Next Article in Journal
Machine Learning for Solar Resource Assessment Using Satellite Images
Next Article in Special Issue
A Critical Review on the Use of Shallow Geothermal Energy Systems for Heating and Cooling Purposes
Previous Article in Journal
Development of Fault Diagnosing System for Ice-Storage Air-Conditioning Systems
Previous Article in Special Issue
Future Technology Mix—Enhanced Geothermal System (EGS) and Carbon Capture, Utilization, and Storage (CCUS)—An Overview of Selected Projects as an Example for Future Investments in Poland
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Down-Hole Electromagnetic Heating of Deep Aquifers for Renewable Energy Storage

by
Samuel O. de Almeida
1,*,
Grigori Chapiro
2,* and
Pacelli L. J. Zitha
3,*
1
Federal Institute of the Southeast of Minas Gerais, Santos Dumont 36240-000, Brazil
2
Department of Mathematics, Federal University of Juiz de Fora, Juiz de Fora 36036-900, Brazil
3
Civil Engineering and Geosciences, Delft University of Technology, 2628 CD Delft, The Netherlands
*
Authors to whom correspondence should be addressed.
Energies 2022, 15(11), 3982; https://doi.org/10.3390/en15113982
Submission received: 30 March 2022 / Revised: 28 April 2022 / Accepted: 30 April 2022 / Published: 28 May 2022
(This article belongs to the Special Issue Geothermal Systems)

Abstract

:
Electromagnetic (EM) heating is an emerging method for storing renewable energy, such as photovoltaic solar and wind electric power, into aquifers. We investigate how the captured energy increases the temperature of a prototypical deep aquifer for a six-month period and then to which extent the stored energy can be recovered during the consecutive six months. Water injected at a constant flow rate is simultaneously heated using a high-frequency electromagnetic microwave emitter operating at the water natural resonance frequency of 2.45 GHz. The coupled reservoir flow and EM heating are described using Darcy’s and the energy balance equations. The latter includes a source term accounting for the EM wave propagation and absorption, modeled separately using Maxwell’s equations. The equations are solved numerically by the Galerkin least-squares finite element method. The approach was validated using EM-heating input data obtained from controlled laboratory experiments and then was applied to the aquifer. We found that after six years of alternate storage and recovery, up to 77% of the injected energy is recovered when considering realistic heat losses estimated from field data. Even when heat losses are increased by a factor of two, up to 69% of the injected energy is recovered in this case. This shows that down-hole EM heating is a highly effective method for storing renewable energies, capable of helping to solve their inherent intermittency.

Graphical Abstract

1. Introduction

In the last decade, the significance of solar and wind energy sources has grown considerably under the societal pressure to curb CO2 emissions imputable to the use of natural hydrocarbons. Solar power plants generate substantial amounts of energy, but mostly during daytime and in summer, and wind-turbine energy output depends on weather conditions. This is in stark contrast with the energy consumption profile in the built environment, which is higher at later parts of the day and in the colder seasons [1]. Resolving this intermittency conundrum of solar and wind energies is essential for a wider acceptance and usage of these renewable energy sources.
Aquifer thermal energy storage (ATES) seeks to increase the efficiency of the terrestrial energy system by storing heated water in the subsurface; see [2]. One of the simplest forms of the ATES system consists of two underground wells that operate cyclically between summer and winter [3]. In summer, cold water is extracted from the aquifer and is generally used for cooling buildings and industrial processes, and then is reinjected into the aquifer through the storage well. In winter, the reverted water flows in any direction possible in the heating buildings and industries. ATES techniques are based on heated water injection; the majority in shallow reservoirs, and a few others in medium/deep aquifers [3,4,5]. Notice that the latter type is more affected by heat losses.
Hybrid energy systems combining low-enthalpy (low-H) geothermal sources, i.e., aquifers with temperatures below 90 °C, with wind or solar energy, were studied by [6,7,8,9]. The geothermal sources can potentially provide an enormous energy storage capacity [10] while enhancing the overall efficiency of the energy supply [11,12]. Nevertheless, there is still a lack of data and models concerning this process for high depths. The present paper follows a similar idea by considering energy storage at deep reservoirs, where thermal losses can play a significant role. Including these types of aquifers increases renewable energy usage.
Subsurface EM heating is not completely new [13,14,15]. It has been a subject of several studies mainly in the context of clean-up of polluted soils [16], and enhanced oil recovery (EOR) [17,18,19,20]. However, it is far from being completely understood, especially in the context of geothermal energy sources. For instance, the impact of EM propagation on the efficiency of energy storage and the role of the various energy losses have not been discussed in detail. Electromagnetic heating is hardly affected by the heterogeneity of the reservoir and allows uniform heat distribution in the medium [21]. In addition, compact and high-efficiency EM heating equipment allows a more straightforward setup inside the well. The experiments reported in [22] showed that heat could be induced from the inside out, while [23] demonstrated that it is possible to obtain heating faster when compared to traditional thermal techniques, resulting in less heat loss due to the greater penetration in the medium. Authors report similar results in [24], stating that the EM heating is advantageous in avoiding thermal losses when compared to traditional thermal methods such as hot water (or steam) injection. Numerical and analytical study [25], considering thermal losses proposed in [26], was performed calculating the heat loss for adjacent extracts considering heat conduction. Similarly, vertical heat losses for over- and underburden formations were modeled through the heat diffusion equation [27].
After the model validation on the data obtained from the laboratory-controlled experiments, we conducted a numerical analysis of the energy storage and recovery from an aquifer based on a coupled EM heating and reservoir flow mathematical model. As illustrated in Figure 1, we apply the model to assess the efficiency in six-month cycle production and storage for a total duration of six years. The paper proceeds with the physical model, numerical details, and results and discussion, followed by conclusions. For completeness, we present the approach validation in the Appendix, although it can also be found in [28].

