct numerical simulation (DNS) approach is used to solve fluid flow equations in a two-dimensional continuous fluid domain between discrete elements (unlike existing fluid flow models).

2.1 DEM for cohesive-frictional materials

DEM simulations are carried out using the 3D explicit solver of discrete element open-source software YADE [71, 72]. With the use of Newton's second law of motion and an explicit time-stepping method, particles in a DEM interact with one another during translational and rotational motions [73]. At the grain contact, the model suggests a cohesive bond with a brittle failure below the critical normal tensile force. Under usual compression, shear cohesion failure causes contact slip and sliding, which are controlled by the Coulomb friction law. The basic governing equations of DEM are presented in Appendix A (Eqs. A1A7) [34,35,36,37,38]. Non-viscous damping is chosen [74] (Eq. A7) in simulations to speed up convergence. In DEM, an arbitrary micro-porosity may be attained due to the possibility of particle overlap. The following material constants are required for DEM simulations: Ec, υc, μc, C, and T (Appendix A). Additionally necessary are the parameters R (sphere’s radius), ρ (mass density), and αd (damping factor). The damping factor is consistently set to αd = 0.08 [35]. The particle contact ratio C/T must be carefully taken into account to properly simulate the distribution of shear and tensile cracks, the relationship between the uniaxial compressive and tensile strength, and the failure mode of specimens (brittle or quasi-brittle) [75].

The process of running several DEM simulations and comparing the results to experimental data from simple tests, such as uniaxial compression, triaxial compression, and simple shear, is typically used to determine the material constants. The DEM model does not account for material softening. The model has been successfully employed by the authors in the modeling of engineering particulate materials, primarily granular materials [76,77,78,79], concretes [34,35,36,37,38, 80, 81], and rocks [32, 75]. The grain damage was not considered in current DEM simulations. This process may be easily taken into account in DEM simulations (by using collections of particles to represent the grains allowing for intra-granular fracture), although their time will be much increased. A shortcoming of DEM is the huge time of computation in industrial-scale problems.

2.2 Fluid flow model

A fundamental idea for streamlining the fluid flow model was put forth in [31] and [32]. The 2D DEM-CFD model developed by the authors is based on the idea that two domains coexist in a physical system: a discrete 3D domain (solid) and a continuous 2D domain (fluid) (Fig. 1). This is in contrast to the basic concept of usual fluid flow networks. It was assumed that the discrete domain was originally 3D due to the need to use a 3D DEM solver in YADE software. Consequently, the initially fluid domain is two-dimensional, while the solid domain consists of a single layer of discrete 3D elements (spheres). The centers of gravity of discrete elements are positioned on a plane (2D surface). To transform discrete 3D elements into a 2D problem, the discrete elements are then projected onto a plane to create circles, which are then discretized into a 2D triangular mesh along with voids [31] (Fig. 1). Finally, after projection and discretization, there are two 2D domains: the fluid domain and the solid domain. Note that the displacement of the spheres in the OZ direction (perpendicular to the plane) is fixed. Therefore, although the 3D DEM solver in YADE is used, the mechanical problem is quasi 2D as in the fluid flow model. In the CFD model, all definitions regarding the geometry of both domains (solid and fluid) are treated as two-dimensional, and the third dimension has a unit size.

Fig. 1
figure 1

Two domains co-existing in one physical system: a domains before projection and discretization and b solid and fluid domains after discrete elements projection and discretization (fluid domain is in red and solid domain is in grey) [31] (colour figure online)

