Next Article in Journal
Peak Shaving Analysis of Power Demand Response with Dual Uncertainty of Unit and Demand-Side Resources under Carbon Neutral Target
Previous Article in Journal
Optimizing the Power Usage of Anti-Sweat Heaters in Glass-Door Refrigerators According to the Dew Point
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Generating 3D Geothermal Maps in Catalonia, Spain Using a Hybrid Adaptive Multitask Deep Learning Procedure

by
Seyed Poorya Mirfallah Lialestani
1,*,
David Parcerisa
1,
Mahjoub Himi
2 and
Abbas Abbaszadeh Shahri
3
1
Department of Mining, Industrial and ICT Engineering, Universitat Politècnica de Catalunya, Av. Bases de Manresa 61-73, 08242 Manresa, Spain
2
Department of Mineralogy, Petrology and Applied Geology, University of Barcelona, 08007 Barcelona, Spain
3
Johan Lundberg AB, 754 50 Uppsala, Sweden
*
Author to whom correspondence should be addressed.
Energies 2022, 15(13), 4602; https://doi.org/10.3390/en15134602
Submission received: 25 May 2022 / Revised: 17 June 2022 / Accepted: 20 June 2022 / Published: 23 June 2022
(This article belongs to the Section H2: Geothermal)

Abstract

:
Mapping the subsurface temperatures can efficiently lead to identifying the geothermal distribution heat flow and potential hot spots at different depths. In this paper, an advanced adaptive multitask deep learning procedure for 3D spatial mapping of the subsurface temperature was proposed. As a result, predictive 3D spatial subsurface temperatures at different depths were successfully generated using geolocation of 494 exploratory boreholes data in Catalonia (Spain). To increase the accuracy of the achieved results, hybridization with a new modified firefly algorithm was carried out. Subsequently, uncertainty analysis using a novel automated ensemble deep learning approach for the predicted temperatures and generated spatial 3D maps were executed. Comparing the accuracy performances in terms of correct classification rate (CCR) and the area under the precision–recall curves for validation and whole datasets with at least 4.93% and 2.76% improvement indicated for superiority of the hybridized model. According to the results, the efficiency of the proposed hybrid multitask deep learning in 3D geothermal characterization to enhance the understanding and predictability of subsurface spatial distribution of temperatures is inferred. This implies that the applicability and cost effectiveness of the adaptive procedure in producing 3D high resolution depth dependent temperatures can lead to locate prospective geothermally hotspot active regions.

1. Introduction

Access to energy resources is a critical worldwide issue. From this point of view, renewable power is emerging as an innovative cost-effective and clean source of energy for the future. Accordingly, the renewable resources are increasingly displacing fossil fuels in the power sector because of their capability to reduce carbon emissions and other types of pollution. However, several marketed renewable resources (e.g., biomass and large hydroelectric dams) due to some problematic tradeoffs may not be suitable to the geoenvironmental concerns [1]. In the field of renewables, geothermal is always available to be tapped and thus in comparison with other resources provides a reliable source of power [2]. Moreover, the low carbon intensity than the solar photovoltaic makes it a useful tool against the advance of climate change. Furthermore, versatility of geothermal resources dedicates particularly useful for promoting economic diversification [3,4].
The view of the increasing trend of the consumed energy in Europe and subsequently worldwide dependency on the external sources attractively moves toward developing the use of renewable energies and, in particular, geothermal resources as an appropriate solution [5,6]. Referring to Chamorro et al., (2014) [7], Spain can significantly be benefited from the high capacity of the natural geothermal resources. This is because in the Iberian Peninsula, the temperature can reach high values at shallow depths. Accordingly, in recent years using shallow geothermal resources have depicted steady growth and popularity in the construction subsector, whereas in power generation due to associated problem with electricity, unequal progress was reported [8,9].
The temperature of geothermal resources originating from the continuously produced heat within the earth can be classified into high (T > 225 °C), intermediated (125–225 °C), and low (T < 125 °C) systems [10]. To achieve this source of energy, the knowledge on subsurface geological hot spots for more precise drilling to access the target temperature is of great importance [11]. Therefore, to assess the well productivity, several parameters such as morphology and temperature of the ground, geothermal gradient, porosity, permeability, fluid salinity, thermal conductivity, and specific heat capacity at various depths play a critical role [12]. However, all this information often is not accessible [13] and also continuous or periodically monitoring the equipped test boreholes with temperature sensors is a time-consuming and costly procedure [4]. This is the reason why the geoengineering characteristics depend on the scale of the project are often approximated with different techniques and computer modelling approaches [14,15,16,17]. Producing 3D conceptual shallow geothermal potentials [18,19], utilizing GIS [20,21,22], spatial data analysis [23,24,25], applying numerical technique [26,27,28,29], integrated of different geophysical prospecting techniques such as magnetotelluric [30,31,32,33,34], gravity [33,35], seismic [31,33], and electrical resistivity [31,36,37], as well as evident geological characteristics [38,39] are some of the carried efforts in Spain, Chile, Pakistan, Iran, India, Nigeria, Indonesia, Denmark, China, Thailand, Italy, Taiwan, Finland and Japan. However, simulating the geothermal resources using numerical techniques due to complexity of the model preparation (natural state properties of the rocks and geothermal system), description of the realistic problem and evaluation of the results as well as inability in providing any insight into generalizations is a very time-consuming task that demands extensive experience. In case of GIS, learning curve can be long where the spatial relationships do not lead to absolute solutions and also the integration with traditional map is difficult [40,41,42]. Such drawbacks in generating high resolution 3D maps due to dedicating more flexibility but higher computational and analytical capabilities can be handled by artificial intelligence (AI) approaches. The skilled AI-based models in geoengineering problems were used to tackle the difficulties in handling the big data [43] and provide physically meaningful relationships within geo-data [44,45,46]. Accordingly, applicability of the AI techniques in the form of artificial neural network, machine/deep learning, evolutionary algorithms, and hybrid structures in producing the predictive 3D subsurface models have been highlighted [46,47,48,49,50,51]. Due to characterized features in creating transferable solutions and learnability from high-level data attributes [52] the feasibility of AI techniques in geothermal modeling [53,54] and compared performance by field prospecting methods [55,56] have been notified in several studies dealing with predicting the location of hot spot structures [57,58,59,60], estimating the temperature distribution [61,62], and potential of geothermal production associated with geological data [63,64].
Referring to literature reviews, the AI models can be optimized through metaheuristic algorithms to find the most proper solutions relative to a set of prioritized criteria or constraints [65,66,67,68]. However, the optimizing performance of these algorithms due to analytical shortcomings cannot be guaranteed [65,69]. This limitation of metaheuristic algorithms with growing interested in optimization methods can be covered using modification process to improve the performance and prevent from computational costs of a rework [69,70,71].
The firefly algorithm (FA) [72] is a developed swarm optimization method based on the attractiveness of fireflies, but suffers from slow convergence and getting trapped in local optimum for multimodal problems [73]. In this paper a new developed version of modified FA by Abbaszadeh Shahri et al., 2022a [69] is used.
In this point of view, Spain, despite the access to high potential of different types of geothermal resources is still far from achieving a generalized utilizing of these renewable systems [74]. However, following the European initiatives, based on the 2011–2020 renewable energy plan, Spain is also working on promising geothermal areas for both power production and direct applications [75]. Due to the lack of worldwide characterized information of subsurface geothermal hotspots [8], developing the modelling techniques for evaluation, and utilizing these resources critically should be considered as one of the most prolific tasks for contributing to the global sustainability [18,76,77,78]. Moreover, the information of subsurface soils often is acquired from the vertical sparse exploratory boreholes in which laterally spatial distribution of predicted parameter is a difficult task [46,47,49]. This implies that evaluating the models aiming to identify the possible hotspots can lead to utilize and decrease the available gaps of this resource with respect to other European nations and would allow Spain to reduce its foreign energy dependency. Such demands crucially motive for developing new computational modelling techniques leading to new frameworks for the future expansion of this energy.
Despite the demonstrated more precise results in hybridized AI techniques with the high-level metaheuristic optimization algorithms [69], no distinguished work dealing on developing multi-task predictive models through platform of intelligence systems in geothermal application is available. Multi-task learning is inherently a multi-objective problem because different tasks may conflict, necessitating a trade-off [79] and thus has received considerable attention in real-world problems. In such learning schemes multiple tasks are solved jointly and sharing inductive bias between tasks. Therefore, referring to the above-mentioned concerns and the progressively emerged interest in developing 3D digital models, an adaptive hybridized multi-task deep learning AI procedure for generating spatial subsurface temperature at different depths in Catalonia, Spain, was developed. Hybridizing process was carried out through a new version of modified firefly algorithm (MFA). The modelling procedure was executed through 496 numbers of sparse data comprising latitude, longitude, elevation, and surface temperature. Accordingly, the temperatures at depths 50 and 150 m were predicted, and the model then was setup to provide estimation for 120 and 180 m, respectively. The evaluated receiver operating characteristic (ROC) supplemented by error analysis showed the applicability and cost effectiveness of the adopted multitask procedure in producing high resolution depth dependent subsurface temperature maps for the study area. Subsequently, an uncertainty analysis using an introduced state-of-the-art approach for the predicted temperature and generated spatial maps were conducted.