2. Physical Model

Consider a solar and wind power electrical power source and a well drilled into an aquifer, as depicted in Figure 1. A source of EM waves operating in the microwave at wavelengths is placed at the bottom of the well to enable energy to be injected directly into the reservoir [29]. This configuration clearly minimizes the heat losses due to energy transfer along the walls of the injection well and maximizes the total amount of energy reaching the reservoir. Our discussion will be confined to the less-evident EM propagation in the reservoir.
To fix ideas, we assume the reservoir to be a homogeneous and isotropic reservoir with porosity ϕ and permeability k and dimensions L × L × h . In addition to the injection well already mentioned, four wells are drilled at the edges of the reservoir so that altogether they form a five-spot configuration (see Figure 1). The well diameter is considered equal to 27.3 cm (or 10 3/4 in), one of the standard casing diameters. Water at temperature θ 0 is injected into the aquifer at a flow rate Q w and a constant pressure gradient, while the microwave emitter irradiates energy with nominal power W. From the physical point of view, we need to solve the problem of the flow of water into the reservoir coupled with the heat transfer with an EM source.
This work’s premise is that the energy stored in the aquifer can be recovered through the four production wells. Since flow and temperature profiles are symmetrical with respect to the central well’s axis, we present only the first quadrant of the simulation.
Note that heating of shallow aquifers should not exceed 30 °C (LT-ATES) due to environmental, economic, and social risks; see [5] for details. We consider deep aquifers with high temperature and pressure, although we limit this study to reservoir temperature below 365.7 °C and pressures over 20 MPa to avoid the possibility of steam formation. When working with deep aquifers, the risks of harm to the environment are lower. Another factor is the presence of multiple rock layers overlying the aquifer, which will limit the vertical heat flow and chemical transport to the surface. In fact, high temperatures occur naturally at these depths in various parts of the world due to relative proximity to the Earth’s core.

2.1. Governing Equations

Fluid flow in the aquifer is governed by mass, momentum, and energy conservation. For porous media flow, the momentum equation reduces to Darcy’s law; see [30,31]. As we deal with a reservoir-scale phenomenon with a significant timescale (months), we assume that flow is incompressible and in the steady state. The mass conservation equations and Darcy’s law can be combined into the following pressure equation:
· k ν p = 0 , where log ν = a 1 + a 2 ( a 3 θ ) ,
where k is the permeability of the porous medium, ν is the viscosity of water depending on temperature θ [32], and p is the local pressure in the reservoir.
The energy balance is expressed by the following heat convection-diffusion equation [20,33,34]:
C t θ t + C w · ( V θ ) = K t Δ θ C l ( θ θ 0 ) + W ,
where velocity V is obtained from Darcy’s equation using pressure field from Equation (1), C t is the system’s specific heat, C w is the total fluid specific heat, K t is the total system thermal conductivity, C l is the thermal losses coefficient, and W is the total electromagnetic energy source. Note that, for simplicity, the thermal expansion of the rock and the fluids are assumed to be negligible, which is an acceptable approximation for the range of temperature variations considered in this study. The temperature in the reservoir rises due to heat conduction, convection, and radiation. Using the relationships between volume fraction, saturation, total Darcy velocity, and flow function, we can write the coefficients as
C t = ϕ ρ w c w + ( 1 ϕ ) ρ s c s ,
C w = ρ w c w + ρ s c s ,
K t = ϕ k w + k s ( 1 ϕ ) .
Note that, consistently with Equation (2), the amount of heat transferred by conduction q (W) in the unit of time inside the material in the vertical direction is given by Fourier thermal conduction law:
q = K t A d z θ ,
where K t is the thermal conductivity (W/(m K)), and A (m2) is the area of the section perpendicular to the flow’s direction. Equating the total energy loss, calculated by using (6), for L × L × h reservoir’s size (length, width, and height), with the same using the linear thermal loss coefficient C l , we arrive at C l = K t / h 2 .

2.2. EM Heating

The detailed derivation of the total electromagnetic energy source W of electromagnetic heating from the Maxwell equations can be found in [28]. Here, we recall the main result, which can be expressed as
W ( r ) = P 0 2 π h 2 α e 2 α ( r r w ) r + e 2 α ( r r w ) r 2 ,
where P 0 is the incidence power, h is the reservoir height, r is the radius, and r w is the wellbore radius. The attenuation factor is given by
α = ω μ 0 μ r ε 0 ε r 2 1 + σ μ ε 0 ε r 2 1 ,
where ω is the angular frequency, μ 0 and μ r are vacuum and relative magnetic permeabilities, ε 0 and ε r are vacuum and relative electric permittivities, and σ is the medium conductivity.

