Abstract

Moving packed-bed heat exchangers in concentrated solar power (CSP) plants involves heat transfer between heated falling particles and supercritical carbon dioxide. The overall effective thermal conductivity of the moving packed bed and particle-side channel contact resistances are still the bottlenecks in achieving the desirable thermal transport levels. To this end, a novel moving packed bed heat exchanger consisting of an Octet lattice packed between the walls of the particle-side channel is proposed in this study. Granular flow analysis in Octet lattice moving packed bed heat exchanger (OLHX) was conducted through experiments and discrete element method (DEM)-based numerical simulations. The experimental images clearly demonstrated stagnation regions upstream of lattice fibers, void regions downstream of the fiber junctions, and wavy-type unobstructed flow on the lateral sides of the fibers. DEM simulations were successful in capturing all these critical flow phenomena. Larger flow velocities were observed on the lateral sides of the fibers in the simulations. Also, when the particles in the silo were emptied, the final images showed an accumulation of particles on the inter-fiber as well as fiber–channel wall junctions. Moreover, the fiber connections resulted in some regions devoid of particle contact on the channel endwall, which means that these regions would suffer from poor thermal exchange. The overall mass flowrate increased with increasing porosity for a fixed particle diameter.

1 Introduction

Supercritical carbon-dioxide (sCO2)-based Brayton cycle is a promising alternative to the traditional steam-powered Rankine cycle because of its potential for achieving high thermal-to-electric conversion efficiency of ∼50% [1]. The overall conversion efficiency of the power cycle can be improved by increasing the working fluid temperature at the turbine inlet (typically ≥700 °C). To this end, the power cycles are coupled with the solar energy field in concentrating solar power plants (CSPs) as shown in Fig. 1 [2].

Fig. 1
Schematic of the falling-particle receiver system coupled with power cycle [2]
Fig. 1
Schematic of the falling-particle receiver system coupled with power cycle [2]
Close modal

The CSPs are an attractive technology as they feed on solar energy, a renewable source, and there has been a continuous effort to ensure their commercial scalability. In CSPs, the heat transfer fluid (HTF) flowing in a closed loop gets heated by the concentrated solar energy in the receiver section, and this thermal energy is eventually transferred to the working fluid (sCO2) of the power cycle via a particle-to-sCO2 heat exchanger. These HTFs also act as thermal energy storage (TES) systems which store the heat energy so that it can be dispatched as and when required, such as, in instances when solar energy is not readily available.

Molten salt and granular media are the most widely known HTFs for CSP application, and several heat exchanger (HX) designs such as shell-and-tube, shell-and-plate, and printed circuits have been investigated in the past based on these HTFs. Du et al. [3] investigated heat transfer characteristics of HITEC (NaNO3–KNO3–NaNO2 eutectic) salt–oil pair in a shell-and-tube HX and provided empirical correlations to predict the thermal performance. He et al. [4] analyzed the turbulent heat transfer characteristics of molten HITEC salt and YD-350 synthetic oil pair in a shell-and-tube HX and reported that no significant heat transfer augmentation occurred when salt was replaced by water as HTF. Qian et al. [5] compared the thermal performance of molten salt and conductive 320-oil in baffled shell-and-tube HX, where molten salt provided higher heat transfer performance. Shell-and-tube HX design is one of the most primitive designs which suffers from fluid leakage, large volume, and inability to sustain high pressures [6]. Printed circuit HXs have considerable advantages over these traditional HXs in terms of specific areas for heat transfer, sustainability of high-pressure drops (up to 60 MPa), and relatively compact sizes [6].

Printed circuit HXs consist of a stack of plates etched with minichannels, where the contact between the plates is ensured through a diffusion bonding process. These minichannels are fabricated in various shapes such as straight, trapezoidal, serpentine, zig-zag, etc. [6]. Wang et al. [7] experimentally studied the heat transfer performance of airfoil finned printed circuit HX with HITEC molten salt on the hot side and YD-325 synthetic oil on the cold side. The heat transfer of the airfoil channel was ∼ 7 times higher than the straight channel due to disruptions of fluid flow by the fins. Nikitin et al. [8] experimentally determined the heat transfer and pressure drop characteristics of sCO2 in a printed circuit HX having zig-zag channels with a semi-circular cross-section profile. The overall heat transfer coefficient for the zig-zag channel was obtained in the range of 300–650 W/m2K. Despite hosting several advantages over traditional shell-and-tube HXs, minichannels of printed circuit HXs have the drawback of potential plugging by the salt. Montes et al. [1] suggested a compact honeycomb HX where the molten salt flowed through a central tube and the sCO2 was present on the outer trapezoidal ducts surrounding the tube. The proposed design prohibited the clogging of channels and avoided large mechanical stresses due to high pressures of sCO2. Molten salts inherit some other limitations such as thermal instability, high melting points, and corrosive behavior [9]. Popular molten metals such as Nitrates have a maximum operating temperature limit of 560 °C, which prohibits the realization of turbine inlet temperatures ≥700 °C. On the other hand, solid particles as heat transfer media can have very high working temperatures (>1000 °C), high heat capacity, low-cost, and non-corrosive behavior [10]. Particle-to-sCO2 heat exchangers have been actively investigated recently due to these advantages and two of the most widely researched concepts are—(a) fluidized bed, and (b) gravity-driven moving packed-bed.

Fluidized bed technology provides higher overall particle-to-wall heat transfer coefficient values, but the cost associated with the machinery required to supply fluidized gas and with the recuperative heat exchange devices to minimize convective heat transfer losses from the gas poses challenges in their commercial applications [11]. Gravity-driven moving packed-bed heat exchangers have become a viable alternative due to their simpler setup and lower operational costs [12]. However, the heat transfer coefficient values of the moving packed-bed HXs are lower than the fluidized beds, which is being countered by developing innovative HX designs by several researchers. Albrecht and Ho [12] investigated a counter-flow multi-bank shell-and-plate HX where sCO2 flowed in a serpentine fashion via parallel plates and the particles moved in between the stacked plates under the influence of gravitational force. It was found that the thermal exchange in such configurations is limited by the low thermal conductivity of the particle bed. The authors in Ref. [12] proposed a reduced order modeling methodology to predict the heat transfer in HX and reported that an overall heat transfer coefficient of 400 W/m2K could be achieved for a particle channel width of 4 mm and particle diameter of 200 µm at CSP plant operating temperatures (500–800 °C). Baumann and Zunft [13] investigated crossflow shell-and-tube HX where the dense granular medium moved downward in a shell enclosing the tubes through which the working fluid of the power cycle passed. In such configurations, the heat transfer performance can deteriorate due to the stagnation zone on the upper vertex and the void zone in the wake of the tubes. The narrowing down of the horizontal tube pitch in Ref. [13] provided an increment in the overall heat transfer due to shorter particle–wall contact time.