2. Materials and Methods

2.1. Study Area and Acquired Datasets

Catalonia is a province of Spain located at the north-east of the Iberian Peninsula, in the western Mediterranean area with a diverse range of geomorphological features from mountainous to coastal landscapes [80,81]. This study is focused in an area of 7942 km2 in the northeastern part of Catalonia, as presented in Figure 1. The northern domain is a mountainous territory corresponding to the eastern part of the Pyrenees and the southern limb tie in with the Catalan Coastal Ranges that are parallel to the Mediterranean shoreline. The area is made up by a Paleozoic continental crust affected by extensive sedimentary basins during Mesozoic and Neogene and a compressive basin of Paleogene age coeval with the development of the Pyrenees and the Catalan Coastal Ranges [82,83].
Previous carried out research implied on ideal geothermal inquiry in several regions of this area such as Vallès, Empordà, La Selva, Penedès, Fossa d’Olot, Plana de Vic, Depressió Central (Lleida), and Maresme. Accordingly, in this study, a number of 496 stochastic datasets including longitude (X), latitude (Y), elevation (Z) and surface temperature (Ts) were derived from the database of the Catalonia Cartographic and Geological Institute (ICGC) [84,85]. The used database in the current paper was constructed from the compiled annually time-dependent information provided by ICGC for the year of 2021 at each measuring station or control point. The acquired datasets for further procedure were randomized and split into 65%, 20% and 15% to establish the training, testing, and validation sets and after compiling were normalized within [0, 1] interval to increase the learning speed and model stability.

2.2. Deep Neural Learning Structure

Multitask learning can increase the generalization of deep neural learning structures (DNLS) by employing domain information contained in the training inputs of related tasks as an inductive bias. In comparison to training the models independently, this learning procedure aims to tackle multiple tasks at the same time to enhance efficiency and prediction accuracy [79,86]. This implies on performing parallel learning with a shared representation, in which each task can help other tasks be learned better [87,88,89].
As shown in Figure 2, DNLS is a subcategory of AI techniques with multiple layers in which without being explicitly can scan the data to search combinable features for faster learning. This ability indicates for ameliorate feature extraction over machine learning and thus supreme performance of DNLS with unstructured data (e.g., texts, pictures, pdf, etc.) as well as exploring new complex attributes that humans might miss [90].
In this structure, the received input signals (x1, x2, …, xn) are passed to the hidden layers and then to the next senior level using adopted weights (wij, wjk):
I n p u t x 1 ,   x n   w i j   H i d d e n   l a y e r s     w j k o j z O u t p u t             Y
Referring to Figure 2, the output of the jth neuron in the zth hidden layer at the tth iteration, o j z ( t ) , subjected to activation function, f j z , is defined as:
o j z ( t ) = f j z ( i = 1 n z 1 w i j z x i z 1 ( t ) + b j z )             ( j n z )  
where nz shows the number of neurons in the zth layer, and b j z denotes the bias which shifts the summed signals received from the neuron.
The outcome of the lth neuron in the mth output layer after t iteration ( Y l ( t ) ), is then calculated using the updated weight by:
Y l ( t ) = i = 1 n m 1 w i j m x i m 1 ( t )           ( l n o )
where no represents the number of neurons in the output layer.
Accordingly, the error of model for entire network (E (t)) and predicted values (e (t)) are expressed as:
E ( t ) = 1 2 j = 1 n ( d j ( t ) o j k ( t ) ) 2  
e ( t ) = y Y ( t )
where dj (t) denotes the desired output of neuron j at the tth iteration and y shows the actual output.
In each iteration, then the weights are updated through the learning rate to minimize the prediction error of using:
Δ w ( t ) = η w ( E ( t ) w o l d )
where η w represent the learning rate.

2.3. The Used Modified FA (MFA)

The original FA [72] has been formulated using the attractiveness of fireflies, their brightness and adjacent distance. Table 1 shows the involved parameters in FA that depending on the problem should properly be tuned through trial-error procedure.
The concept of this algorithm is based on the I as the objective function in such way that a firefly is attracted by the brighter one (new solution). As a result of moving toward each other, the distance between the fireflies is updated. Accordingly, the position of the moved firefly after t iteration and implemented update step size of (xjxi) reflects a new solution ( x i n e w ) that further should be evaluated by the fitness function (FT) in the population by:
x i n e w = x i t + β 0 e γ r i j 2 ( x j x i ) + α ( r a n d 0.5 )
where i, j denotes the index of fireflies and the rand function corresponds a random number of solutions. Using this iterative process only one solution with the lowest FT will be kept [91]. Therefore, each new generated solution, x i n e w showing the updated location of firefly is governed by α and γ. Accordingly, the smaller γ, the faster convergence history but the greater β between fireflies. On the other hand, the larger α connote to increasing the range of random motion of fireflies and thereby the slower convergence. In the modified version presented by [69], Equation (7) was updated through the brightness expectation value, using:
x i m o d i f i e d = x i + β 0 e γ t r i j 2 ( x j x i ) + α t r i j ( r a n d 0.5 )
Accordingly, α and γ then adaptively are tuned and updated using the variance of the population of brightness (Var (I)) by:
{ γ t = γ 0 + e k . V a r ( I ) ( γ u γ 0 ) α t = α 0 e k . V a r ( I ) | ( α u α 0 ) | α 0 > α u , γ u > γ 0 , k > 0
where, subscripts 0 and u denotes the initial and ultimate values. Using this modification, the convergence speed and the computational time are boosted and strengthened [69].