3. Numerical Methods

We solve the system of Equations (1) and (2) to simulate the heating of a 100 m × 100 m aquifer with five wells in a five-spot configuration.
The two cases having practical significance considered are (a) continuous water injection and EM heating, where EM heater is placed inside the vertical injection well located in the center of the aquifer, and (b) alternated water injection and EM heating; in this case, the wells are alternatively switched from injectors to producers and vice versa (see Figure 2). Both configurations were simulated using parameters described in Table 1, which were obtained from the literature [20,32,35,36].
We use an efficient numerical scheme based on the Galerkin least-squares finite element method (GLS) to solve Equations (1) and (2) [37]. GLS possesses the accuracy of the standard Galerkin method and the stability of the least-squares method. It consists of determining the equations’ variational formulation, approximating the solution by a different family of polynomial base functions when high convection occurs alongside low diffusion.
A staggered algorithm is employed to solve the system of partial differential equations by decoupling it into subsystems solved sequentially. This scheme is commonly used when equations express different physics; see [38].
Following this technique, we solve Equation (1) considering the injection well located at coordinate (0, 0) and the production well, located at coordinate (50, 50), at the opposite end following the aquifers’ diagonal. Both wells are maintained at a constant flux Q w , implemented as a constant flux Neumann boundary condition at injection/recovery wells and null flux Neumann at the rest of the boundary. As a result, we obtain the velocity field V. In the next step, we use V to solve Equation (2), considering the initial condition given by θ 0 with Dirichlet boundary condition equal to θ 0 in the injection well and null flux Neumann condition for the rest of the boundary. Here, we use the coefficients C w , C t , and K t given by Equations (3)–(5), and the volumetric heat dissipation W given by Equations (7) and (8). As a result, we obtain a temperature profile θ . The procedure restarts by using θ to solve Equation (1), at the next time step, through implicit Euler’s method.
The maximum relative error for the numerical method was obtained by comparing the energy balance between the power P of the EM emitter with the energy stored, recovered, and lost in the reservoir, obtained by the space-time integral of the first four terms of Equation (2). For a six-year time, the relative error of numerical simulations was estimated to be 1.2%.

4. Results and Discussion

Here, we present the results obtained in the two cases considered in Section 3. Our discussion will focus on the amount of energy stored in the aquifer and, especially for the second case, on the energy that can be recovered upon switching to the stored energy production. All simulations use parameter values described in Table 1.

4.1. Continuous Water Injection under EM Heating

Water at constant temperature of 35 °C is injected with constant Darcy’s velocity of 4 × 10 5 (m/s), which corresponds to the pressure gradient of approximately 0.07 MPa. Since the flow and temperature are symmetrical with respect to the central aquifer well, we present the reservoir temperature distribution only for the first quadrant in Figure 3. Figure 3a shows the 3D map of the obtained temperature distribution after 6 months’ continuous injection.
The temperature profiles in Figure 4a correspond to the diagonal section in Figure 3 for different times. Solid lines were obtained without thermal losses, while dashed lines were obtained with C l = 0.02 . As expected, after the heat wave moves towards the injection well, we notice a stabilization in the reservoir temperature over time. Considering small heat losses, the temperature profile shows deviation from the original one after three months, as can be observed in Figure 3a. The change in the temperature of the reservoir is stable during the energy storage cycle. As can be observed in Figure 3a, the traveling heat profile remains uniform over time with decreasing velocity. When thermal losses are considered, a small flattening of the heat front is observed. This temperature stability around the well allows a controlled and effective recovery in the subsequent cycle.
Temperature control is essential to avoid the formation of water vapor near the EM emitter as it has a lower thermal capacity and can cause overheating. The simulations show that the temperature rises considerably with decreasing the wellbore radius, and it decreases with the increase in water injection speed. Thus, controlling the emitter power and the injection flow rate is essential to maintain the temperature below 365.7 °C.

4.2. Energy Recovery

Following the case (b) from Section 3, we simulate the alternated water injection and EM heating in six-month cycles. The wells are alternatively switched from injectors to producers and vice versa. Figure 3b,d show the 3D temperature distribution inside the reservoir after 12 months (end of the first recovery cycle) and 72 months (end of the last recovery cycle). Figure 3c shows the 3D map of the obtained temperature distribution after 66 months at the end of the last storage cycle. The temperature profiles in Figure 4b,d correspond to the diagonal section of the reservoir part plotted in Figure 3b,d, respectively. The temperature profiles in Figure 4c correspond to the diagonal section in Figure 3c for different times during the last storage cycle. As before, these profiles were plotted neglecting thermal losses ( C l = 0 ), using thermal losses estimated for this reservoir in Section 2.2 ( C l = 0.02 ) and using double thermal losses ( C l = 0.04 ). The total energy balance for the six-year simulation is shown in Figure 5, where the time is measured in semesters. Electromagnetic waves emission is constant for six months during renewable energy storage and switched off for the next six months during the energy recovery process. The red curves show the energy recovered from the aquifer in semiannual cycles. The recovery is made during the six months in which the electromagnetic emitter is turned off, being interrupted during the electromagnetic emission period. For instance, in the first year, the total amount of energy injected into the reservoir is 287.71 GJ, while the energy recovered is 222.40 GJ. This implies that part of the energy remains in the reservoir ( C l 0 ) or is lost ( C l > 0 ). The corresponding energy balance is discussed in more detail in the next section.