The formation of the stagnation and void zones around the tubes in a shell-and-tube HX can be reduced by adopting more sophisticated tube designs as compared to the traditional circular ones. Tian et al. [14] proposed an elliptical-like combined tube shape constructed from the arrangement of three circular tubes. The stagnation and cavity zones in the novel elliptical-like tube were significantly lower than those in the conventional circular tubes resulting in the outside tube heat transfer performance due to particles similar to pure-elliptical configuration. The inside tube heat transfer of the elliptical-like configuration was compared with the corresponding elliptical- and circular-only configurations using the standard Gnielinski equation. For the same mass flowrate, the heat transfer coefficient was the highest for the elliptical-like configuration. In a recent study, Tian et al. [15] proposed a trapezoidal corrugated modification to the standard plates in a shell-to-plate HX where the trapezoidal bulge was reported to significantly improve the heat transfer between the particles and plate. The trapezoidal bulge enhanced particle–particle mixing and destroyed the thermal boundary layer development resulting in improved heat transfer in the downstream regions. The base angle of the trapezoidal bulge was reported to have a significant influence on the particle-to-wall heat transfer. Guo et al. [16] investigated the influence of the tube inclination on the particle-to-tube heat transfer. The contact resistance first increased and then decreased with the increment in the inclination angle. An optimum inclination angle of 15–37.5 deg was suggested by the authors to slow down the development of penetration resistance.

As discussed earlier, several novel HX designs have been explored in the past to improve the efficiency of the CSP plants. The granular flow and thermal exchange characteristics are being actively investigated through experimental and numerical means. Owing to the large capital investment required to conduct experiments, computational procedures have become a relatively more economical tool for testing and optimizing the new HX configurations and understating the intrinsic fundamental granular flow behavior. The discrete element method has become a very reliable and robust model in recent years to predict the macroscopic behavior of any granular system by resolving the particle–scale interactions. The concept of discrete element method (DEM) was first proposed by Cundall and Strack [17]. This method uses a Lagrangian framework to explicitly solve the trajectories of particles by evaluating the interparticle forces resulting from contact. The DEM theory is being currently employed by several researchers to model the granular interactions, and a discussion of some relevant studies is presented later.

Bartsch and Zunft [18] used DEM to model the dense granular flow field around the circular tubes in a moving packed-bed HX. The simulated results were in good agreement with the velocity profiles obtained from the particle image velocimetry (PIV) experiments. Yarrington et al. [19] measured the granular properties required as input in DEM simulations at elevated temperatures and consequently investigated its impact on the particles flowing down on an inclined plane. The free surface velocity magnitudes, flow volume fraction, average velocity magnitude, and mass flux contours were analyzed. Besides the simulation of moving packed-bed HXs in regard to the CSP plant applications, DEM is being extensively used to model the mixing and blending of particles in a rotating drum for chemical, pharmaceutical, ceramic, and fertilizer industries [2022]. DEM is used in combination with computational fluid dynamics (CFD) to model the particle flow behavior in fluidized beds as well [2325]. Guo et al. [26] employed DEM to investigate heat transfer in moving packed-bed tube HX, where the tubes were subjected to oscillations to improve the heat transfer. Tube oscillations improved thermal exchange by accelerating the particle update and improving the solid fraction in local zones. However, stronger oscillations led to the separation of particles from the tube, thereby deteriorating the heat transfer. Tian et al. [27] also investigated the effect of pin-fin and its oscillation on the gravity-driven granular flow. Several other researchers in the past have developed DEM codes to model the complex particle–particle heat transfer phenomenon [2830].

The particle-to-sCO2 heat transfer is influenced by several factors such as particle-side bulk thermal conductivity, particle–wall contact resistance, particle size, channel width, interstitial gas, particle velocity, and particle residence time [3133]. As discussed earlier, several types of HX concepts have been reported which can potentially augment the thermal exchange from the particle-side channel. In this study, the authors propose a novel gravity-driven moving packed-bed HX with an Octet lattice framework packed between the plates of the channel on the falling-particle side. Octet has been investigated extensively in the past for its superior mechanical properties [34,35]. The past work by our group focused on experimentally and numerically analyzing the convective heat transfer augmentation capabilities of fiber-based lattices from both Cubic and Octahedron families in forced convection channel configurations with both air and water as the working fluid [3640]. A thorough flow and heat transfer investigation has been conducted to characterize the flow patterns and thermal exchange dynamics rendered by four different topologies, namely, Octet, Tetrakaidecahedron, Cube, and FD-Cube, subjected to both laminar and turbulent flow conditions. Extensive experimental and numerical investigation was conducted recently to establish the stagnant thermal conductivities of the aforementioned lattices with water and air filled in their void volume [41,42]. Amongst these four investigated topologies, Octet provided the highest interfacial heat transfer coefficient values on the fibers whereas effective thermal conductivity was found to be independent of lattice topology for porosity ≥0.8. From prior investigations, it has been well established that Octet exhibits superior interfacial heat transfer and mechanical properties [43]. These previous findings are a motivation to extend the utilization of unique fiber-based Octet topology to granular flow systems employed in CSP plants. Recently, Aider et al. [44] experimentally evaluated the convective heat transfer coefficient of an Octet lattice moving packed-bed heat exchanger (OLHX) additively manufactured in 420 stainless steel and found that the particle flow through OLHX channel provided a heat transfer coefficient of 400 W/m2K at lattice porosity of 0.88. Moreover, the effective thermal conductivity of Octet–particles combination configurations was about 4–8 times higher than the corresponding particles-only configurations. OLHX therefore has the potential to reduce the particle-side thermal resistance and provide enhanced heat transfer owing to a large wetted specific area. The current study aims at experimentally visualizing the flow of particles and eventually developing a DEM framework to simulate the granular flow through OLHX. The mass flowrates of the particles under free-fall through OLHX were measured through experiments. The results from DEM simulations were compared to the experimental data in terms of the mass flowrate as well as local flow features. The particle-side heat transfer coefficient is critical in determining the viability of granular-based heat exchange technology in CSP plants as dictated by the $/kWh metrics, and the proposed OLHX can be optimized to meet these cost and power requirements of power plants.

2 Experimental Setup and Computational Domain

Figure 2 shows the schematic of the experimental setup used for granular flow visualization in OLHX, and Fig. 3 provides the details of the corresponding computational domain used in the DEM simulations. The OLHX consisted of a single-cell thick Octet lattice packed in an enclosure made of 12.7-mm-thick acrylic sheets. The acrylic panels provided the required transparent access to record the granular flow motion using a GoPro camera at 30 Hz. The setup consisted of a large hopper formed by bending mild steel sheets riveted at junctions to act as buffer storage for the desired number of particles before they were released to flow through the lattice under the influence of gravity. At the beginning of every test-run (t = 0), the lattice exit (shown in Fig. 2) was sealed with cardboard sheet, and the particles were poured from the top into the hopper such that the whole Octet lattice section and a certain known height of the hopper were charged with stagnant particles. The cardboard sheet at the lattice exit was then instantly removed to allow the particles to fall into a container underneath the exit. A stopwatch was used to monitor the time during the free-fall and the time when all the particles had drained out of the OLHX was recorded. The particles collected underneath the OLHX exit were then measured on a high-precision weighing scale to evaluate the mass of the particles. The measured mass of the particles and the recorded time were finally used to evaluate the time-averaged mass flowrate of particles. Each run was repeated several times to ensure the repeatability of results and accuracy of the procedure adopted.