In the 2D fluid domain, the centroids of the triangles are connected by channels and form a fluid flow network. To get a more precise distribution of pressures, fluid-phase fractions, and densities, a remeshing technique discretizes the overlapping circles, sets the contact lines, and eliminates the overlapping areas [31, 32]. The contact forces are estimated based on the pressure and shear stress for the specimen thickness equal to the mean particle diameter. The basic equations of the fluid flow model are given in Appendix B (Eqs. B1B28). The mass change in triangular cells is correlated with the density change in a fluid phase, which results in changes in pressure. Because of this, triangles do not obey the equation for the conservation of momentum, but their mass is nevertheless conserved throughout their whole volume. This procedure is carried out using an explicit formulation for every VP in the fluid flow network, called a Virtual Pore Network (VPN). Initially, the liquid and gas might be present in the material and pre-existing discontinuities. Two distinct channels are introduced by VPN [31]:

  1. (A)

    Artificial ‘S2S’ channels are those that connect discrete concrete elements that are in contact with one another and

  2. (B)

    Actual ‘T2T’ channels are those that link grid triangles in pores that are in contact with one another along a common edge.

The channel length is thought to be the distance between the gravity centers of adjacent grid triangles. The mass flow rate in channels is calculated by solving continuity and momentum equations for laminar flows of an incompressible fluid. The fluid flow in the network only serves to estimate the fluid's mass flow rate as it passes through the triangle cell's face (edge).

Three flow regimes are identified in the channels [31]: (a) single gas phase flow with a gas phase fraction, (b) single liquid phase flow with a liquid phase fraction, and (c) two-phase flow (liquid and gas). In the flow regime (a) and (b) (single phase fluid flow) the Poiseuille equation [82, 83] is used to compute the mass flow rate. The two-phase fluid flow driven by a pressure gradient in adjacent VPs exhibits similar behavior to a two-phase fluid flow of two immiscible and incompressible fluids in channels (Fig. 2).

Fig. 2
figure 2

Two-layer fluid flow in channels ‘S2S’ and ‘T2T’ (h—channel aperture, L—channel length, ‘q’—liquid and ‘p’—gas) [32]

The gas phase fraction and the liquid phase fraction are denoted by the symbols \({\alpha }_{p}\) and αq, respectively, and they sum to 1. They are local parameters of the single cell in the mesh. When defining the initial conditions, the parameters can be the same for each grid cell. The liquid–gas interface is parallel to the channel plates and is calculated as the average of the liquid phase fraction in adjacent cells. However, the fraction of liquid and gas phases changes over time, as well as the position of the phase interface in the channel. Gravity's effects are disregarded. While the volumetric flow rates of the fluid phases are unknown, the interface position is connected to fractions of the fluid phases in adjacent VPs. Equations of continuity and momentum describe the flow in each phase.

VPs, unlike the channel flow model, assume that the fluid is compressible, which means that the density of the fluid phases can change in space and time. The fluid pressure can reach 70 MPa in specific situations, such as during the hydraulic fracturing process. The gas phase exceeds the critical point and becomes a supercritical fluid under these conditions. Therefore, for both fluid phases in VPs, the Peng-Robinson equation of state [69] is used to describe fluid behavior. Assuming that both fluid phases have the same pressure (as in the Euler model), the pressure equation is solved to determine the density of the fluid phase. The mass conservation equation for both phases is applied to compute the density of the liquid/gas phase. Finally, the cell pressure (VP) is calculated from the Peng-Robinson equation of state. The basic equations of the fluid flow model in pores (VP) are given in Appendix B (Eqs. B10B28).

Four major phases make up the numerical algorithm:

  1. (a)

    by resolving momentum and continuity equations, one can calculate the mass flow rate for each phase of fluid flowing through the cell faces (in channels encircling VP),

  2. (b)

    using equations of state and continuity to calculate the phase fractions and their densities in VPs,

  3. (c)

    utilizing the equation of state to determine the pressure in VPs,

  4. (d)

    updating material properties (e.g. liquid and gas densities in each cell grid).