4.3. Energy Balance

From Figure 5, one can see that there is a difference between emitted and recovered energy. In order to evaluate this difference, Table 2 shows the relative amounts of energy recovered, stored, and lost in one-year cycles corresponding to different values of thermal losses. The data indicate that the fraction of the recovered energy increases due to the reservoir heating up. This is clearly illustrated by the ideal case (without thermal losses) where the stored energy diminishes from 18.9% to 13.2% throughout the simulations (6 years). The effect of thermal losses leads to a reduction of the recovery energy but does affect the energy stored. For the heat losses upper bound, we consider the value C l = 0.04 , double the value estimated for the considered parameters. For such high thermal losses coefficient, we note that more than 69% of emitted energy can be recovered, indicating the potential efficiency of the proposed technology.
Notice that the energy required to heat the reservoir remains approximately the same, independent of the emitter power. Thus, the relative residual stored energy will be more considerable for the lower EM energy emissions, consequently diminishing the relative recovered energy.
The comparison between the proposed technique and a simple thermal energy injection rely on heat losses in the injection/production wells. In Section 4.4, we estimate the heat losses for different well sizes used in industry and different sealing materials. The estimated significant heat losses justify the EM heating technique.

4.4. Wellbore Heat Losses

Several works in the literature discuss the thermal losses in injection/production wells where water is preheated on the surface and later injected into reservoirs [39,40,41,42]. The authors concluded that thermal losses could be significant depending on the geometry and materials that make up the well and its surroundings.
The mathematical model based on Fourier’s law describes the heat flux Q H ( W ) in the injection/production wellbore, which, through the tubing surface and the external composition layers (steel, cement, and rock formation), is given by
Q H = 2 π r t o U t o ( θ h θ s ) Δ H ,
where r t o is the tubing outside radius, h is the depth, and θ s is the formation steady temperature. The overall heat transfer U t o and cement/formation interface temperature θ h are presented in [40] as
U t o = r t o ln ( r t o / r h ) k c 1 and θ h = θ w f ( t ) + θ s k s / ( r t o U t o ) f ( t ) + k s / ( r t o U t o ) ,
where r h is tubing internal radius, θ h is the cement/formation interface temperature, k c is cement thermal conductivity, k s is formation thermal conductivity, and f ( t ) = ln ( 2 ψ t / r c ) + 0.29 is transient heat conduction function, where ψ is thermal diffusivity of the Earth and t is time. Following this procedure, we estimated the thermal losses for different types of wellbore (see Table 3).
The estimation results suggest that thermal losses in the well can be significant when the tubing diameter increases and the thickness of the cement used to fix and insulate the wellbore is thinner. In addition, the thermal conductivity of the cement and the rock formation around the well contribute to the loss of energy storage efficiency. The estimated thermal losses could exceed 10% of the energy. The electromagnetic heating method is an alternative to avoid the thermal losses during the injection process; it allows thermal energy to be stored more effectively.

5. Conclusions

We investigated heat storage into deep high-temperature aquifers by numerical simulations validated by controlled laboratory experiments. The processes are reminiscent of high-temperature energy storage (HT-ATES). The following main conclusions can be drawn from this study:
  • The proposed EM heating technique could be used to produce a stable thermal front, spreading the energy through the aquifer with temperatures below the boiling point.
  • Our simulation results show that even considering excessive thermal losses, the EM-heating-assisted water flooding can recover up to 70% of the stored energy. In terms of efficiency, this value is also comparable with the low-temperature ATES as reported in the literature.
  • The energy balance estimate in the wellbore shows that down-hole EM heating can reduce energy losses for the deep aquifers (>1000 m), which can be notable if the water is heated at the surface.
In summary, the results discussed in this work present the use of EM-heating-assisted water flooding as a promising alternative for the storage of renewable energies in deep aquifers with high temperatures (HT-ATES).

Author Contributions

Conceptualization, G.C. and P.L.J.Z.; methodology, G.C. and S.O.d.A.; software, S.O.d.A.; validation, S.O.d.A., G.C. and P.L.J.Z.; formal analysis, S.O.d.A. and G.C.; investigation, S.O.d.A.; writing—original draft preparation, S.O.d.A., G.C. and P.L.J.Z.; writing—review and editing, S.O.d.A., G.C. and P.L.J.Z.; supervision, G.C. and P.L.J.Z. All authors have read and agreed to the published version of the manuscript.

Funding

G.C. was supported in part by CNPq grants 303245/2019-0, 405366/2021-3 and FAPEMIG grant APQ-00405-21.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank Thiago Quinelato for the assistance with the numerical algorithms.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study.