Fig. 2
(a) Schematic of experimental setup used for granular flow visualization (not drawn to scale) and (b) additively manufactured Octet lattice packed between enclosure of acrylic glass
Fig. 2
(a) Schematic of experimental setup used for granular flow visualization (not drawn to scale) and (b) additively manufactured Octet lattice packed between enclosure of acrylic glass
Close modal
Fig. 3
Geometrical dimensions of the hopper and lattice section in the computational domain
Fig. 3
Geometrical dimensions of the hopper and lattice section in the computational domain
Close modal

The Octet lattices used in the experiments were additively manufactured in resin using an in-house Form 3 SLA printer (from Formlabs). The total length of each lattice in the mean-flow direction was 100 mm, and the width of the lattice was 50 mm (shown in Fig. 3). The lattice was single-cell thick where each repeating unit cell had dimensions of 10 mm × 10 mm × 10 mm, resulting in total ten cells along the streamwise direction and five along the channel span. Different views of the Octet topology are also shown in Fig. 3. Three porosities of the Octet lattice were investigated experimentally and numerically, namely, 0.75, 0.80, and 0.88. The corresponding fiber diameters were 1.55 mm, 1.36 mm, and 1 mm, respectively. The dimensional scales of the lattice section in the computational domain were exactly the same as that of the experiments. However, the large computational time and memory requirements prohibited the simulation of the exact same mass of particles as filled during experiments. A relatively lower mass of particles was simulated (due to the computational cost of DEM) due to which a scaled-down version of the hopper was modeled as shown in Fig. 3. The hopper in the computational domain resembled a frustum with a rectangular base. The width and depth of the base where it connected with the lattice section were 50 mm and 10 mm, respectively, in order to completely conform to the lattice entrance. The total height of the hopper was 120 mm which opened to a rectangular cross section of 80 mm × 40 mm on the top.

Unlike experiments, the covering lid was placed at the junction of the frustum base and the lattice entrance (shown in red color in Fig. 3) in computations. The lattice was bounded by walls of zero-thickness on the four sides in simulations. In the first step, the particles were injected from the top of the hopper and allowed to accumulate down on the lid under the influence of gravity. When the required amount of particle mass was injected into the hopper, the particles were allowed to settle down for some time to ensure that the mass flowrate during free-fall was calculated from the initial resting state of the particle just like in experiments. Figure 4 shows the resting particles filled in the hopper before the free-fall. Once the particles were settled, the lid below them was removed, and they were allowed to fall through the Octet lattice under the gravitational force.

Fig. 4
Particles settled in the hopper before being discharged through the Octet lattice
Fig. 4
Particles settled in the hopper before being discharged through the Octet lattice
Close modal

3 Particle Characteristics

Three different batches of sintered bauxite CARBOBEAD particles were experimentally investigated. Figure 5 shows the distribution profiles from the particle-size analysis performed using Anton Paar PSA 1190 LD. The investigated particles and the corresponding mean diameters are provided in Table 1. The flow of particles (three different diameters) was experimentally investigated for three different Octet lattice porosities; however, the simulations were performed using only CARBOBEAD-CP 30/60 particles for all three porosities of the lattice.

Fig. 5
Particle-size distribution profiles
Fig. 5
Particle-size distribution profiles
Close modal
Table 1

Mean diameter values of three different particle types experimentally investigated

Particle typeMean diameter
CARBOBEAD-CP 40/100∼266 µm
CARBOBEAD-CP 30/60∼397 µm
CARBOBEAD-HSP 20/40∼966 µm
Particle typeMean diameter
CARBOBEAD-CP 40/100∼266 µm
CARBOBEAD-CP 30/60∼397 µm
CARBOBEAD-HSP 20/40∼966 µm

4 Discrete Element Method Procedure

In the present study, the DEM simulations were performed using the open-source code liggghts (LAMMPS Improved for General Granular and Granular Heat Transfer Simulations) [45], which predicts the motion of a particle using the Newtons’ law of motion as given below [46].

For the translational motion:
(1)
For angular motion:
(2)
where v is the velocity vector, F is the net contact force acting on the particle, ω is the angular velocity vector, T is the net torque, and I is the moment of inertia. The particle–particle collision was modeled using “soft sphere” scheme where the deformation was approximated through normal and tangential interparticle overlaps. These overlaps were then used to evaluate the normal and tangential forces acting on the particle. Figure 6 represents the interactive-forces schematic between contacting particles through a spring-dashpot model [19,47]. Energy dissipation in the system is represented through the dashpot connected in parallel to the elastic springs. The sliding element placed in series with the spring in the tangential direction allows for account for the sliding of particles with each other while limiting the magnitude of tangential force.
Fig. 6
Spring-dashpot model for interaction in the normal and tangential directions
Fig. 6
Spring-dashpot model for interaction in the normal and tangential directions
Close modal
The Hertz–Mindlin contact model was specified during the simulations which calculates the interparticle forces as follows, provided the distance (D) between the two interacting particles of radii (Ri and Rj) is less than the contact distance d = (Ri + Rj):
(3)
The normal force had the contribution from spring and damping forces whereas the tangential forces were formulated using shear and damping forces. The values of elastic constants (kn and kt) and viscoelastic damping constants (γn and γt) were evaluated using the material properties and particle overlaps (δn and δt) as follows:
(4)
(5)
(6)
(7)
where
(8)
(9)
(10)
(11)
(12)
(13)
(14)

In the above-given formulae, Y is Young's modulus, G is the shear modulus, ν is Poisson's ratio, and e represents the coefficient of restitution. The superscript asterisk (*) refers to the effective properties. The shear force accounting for the tangential displacement between two interacting particles for the duration of contact time is a “history effect.” The tangential overlap was truncated to fulfill the criteria FtμsFn, where μs is the coefficient of static sliding friction and Fn is the normal force.

For the angular motion, the net torque is contributed by the torque from the tangential force and the rolling friction. Two rolling friction models from the current code library, namely, constant directional torque (CDT), and elastic–plastic spring-dashpot (EPSD2) were used in the current study.

The CDT model adds rolling friction torque as follows:
(15)
where ωs,ij is the projection of relative angular velocity (ωiωj) into the shear plane. For the CDT model, the coefficient of rolling friction (µr) must be specified. The EPSD2 model is determined as following over multiple time-steps without using the viscous damping contribution
(16)
where Δθr is the incremental relative rotation between the interacting particles. The maximum magnitude of spring torque is limited by the full mobilization torque determined using normal force and coefficient of rolling friction:
(17)

The correct evaluation of forces as per the equations presented in this section requires accurate a priori determination of several input properties such as Young's modulus, Poisson's ratio, and coefficients of sliding friction, rolling friction, and restitution. The properties and other numerical inputs adopted in the simulations of the current study are discussed in Sec. 5.

5 Numerical Setup

The parameters provided as input to evaluate particle–particle, particle–fiber, and particle–channel wall interactions are provided in Table 2. Both the fiber surfaces and channel walls were provided with the same material properties. The variables particle–particle sliding friction, rolling friction, and restitution coefficients were adopted from a recent study by Bagepalli et al. [48], where the authors experimentally measured the mechanical properties of CARBOBEAD-CP 30/60 proppants. The authors also reported the contact properties of particles against alumina- and resin-walls. The coefficient of sliding friction and rolling friction between particle–alumina wall was 1.0 and 0.41, respectively, whereas the coefficient values for the particle–resin wall were 0.80 and 0.40, respectively. The particle–wall sliding friction values of 0.25 and 0.30 were used in DEM simulations by Gallego et al. [49] when the walls were considered to be made of steel and methacrylate, respectively, while the glass beads were the falling granular medium.