3. Developing Hybrid Adaptive Procedure

The hybridizing procedure benefits from integration of different predictive AI model with optimization methods to capture more accurate output and thus higher performance. Therefore, adopting appropriate procedure to apprehend an optimum multitask topology for predictive unit is critical to avoid from over-fitting problem, early convergence and not getting stuck in local minima. Figure 3 depicts the block-based hybrid adaptive DNLS-MFA procedure subjected to iterative constructive technique that using several inner nested loops was programmed in C++. In this approach, 80% of randomized data is used for training. Through the adopted k-fold checking (here set for 10 times), the data then internally are re-randomized into 60% and 20% for training and testing. The remained 20% of unseen data was used for model validation. This iterate-based adaptive procedure was configured for leading to an optimum multitask DNSL with a proper adjustment of the internal characteristics and corresponding hyper parameters. To reduce the computational time, the number of maximum hidden layers and used neurons were limited to 3 and 30, respectively. The learning rate also was managed within [0.001, 1.000] interval with a step size of 0.05. The first priority to terminate the training process is to achieve the minimum network root mean square error (RMSE) and if not satisfied then the number of iterations (t) will active as set for 500 in this study.
Figure 4a shows the results of series dynamically monitored and analyzed architectures depend upon the implemented training algorithm (TA), learning rate and activation function (AF). According to Figure 4b, the DNLS topology with structure of 4-15-5-2 trained by limited memory quasi-Newton (LMQN) subjected to hyperbolic tangent (Hyt) activation function can be selected as optimal. In the given topology, 4 denotes the number of used inputs (X, Y, Z, Ts), 15 and 5 express the number of managed neurons in first and second hidden layers and 2 is assigned to the multiple outputs comprising the temperature at depths 50, 150 m (T50, T150), respectively.
As presented in Figure 3, the selected optimum topology (Figure 4b) is hybridized with the MFA. Referring to Table 1 and Equations (7)–(9), the updated location of firefly should be tuned using α, β and γ. However, appropriate setting of these parameters is not easy task. The most representative values of α, β and γ then were captured through parametric regularizing procedure. Figure 5a shows the result of a series of analyses subjected to RMSE with two fixed parameters and one variable. The results showed that the values of 1, 0.5, 0.2, 0.5 corresponding to γ, β, α, β0 (Figure 5a) and 30 for population of fireflies (Figure 5b) can be the optimal. Accordingly, the predictability of the model using adaptive DNLS (Figure 5c,d) and the hybrid DNLS-MFA (Figure 5e,f) were presented and compared.
Traditional 3D modelling techniques are subjective and limited to the knowledge and experiences of experts in the selection of assumptions and parameters. For subsurface geothermal purposes, therefore, creating 3D depth-dependent digitized spatial predictive maps using identified hybrid optimum DNLS model play an important priority. As a result, for the study area, the created 3D maps using the hybrid DNLS-MFA at the surface and depth of 50 and 150 m are presented (Figure 6a–c). The applicability of the DNLS-MFA then was examined using unlabelled data for two other depths. Accordingly, the elevation was replaced by converted depths for 120 and 180 m with respect to the ground surface elevation with the same used input data. The model then was trained for the whole datasets through the saved weight database and the visualized results were reflected in Figure 6d,e. According to the documented results by Colmenar-Santos et al. (2016) [92] from the report of Institute for the Diversification and Saving of Energy (IDEA), the range of predicted subsurface temperatures show appropriate compatibilities. The predicted subsurface using the applied model at depth of 150 m dedicated about 33 °C, where considering the nonlinearity and heterogeneity it is expected to fall around 70 at 300 m depth. Therefore, referring to [92], the subsurface temperature in the study area corresponding to the depths between 300–2500 m will fall within the interval of [70 °C, 140 °C].

4. Discussion and Validation

Generally, the statistical error metrics are used to evaluate the performance of the complex models through pair-wise-matched of observations and predictions [93]. In this study, as presented in Table 2 the models were assessed by means of the mean absolute percentage error (MAPE), RMSE, index of agreement (IA) [94] and R2. The better performance was selected by higher IA, R2 accompanying with lower MAPE and RMSE. Table 2 reflects the comparison of the applied metrics for the 15% of unseen randomized validation datasets through the weight database.
Using confusion matrix [95], the predictability performance of developed models is analyzed, where each diagonal entry of [aij] as the number of records for ith and jth category of target and corresponding network output would be non-zero. In this study, the multiclass confusion matrixes for optimum and hybridized DNLS models using validation datasets were established (Table 3). The cells with value of 0 shows there is no predicted result correspond to target output. This can be an indicator of appropriate predictability of developed models. Accordingly, the compared qualitative characteristics in classification tasks using correct classification rate (CCR) (Table 4) showed at least 4.2% improvement in the predictability level of hybrid DNLS than the optimum topology.
The presented confusion matrix (Table 3) then was used to plot the area under the curve of receiver operating characteristics (AUCROC). The ROC is a 2D graphical probability approach to assess the overall performance of a model, where the greater AUC the higher capability in diagnosed classes [96,97]. Precision shows the true predictions in each class and recall reflects the ability of model in identifying the actual positives. Therefore, the predictability of a model in a full picture at different thresholds can be quantified using the precision–recall ROC curves. Referring to Figure 7, the achieved accuracies of 90.7% and 86.3% using hybrid DNLS-MFA for T50 and T150 demonstrated more reliable outputs than optimum DNLS.
In computational systems the uncertainty analysis for the new individual upcoming observations can be estimated using prediction interval (PI) [98]. One of the key benefits of PI is referred to dedicating a range of weights allowing for insight into how accurate the predicted weight is likely to be [99]. The PI is always wider than a confidence interval because it considers the uncertainty in both predicting the population mean and the random variation of the individual values [100]. In this study, the PI of the hybrid DNLS-MFA at the level of 95% was estimated using a state-of-the-art automated random deactivating weight approach [101], as shown in Figure 8. The applied approach [101] subjected to hybrid DNLS-MFA for T50 and T150 showed narrower distribution of PI, leading to the identification of the potential hot spot sub-areas with lower uncertainty.