3 Input data for 2D DEM/CFD simulations

Typically, mechanical (DEM) and permeability and sorptivity tests (CFD) are used to calibrate DEM and CFD separately [31, 32, 41]. For the numerical calculations in the first calculation stage, a simple one-phase model for concrete made out of spheres with varying diameters was employed to roughly depict quasi-static concrete's performance in uniaxial compression as compared to laboratory tests. The diameters of spheres d were arbitrarily set to range from 1.0 to 1.8 mm, with a mean diameter of d50 = 1.4 mm. For uniaxial compression tests with dry and wet specimens (containing about 1600 spheres), a quadratic specimen of 50 × 50 mm2 was assumed (see Fig. 4, Sect. 4). The specimen thickness was always equal to the mean grain diameter. Horizontal boundaries at the top and bottom were frictionless and smooth. Vertical boundaries were free to move. One grain at the bottom midpoint was fixed. A continuous lowering of the top specimen boundary caused deformation in the specimen. There were no pre-existing macro-pores or cracks in the specimen. The initial micro-porosity was p = 5% (being equivalent to typical concrete [37, 81]). The stress–strain curve and the deformed specimen are presented in Figs. 3 and 4 of Sect. 4. The estimated compressive strength (36 MPa), the corresponding vertical normal strain (0.23%), the elastic modulus (20 GPa), and the crack pattern are in agreement with the experimental ones for usual concrete [84]. The more realistic response of concrete due to uniaxial compression, described as a four-phase material at the mesoscale, was shown in [37].

Fig. 3
figure 3

Pure DEM simulation results for one-phase concrete specimen during uniaxial compression-relationship between vertical normal stress σy and vertical normal strain εy with marked characteristic points 1–6

Fig. 4
figure 4

Pure DEM simulation results for one-phase concrete specimen during uniaxial compression: evolution of fracture pattern a εy = 0.12%, b εy = 0.15%, c εy = 0.18%, d εy = 0.19%, e at peak stress (εy = 0.235%), and f slightly after peak stress (blue particles have constant vertical velocity and green particles are blocked in vertical direction, displacements are magnified by factor 30) (colour figure online)

For permeability simulations with pure CFD, a quadratic non-deformable specimen of 10 × 10 mm2 was chosen as in [41]. The macroscopic permeability coefficient κ of the concrete specimen, calculated with Darcy's law, was 4.0 × 10−16 m2 for the values of ℎinf, ℎ0, γ, and β (Eqs. B1 and B2 in Appendix B) given in Table 1 [41]. This value is in agreement with experimental results for concretes and mortars [41]. The larger values of ℎinf, ℎ0, γ, and β caused the permeability coefficient κ higher. An arbitrary permeability coefficient may be, thus, assumed in numerical calculations.

Table 1 Basic material constants assumed for concrete, fluid and gas in DEM/CFD calculations

The basic material constants assumed for concrete specimens in the coupled DEM-CFD calculations are given in Table 1. The adaptive time step in DEM and CFD was always defined [31]. The maximum time step was limited to 1·10−8 s. The computation time of one simulation was about 14 h on a computer with two Intel Xeon Platinum processors 8280 (2.70 GHz). The computational cost of the simulation was relatively high because the existing DEM-CFD model was parallelized on threads only but not in a distributed mode (on cluster computer nodes).

4 Pure DEM simulation results

The specimen 50 × 50 mm2 was assumed for mechanical computations (Sect. 3). In Figs. 3, 4, 5, 6 and 7, DEM's numerical findings for concrete defined as a one-phase material, composed of spheres with diameters of 1.0–1.8 mm (without distinguishing aggregate, mortar, ITZs, and macro-pores [36,37,38]), are displayed in uniaxial compression. Figure 3 illustrates the stress–strain diagram. On the diagram, some of the characteristic points ‘1’– ‘6’ were marked. Figures 4 and 5 show the progression of the fracture patterns and broken particle contacts (marked by lines tangential to the contact surfaces) for the various vertical normal strains (points 1–6). The development of the relative broken contact number is depicted in Fig. 6. Figure 7 shows the distribution of the compressive and tensile contact forces at the maximum vertical normal stress (point ‘5’ in Fig. 3).