Table 2

Properties of the particles used in simulations (the wall here refers to both the Octet fibers and surfaces of the bounding channel)

PropertiesValue
Particle diameter400 µm
Density3071 kg/m3
Young's modulus1 × 107 Pa
Poisson's ratio0.28
Particle–particle sliding friction coefficient0.53
Particle–particle rolling friction coefficient0.37
Particle–particle coefficient of restitution0.52
Particle–wall sliding friction coefficient0.3, 0.5, 1.0
Particle–wall rolling friction coefficient0.41
Particle–wall coefficient of restitution0.41
PropertiesValue
Particle diameter400 µm
Density3071 kg/m3
Young's modulus1 × 107 Pa
Poisson's ratio0.28
Particle–particle sliding friction coefficient0.53
Particle–particle rolling friction coefficient0.37
Particle–particle coefficient of restitution0.52
Particle–wall sliding friction coefficient0.3, 0.5, 1.0
Particle–wall rolling friction coefficient0.41
Particle–wall coefficient of restitution0.41

In the current study, alumina particles interacted with steel surface in the hopper, resin wall in the lattice volume, and acrylic glass on the bounding walls in the experimental setup. However, all the surfaces in a simulation were attributed the same sliding friction coefficient value for particle interaction. Three different cases of particle–wall sliding friction (0.3, 0.5, and 1.0) covering a wide range of values as reported in the literature discussed earlier were simulated in order to calibrate it against the current experimental data.

Time-step size (δt) for Hertzian contact model in DEM simulations is based on the Rayleigh time as provided by the following theoretical expression [50,51]:
(18)

The considered time-step (δt) should be smaller than the particle's contact duration [52]. A common guideline is to adopt δt of about 20% of tRay for a stable solution; however, larger time-steps can be adopted if the achieved particle velocities in the computational domain are not expected to be very large. In the current study, δt = 5 × 10−6 was specified in all the simulations. The value of Young's modulus significantly influences the time-step size because larger the Young's modulus smaller is the time-step required to resolve particle–scale interactions. Researchers generally use a reduced elastic modulus magnitude as compared to the actual Young's modulus of particles [46,52] to reduce the computational time. Zhou et al. [53] investigated the sensitivity of angle of repose of monosized spheres to various factors such as particle characteristics, material properties, and geometrical constraints. Young's modulus range of 105–108 N/m2 was investigated which showed no obvious effect on the angle of repose. The major controlling factors for translational and rotational motion of particles were found to be sliding and rolling frictions. Yan et al. [54] performed a parametric multi-level sensitivity analysis to understand the effect of DEM input particle properties on the bulk response for system involving discharge of particles from a flat bottom cylindrical container onto a plate. With all other parameter fixed, change in Young's modulus from 0.02 GPa to 200 GPa demonstrated small influence on the final shape of pile of material and average velocity of particles near the orifice. However, computational time increased from 0.4 h for Young's modulus value of 0.02 GPa with time-step size of 5 × 10−6 s to 240 h for Young's modulus value of 200 GPa at time-step size of 5 × 10−8 s. Xu et al. [55] investigated the effects of material properties on granular flow in a silo. The authors found that material modulus with artificial value of 1/1000 of the actual value yielded no significant difference in flow pattern and discharge rate suggesting that soft-sphere approach (lower modulus values) can be used to predict of behavior of hard particles to save computational time.

Yarrington et al. [19] and Bagepalli et al. [48] reported Young's modulus value of alumina particles to be ∼ 209 GPa which would require a very small time-step. Instead, Young's modulus of 1 × 107 Pa was used in the current study to reduce the overall computational cost.

In DEM simulations, the neighbor list of potential contacts is built periodically which is checked at every time-step, and the pairs of particles that are too far away from each other to be in contact are excluded. The pair of interacting particles can be included in the list based on a certain range as given by the following formula [45]:
(19)
where s is the skin parameter that is specified by the user. All interacting pairs within the cutoff distance provided by Eq. (19) are stored in the list which is periodically updated to account for the relocation of particles in a domain upon interaction. The larger the skin parameter, the less often the neighbor list requires rebuilding; however, a greater number of pairs must be checked every time-step for any possible interactions. The skin parameter used in the current study was 800 µm, i.e., twice the particle diameter. A 100 g of total particle mass was injected into the hopper before the free-fall. The mass-flowrate profiles were obtained from simulations at the exit plane of the Octet lattice.

6 Results and Discussion

6.1 Calibration of Sliding Friction Coefficient for Discrete Element Method Simulations.

As discussed in Sec. 5, three different values (0.3, 0.5, and 1.0) of particle–wall sliding friction coefficient were investigated numerically. The mass-flowrate characteristics of CP 30/60 particles using these three values in OLHX having a porosity of 0.88 is presented in Fig. 7, where all the simulation results correspond to the EPSD2 rolling friction model. The time t = 0 in these plots corresponded to the initiation of injection into the hopper. The same amount of time was provided for particle filling and resting in all the simulations before they were released for the free-fall. The mass flowrate measured from the experiments for the considered lattice porosity was 38.98 g/s. The simulated mass of the particles falling through OLHX provided a steady-mass flowrate for a considerable time indicating that 100 g was sufficient to analyze the flow behavior without the start- and end-effects. Significant deviation in mass flowrates was observed for the three different particle–wall sliding friction coefficients. As the coefficient of sliding friction decreased, the predicted mass-flowrate value increased. For the current setup, the best agreement with the experimental data was provided by a friction coefficient value of 0.5. Therefore, this value was used for the rest of the simulations.

Fig. 7
Comparison of mass flowrates obtained from DEM simulations for three different particle–wall sliding friction coefficients against the experimental data (mass flowrate was evaluated on the exit plane of the Octet lattice shown in Fig. 2 where the following quantities were remained constant over three cases of varying sliding frictions: lattice porosity = 0.88, particle diameter = 400 µm, Young's modulus = 1 × 107 Pa, rolling friction model = EPSD2, and particle–particle coefficient of restitution = 0.52).
Fig. 7
Comparison of mass flowrates obtained from DEM simulations for three different particle–wall sliding friction coefficients against the experimental data (mass flowrate was evaluated on the exit plane of the Octet lattice shown in Fig. 2 where the following quantities were remained constant over three cases of varying sliding frictions: lattice porosity = 0.88, particle diameter = 400 µm, Young's modulus = 1 × 107 Pa, rolling friction model = EPSD2, and particle–particle coefficient of restitution = 0.52).
Close modal

6.2 Rolling Friction Models.

Once the coefficient of sliding friction was fixed as 0.5, the effect of the rolling friction model on the mass flowrate was analyzed. Figure 8 shows the temporal variation of the mass flowrates of CP 30/60 particles in OLHX having porosity of 0.88 for two rolling friction models, namely, EPSD2 and CDT. No significant difference in the mass-flowrate characteristics was observed for the two rolling friction models. During the steady-rate condition, i.e., between 3 s and 4.5 s, the value of EPSD2 was slightly lower and CDT was slightly higher than the mean experimental value. Since the choice of the rolling friction model did not significantly influence the prediction of final mass-flowrate value, EPSD2 was chosen to conduct further simulations.