Appendix A. Model Validation

This validation procedure can be found in [28] (we present it here for the sake of completeness), where analytical solutions were presented and compared with laboratory experiments and numerical simulations. Analytical solutions for a similar system of equations describing the thermal-hydro coupling effect during energy extraction in geothermal reservoirs can be found in the literature (see [43,44]).
In order to validate the model presented in this paper, we retrace the laboratory-scale experiments of water injection in a porous medium under high-frequency EM heating [20]. The experiments were conducted in the presence of oil, but oil plays a passive role in comparing the model and experiments. A conventional EM microwave emitter, a conical EM wave guide, and the core-holder were placed inline in the Faraday cage to contain the radiation. The wave guide’s conical shape was specially designed to direct the emitted microwaves towards the inlet section of the core. The core structure was formed by a cylindrical case around the porous medium composed of Bentheimer Sandstone. A high-precision pump was used to inject water into the core. Fiber-optic temperature sensors were placed at a distance of 2, 5, 9, and 14 cm from the injection point.
From the set of experiments carried out, two were selected for the validation of our model. Table A1 summarizes all parameters of these experiments, which were also used in simulations. The only parameter that was not provided is the electromagnetic attenuation constant. Thus, using the data obtained by the temperature sensors and the relative data process, we can estimate its value as α = 2.8 1/m.
Table A1. Physical parameters used in modeling validation.
Table A1. Physical parameters used in modeling validation.
SymbolPhysical QuantityExp AExp BUnit (SI)
LReservoir length 0.17 0.17 [m]
ACross-section area 1.13 × 10 3 1.13 × 10 3 [m2]
θ 0 Initial temperature 308.15 308.15 [K]
α EM energy absorption 4.8 2.8 [1/m]
PEM emitters power4540[W]
DDarcy velocity0 2.8 × 10 4 [m/s]
C t System specific heat 2.253 5.058 [MJ/m3·K]
C w Total fluid specific heat 3.52 0.924 [MJ/m3·K]
K t Total system thermal conduc. 1.824 1.9216 [W/m·K]
C l Thermal losses coefficient 0.0 7.0 [MJ/m3·K]
Experiment A:
In this experiment, the porous medium is completely saturated with water at room temperature. After stopping the water injection, the electromagnetic wave emitter is turned on, and the optical sensors measure the temperature over time without water flooding. For numerical simulations, we disregard the convective term of Equation (1). The boundary conditions consider the Neumann condition null on both borders. For the initial condition, we consider water’s ambient temperature. The specific parameters used in numerical simulations are given in Table A1, and the results are shown in Figure A1.
Figure A1. Temperature profiles for Experiment A (EM heating with no water flooding). The continuous lines represent the model’s solution, and the dots represent data obtained by optical thermometers [20].
Figure A1. Temperature profiles for Experiment A (EM heating with no water flooding). The continuous lines represent the model’s solution, and the dots represent data obtained by optical thermometers [20].
Energies 15 03982 g0a1
Experiment B:
In this experiment, the porous medium is saturated with water and oil. Then, water is injected at a constant flow simultaneously with electromagnetic heating. For the numerical simulation, we considered the ambient water temperature as the initial condition. The boundary conditions were defined considering the Neumann condition null at the outlet and Dirichlet with a linear increase at the inlet. The specific parameters used in numerical simulations are given in Table A1, and the results are shown in Figure A2.
Notice that in Figure A1 there are differences between the experiment data and numerical simulations. The explanation, according to authors [20], is that the air inside the Faraday cage around the core begins to heat up quickly, according to Fourier’s law. They believe that this difference decreases according to the stabilization of the air temperature inside the Faraday cage with time. We believe that on the reservoir scale, especially for deep aquifers, this temperature variation is less significant.
Taking into account the experimental errors, one can observe that the proposed model captures the correct qualitative behavior observed in the experiments. This supports the overall approach adopted in our paper.
Figure A2. Temperature profiles for Experiment B (EM heating and flooding). The continuous lines represent the model’s solution, and the dots represent data obtained by optical thermometers [20].
Figure A2. Temperature profiles for Experiment B (EM heating and flooding). The continuous lines represent the model’s solution, and the dots represent data obtained by optical thermometers [20].
Energies 15 03982 g0a2