Fig. 5
figure 5

Pure DEM simulation results for one-phase concrete specimen during uniaxial compression: distribution of broken contacts (marked by lines tangential to contact surfaces) in non-deformed specimen for different vertical normal strain εy a εy = 0.12%, b εy = 0.15%, c εy = 0.18%, d εy = 0.19%, e εy = 0.235% (at peak stress), and f slightly after peak stress (broken contacts are marked in red) (colour figure online)

Fig. 6
figure 6

Pure DEM simulation results for one-phase concrete specimen during uniaxial compression: evolution of relative number N of broken contacts in (%)

Fig. 7
figure 7

Pure DEM simulation results for one-phase concrete specimen during uniaxial compression: distribution of contact forces at peak stress (εy = 0.235%) a compressive forces and b tensile forces (force thickness denotes force magnitude)

For a vertical normal strain of εy = 0.235%, the estimated compressive strength was fc = 36 MPa and the elastic modulus was E = 20 GPa (Fig. 3). The elastic range was up to εy = 0.11%. Due to several simplifications assumed in the calculations to accelerate the simulations (e.g. one-phase material composed of spheres, narrow particle diameter size range, a small number of spheres, and 2D conditions), the stress–strain curve was solely approximately reproduced. The calculated specimen's pre-peak response was too linear and its post-peak response was too brittle (Fig. 3). Without those 3 first simplifications, the concrete behavior (stiffness, strength, brittleness, and fracture) can be realistically reproduced in simulations of uniaxial compression [37, 38]. Jumps on the stress–strain curve in Fig. 3 were caused by crack formation. A few almost vertical cracks that appeared during uniaxial compression were the failure mode (Figs. 4 and 5). Long before the peak stress, at around εy = 0.0582%, the first micro-cracks occurred in the specimen (Fig. 6). More than 25% of contacts were already damaged at the maximum vertical stress (εy = 0.235%). The rate of contact damage progressively rose during compression. After the peak stress for εy > 0.235%, it reached its maximum rate (Fig. 6). A network of vertical normal contact forces conveyed the external vertical load (Fig. 7a). Too weak tensile contact forces led to specimen damage (Fig. 7b).

5 DEM/CFD simulation results

The same concrete specimen 50 × 50 mm2 was assumed for hydro-mechanical computations. The initial fluid (water and gas) pressure in the specimen was Po = 0.1 MPa. The impact of various initial fluid volume fractions αq (αq = 0.3, 0.7, and 1.0) on the stress–strain curve during uniaxial compression is shown in Fig. 8.

Fig. 8
figure 8

Pure DEM and coupled DEM-CFD simulation results for one-phase concrete specimen during uniaxial compression with different initial fluid volume fraction αq: relationship between vertical normal stress σy and vertical normal strain εy: a pure DEM (Fig. 3), b DEM-CFD with αq = 0.3, c DEM-CFD with αq = 0.7 and d DEM-CFD with αq = 1.0

The numerical results demonstrate that when the initial fluid volume fraction αq increased, the compressive strength fc and the related vertical normal strain εy decreased. The reduction of fc with increasing fluid volume fraction αq was approximately linear, agreeing with experiments [11, 13, 14]. The initial material stiffness remained the same up to εy = 0.1% due to low pore fluid pressures (see Fig. 18a) caused by low specimen porosity [30] and the lack of cracks. This numerical outcome is in agreement with the experiment in [13] but contrasts with some laboratory tests wherein the elastic modulus rose [11, 20] or declined [21] with a higher saturation degree. In comparison to the results for a dry specimen (αq = 0) (curve 'a' in Fig. 8), the vertical normal strain was 20% lower for a fully saturated specimen (αq = 1.0) and the compressive strength was 15% lower. Due to the initial porosity's low value of 5%, those reductions were relatively small for the wet specimen [11, 30]. For a partially saturated specimen, the brittleness was slightly higher. The curves σy = f(εy) had a similar shape regardless of αq.