Fig. 8
Comparison of mass flowrates obtained from DEM simulations for two different rolling friction models (mass flowrate was evaluated on the exit plane of the Octet lattice shown in Fig. 2 where the following quantities were remained constant over the two cases of varying rolling friction models: lattice porosity = 0.88, particle diameter = 400 µm, Young's modulus = 1 × 107 Pa, coefficient of sliding friction = 0.5, and particle–particle coefficient of restitution = 0.52).
Fig. 8
Comparison of mass flowrates obtained from DEM simulations for two different rolling friction models (mass flowrate was evaluated on the exit plane of the Octet lattice shown in Fig. 2 where the following quantities were remained constant over the two cases of varying rolling friction models: lattice porosity = 0.88, particle diameter = 400 µm, Young's modulus = 1 × 107 Pa, coefficient of sliding friction = 0.5, and particle–particle coefficient of restitution = 0.52).
Close modal

6.3 Instantaneous Flow Characteristics.

Figure 9 shows the image of particles taken during experiments at instants when the particles were in steady-fall condition. Also presented is the zoomed-in image of the front side of the acrylic panel of OLHX having 0.88 porosity with CP 30/60 particles to better understand the granular motion.

Fig. 9
Instantaneous images of the free-fall of particles in three different lattices taken during experiments
Fig. 9
Instantaneous images of the free-fall of particles in three different lattices taken during experiments
Close modal

From the front view, three distinct flow characteristics were observed, i.e., (a) stagnant zone upstream of the fiber junctions, (b) cone-shaped void zone in the wake of the fibers, and (c) unobstructed wavy-type motion of particles on lateral sides of the fibers. The upstream stagnation of particles and downstream void due to particle separation is a common characteristic of granular flow around tubes in a shell-and-tube type heat exchanger. The downstream regions of fibers typically remain devoid of contact with the heated particles leading to weaker heat transfer. The overall conjugate heat transfer coefficient of OLHX depends on the effective thermal conductivity of the system and interfacial heat transfer coefficient values. The interfacial heat transfer coefficient depends on the flow fields rendered by the moving media in the lattice. The local heat transfer coefficient distribution around a discrete arrangement of circular tubes arranged normally and inclined to the mean-flow direction is reported in the literature. However, the current OLHX represents a relatively complex scenario where all the cylindrical fibers of the repeating unit cell are placed at a different angle to the encountering particles. The solid fibers attached to the bounding walls when fabricated in metal will develop thermal gradients internally leading to interesting thermal exchange dynamics with the particles. Figure 10 shows the images of OLHX (ɛ = 0.88) during the discharge of particles CP 30/60 colored by the local flow velocities obtained from the simulations. The particles on the lateral side of the fibers had relatively larger flow velocities which corresponded to the speeding wavy motion in Fig. 9. The zoomed-in view in Fig. 10(b) clearly demonstrates regions underneath the fibers junctions which were devoid of the particle contact. It is noted that not only the downstream fiber regions suffered from poor contact but also certain sections of the front- and side-acrylic walls (Fig. 10(c)). Therefore, the heat transfer coefficient values would have non-uniform distribution on the walls connected to the sCO2-side channel exhibiting regions of locally high and low heat transfer magnitudes. Moreover, the fiber–fiber junctions and fiber–channel wall junctions provided accommodation to the particles for accumulation as demonstrated in the post-free-fall image in Fig. 11. The particles collected in the junctions would not allow these spots to be updated with the fresh heated particles which will affect the heat transfer negatively in these regions over time.

Fig. 10
(a) Instantaneous image of the free-fall of particles from DEM simulations with particles colored with their local flow velocities, (b) zoomed-in view of the front side of the channel, and (c) zoomed-in side view of the channel
Fig. 10
(a) Instantaneous image of the free-fall of particles from DEM simulations with particles colored with their local flow velocities, (b) zoomed-in view of the front side of the channel, and (c) zoomed-in side view of the channel
Close modal
Fig. 11
Image of the particles taken after the completion of the DEM simulation showing accumulation in the junction regions
Fig. 11
Image of the particles taken after the completion of the DEM simulation showing accumulation in the junction regions
Close modal

6.4 Effect of Porosity on the Mass Flowrate.

The temporal evolution of mass-flowrate profiles for CP 30/60 particles at three different lattice porosities is provided in Fig. 12. For the same total mass of particles, the mass flowrate was maximum for the highest porosity value of 0.88. Therefore, the particles drained out of 88% porous OLHX in the least time span. The lowest porosity value of 0.75 provided the maximum hindrance to the free flow of the particles resulting in the maximum discharge time.

Fig. 12
Temporal evolution of mass-flowrate profiles for three lattice porosities obtained from DEM simulations (mass flowrate was evaluated on the exit plane of the Octet lattice shown in Fig. 2 where the following quantities were remained constant over the three cases of varying lattice porosities: particle diameter = 400 µm, Young's modulus = 1 × 107 Pa, rolling friction model = EPSD2, coefficient of sliding friction = 0.5, and particle–particle coefficient of restitution = 0.52).
Fig. 12
Temporal evolution of mass-flowrate profiles for three lattice porosities obtained from DEM simulations (mass flowrate was evaluated on the exit plane of the Octet lattice shown in Fig. 2 where the following quantities were remained constant over the three cases of varying lattice porosities: particle diameter = 400 µm, Young's modulus = 1 × 107 Pa, rolling friction model = EPSD2, coefficient of sliding friction = 0.5, and particle–particle coefficient of restitution = 0.52).
Close modal

Figure 13(a) shows mass flowrates obtained from the experimental runs of three different particle batches as a function of lattice porosity. The trends with respect to the porosity variation were similar for particles of different diameters, where the mass flowrate increased with increasing porosity. The difference in mass flowrates of CP 40/100 and 30/60 particles remained almost the same across the entire porosity range. The gap between mass flowrates for HSP 20/40 particles with respect to the other batches decreased as the lattice porosity increased. The larger porosity of the lattices ensures greater voidage volume for the particles to fall freely resulting in increased flow velocities in the domain resulting in relatively larger mass flowrates. However, it should be noted that the stagnant effective thermal conductivity would decrease with increasing porosity too. Figure 13(b) shows the comparison of experimental mean mass-flowrate data with the numerically predicted values as a function of porosity for CP 30/60 particles. Overall, the predicted values were in good agreement with the corresponding experimental data, especially for the higher porosities. The simulated mass had spherical particles of uniform diameter as opposed to the distribution profile provided in Fig. 5. This is expected to influence the mass-flowrate values predicted via simulations.

Fig. 13
(a) Experimentally obtained mass flowrates for three different particle batches as a function of lattice porosity, (b) comparison of mean mass flowrates obtained from the experiments and DEM simulations for CP 30/60 particles as a function of lattice porosity
Fig. 13
(a) Experimentally obtained mass flowrates for three different particle batches as a function of lattice porosity, (b) comparison of mean mass flowrates obtained from the experiments and DEM simulations for CP 30/60 particles as a function of lattice porosity
Close modal

6.5 Effect of Particle Diameter on the Mass Flowrate.

Figure 14 shows the experimentally obtained mass flowrates for three different porosities of the Octet lattice as a function of particle diameter. At a fixed porosity, the mass flowrate diminished with increasing particle diameter. The smaller diameter particles at any fixed porosity will have a tendency to movei faster in the porous volume and subsequently have more frequent interactions with the fibers of the lattice oriented at different angles. In the case of OLHX where the fiber network can be very dense for lower porosities, the scale of particles with respect to the fiber dimensions can be a very critical parameter of investigation. The previous study by the authors [44] showed that the particles with smaller diameters provided a higher overall heat transfer coefficient for OLHX at a fixed porosity.