5. Conclusions

Mapping the subsurface temperature in complex structural settings using advanced DNLS can dedicate a viable tool for exploration analyses and thus drilling phases. Refer to this motivation, an adaptive hybrid multitask DNSL-MFA procedure for producing the 3D predictive spatial maps of the temperature-at-depths was developed and applied on 494 geo-location and surface temperature datasets in a part of Catalonia, Spain. Compared with DNLS, the predicted maps at depths of 50 and 150 m using the introduced hybrid DNSL-MFA scheme showed 90.7% and 86.3% accuracy performances leading to 6.24% and 4.93% improvements, respectively. The evaluations then were supplemented by error criteria and uncertainty analysis, where the narrower distribution of PI in hybrid DNLS-MFA showed higher reliability in identifying the most compatible potentials in the study area. Accordingly, the predictability of the hybrid model was then examined with unlabelled data at depths 120 and 180 m, in which the created 3D maps showed appropriate responses to categorized subsurface temperatures. Using the generated 3D visualized maps not only the spatial extension of the interested areas at investigated depth are recognized but also facilitate the interpretability of subsurface heating conditions and, consequently, reducing the geothermal exploration costs. This implies on applicability of produced maps in reflecting the most favorable hotspot locations for further geothermal analyses and thus better site characterizing for field and real-world applications. The presented hybrid DNLS-MFA is preferred because the produced digitized 3D spatial maps can be updated using new geoscientific data to improve serving for deeper insights into the geothermal resource systems. However, more knowledge on potential subsurface hot spots is needed to enhance geothermal energy use and public awareness.