5.1 Mechanical results for wet specimens

For the fully fluid-saturated specimen (αq = 1.0), the DEM-CFD findings are given in Figs. 9, 10, 11 like the pure DEM results (Figs. 4, 5, 6). The developments of the fracture patterns and broken contacts are illustrated in Figs. 9 and 10. Figure 11 shows how the damaged contact number changed over specimen deformation.

Fig. 9
figure 9

DEM-CFD simulation results for one-phase concrete specimen (αq = 1.0) during uniaxial compression: evolution of fracture pattern a εy = 0.12%, b εy = 0.15%, c εy = 0.18%, d at peak stress (εy = 0.19%) and e slightly after peak stress (displacements are magnified by factor 30)

Fig. 10
figure 10

DEM-CFD simulation results for one-phase concrete specimen (αq = 1.0) during uniaxial compression: distribution of broken contacts (marked by lines tangential to contact surfaces) in non-deformed specimen for different vertical normal strain εy a εy = 0.12%, b εy = 0.15%, c εy = 0.18%, d at peak stress (εy = 0.19%) and e slightly after peak stress (broken contacts are marked in red) (colour figure online)

Fig. 11
figure 11

Pure DEM (curve ‘a’) and coupled DEM-CFD simulation result (curve ‘b’) for one-phase concrete specimen (with αq = 1.0) during uniaxial compression: A evolution of relative number N of broken contacts with vertical normal strain εy and B zoom on onset of contact breakage

The wet specimen's behavior mirrored that of the dry specimen (Figs. 4 and 5). However, the wet specimen was slightly more damaged for εy = 0.19% and more and more damaged after the peak stress (Figs. 9, 10, 11) than the dry specimen. The first microcracks appeared a little bit earlier in the wet specimen as compared to the dry specimen (εy = 0.0578% versus εy = 0.0582%) (Fig. 11).

Figures 12 and 13 present some results for the partially fluid-saturated specimen (αq = 0.7): the development of the fracture patterns (Fig. 12) and the evolution of broken contacts (Fig. 13). The wet specimen's behavior with αq = 0.7 and αq = 1.0 was similar, however, the specimen with αq = 0.7 was less damaged than this with αq = 1.0.

Fig. 12
figure 12

DEM-CFD simulation results for one-phase partially fluid-saturated concrete specimen (αq = 0.7) during uniaxial compression: evolution of fracture pattern a εy = 0.12%, b εy = 0.15%, c εy = 0.18%, d εy = 0.19%, e at peak stress (εy = 0.215%) and f slightly after peak stress (displacements are magnified by factor 30)

Fig. 13
figure 13

Pure DEM (curve’a’) and DEM-CFD simulation result for one-phase partially fluid-saturated concrete specimen with αq = 0.7 (curve ‘b’) during uniaxial compression: A evolution of relative number N of broken contacts with vertical normal strain εy and B zoom on onset of contact breakage

5.2 Fluid flow results for fully saturated specimen

The distributions of the high (> 0.1 MPa—initial pore fluid pressure) and low pore fluid pressure (≤ 0.1 MPa—initial pore fluid pressure) in the fully saturated specimen (αq = 1.0) are shown in Figs. 14, 15, 16 for a vertical normal strain εy = 0.05% (Fig. 14), εy = 0.10% (Fig. 15) and εy = 0.19% (peak stress) (Fig. 16). Figure 17 shows the distribution of pore fluid velocities in the specimen for a vertical normal strain εy = 0.5% (Fig. 17a), εy = 0.10% (Fig. 17b) and εy = 0.19% (peak stress) (Fig. 17c). The evolutions of the average and maximum pore fluid pressures during uniaxial compression are depicted in Fig. 18. Figure 19 depicts the evolution of fluid pressure in one arbitrary pore.