Fig. 14
Experimentally obtained mass flowrates for three different lattice porosities as a function of particle diameter
Fig. 14
Experimentally obtained mass flowrates for three different lattice porosities as a function of particle diameter
Close modal

7 Conclusions

The experimental and DEM-based numerical investigation is conducted for flow in a novel Octet lattice moving packed-bed heat exchanger (OLHX) in this study with potential applications in CSP-based power generation. The OLHX packed beds demonstrate high effective thermal conductivity and overall heat transfer coefficients; therefore, it is imperative to gain in-depth knowledge of how it modifies the granular flow field. In the present study, experimental flow visualization was conducted on three different porosities, namely, 0.75, 0.80, and 0.88, of OLHX for three different CARBOBEAD particle batches—(a) CP 40/100, (b) CP 30/60, and (c) HSP 20/40. The DEM framework to simulate the granular flow through the OLHX was then developed in the open-source ligggths software, the results from which provided good agreement with the corresponding experimental data. This framework will be further developed to investigate the heat transfer in complex OLHX configurations. Major conclusions from this flow analysis study are highlighted as follows:

  • For a fixed particle diameter, the mass flowrate increased with increasing OLHX porosity. For a fixed porosity of the OLHX, the mass flowrate decreased with increasing particle diameter. The enhanced particle flow velocities coupled with frequent interactions with the fiber network are expected to provide high overall heat transfer coefficient values.

  • The experimental images showed cone-shaped void zones downstream of fibers, stagnation areas upstream of junctions, and wavy-type unobstructed motion around fibers. The DEM simulations were able to capture all these important flow features of the experimental results.

  • DEM images clearly demonstrated regions of the fiber network as well as channel walls which were completely devoid of the particle contact. These regions will provide very low thermal exchange opportunity leading to non-uniform heat transfer coefficient distribution on the fibers as well as channel walls in contact with the working fluid of the power cycle.

  • The images of the OLHX domain post-completion of simulations showed the accumulation of the particles on the junctions of fibers with other fibers as well as with the channel walls. These regions did not have very frequent updates of the fresh incoming particles which can prove to be the hot spots for poor thermal exchange over time.

  • The DEM simulations were conducted with the assumption of particles being spherical and having uniform fixed diameters which was different from the measured morphological characteristics of the particles used in experiments. Modeling of real-like particle distribution can further improve the accuracy of the DEM simulations.

Acknowledgment

This material is based upon work supported by the U.S. Department of Energy's Office of Energy Efficiency and Renewable Energy (EERE) under the Solar Energy Technologies Office Award Number DE-EE0009377. This report was prepared as an account of work sponsored by an agency of the United States Government. Neither the United States Government nor any agency thereof, nor any of their employees, makes any warranty, express or implied, or assumes any legal liability or responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately owned rights. Reference herein to any specific commercial product, process, or service by trade name, trademark, manufacturer, or otherwise does not necessarily constitute or imply its endorsement, recommendation, or favoring by the U.S. Government or any agency thereof. The views and opinions of authors expressed herein do not necessarily state or reflect those of the United States Government or any agency thereof.

Conflict of Interest

There are no conflicts of interest.

Data Availability Statement

The datasets generated and supporting the findings of this article are obtainable from the corresponding author upon reasonable request.