Author Contributions

Conceptualization, S.P.M.L., D.P. and M.H.; methodology, S.P.M.L. and A.A.S.; code development and visualization, S.P.M.L. and A.A.S.; validation, S.P.M.L. and A.A.S.; formal analysis, S.P.M.L., D.P., M.H. and A.A.S.; data curation, S.P.M.L.; writing—original draft preparation, S.P.M.L.; writing—review by A.A.S., D.P. and M.H.; thoroughly was edited by native expert. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest. There is no funder role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Lund, J.W. Characteristics, development and utilization of geothermal resources. Geo-Heat Cent. Q. Bull. Klamath Falls Inst. Technol. 2007, 28, 140–147. [Google Scholar]
  2. Tester, J.W.; Anderson, B.J.; Batchelor, B.J.; Blackwell, D.D.; DiPippo, R.; Drake, E.M.; Garnish, J.; Livesay, B.; Moore, M.C.; Nichols, K.; et al. The Future of Geothermal Energy. In Impact of Enhanced Geothermal Systems (EGS) on the United States in the 21st Century; Renewable Energy and Power Department, Idaho National Laboratory: Idaho Falls, ID, USA, 2006. Available online: http://geothermal.inel.gov (accessed on 24 May 2022).
  3. McLarty, L.; Reed, M.J. The US geothermal industry: Three decades of growth. Energy Sources 1992, 14, 443–455. [Google Scholar] [CrossRef]
  4. Lesmana, A.; Winofa, N.C.; Pratama, H.B.; Ashat, A.; Saptadji, N.M. Preliminary financial modelling with probabilistic approach for geothermal development project in Indonesia. IOP Conf. Ser. Earth Environ. Sci. 2020, 417, 012024. [Google Scholar] [CrossRef]
  5. Antics, M.; Sanner, B. Status of geothermal energy use and resources in Europe. In Proceedings of the European Geothermal Congress, Unterhaching, Germany, 30 May–1 June 2007. [Google Scholar]
  6. Lund, J.W.; Toth, A.N. Direct utilization of geothermal energy 2020 worldwide review. In Proceedings of the World Geothermal Congress, Reykjavik, Iceland, 24–27 October 2021. [Google Scholar]
  7. Chamorro, C.R.; Garcia-Cuesta, J.L.; Modejar, M.E.; Linares, M.M. An estimation of the enhanced geothermal system potential for the Iberian Peninsula. Renew. Energy 2014, 66, 1–14. [Google Scholar] [CrossRef]
  8. Sanchez-Guzman, J.; García de la Noceda, C. The evolution of geothermal energy in Spain-country update (2005–2009). In Proceedings of the World Geothermal Congress, Bali, Indonesia, 25–29 April 2010. [Google Scholar]
  9. Arrizabalaga, I.; De Gregorio, M.; De Santiago, C.; García de la Noceda, C.; Pérez, P.; Urchueguía, J.F. Geothermal energy use, country update for Spain. In Proceedings of the European Geothermal Congress, Den Haag, The Netherlands, 11–14 June 2019. [Google Scholar]
  10. Hochstein, M.P.; Brown, P.R.L. Surface manifestations of geothermal systems with volcanic heat sources. In Encyclopedia of Volcanoes; Sigurdsson, H., Ed.; Academic Press: San Diego, CA, USA, 2000; pp. 835–855. [Google Scholar]
  11. Montanari, D.; Lupi, M.; Calcagno, P. Geothermal syatems: Interdisciplinary approach for an effective exploration. Geofluids 2019, 2019, 9895659. [Google Scholar] [CrossRef] [Green Version]
  12. Omodeo-Salé, S.; Eruteya, O.E.; Cassola, T.; Baniasad, A.; Moscariello, A. A basin thermal modelling approach to mitigate geothermal energy exploration risks: The St. Gallen case study (eastern Switzerland). Geothermics 2020, 87, 101876. [Google Scholar] [CrossRef]
  13. Matsumoto, M. An approach for estimating geothermal reservoir productivity under access limitations associated with snowy and mountainous prospects. In Geothermal Energy; Ismail, B.I., Ed.; IntechOpen Press: London, UK, 2021. [Google Scholar] [CrossRef]
  14. Hochstein, M.P. Assessment and modelling of geothermal reservoirs (small utilization schemes). Geothermics 1998, 17, 15–49. [Google Scholar] [CrossRef]
  15. Ilyasov, M.; Ostermann, I.; Punzi, A. Modeling Deep Geothermal Reservoirs: Recent Advances and Future Problems. In Handbook of Geomathematics; Freeden, W., Nashed, M.Z., Sonar, T., Eds.; Springer: Berlin/Heidelberg, Germany, 2010. [Google Scholar] [CrossRef]
  16. Gumming, W. A conceptual model approach to the geophysical exploration of permeable geothermal reservoirs that considers context and uncertainty. SEG Tech. Program Expand. Abstr. 2009, 4326–4330. [Google Scholar] [CrossRef]
  17. Gao, J.; Shi, Q. A new mathematical modeling approach for thermal exploration efficiency under different geothermal well layout conditions. Sci. Rep. 2021, 11, 22930. [Google Scholar] [CrossRef]
  18. Zhu, Z.; Lei, X.; Xu, N.; Shao, D.; Jiang, X.; Wu, X. Integration of 3D geological modeling and geothermal field analysis for the evaluation of geothermal reserves in the northwest of Beijing Plain, China. Water 2020, 12, 638. [Google Scholar] [CrossRef] [Green Version]
  19. Fulignati, P.; Marianelli, P.; Sbrana, A.; Ciani, V. 3D geothermal modelling of the Mount Amiata Hydrothermal system in Italy. Energies 2014, 7, 7434–7453. [Google Scholar] [CrossRef] [Green Version]
  20. Noorollahi, Y.; Itoi, R.; Fujii, H.; Tanaka, T. GIS model for geothermal resource exploration in Akita and Iwate prefectures, northern Japan. Comput. Geosci. 2007, 33, 1008–1021. [Google Scholar] [CrossRef]
  21. Schiel, K.; Baume, O.; Caruso, G.; Leopold, U. GIS-based modelling of shallow geothermal energy potential for CO2 emission mitigation in urban areas. Renew. Energy 2016, 86, 1023–1036. [Google Scholar] [CrossRef]
  22. Meng, F.; Liang, X.; Xiao, C.; Wang, G. Geothermal resource potential assessment utilizing GIS-based multi criteria decision analysis method. Geothermics 2021, 89, 101969. [Google Scholar] [CrossRef]
  23. Tende, A.W.; Aminu, M.D.; Gajere, J.N. A spatial analysis for geothermal energy exploration using bivariate predictive modelling. Scentific Rep. 2021, 11, 19755. [Google Scholar] [CrossRef]
  24. Moghaddam, M.K.; Noorollahi, Y.; Samadzadegan, F.; Sharifi, M.A.; Itoi, R. Spatial data analysis for exploration of regional scale geothermal resources. J. Volcanol. Geotherm. Res. 2013, 266, 69–83. [Google Scholar] [CrossRef]
  25. Carranza, E.J.M.; Wibowo, H.; Barritt, S.D.; Sumintadireja, P. Spatial data analysis and integration for regional-scale geothermal potential mapping, West Java, Indonesia. Geothermics 2008, 37, 267–299. [Google Scholar] [CrossRef]
  26. Makasis, N.; Kreitmair, M.J.; Bidarmaghz, A.; Farr, G.J.; Sceidegger, J.M.; Choudhary, R. Impact of simplifications on numerical modelling of the shallow subsurface at city-scale and implications for shallow geothermal potential. Sci. Total Environ. 2021, 791, 148236. [Google Scholar] [CrossRef]
  27. Pola, M.; Fabbri, P.; Piccinini, L.; Zampieri, D. Conceptual and numerical models of a tectonically-controlled geothermal system: A case study of the Euganean Geothermal System, Northern Italy. Cent. Eur. Geol. 2015, 58, 129–151. [Google Scholar] [CrossRef]
  28. Arola, T.; Korhonen, K.; Martinkauppi, A.; Leppäharju, N.; Hakala, P.; Ahonen, L.; Pashkovskii, M. Creating shallow geothermal potential maps for Finland. In Proceedings of the European Geothermal Congress, Den Haag, The Netherlands, 11–14 June 2019. [Google Scholar]
  29. Balling, N.; Poulsen, S.E.; Fuchs, S.; Mathiesen, A.; Bording, T.S.; Nielsen, S.B.; Nielsen, L.H. Development of a numerical 3D geothermal model for Denmark. In Proceedings of the European Geothermal Congress, Strasbourg, France, 19–24 September 2016. [Google Scholar]