References

  1. Dorsey-Palmateer, R. Effects of wind power intermittency on generation and emissions. Electr. J. 2019, 32, 25–30. [Google Scholar] [CrossRef]
  2. Dickinson, J.S.; Buik, N.; Matthews, M.C.; Snijders, A. Aquifer thermal energy storage: Theoretical and operational analysis. Géotechnique 2009, 59, 249–260. [Google Scholar] [CrossRef]
  3. Sommer, W. Modelling and Monitoring of Aquifer Thermal Energy Storage: Impacts of Soil Heterogeneity, Thermal Interference and Bioremediation. Ph.D. Thesis, Wageningen University, Wageningen, The Netherlands, 2015. [Google Scholar]
  4. Kallesøe, A.; Vangkilde-Pedersen, T. HEATSTORE–Underground Thermal Energy Storage (UTES)-State of the Art, Example Cases and Lessons Learned. In Proceedings of the World Geothermal Congress, Reykjavik, Iceland, 24–27 October 2020. [Google Scholar]
  5. Fleuchaus, P.; Schüppler, S.; Bloemendal, M.; Guglielmetti, L.; Opel, O.; Blum, P. Risk analysis of High-Temperature Aquifer Thermal Energy Storage (HT-ATES). Renew. Sustain. Energy Rev. 2020, 133, 110153. [Google Scholar] [CrossRef]
  6. Kastner, O.; Norden, B.; Klapperer, S.; Park, S.; Urpi, L.; Cacace, M.; Blöcher, G. Thermal solar energy storage in Jurassic aquifers in Northeastern Germany: A simulation study. Renew. Energy 2017, 104, 290–306. [Google Scholar] [CrossRef]
  7. Schmidt, T.; Müller-Steinhagen, H. Die solar unterstützte Nahwärmeversorgung mit saisonalem Aquifer-Wärmespeicher in Rostock-Ergebnisse nach vier Betriebsjahren. In Symposium Erdgekoppelte Wärmepumpen; Geothermische Fachtagung: Landau, Germany, 2004. [Google Scholar]
  8. Kangas, M.; Lund, P. Modeling and simulation of aquifer storage energy systems. Sol. Energy 1994, 53, 237–247. [Google Scholar] [CrossRef]
  9. Ghaebi, H.; Bahadori, M.; Saidi, M. Performance analysis and parametric study of thermal energy storage in an aquifer coupled with a heat pump and solar collectors, for a residential complex in Tehran, Iran. Appl. Therm. Eng. 2014, 62, 156–170. [Google Scholar] [CrossRef]
  10. Badakhshan, S.; Hajibandeh, N.; Shafie-khah, M.; Catalão, J. Impact of solar energy on the integrated operation of electricity-gas grids. Energy 2019, 183, 844–853. [Google Scholar] [CrossRef]
  11. Lau, D.; Song, N.; Hall, C.; Jiang, Y.; Lim, S.; Perez-Wurfl, I.; Ouyang, Z.; Lennon, A. Hybrid solar energy harvesting and storage devices: The promises and challenges. Mater. Today Energy 2019, 13, 22–44. [Google Scholar] [CrossRef]
  12. Wilberforce, T.; Baroutaji, A.; El Hassan, Z.; Thompson, J.; Soudan, B.; Olabi, A.G. Prospects and challenges of concentrated solar photovoltaics and enhanced geothermal energy technologies. Sci. Total Environ. 2019, 659, 851–861. [Google Scholar] [CrossRef] [Green Version]
  13. Bera, A.; Babadagli, T. Status of electromagnetic heating for enhanced heavy oil/bitumen recovery and future prospects: A review. Appl. Energy 2015, 151, 206–226. [Google Scholar] [CrossRef]
  14. Shafiai, S.; Gohari, A. Conventional and electrical EOR review: The development trend of ultrasonic application in EOR. J. Pet. Explor. Prod. Technol. 2020, 10, 2923–2945. [Google Scholar] [CrossRef]
  15. Sivakumar, P.; Krishna, S.; Hari, S.; Vij, R. Electromagnetic heating, an eco-friendly method to enhance heavy oil production: A review of recent advancements. Environ. Technol. Innov. 2020, 20, 101100. [Google Scholar] [CrossRef]
  16. Hasanvand, M.Z.; Golparvar, A. A Critical Review of Improved Oil Recovery by Electromagnetic Heating. Pet. Sci. Technol. 2014, 32, 631–637. [Google Scholar] [CrossRef]
  17. Pizarro, J.; Trevisan, O. Electrical heating of oil reservoirs: Numerical simulation and field test results. J. Pet. Technol. 1990, 42, 1–320. [Google Scholar] [CrossRef]
  18. Sahni, A.; Kumar, M.; Knapp, R.B. Electromagnetic heating methods for heavy oil reservoirs. In Proceedings of the Western Regional Meeting, SPE/AAPG, Long Beach, CA, USA, 19–23 June 2000; pp. 19–23. [Google Scholar]
  19. Chhetri, A.; Islam, M. A critical review of electromagnetic heating for enhanced oil recovery. Pet. Sci. Technol. 2008, 26, 1619–1631. [Google Scholar] [CrossRef]
  20. Paz, P.; Hollmann, T.; Kermen, E.; Chapiro, G.; Slob, E.; Zitha, P. EM Heating-Stimulated Water Flooding for Medium–Heavy Oil Recovery. Transp. Porous Media 2017, 119, 57–75. [Google Scholar] [CrossRef] [Green Version]