Nomenclature

e =

coefficient of restitution

g =

gravitational force

m =

mass of the particle

s =

skin parameter

v =

velocity

x =

particle position

F =

force acting on the particle

G =

shear modulus

I =

moment of inertia

R =

radius of the particle

T =

torque acting on the particle

Y =

Young's modulus

df =

fiber diameter

dp =

mean particle diameter

kn =

elastic constant for normal contact

kt =

elastic constant for tangential contact

tRay =

Rayleigh time-step criteria

vn =

normal component of the relative velocity

vt =

tangential component of the relative velocity

Sn =

material constant in normal direction

St =

material constant in tangential direction

Greek Symbols

β =

material constant used to calculate interacting forces in LIGGGHTS

ɛ =

porosity

δn =

normal overlap distance between two particles

δt =

tangential displacement between two particles

γn =

viscoelastic damping constant for normal contact

γt =

viscoelastic damping constant for tangential contact

μs =

sliding friction coefficient

μr =

rolling friction coefficient

ν =

Poisson's ratio

ρ =

density of the particle

ω =

angular velocity

Superscript

* =

effective properties

Abbreviations

CDT =

constant direction torque

CSP =

concentrated solar power plant

DEM =

discrete element method

EPSD =

elastic–plastic spring dashpot

HTF =

heat transfer fluid

HX =

heat exchanger

OLHX =

Octet lattice moving packed-bed heat exchanger

sCO2 =

supercritical carbon-dioxide

TES =

thermal energy storage

References

1.
Montes
,
M. J.
,
Linares
,
J. I.
,
Barbero
,
R.
, and
Moratilla
,
B. Y.
,
2020
, “
Optimization of a New Design of Molten Salt-to-CO2 Heat Exchanger Using Exergy Destruction Minimization
,”
Entropy
,
22
(
8
), p.
883
.
2.
Mehos
,
M.
,
Turch
,
C.
,
Vidal
,
J.
,
Wagner
,
M.
,
Ma
,
Z.
,
Ho
,
C.
,
Kolb
,
W.
,
Andraka
,
C.
, and
Kruizenga
,
A.
,
2017
, “Concentrating Solar Power Gen3 Demonstration Roadmap (No. NREL/TP-5500-67464),” National Renewable Energy Lab.(NREL), Golden, CO.
3.
Du
,
B. C.
,
He
,
Y. L.
,
Qiu
,
Y.
,
Liang
,
Q.
, and
Zhou
,
Y. P.
,
2018
, “
Investigation on Heat Transfer Characteristics of Molten Salt in a Shell-and-Tube Heat Exchanger
,”
Int. Commun. Heat Mass Transfer
,
96
, pp.
61
68
.
4.
He
,
Y. L.
,
Zheng
,
Z. J.
,
Du
,
B. C.
,
Wang
,
K.
, and
Qiu
,
Y.
,
2016
, “
Experimental Investigation on Turbulent Heat Transfer Characteristics of Molten Salt in a Shell-and-Tube Heat Exchanger
,”
Appl. Therm. Eng.
,
108
, pp.
1206
1213
.
5.
Qian
,
J.
,
Kong
,
Q. L.
,
Zhang
,
H. W.
,
Zhu
,
Z. H.
,
Huang
,
W. G.
, and
Li
,
W. H.
,
2017
, “
Experimental Study for Shell-and-Tube Molten Salt Heat Exchangers
,”
Appl. Therm. Eng.
,
124
, pp.
616
623
.
6.
Huang
,
C.
,
Cai
,
W.
,
Wang
,
Y.
,
Liu
,
Y.
,
Li
,
Q.
, and
Li
,
B.
,
2019
, “
Review on the Characteristics of Flow and Heat Transfer in Printed Circuit Heat Exchangers
,”
Appl. Therm. Eng.
,
153
, pp.
190
205
.
7.
Wang
,
W. Q.
,
Qiu
,
Y.
,
He
,
Y. L.
, and
Shi
,
H. Y.
,
2019
, “
Experimental Study on the Heat Transfer Performance of a Molten-Salt Printed Circuit Heat Exchanger With Airfoil Fins for Concentrating Solar Power
,”
Int. J. Heat Mass Transfer
,
135
, pp.
837
846
.
8.
Nikitin
,
K.
,
Kato
,
Y.
, and
Ngo
,
L.
,
2006
, “
Printed Circuit Heat Exchanger Thermal–Hydraulic Performance in Supercritical CO2 Experimental Loop
,”
Int. J. Refrig.
,
29
(
5
), pp.
807
814
.
9.
Encinas-Sánchez
,
V.
,
Lasanta
,
M. I.
,
De Miguel
,
M. T.
,
García-Martín
,
G.
, and
Pérez
,
F. J.
,
2021
, “
Corrosion Monitoring of 321H in Contact With a Quaternary Molten Salt for Parabolic Trough CSP Plants
,”
Corros. Sci.
,
178
, p.
109070
.
10.
Tian
,
X.
,
Yang
,
J.
,
Guo
,
Z.
,
Wang
,
Q.
, and
Sunden
,
B.
,
2020
, “
Numerical Study of Heat Transfer in Gravity-Driven Dense Particle Flow Around a Hexagonal Tube
,”
Powder Technol.
,
367
, pp.
285
295
.
11.
Albrecht
,
K. J.
, and
Ho
,
C. K.
,
2019
, “
Heat Transfer Models of Moving Packed-Bed Particle-to-sCO2 Heat Exchangers
,”
ASME J. Sol. Energy Eng.
,
141
(
3
), p.
031006
.
12.
Albrecht
,
K. J.
, and
Ho
,
C. K.
,
2019
, “
Design and Operating Considerations for a Shell-and-Plate, Moving Packed-Bed, Particle-to-sCO2 Heat Exchanger
,”
Sol. Energy
,
178
, pp.
331
340
.
13.
Baumann
,
T.
, and
Zunft
,
S.
,
2015
, “
Development and Performance Assessment of a Moving Bed Heat Exchanger for Solar Central Receiver Power Plants
,”
Energy Procedia
,
69
, pp.
748
757
.
14.
Tian
,
X.
,
Guo
,
Z.
,
Jia
,
H.
,
Yang
,
J.
, and
Wang
,
Q.
,
2021
, “
Numerical Investigation of a New Type Tube for Shell-and-Tube Moving Packed Bed Heat Exchanger
,”
Powder Technol.
,
394
, pp.
584
596
.
15.
Tian
,
X.
,
Zhu
,
F.
,
Guo
,
Z.
,
Zhang
,
J.
,
Yang
,
J.
, and
Wang
,
Q.
,
2022
, “
Numerical Investigation of Gravity-Driven Particle Flow Along the Trapezoidal Corrugated Plate for a Moving Packed bed Heat Exchanger
,”
Powder Technol.
,
405
, p.
117526
.
16.
Guo
,
Z.
,
Yang
,
J.
,
Tan
,
Z.
,
Tian
,
X.
, and
Wang
,
Q.
,
2021
, “
Numerical Study on Gravity-Driven Granular Flow Around Tube Out-Wall: Effect of Tube Inclination on the Heat Transfer
,”
Int. J. Heat Mass Transfer
,
174
, p.
121296
.
17.
Cundall
,
P. A.
, and
Strack
,
O. D.
,
1979
, “
A Discrete Numerical Model for Granular Assemblies
,”
Geotechnique
,
29
(
1
), pp.
47
65
.
18.
Bartsch
,
P.
, and
Zunft
,
S.
,
2019
, “
Granular Flow Around the Horizontal Tubes of a Particle Heat Exchanger: DEM-Simulation and Experimental Validation
,”
Sol. Energy
,
182
, pp.
48
56
.
19.
Yarrington
,
J. D.
,
Bagepalli
,
M. V.
,
Pathikonda
,
G.
,
Schrader
,
A. J.
,
Zhang
,
Z. M.
,
Ranjan
,
D.
, and
Loutzenhiser
,
P. G.
,
2021
, “
Numerical Analyses of High Temperature Dense, Granular Flows Coupled to High Temperature Flow Property Measurements for Solar Thermal Energy Storage
,”
Sol. Energy
,
213
, pp.
350
360
.
20.
Soni
,
R. K.
,
Mohanty
,
R.
,
Mohanty
,
S.
, and
Mishra
,
B. K.
,
2016
, “
Numerical Analysis of Mixing of Particles in Drum Mixers Using DEM
,”
Adv. Powder Technol.
,
27
(
2
), pp.
531
540
.
21.
Liu
,
P. Y.
,
Yang
,
R. Y.
, and
Yu
,
A. B.
,
2013
, “
DEM Study of the Transverse Mixing of Wet Particles in Rotating Drums
,”
Chem. Eng. Sci.
,
86
, pp.
99
107
.
22.
Xiao
,
X.
,
Tan
,
Y.
,
Zhang
,
H.
,
Deng
,
R.
, and
Jiang
,
S.
,
2017
, “
Experimental and DEM Studies on the Particle Mixing Performance in Rotating Drums: Effect of Area Ratio
,”
Powder Technol.
,
314
, pp.
182
194
.
23.
Deb
,
S.
, and
Tafti
,
D. K.
,
2014
, “
Two and Three Dimensional Modeling of Fluidized Bed With Multiple Jets in a DEM–CFD Framework
,”
Particuology
,
16
, pp.
19
28
.