  30. Yadav, K.; Shah, M.; Sircar, A. Application of magnetotelluric (MT) study for the identification of shallow and deep aquifers in Dholera geothermal region. Groundw. Sustain. Dev. 2021, 11, 100472. [Google Scholar] [CrossRef]
  31. Chiang, C.W.; Yang, Z.X.; Chen, C.C.; Yeh, E.C.; Chen, C.S.; Wang, C.Y. Potential geothermal structure inferred from the electrical resistivity and seismic reflection models in the western Ilan Plain, NE Taiwan. Geothermics 2021, 94, 102124. [Google Scholar] [CrossRef]
  32. Di Paolo, J.F.; Ledo, J.; Slezak, K.; Van Dorth, D.M.; Pérez, I.C.; Pérez, N. La Palma Island (Spain) geothermal system revealed by 3D magnetotelluric data inversion. Sci. Rep. 2020, 10, 18181. [Google Scholar] [CrossRef] [PubMed]
  33. Mitjanas, G.; Marcuello, A.; Ledo, J.; Macau, A.; Alías, G.; Queralt, P.; Bellmunt, F.; Rivero, L.I.; Gabas, A.; Benjumea, B.; et al. Integrated seismic ambient noise, magnetotellurics and gravity data for the 2D interpretation of the Vall`es basin structure in the geothermal system of La Garriga-Samalús (NE Spain). Geothermics 2021, 93, 102067. [Google Scholar] [CrossRef]
  34. Amatyakul, P.; Prommakorn, N.; Wood, S.H.; Arunwan, T.R.; Vachiratienchai, C.; Chanapiwat, P.; Siripunvaraporn, W. An assessment of a shallow geothermal reservoir of Mae Chan hot spring, northern Thailand via magnetotelluric surveys. Geothermics 2021, 95, 102137. [Google Scholar] [CrossRef]
  35. Zaher, M.A.; Saibi, H.; Mansour, K.; Khalil, A.; Soliman, M. Geothermal exploration using airborne gravity and magnetic data at Siwa Oasis, Western Desert, Egypt. Renew. Sustain. Energy Rev. 2018, 82, 3824–3832. [Google Scholar] [CrossRef]
  36. Manzella, A.; Viezzoli, A.; Menghini, A.; Santilano, A.; Donato, A.; Montanari, D.; Maggi, S.; Rizzo, E. Shallow geothermal exploration using SkyTEM data: The VIGOR experiment. In Proceedings of the European Geothermal Congress EGC, Pisa, Italy, 3–7 June 2013. [Google Scholar]
  37. Nieto, I.M.; Martín, A.F.; García, P.C.; Blázquez, C.S.; Aguilera, D.G.; García, J.C. Geophysical Prospecting for Geothermal Resources in the South of the Duero Basin (Spain). Energies 2020, 13, 5397. [Google Scholar] [CrossRef]
  38. Navarro, A.; Carulla, N. Evaluation of geothermal potential in the vicinity of the flooded Sierra Almagrera Mines (Almeria, SE Spain). Mine Water Environ. 2018, 37, 137–150. [Google Scholar] [CrossRef]
  39. Daniele, L.; Taucare, M.; Viguier, B.; Arancibia, G.; Aravena, D.; Roquer, T.; Sepúlveda, J.; Molina, E.; Delgado, A.; Muñoz, M.; et al. Exploring the shallow geothermal resources in the Chilean Southern Volcanic Zone: Insight from the Liquiñe thermal springs. J. Geochem. Explor. 2020, 218, 106611. [Google Scholar] [CrossRef]
  40. Mallet, J.L. Discrete smooth interpolation in geometric modeling. Comp. Comput.-Aided Des. 1992, 24, 178–191. [Google Scholar] [CrossRef]
  41. Zhu, J.; Wright, G.; Wang, J.; Wang, X. A critical review of the integration of geographic information system and building information modelling at the data level. ISPRS Int. J. Geo-Inf. 2018, 7, 66. [Google Scholar] [CrossRef] [Green Version]
  42. Choi, Y.; Baek, J.; Park, S. Review of GIS-based applications for mining: Planning, operation, and environmental management. Appl. Sci. 2020, 10, 2266. [Google Scholar] [CrossRef] [Green Version]
  43. Chen, X.W.; Lin, X. Big data deep learning: Challenges and perspectives. IEEE Access 2014, 2, 514–525. [Google Scholar] [CrossRef]
  44. Karpatne, A.; Ebert-Uphoff, I.; Ravela, S.; Babaie, H.A.; Kumar, V. Machine learning for the geosciences: Challenges and opportunities. IEEE Trans. Knowl. Data Eng. 2018, 31, 1544–1554. [Google Scholar] [CrossRef] [Green Version]
  45. Spina, R. Big data and artificial intelligence analytics in geosciences: Promises and potential. GSA Today Archive. Geol. Soc. Am. 2019, 29, 42–43. [Google Scholar] [CrossRef]
  46. Abbaszadeh Shahri, A.; Shan, C.; Zall, E.; Larsson, S. Spatial distribution modeling of subsurface bedrock using a developed automated intelligence deep learning procedure: A case study in Sweden. J. Rock Mech. Geotech. Eng. 2021, 13, 1300–1310. [Google Scholar] [CrossRef]
  47. Ghaderi, A.; Abbaszadeh Shahri, A.; Larsson, S. An artificial neural network-based model to predict spatial soil type distribution using piezocone penetration test data (CPTu). Bull. Eng. Geol. Environ. 2019, 78, 4579–4588. [Google Scholar] [CrossRef] [Green Version]
  48. Abbaszadeh Shahri, A.; Larsson, S.; Renkel, C. Artificial intelligence models to generate visualized bedrock level: A case study in Sweden. Modeling Earth Syst. Environ. 2020, 6, 1509–1528. [Google Scholar] [CrossRef]
  49. Abbaszadeh Shahri, A.; Kheiri, A.; Hamzeh, A. Subsurface topographic modeling using geospatial and data driven algorithm. ISPRS Int. J. Geo-Inf. 2021, 10, 341. [Google Scholar] [CrossRef]
  50. Nguyen, X.C.; Nguyen, T.T.H.; Le, Q.V.; Le, P.C.; Srivastav, A.L.; Pham, Q.B.; Nguyen, P.M.; La, D.D.; Rene, E.R.; Ngo, H.H.; et al. Developing a new approach for design support of subsurface constructed wetland using machine learning algorithms. J. Environ. Manag. 2022, 301, 113868. [Google Scholar] [CrossRef]
  51. Li, H.; Misra, S. Robust machine-learning workflow for subsurface geomechanical characterization and comparison against popular empirical correlations. Expert Syst. Appl. 2021, 177, 114942. [Google Scholar] [CrossRef]
  52. Schmidhuber, J. Deep learning in neural networks: An overview. Neural Netw. 2015, 61, 85–117. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Harp, D.R.; O’Malley, D.; Yan, B.; Pawar, R. On the feasibility of using physics-informed machine learning for underground reservoir pressure management. Expert Syst. Appl. 2021, 178, 115006. [Google Scholar] [CrossRef]
  54. Chen, Y.; Zhang, D. Physics-constrained deep learning of geomechanical logs. IEEE Trans. Geosci. Remote Sens. 2020, 58, 5932–5943. [Google Scholar] [CrossRef]
  55. Duplyakin, D.; Beckers, K.; Siler, D.; Martin, M.J.; Johnston, H. Modeling subsurface performance of a geothermal reservoir using machine learning. Energies 2022, 15, 967. [Google Scholar] [CrossRef]
  56. Witter, J.B.; Trainor-Guitton, W.J.; Siler, D.L. Uncertainty and risk evaluation during the exploration stage of geothermal development: A review. Geothermics 2019, 78, 233–242. [Google Scholar] [CrossRef]
  57. Singh, U.K.; Tiwarib, R.K.; Singh, S.B. Neural network modeling and prediction of resistivity structures using VES Schlumberger data over a geothermal area. Comput. Geosci. 2013, 52, 246–257. [Google Scholar] [CrossRef]
  58. Yadav, A.; Yadav, K.; Sircar, A. Feedforward neural network for joint inversion of geophysical data to identify geothermal sweet spots in Gandhar, Gujarat, India. Energy Geosci. 2021, 2, 189–200. [Google Scholar] [CrossRef]
  59. Gudmundsdottir, H.; Horne, R.N. Prediction modeling for geothermal reservoirs using deep learning. In Proceedings of the 45th Workshop on Geothermal Reservoir Engineering, SGP-TR-216, Stanford, CA, USA, 10–12 February 2020. [Google Scholar]
  60. Kaftana, I.; Salka, M.; Senolb, Y. Evaluation of gravity data by using artificial neural networks case study: Seferihisar geothermal area (Western Turkey). J. Appl. Geophys. 2011, 75, 711–718. [Google Scholar] [CrossRef]