  21. Cerutti, A.; Bandinelli, M.; Bientinesi, M.; Petarca, L.; De Simoni, M.; Manotti, M.; Maddinelli, G. A New Technique for Heavy Oil Recovery Based on Electromagnetic Heating: System Design and Numerical Modelling. Chem. Eng. Trans. 2013, 32, 1255–1260. [Google Scholar]
  22. Jha, A.K.; Joshi, N.; Singh, A. Applicability and Assessment of Micro-Wave Assisted Gravity Drainage (MWAGD) Applications in Mehsana Heavy Oil Field. In Proceedings of the SPE Heavy Oil Conference and Exhibition, Kuwait City, Kuwait, 12–14 December 2011; pp. 11–14. [Google Scholar]
  23. Mukhametshina, A.; Martynov, E. Electromagnetic Heating of Heavy Oil and Bitumen: A Review of Experimental Studies and Field Applications. J. Pet. Eng. 2013, 2013, 476519. [Google Scholar] [CrossRef]
  24. Eskandari, S.; Jalalalhosseini, S.M.; Mortezazadeh, E. Microwave Heating as an Enhanced Oil Recovery Method—Potentials and Effective Parameters. Energy Sources Part A Recovery Util. Environ. Eff. 2015, 37, 742–749. [Google Scholar] [CrossRef]
  25. Soliman, M. Approximate solutions for flow of oil heated using microwaves. J. Pet. Sci. Eng. 1997, 18, 93–100. [Google Scholar] [CrossRef]
  26. Rubinstein, L. The total heat losses in injection of hot liquid Into a stratum. Neft. Gaz 1959, 9, 41. [Google Scholar]
  27. Carrizales, M.A.; Lake, L.W.; Johns, R.T. Multiphase fluid flow simulation of heavy oil recovery by electromagnetic heating. In SPE IOR Symposium; OnePetro: Richardson, TX, USA, 2010. [Google Scholar]
  28. Almeida, S.O.; Zitha, P.L.J.; Chapiro, G. A method for analyzing electromagnetic heating assisted water flooding process for heavy oil recovery. Transp. Porous Media 2021. [Google Scholar] [CrossRef]
  29. Fanchi, J.R. Feasibility of reservoir heating by electromagnetic irradiation. In Proceedings of the SPE Annual Technical Conference and Exhibition, New Orleans, LA, USA, 23–26 September 1990. [Google Scholar]
  30. Darcy, H. Les Fontaines Publiques de la Ville de Dijon; Victor Dalmond: Paris, France, 1856. [Google Scholar]
  31. Whitaker, S. The equations of motion in porous media. Chem. Eng. Sci. 1966, 21, 291–300. [Google Scholar] [CrossRef]
  32. Viswanath, D.; Ghosh, T.; Prasad, D.; Dutt, N.; Rani, K. Viscosity of Liquids: Theory, Estimation, Experiment, and Data; Springer: Dordrecht, The Netherlands, 2007. [Google Scholar]
  33. Chen, Z.; Huan, G.; Ma, Y. Computational Methods for Multiphase Flows in Porous Media (Computational Science and Engineering 2); SIAM: Philadelphia, PA, USA, 2006. [Google Scholar]
  34. Kaviany, M. Principles of Heat Transfer in Porous Media; Springer: Berlin/Heidelberg, Germany, 1991. [Google Scholar]
  35. Sudiko, M. Elementes of Electromagnetics; Oxford University Press: New York, NY, USA, 2014. [Google Scholar]
  36. Stratton, J.A. Electromagnetic Theory; McGraw-Hill: New York, NY, USA, 1941. [Google Scholar]
  37. Hughes, T.; Franca, P.; Hulbert, G. A new finite element formulation for computational fluid dynamics: VIII. The galerkin/least-squares method for advective-diffusive equations. Comput. Methods Appl. Mech. Eng. 1989, 73, 173–189. [Google Scholar] [CrossRef]
  38. Settari, A.; Walters, D. Advances in Coupled Geomechanical and Reservoir Modeling with Applications to Reservoir Compaction. SPE J. 2001, 6, 334–342. [Google Scholar] [CrossRef]
  39. Ramey, H.J., Jr. Wellbore Heat Transmission. J. Pet. Technol. 1962, 14, 427–435. [Google Scholar] [CrossRef]
  40. Willhite, G. Over-All Heat Transfer Coefficients in Steam And Hot Water Injection Wells. J. Pet. Technol. 1967, 19, 607–615. [Google Scholar] [CrossRef]
  41. Cheng, W.L.; Han, B.B.; Nian, Y.L.; Wang, C.L. Study on wellbore heat loss during hot water with multiple fluids injection in offshore well. Appl. Therm. Eng. 2016, 95, 247–263. [Google Scholar] [CrossRef]
  42. Ameli, F.; Alashkar, A.; Hemmati-Sarapardeh, A. Thermal Recovery Processes; Elsevier: Amsterdam, The Netherlands, 2018; pp. 139–186. [Google Scholar]
  43. Fan, Z.; Parashar, R. Analytical Solutions for a Wellbore Subjected to a Non-isothermal Fluid Flux: Implications for Optimizing Injection Rates, Fracture Reactivation, and EGS Hydraulic Stimulation. Rock Mech. Rock Eng. 2019, 52, 4715–4729. [Google Scholar] [CrossRef]