Fig. 14
figure 14

DEM-CFD results for fully fluid-saturated specimen (αq = 1.0): distribution of high a and low b pore fluid pressures in concrete specimen for vertical strain εy = 0.05%

Fig. 15
figure 15

DEM-CFD results for fully fluid-saturated specimen (αq = 1.0): distribution of high a and low b pore fluid pressures in concrete specimen for vertical strain εy = 0.10%

Fig. 16
figure 16

DEM-CFD results for fully fluid-saturated specimen (αq = 1.0): distribution of high a and low b pore fluid pressures in concrete specimen for vertical strain εy = 0.19% (peak stress)

Fig. 17
figure 17

DEM-CFD results for fully fluid-saturated specimen (αq = 1.0): distribution of fluid velocity vectors specimen for vertical normal strain εy: a εy = 0.05%, b εy = 0.10% and c εy = 0.19% (peak stress)

Fig. 18
figure 18

DEM-CFD results for fully fluid-saturated specimen (with αq = 1.0): evolution of average a and maximum pore fluid pressure b against vertical normal strain εy during uniaxial compression

Fig. 19
figure 19

DEM-CFD results for fully fluid-saturated specimen (αq = 1.0): evolution of pore fluid pressure in vicinity of maximum fluid pressure (point ‘P1’) against vertical normal strain εy: a location of measurement points ‘1’ and ‘2’, and b evolution of fluid pressure in point ‘P2’ in vicinity of point ‘P1’

The specimen's pore fluid pressure distribution in the pores was noticeably non-uniform (Figs. 14, 15, 16). There were large areas with high fluid pore pressure (> 0.1 MPa—initial pore fluid pressure) and a few with low fluid pore pressure (≤ 0.1 MPa—initial pore fluid pressure). Different velocities of fluid flow were observed in the specimen (Fig. 17). The maximum pore fluid velocity was found to be 0.012 m/s (Fig. 17) and its location matched the location of the maximum fluid pressure.

The average fluid pore pressure initially increased due to the specimen compaction induced by a compression process and then reached an asymptote (Fig. 18a). Its maximum value was about 0.12 MPa. However, the maximum fluid pore pressure initially increased, then decreased and reached an asymptote (Fig. 18b). There were two small fluid pressure peaks around the vertical strain of 0.13% (Fig. 18b). They were caused by a momentary increase in sphere overlap, which resulted in a transient decrease in pore volume and an increase in fluid pressure. The maximum pore fluid pressure was locally about 0.8 MPa for εy = 0.05%. Later, as microcracks started to appear in the specimen, the maximum pore fluid pressure lowered (down to 0.2 MPa). The biggest decline in pore fluid maximum pressure was seen in the region of εy = 0.05–0.085% (Fig. 18b). The fluid pressures were greater than the initial one equal to 0.1 MPa in the entire εy-range. Some suction pressures were noticeable in the specimen at some places (Figs. 14 and 19). The expansion of the crack intensity by the pore fluid pressure’s increase over the initial one reduced compressive strength for wet specimens.

The increase of fluid pressure to the maximum value (during the entire simulation) was observed only in one pore (point ‘P1’ in Fig. 19a). Moving away from this point (e.g. point ‘P2’ in Fig. 19a), the maximum fluid pressure decreased and the changes in pressure over time smoothed out. This phenomenon was caused by a random arrangement of discrete elements. The movement in the uniaxial compression test caused discrete elements located along the upper edge of the specimen to move in different directions. By chance, the two spheres began to overlap. Consequently, the volume of pores located between these spheres began to decrease relatively quickly increasing fluid pressure. This effect was local, and the pressure of the fluid surrounding the pore decreased very quickly (Fig. 19b) at the same time step. The maximum fluid pressure at point ‘P1’ occurred slightly earlier (for vertical normal stress εy = 0.0564%) than at point ‘P2’ (for vertical normal stress εy = 0.066%). While the increase in fluid pressure at point ‘P1’ was caused by the deformation of the material skeleton, the increase in fluid pressure at point ‘P2’ was mainly caused by the flow of fluid from point ‘P1’ (higher pressure) to point ‘P2’ (lower pressure). The fluid flow took some time to increase the fluid pressure at P2 to its maximum value.