  61. Spichak, V. Estimating temperature distributions in geothermal areas using a neuronet approach. Geothermics 2006, 35, 181–197. [Google Scholar] [CrossRef] [Green Version]
  62. Shahdi, A.; Lee, S.; Karpatne, A.; Nojabaei, B. Exploratory analysis of machine learning methods in predicting subsurface temperature and geothermal gradient of Northeastern United States. Geotherm. Energy 2021, 9, 18. [Google Scholar] [CrossRef]
  63. Siler, D.L.; Pepin, J.D.; Vesselinov, V.V.; Mudunuru, M.K.; Ahmed, B. Machine learning to identify geologic factors associated with production in geothermal fields: A case-study using 3D geologic data, Brady geothermal field, Nevada. Geotherm. Energy 2021, 9, 17. [Google Scholar] [CrossRef]
  64. Bourhis, P.; Cousin, B.; Loria, A.F.R.; Laloui, L. Machine learning enhancement of thermal response tests for geothermal potential evaluations at site and regional scales. Geothermics 2021, 95, 102132. [Google Scholar] [CrossRef]
  65. Yang, X.S. Metaheuristic optimization: Algorithm analysis and open problems. In Experimental Algorithms, SEA; Pardalos, P.M., Rebennack, S., Eds.; Lecture Notes in Computer Science; Springer: Berlin/Heidelberg, Germany, 2011; Volume 6630. [Google Scholar] [CrossRef] [Green Version]
  66. Mansouri, A.; Aminnejad, B.; Ahmadi, H. Introducing modified version of penguins search optimization algorithm (PeSOA) and its application in optimal operation of reservoir systems. Water Supply 2018, 18, 1484–1496. [Google Scholar] [CrossRef]
  67. Asheghi, R.; Abbaszadeh Shahri, A.; Khorsand Zak, M. Prediction of uniaxial compressive strength of different quarried rocks using metaheuristic algorithm. Arab. J. Sci. Eng. 2019, 44, 8645–8659. [Google Scholar] [CrossRef]
  68. Oliveira, P.M.; Pires, E.J.S.; Boaventura-Cunha, J.; Pinho, T.M. Review of nature and biologically inspired metaheuristics for greenhouse environment control. Trans. Inst. Meas. Control 2020, 42, 2338–2358. [Google Scholar] [CrossRef]
  69. Abbaszadeh Shahri, A.; Khorsand Zak, M.; Abbaszadeh Shahri, H. A modified firefly algorithm applying on multi-objective radial-based function for blasting. Neural Comput. Appl. 2022, 34, 2455–2471. [Google Scholar] [CrossRef]
  70. Shi, Y.; Eberhart, R. A modified particle swarm optimizer. In Proceedings of the IEEE, International Conference on Evolutionary Computation Proceedings, IEEE World Congress on Computational Intelligence (Cat. No. 98TH8360), Anchorage, AK, USA, 4–9 May 1998. [Google Scholar] [CrossRef]
  71. Faritha, B.A.; Chandrasekar, C. An optimized approach of modified bat algorithm to record deduplication. Int. J. Comput. Appl. 2013, 62, 10–15. [Google Scholar] [CrossRef]
  72. Yang, X.S. Nature-Inspired Metaheuristic Algorithms; Luniver Press: Cambridge, UK, 2008. [Google Scholar]
  73. Chou, J.S.; Ngo, N.T. Modified firefly algorithm for multidimensional optimization in structural design problems. Struct. Multidiscip. Optim. 2017, 55, 2013–2028. [Google Scholar] [CrossRef]
  74. Blázquez, C.S.; Maté-González, M.A.; Nieto, I.M.; Martín, A.F.; González-Aguilera, D. Assessment of the geothermal potential in the region of Ávila (Spain): An integrated and interactive thermal approach. Geothermics 2022, 98, 102294. [Google Scholar] [CrossRef]
  75. PANER. Plan de Acción Nacional de Energías Renovables de España (2011–2020); Spanish Department of Industry Tourism and Commerce: Madrid, Spain, 2010. [Google Scholar]
  76. Ryan, G.A.; Peacock, J.R.; Shalev, E.; Rugis, J. Montserrat geothermal system: A 3D conceptual model. Geophys. Res. Lett. 2013, 40, 2038–2043. [Google Scholar] [CrossRef]
  77. Witter, J.B.; Siler, D.L.; Faulds, J.E.; Hinz, N.H. 3D geophysical inversion modelling of gravity data to test the 3D geologic model of the Bradys geothermal area, Nevada, USA. Geotherm. Energy 2016, 4, 14. [Google Scholar] [CrossRef] [Green Version]
  78. Milicich, S.D.; Pearson-Grant, S.C.; Alcaraz, S.; White, P.A.; Tschritter, C. 3D Geological modelling of the Taupo Volcanic Zone as a foundation for a geothermal reservoir model. N. Z. J. Geol. Geophys. 2018, 61, 79–95. [Google Scholar] [CrossRef]
  79. Caruana, R. Multitask learning. Mach. Learn. 1997, 28, 41–75. [Google Scholar] [CrossRef]
  80. Puigdefàbregas, C.; Muñoz, J.A.; Vergés, J. Thrusting and foreland basin evolution in the southern Pyrenees. In Thrust Tectonics; Springer: Dordrecht, The Netherlands, 1992; pp. 247–254. [Google Scholar] [CrossRef]
  81. Lladó, X.M.; Vilà, M.; Martí, V.; Rovira, M.; Domènech, J.A.; Pablo, J.D. Trace Element Distribution in Topsoils in Catalonia: Background and Reference Values and Relationship with Regional Geology. Environ. Eng. Sci. 2008, 25, 863–878. [Google Scholar] [CrossRef]
  82. Bartrina, M.T.; Cabrera, L.; Jurado, M.J.; Guimerà, J.; Roca, E. Evolution of the central Catalan margin of the Valencia trough (western Mediterranean). Tectonophysics 1992, 203, 219–247. [Google Scholar] [CrossRef]
  83. Roca, E. The northwest-Mediterranean basin (Valencia Trough, Gulf of Lions and Liguro-Provencal basins): Structure and geo-dynamic evolution. Mémoires du Muséum Natl. d’Histoire Nat. 2001, 168, 671–706. [Google Scholar]
  84. ICGC Shallow Geothermal Energy. Geoindex (Viewer & Geoservices WMS), February 2021, Institut Cartogràfic i Geològic de Catalunya, Barcelona. Available online: https://www.icgc.cat/ (accessed on 1 June 2021).
  85. ICGC Deep Geothermal Energy. Geoindex (Viewer & Geoservices WMS), October 2020, Institut Cartogràfic i Geològic de Catalunya, Barcelona. Available online: https://www.icgc.cat/ (accessed on 1 June 2021).
  86. Baxter, J. A model of inductive bias learning. J. Artif. Intell. Res. 2000, 12, 149–198. [Google Scholar] [CrossRef]
  87. Ando, R.; Zhang, T. A framework for learning predictive structures from multiple tasks and unlabeled data. J. Mach. Learn. Res. 2005, 6, 1817–1853. [Google Scholar]
  88. Ruder, S. An overview of multi-task learning in deep neural networks. arXiv 2017, arXiv:1706.05098v1. [Google Scholar]
  89. Zhang, Y.; Yang, Q. A survey on multi-task learning. IEEE Trans. Knowl. Data Eng. 2021. [Google Scholar] [CrossRef]
  90. LeCun, Y.; Bengio, Y.; Hinton, G. Deep learning. Nature 2015, 521, 436–444. [Google Scholar] [CrossRef] [PubMed]
  91. Fister, I.; Yang, X.S.; Brest, J. A comprehensive review of firefly algorithms. Swarm Evol. Comput. 2013, 13, 34–46. [Google Scholar] [CrossRef] [Green Version]
  92. Colmenar-Santos, A.; Folch-Calvo, M.; Rosales-Asensio, E.; Borge-Diez, D. The geothermal potential in Spain. Renew. Sustain. Energy Rev. 2016, 56, 865–886. [Google Scholar] [CrossRef]
  93. Willmott, C.J. On the validation of models. Phys. Geogr. 1981, 2, 184–194. [Google Scholar] [CrossRef]
  94. Willmott, C.J. On the evaluation of model performance in physical geography. Spat. Stat. Models 1984, 40, 443–460. [Google Scholar] [CrossRef]
  95. Stehman, S. Selecting and interpreting measures of thematic classification accuracy. Remote Sens. Environ. 1997, 62, 77–89. [Google Scholar] [CrossRef]
  96. Bradley, A.P. The use of the area under the ROC curve in the evaluation of machine learning algorithms. Pattern Recognit. 1997, 30, 1145–1159. [Google Scholar] [CrossRef] [Green Version]