24.
Li
,
T.
, and
Guenther
,
C.
,
2012
, “
MFIX-DEM Simulations of Change of Volumetric Flow in Fluidized Beds Due to Chemical Reactions
,”
Powder Technol.
,
220
, pp.
70
78
.
25.
Xu
,
Y.
,
Li
,
T.
,
Musser
,
J.
,
Liu
,
X.
,
Xu
,
G.
, and
Rogers
,
W. A.
,
2017
, “
CFD-DEM Modeling the Effect of Column Size and bed Height on Minimum Fluidization Velocity in Micro Fluidized Beds With Geldart B Particles
,”
Powder Technol.
,
318
, pp.
321
328
.
26.
Guo
,
Z.
,
Zhang
,
S.
,
Tian
,
X.
,
Yang
,
J.
, and
Wang
,
Q.
,
2020
, “
Numerical Investigation of Tube Oscillation in Gravity-Driven Granular Flow With Heat Transfer by Discrete Element Method
,”
Energy
,
207
, p.
118203
.
27.
Tian
,
X.
,
Yang
,
J.
,
Guo
,
Z.
, and
Wang
,
Q.
,
2021
, “
Numerical Investigation of Gravity-Driven Granular Flow Around the Vertical Plate: Effect of Pin-Fin and Oscillation on the Heat Transfer
,”
Energies
,
14
(
8
), p.
2187
.
28.
Peng
,
Z.
,
Doroodchi
,
E.
, and
Moghtaderi
,
B.
,
2020
, “
Heat Transfer Modelling in Discrete Element Method (DEM)-Based Simulations of Thermal Processes: Theory and Model Development
,”
Prog. Energy Combust. Sci.
,
79
, p.
100847
.
29.
Batchelor
,
G. K.
, and
O'brien
,
R. W.
,
1977
, “
Thermal or Electrical Conduction Through a Granular Material
,”
Proc. R. Soc. A
,
355
(
1682
), pp.
313
333
.
30.
Wang
,
S.
,
Luo
,
K.
,
Hu
,
C.
,
Lin
,
J.
, and
Fan
,
J.
,
2019
, “
CFD-DEM Simulation of Heat Transfer in Fluidized Beds: Model Verification, Validation, and Application
,”
Chem. Eng. Sci.
,
197
, pp.
280
295
.
31.
Denloye
,
A. O. O.
, and
Botterill
,
J. S. M.
,
1977
, “
Heat Transfer in Flowing Packed Beds
,”
Chem. Eng. Sci.
,
32
(
5
), pp.
461
465
.
32.
Sullivan
,
W. N.
, and
Sabersky
,
R. H.
,
1975
, “
Heat Transfer to Flowing Granular Media
,”
Int. J. Heat Mass Transfer
,
18
(
1
), pp.
97
107
.
33.
Spelt
,
J. K.
,
Brennen
,
C. E.
, and
Sabersky
,
R. H.
,
1982
, “
Heat Transfer to Flowing Granular Material
,”
Int. J. Heat Mass Transfer
,
25
(
6
), pp.
791
796
.
34.
Lv
,
W.
,
Li
,
D.
, and
Dong
,
L.
,
2020
, “
Study on Mechanical Properties of a Hierarchical Octet-Truss Structure
,”
Compos. Struct.
,
249
, p.
112640
.
35.
Deshpande
,
V. S.
,
Fleck
,
N. A.
, and
Ashby
,
M. F.
,
2001
, “
Effective Properties of the Octet-Truss Lattice Material
,”
J. Mech. Phys. Solids
,
49
(
8
), pp.
1747
1769
.
36.
Kaur
,
I.
, and
Singh
,
P.
,
2020
, “
Flow and Thermal Transport Through Unit Cell Topologies of Cubic and Octahedron Families
,”
Int. J. Heat Mass Transfer
,
158
, p.
119784
.
37.
Kaur
,
I.
, and
Singh
,
P.
,
2021
, “
Numerical Investigation on Conjugate Heat Transfer in Octet-Shape-Based Single Unit Cell Thick Metal Foam
,”
Int. Commun. Heat Mass Transfer
,
121
, p.
105090
.
38.
Kaur
,
I.
, and
Singh
,
P.
,
2022
, “
Direct Pore-Scale Simulations of Fully Periodic Unit Cells of Different Regular Lattices
,”
ASME J. Heat Transfer-Trans. ASME
,
144
(
2
), p.
022702
.
39.
Aider
,
Y.
,
Kaur
,
I.
,
Cho
,
H.
, and
Singh
,
P.
,
2022
, “
Periodic Heat Transfer Characteristics of Additively Manufactured Lattices
,”
Int. J. Heat Mass Transfer
,
189
, p.
122692
.
40.
Kaur
,
I.
,
Aider
,
Y.
,
Nithyanandam
,
K.
, and
Singh
,
P.
,
2022
, “
Thermal-Hydraulic Performance of Additively Manufactured Lattices for Gas Turbine Blade Trailing Edge Cooling
,”
Appl. Therm. Eng.
,
211
, p.
118461
.
41.
Kaur
,
I.
,
Mahajan
,
R. L.
, and
Singh
,
P.
,
2023
, “
Generalized Correlation for Effective Thermal Conductivity of High Porosity Architectured Materials and Metal Foams
,”
Int. J. Heat Mass Transfer
,
200
, p.
123512
.
42.
Wang
,
N.
,
Kaur
,
I.
,
Singh
,
P.
, and
Li
,
L.
,
2021
, “
Prediction of Effective Thermal Conductivity of Porous Lattice Structures and Validation With Additively Manufactured Metal Foams
,”
Appl. Therm. Eng.
,
187
, p.
116558
.
43.
Dong
,
L.
, and
Wadley
,
H.
,
2015
, “
Mechanical Properties of Carbon Fiber Composite Octet-Truss Lattice Structures
,”
Compos. Sci. Technol.
,
119
, pp.
26
33
.
44.
Aider
,
Y.
,
Cho
,
H.
, and
Singh
,
P.
,
2022
, “
Convective Heat Transfer Potential of Particles/Airflow Through Single Cell Thick Additively Manufactured Octet-Shaped Lattice Frame Material
,”
Heat Transfer Summer Conference
,
Philadelphia, PA
,
July 11–13
, Vol.
85796
, p.
V001T06A004
.
45.
Kloss
,
C.
,
Goniva
,
C.
,
Hager
,
A.
,
Amberger
,
S.
, and
Pirker
,
S.
,
2012
, “
Models, Algorithms and Validation for Opensource DEM and CFD–DEM
,”
Prog. Comput. Fluid Dyn.
,
12
(
2-3
), pp.
140
152
.
46.
Jahani
,
M.
,
Farzanegan
,
A.
, and
Noaparast
,
M.
,
2015
, “
Investigation of Screening Performance of Banana Screens Using LIGGGHTS DEM Solver
,”
Powder Technol.
,
283
, pp.
32
47
.
47.
Deb
,
S.
, and
Tafti
,
D.
,
2014
, “
Investigation of Flat Bottomed Spouted Bed With Multiple Jets Using DEM–CFD Framework
,”
Powder Technol.
,
254
, pp.
387
402
.
48.
Bagepalli
,
M. V.
,
Yarrington
,
J. D.
,
Schrader
,
A. J.
,
Zhang
,
Z. M.
,
Ranjan
,
D.
, and
Loutzenhiser
,
P. G.
,
2020
, “
Measurement of Flow Properties Coupled to Experimental and Numerical Analyses of Dense, Granular Flows for Solar Thermal Energy Storage
,”
Sol. Energy
,
207
, pp.
77
90
.
49.
Gallego
,
E.
,
Fuentes
,
J. M.
,
Wiącek
,
J.
,
Villar
,
J. R.
, and
Ayuga
,
F.
,
2019
, “
DEM Analysis of the Flow and Friction of Spherical Particles in Steel Silos With Corrugated Walls
,”
Powder Technol.
,
355
, pp.
425
437
.
50.
Washino
,
K.
,
Chan
,
E. L.
,
Miyazaki
,
K.
,
Tsuji
,
T.
, and
Tanaka
,
T.
,
2016
, “
Time Step Criteria in DEM Simulation of Wet Particles in Viscosity Dominant Systems
,”
Powder Technol.
,
302
, pp.
100
107
.
51.
Burns
,
S. J.
,
Piiroinen
,
P. T.
, and
Hanley
,
K. J.
,
2019
, “
Critical Time Step for DEM Simulations of Dynamic Systems Using a Hertzian Contact Model
,”
Int. J. Numer. Methods Eng.
,
119
(
5
), pp.
432
451
.
52.
Paulick
,
M.
,
Morgeneyer
,
M.
, and
Kwade
,
A.
,
2015
, “
Review on the Influence of Elastic Particle Properties on DEM Simulation Results
,”
Powder Technol.
,
283
, pp.
66
76
.
53.
Zhou
,
Y. C.
,
Xu
,
B. H.
,
Yu
,
A. B.
, and
Zulli
,
P.
,
2001
, “
Numerical Investigation of the Angle of Repose of Monosized Spheres
,”
Phys. Rev. E
,
64
(
2
), p.
021301
.
54.
Yan
,
Z.
,
Wilkinson
,
S. K.
,
Stitt
,
E. H.
, and
Marigo
,
M. J. C. P. M.
,
2015
, “
Discrete Element Modelling (DEM) Input Parameters: Understanding Their Impact on Model Predictions Using Statistical Analysis
,”
Comput. Part. Mech.
,
2
(
3
), pp.
283
299
.
55.
Xu
,
Y.
,
Kafui
,
K. D.
,
Thornton
,
C.
, and
Lian
,
G.
,
2002
, “
Effects of Material Properties on Granular Flow in a Silo Using DEM Simulation
,”
Part. Sci. Technol.
,
20
(
2
), pp.
109
124
.