The material response was similar in other DEM-CFD results with a partially saturated specimen for the two various initial fluid volume fractions αq (αq = 0.3 and 0.7). The average fluid pressure was around 0.102 MPa (αq = 0.7) (Fig. 20) and 0.1001 MPa (αq = 0.3). The maximum local fluid pressure was around 0.4 MPa (αq = 0.7) and 0.15 MPa (αq = 0.15). The compressive strength increased as αq declined. The flow velocity also diminished with decreasing αq.

Fig. 20
figure 20

DEM-CFD results for partially fluid-saturated concrete specimen (with αq = 0.7): evolution of average pore fluid pressure against vertical normal strain εy during uniaxial compression

5.3 Effect of viscosity

In the scenario with fully fluid-saturated concrete (αq = 1), the impact of viscosity μq on the stress–strain curve is illustrated in Fig. 21. The DEM-CFD analyses were conducted using a viscosity that was two-fold lower (μq = 5.01 10−4 (Pa s)) as compared to the basic one (Table 1).

Fig. 21
figure 21

DEM-CFD simulation results for one-phase fully saturated concrete specimen (αq = 1.0) during uniaxial compression with fluid dynamic viscosity μq: relationship between vertical normal stress σy and vertical normal strain εy for two different fluid viscosities: a μq = 10.02 10−4 Pa s and b μq = 5.01 10−4 Pa s

With falling viscosity, the specimen's compressive strength rose as in the experiment [28]. The strength rise was about 10% (Fig. 21). The lower maximum pore fluid pressure of 0.4 MPa happened for the lower viscosity μq = 5.01 10−4 (Pa s).

6 Summary and conclusions

This work simulates the behavior of partially fluid-saturation concrete in uniaxial compression under 2D isothermal conditions using an improved fully coupled DEM/CFD-based hydro-mechanical technique. The method takes into account the two-phase laminar flow of fluid in the flow network of predefined channels and discretizes the geometry of pores/cracks with accuracy. The method demonstrated its capacity to evaluate concrete's compressive strength and fracture, and pore fluid pressures of wet concrete specimens. The impacts of both fluid saturation and fluid viscosity on the concrete strength and fracture process, and pore fluid pressures were studied, which proved to be significant. The numerical results showed qualitative agreement with laboratory tests in the literature regarding strength reduction with increasing fluid saturation and decreasing viscosity. The 2D simulations for concrete specimens with initial porosity of 5% lead to the following conclusions:

  • The compressive strength of concrete for fully fluid-saturated specimens was lower by 15% as compared to dry ones. It decreased with higher fluid saturation. The reduction was almost linear with increasing fluid saturation. The compressive strength of concrete for fully fluid-saturated specimens also dropped by 10% when dynamic fluid viscosity was two-fold higher.

  • The material behavior depended on the pore fluid pressure’s increase in pores over the initial one caused by initial specimen compaction induced by the compressive load. The pore fluid pressures promoted a fracture process during fluid migration through pores and cracks (higher contact forces occurred) that decreased the compressive strength. The higher the pore fluid pressure, the higher the compressive strength reduction and fluid migration rate.

  • Viscous fluid flow in pores and cracks had a higher influence on the quasi-static compressive strength than the pore volume changes themselves.

  • In the specimen, the effect of pore fluid pressure in uniaxial compression was the greatest during the initial specimen compaction phase when no micro-cracks were forming. Later, when cracks occurred, the average pore fluid pressures approached a constant value, and the maximum ones decreased and approached next a constant value.