  97. Fawcett, T. An introduction to ROC analysis. Pattern Recognit. Lett. 2006, 27, 861–874. [Google Scholar] [CrossRef]
  98. Meade, N.; Islam, T. Prediction intervals for growth curve forecasts. J. Forecast. 1995, 14, 413–430. [Google Scholar] [CrossRef]
  99. Hastie, T.; Tibshirani, R.; Friedman, J. The Elements of Statistical Learning, 2nd ed.; Springer: New York, NY, USA, 2001. [Google Scholar] [CrossRef]
  100. Lawless, J.F.; Fredette, M. Frequentist prediction intervals and predictive distributions. Biometrika 2005, 92, 529–542. [Google Scholar] [CrossRef]
  101. Abbaszadeh Shahri, A.; Shan, C.; Larsson, S. A novel approach to uncertainty analysis using automated predictive deep learning in groundwater table modelling. Nat. Res. Resour. 2022, 31, 1351–1373. [Google Scholar] [CrossRef]
Figure 1. The location of study area in Spain (A) and the corresponding geological settings (B).
Figure 1. The location of study area in Spain (A) and the corresponding geological settings (B).
Energies 15 04602 g001
Figure 2. The topology of DNSL (the used parameters are described in the text).
Figure 2. The topology of DNSL (the used parameters are described in the text).
Energies 15 04602 g002
Figure 3. Block diagram of developed adaptive hybridized multitask learning procedure to capture the optimum topology.
Figure 3. Block diagram of developed adaptive hybridized multitask learning procedure to capture the optimum topology.
Energies 15 04602 g003
Figure 4. Performance of series monitored topologies as a function of RMSE subjected to different TA and AF (a), and examined topologies to find the optimum model corresponding to 20 neurons (b) (TALMQN: limited memory quasi-Newton, CGD: conjugate gradient descent, QN: quasi-Newton, QP: quick propagation; AFLog: logistic, Hyt: hyperbolic tangent, Relu: rectified linear unit).
Figure 4. Performance of series monitored topologies as a function of RMSE subjected to different TA and AF (a), and examined topologies to find the optimum model corresponding to 20 neurons (b) (TALMQN: limited memory quasi-Newton, CGD: conjugate gradient descent, QN: quasi-Newton, QP: quick propagation; AFLog: logistic, Hyt: hyperbolic tangent, Relu: rectified linear unit).
Energies 15 04602 g004
Figure 5. Adjusting the optimal values of MAF parameters (a), selected population of fireflies using convergence histories (b), and compared predictability of the optimum and hybrid DNLS model (cf).
Figure 5. Adjusting the optimal values of MAF parameters (a), selected population of fireflies using convergence histories (b), and compared predictability of the optimum and hybrid DNLS model (cf).
Energies 15 04602 g005
Figure 6. The spatial distribution of surface temperature (a) and corresponding 3D predictive map of the study area using hybrid DNLS-MFA model at the depths 50 m (b), 120 m (c), 150 m (d) and 180 m (e).
Figure 6. The spatial distribution of surface temperature (a) and corresponding 3D predictive map of the study area using hybrid DNLS-MFA model at the depths 50 m (b), 120 m (c), 150 m (d) and 180 m (e).
Energies 15 04602 g006
Figure 7. Comparing the predictability level of optimum multitask and hybrid DNLS models using the AUCROC of conducted precision–recall curves.
Figure 7. Comparing the predictability level of optimum multitask and hybrid DNLS models using the AUCROC of conducted precision–recall curves.
Energies 15 04602 g007
Figure 8. Analyzed uncertainty of hybrid DNSL-MFA for T50 (a) and T150 (b) using automated random deactivating weight approach according to Ref. [101].
Figure 8. Analyzed uncertainty of hybrid DNSL-MFA for T50 (a) and T150 (b) using automated random deactivating weight approach according to Ref. [101].
Energies 15 04602 g008
Table 1. Required parameters to tune the FA.
Table 1. Required parameters to tune the FA.
Energies 15 04602 i001DescriptionParameterAdvised RangesNotation
brightnessI-objective function
attractivenessβ--
distance between two fireflies i, jrij--
absorption coefficientγ[0.1–10]-
trade off constant in randomized movementα[0, 1]randomization parameter
attractiveness at distance 0 (rij = 0)β0normally 1-
number of generationsGen-iteration
problem dimensionD-dimension
number of firefliesn-population size
β, β0, α and γ also can be organized based on parametric investigations.
Table 2. Evaluated error metrics of the predicted temperatures using validation datasets.
Table 2. Evaluated error metrics of the predicted temperatures using validation datasets.
Optimum DNLSHybrid DNLS-MFA
T50T150T50T150
R20.870.860.920.91
RMSE0.491.000.370.79
MAPE2.404.001.803.2
IA0.960.960.980.97
Table 3. Confusion matrix of optimum models using validation datasets.
Table 3. Confusion matrix of optimum models using validation datasets.
Target
Output
Output of Optimum DNLS: 50 m DepthResults
<12.0112.01–12.6912.69–13.3813.38–14.0614.06–14.7514.75–15.4315.43–16.1116.11–16.8016.80–17.4817.48–18.1718.17–18.85>18.85TotalTrueFalse
12.01–12.69 2 1 321
12.69–13.38 0 000
13.38–14.06 1 110
14.06–14.75 1 110
14.75–15.43 21 321
15.43–16.11 171 1073
16.11–16.80 1611 18162
16.80–17.48 1182 21183
17.48–18.17 181 1082
18.17–18.85 151752
Note020213818201261746014
Output of Hybrid DNLS: 50 m DepthResults
12.01–12.69 21 321
12.69–13.38 0 000
13.38–14.06 1 110
14.06–14.75 01 101
14.75–15.43 3 330
15.43–16.11 181 1082
16.11–16.80 171 18171
16.80–17.48 201 21201
17.48–18.17 181 1082
18.17–18.85 61761
Note0211059172297174659
Target
Output
Output of Optimum DNLS: 150 m DepthResults
<13.6613.66–14.7314.73–15.815.8–16.8716.87–17.9417.94–19.0119.01–20.0820.08–21.1521.15–22.2222.22–23.2923.29–24.36>24.36TotalTrueFalse
13.66–14.73111 312
14.73–15.8 31 431
15.8–16.87 14 1 642
16.87–17.94 4 1 541
17.94–19.01 18 1 1082
19.01–20.08 11101 13103
20.08–21.15 110 1 12102
21.15–22.22 1 8 981
22.22–23.29 1 51 752
23.29–24.36 41541
Note115561013137751745717
Output of Hybrid DNLS: 150 m depthResults
13.66–14.7312 321
14.73–15.8 31 431
15.8–16.87 141 642
16.87–17.94 41 541
17.94–19.01 91 1091
19.01–20.08 1111 13112
20.08–21.15 11 1 12111
21.15–22.22 18 981
22.22–23.29 151 752
23.29–24.36 31532
Note124561012139642746014
Table 4. Comparing the CCR and model improvement using validation data.
Table 4. Comparing the CCR and model improvement using validation data.
ModelCCR (%)Improved Progress
Optimum DNLS-T5081.16.24%
Hybrid DNLS-T5086.5
Optimum DNLS-T15077.14.93%
Hybrid DNLS-T15081.1
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mirfallah Lialestani, S.P.; Parcerisa, D.; Himi, M.; Abbaszadeh Shahri, A. Generating 3D Geothermal Maps in Catalonia, Spain Using a Hybrid Adaptive Multitask Deep Learning Procedure. Energies 2022, 15, 4602. https://doi.org/10.3390/en15134602

AMA Style

Mirfallah Lialestani SP, Parcerisa D, Himi M, Abbaszadeh Shahri A. Generating 3D Geothermal Maps in Catalonia, Spain Using a Hybrid Adaptive Multitask Deep Learning Procedure. Energies. 2022; 15(13):4602. https://doi.org/10.3390/en15134602

Chicago/Turabian Style

Mirfallah Lialestani, Seyed Poorya, David Parcerisa, Mahjoub Himi, and Abbas Abbaszadeh Shahri. 2022. "Generating 3D Geothermal Maps in Catalonia, Spain Using a Hybrid Adaptive Multitask Deep Learning Procedure" Energies 15, no. 13: 4602. https://doi.org/10.3390/en15134602

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