  44. Fan, Z.; Parashar, R.; Jin, Z.H. Impact of convective cooling on pore pressure and stresses around a borehole subjected to a constant flux: Implications for hydraulic tests in an enhanced geothermal system reservoir. Interpretation 2020, 8, SG13–SG20. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the proposed scheme.
Figure 1. Schematic representation of the proposed scheme.
Energies 15 03982 g001
Figure 2. Schematic representation of the proposed energy storage through EM heating.
Figure 2. Schematic representation of the proposed energy storage through EM heating.
Energies 15 03982 g002
Figure 3. The temperature distribution inside the aquifer at different times: (a,b) correspond to the storage cycles; (c,d) correspond to the recovery cycles.
Figure 3. The temperature distribution inside the aquifer at different times: (a,b) correspond to the storage cycles; (c,d) correspond to the recovery cycles.
Energies 15 03982 g003
Figure 4. The temperature distribution along the diagonal section of the aquifer for different times. Storage cycles: (a,c). Recovery cycles: (b,d).
Figure 4. The temperature distribution along the diagonal section of the aquifer for different times. Storage cycles: (a,c). Recovery cycles: (b,d).
Energies 15 03982 g004aEnergies 15 03982 g004b
Figure 5. Comparison between the total emitted EM energy and recovered energy over time.
Figure 5. Comparison between the total emitted EM energy and recovered energy over time.
Energies 15 03982 g005
Table 1. Physical parameters used in calculations.
Table 1. Physical parameters used in calculations.
SymbolPhysical QuantityValueUnit (SI)
ϕ porosity0.220[-]
s w water saturation1[-]
ν w water viscosity 1.002 × 10 3 [Pa·s]
ρ s porous media density 2.650 × 10 3 [Kg/m3]
ρ w water density 0.997 × 10 3 [Kg/m3]
c s PM specific heat capacity 0.920 × 10 3 [J/Kg·K]
c w water specific heat capacity 4.2 × 10 3 [J/Kg·K]
k s PM thermal conductivity2.30[W/m·K]
k w water thermal conductivity0.58[W/m·K]
Lreservoir length and width100[m]
Q w flux rate 4 × 10 5 [m/s]
Across section1.7[m2]
hreservoir height5[m]
Ppower of the EM emitter18.5[kW]
θ 0 initial temperature308.15[K]
kpermeability700[mD]
a 1 constant Equation (1)−4.53[-]
a 2 constant Equation (1)−220.53[-]
a 3 constant Equation (1)149.39[-]
ε 0 vacuum electric permittivity 8.854 × 10 12 [F/m]
ε r relative electric permittivity81[F/m]
μ 0 vacuum magnetic permeability 4 π × 10 7 [H/m]
μ r relative magnetic permeability1[H/m]
σ medium conductivity0.02[S/m]
ω angular frequency 2 π f [rad/s]
α water EM energy absorption 4.185 × 10 1 [1/m]
C t system specific heat2.822[MJ/m3·K]
C w total fluid specific heat6.625[MJ/m3·K]
K t tot. system thermal conductivity1.9216[W/m·K]
C l thermal losses0.02[W/m3·K]
Table 2. Comparison of relative amounts of recovered, residual stored, and lost energy.
Table 2. Comparison of relative amounts of recovered, residual stored, and lost energy.
C l = 0 C l = 0.02 C l = 0.04
Yr.Rec.R. St.LostRec.R. St.LostRec.R. St.Lost
181.1%18.9%0%73.8%15.5%10.7%67.4%12.7%19.9%
283.2%16.8%0%75.2%12.6%12.2%68.3%9.6%22.1%
384.5%15.5%0%76.0%10.7%13.3%68.8%7.6%23.6%
485.5%14.5%0%76.6%9.2%14.2%69.2%6.2%24.6%
586.2%13.8%0%76.9%8.1%15.0%69.4%5.2%25.4%
686.8%13.2%0%77.3%7.2%15.5%69.5%4.5%26.0%
Table 3. Heat losses in the wellbore during 6 months (one injection cycle).
Table 3. Heat losses in the wellbore during 6 months (one injection cycle).
Diameters [in]Th. Conductivity [W/m.K]Depth [m]Heat Losses %
D tub D cas K s K c H HL
57 3/51.40.210001.47
7 3/59 3/41.40.210002.18
9 3/4202.80.420002.35
57 3/52.80.420005.90
9 3/4121.40.210008.85
9 3/4122.80.4200010.6
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Almeida, S.O.d.; Chapiro, G.; Zitha, P.L.J. Down-Hole Electromagnetic Heating of Deep Aquifers for Renewable Energy Storage. Energies 2022, 15, 3982. https://doi.org/10.3390/en15113982

AMA Style

Almeida SOd, Chapiro G, Zitha PLJ. Down-Hole Electromagnetic Heating of Deep Aquifers for Renewable Energy Storage. Energies. 2022; 15(11):3982. https://doi.org/10.3390/en15113982

Chicago/Turabian Style

Almeida, Samuel O. de, Grigori Chapiro, and Pacelli L. J. Zitha. 2022. "Down-Hole Electromagnetic Heating of Deep Aquifers for Renewable Energy Storage" Energies 15, no. 11: 3982. https://doi.org/10.3390/en15113982

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

Article Metrics

Back to TopTop