Next Article in Journal
Analyzing HDPE Thermal and Catalytic Degradation in Hydrogen Atmosphere: A Model-Free Approach to the Activation Energy
Previous Article in Journal
Investigation of Calcium Phosphate Catalysts in Sodium Borohydride Methanolysis for Improved Hydrogen Production
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Prediction of Oxygen Evolution Activity for FeCoMn Oxide Catalysts via Machine Learning

by
Lei Zhang
1,
Jinfei Hou
1,
Honglin Ji
1,
Dan Meng
1,
Jian Qi
2,* and
Xiaoguang San
1,*
1
College of Chemical Engineering, Shenyang University of Chemical Technology, Shenyang 110142, China
2
State Key Laboratory of Biochemical Engineering, Institute of Process Engineering, Chinese Academy of Sciences, Beijing 100190, China
*
Authors to whom correspondence should be addressed.
Catalysts 2024, 14(8), 513; https://doi.org/10.3390/catal14080513
Submission received: 24 June 2024 / Revised: 30 July 2024 / Accepted: 5 August 2024 / Published: 8 August 2024
(This article belongs to the Section Catalytic Materials)

Abstract

:
Electrolytic hydrogen production from water is a promising approach for obtaining clean energy. The development of efficient oxygen evolution reaction (OER) electrocatalysts is crucial for the generation of hydrogen through water electrolysis. Transition metal oxides, such as Fe, Co, and Mn, have shown potential as efficient OER electrocatalysts for water splitting. However, accurately predicting their electrocatalytic performance in complex compositional spaces remains a challenge, impeding the precise design of compositions and processes for optimal performance. Herein, a machine learning-based method is proposed for predicting the OER activity of (FeCoMn)Ox catalysts across a wide range of compositions. Physical features that are highly relevant to the OER overpotential (OP) are identified and analyzed. The random forest algorithm is successfully used to establish the relationship between composition and overpotential. The model demonstrates good accuracy in predicting the outcomes of new experiments, with a mean relative error (MRE) of 9.3%. The features based on covalent radius (RC) and the number of electrons in the outermost d orbitals (DEs) are the primary factors. Their variances (δRC and δDE) exhibit a linearly decreasing relationship with the overpotential (OP), providing direct guidance for designing OP-oriented components. This work presents an effective and innovative approach for predicting and analyzing the physical factors of transition metal oxide electrocatalysts, which can enhance the design of highly catalytic materials for electrocatalysis.

Graphical Abstract

1. Introduction

As the global demand for renewable energy continues to increase, the utilization of water electrolysis technology, which converts electrical energy into hydrogen and oxygen, has gained significant attention [1,2,3]. In this process, the oxygen evolution reaction (OER) is a key step that involves a 4e-transfer and exhibits slow reaction kinetics [4,5,6,7]. Thus, the development of efficient OER electrocatalysts is crucial for the sustainable advancement of clean energy. However, the widespread application of traditional noble metal catalysts (such as RuO2 and IrO2) for the OER is limited due to their scarcity and high cost [8,9,10,11]. As a result, there is a growing focus on the search for efficient, cost-effective, and sustainable non-noble metal catalysts, making it a research hotspot.
Transition metal-based catalysts, including the oxides, hydroxides, phosphides, and borides of Mn, Fe, Co, and Ni, have shown significant promise for OER due to their high catalytic performance and potential as alternatives to noble metal catalysts [12]. However, these catalysts often face challenges, such as limited active sites and suboptimal electrocatalytic activity, which hinder their widespread application. Transition metal oxides (TMOs) possess intricate inherent characteristics that contribute to a diverse range of catalytically influential properties stemming from their unique electronic structures that can be tailored through compositional adjustments and structural modifications. The performance of TMOs in electrochemical applications is also greatly influenced by testing conditions, such as voltage ranges, current densities, and mass loading. In recent years, various techniques have been developed to optimize composition and structure, such as controlling the morphology of the catalyst to expose more active sites [13,14,15,16] or regulating mass loading to achieve the optimal balance of catalytic efficiency and stability [17,18,19]. Furthermore, heteroatoms are incorporated into monometallic metal oxides or cation substitution to create binary, ternary, and other multi-metal oxide catalysts [20]. This approach, particularly with first-row transition metals such as Fe, Ni, Co, Mn, and Cu, has demonstrated potential in modifying the electronic structure, enhancing conductivity, improving hydrogen adsorption energy, and increasing the number of active sites [21,22,23,24,25]. Pu et al. prepared 12% Ni-Co3O4 (the molar ratio of Ni to Co is 12%), which rendered the best activity [26]. This improvement is attributed to the unique synergistic effect between Co and Ni, as well as the modified electronic structure. Li et al. prepared a series of Fe-substituted cobalt spinel oxides, FexNi1−xCo2O4 (x = 0, 0.1, 0.25, 0.5, 0.75, 0.9, 1), where Fe0.5Ni0.5Co2O4 exhibited enhanced OER activity [27]. This improvement can be attributed to the synergistic effect of ternary mixed metals, particularly the incorporation of Fe and the uniform nanowire support on nickel foam (NF). Therefore, the optimization of the elemental composition in binary, ternary, and multi-metal oxide catalysts is an efficient way to enhance their catalytic performance [28,29,30]. The vast design space of binary, ternary, and multi-metal oxides makes it challenging to screen all potential catalysts and identify the optimal composition for catalytic performance. Traditional methods involve synthesizing and testing numerous compositions, which is both time-consuming and resource-intensive. This approach often leads to a bottleneck in the discovery process.
Machine learning (ML) is a powerful statistical technique that employs algorithms to map input data to desired information [31,32,33,34]. Data mining technology can be used to analyze the nonlinear relationship between the catalyst structure (geometric and electronic structure) and performance, thereby quickly predicting the performance of multiple catalysts and accelerating research progress [35,36,37,38]. These methods have proven to be valuable in accelerating the screening and discovery of new materials, particularly in fields such as biomaterials, electrocatalysis, and batteries [39]. However, the effectiveness of ML-based methods relies heavily on the availability and richness of data. Fortunately, there is a wealth of experimental and theoretical data in the literature and databases related to catalyst design and optimization, which can provide valuable insights for the development of new catalysts.
In this work, we introduced a data-driven approach to predict the overpotential (OP) of (Fe-Co-Mn)Ox catalysts using machine learning algorithms. By considering valence electrons, relative atomic mass, atomic number, the non-bonding atomic radius, covalent radius, first ionization energy, Pauling electronegativity, and the number of outermost d-orbital electrons, various physical features were successfully correlated with overpotential. According to the optimal value of the mean square error and mean absolute error, the random forest (RF) algorithm was chosen as the best model. This model was used to efficiently screen the Fe-Co-Mn ternary catalyst system and identify optimal composition ratios [40]. In order to verify the results of machine learning in the new experiments, we prepared five iron–cobalt–manganese oxide catalysts with different ratios using a hydrothermal method and tested their OER catalytic performance. The experimental results were basically consistent with machine learning prediction results, with a mean relative error (MRE) of 9.3%. This work provides an efficient new approach to predicting and designing transition metal oxide OER electrocatalysts.

2. Results and Discussion

2.1. Data Collection

All the data used in this work were collected and screened from Ahmed’s published research through high-throughput experiments [41]. We considered oxide catalysts belonging to the FexCoyMnz system, where the molar fractions of each element (x, y, and z) were subject to the constraint x + y + z = 100%. The dataset comprises 66 entries. These entries cover the elemental compositions of various (Fe-Co-Mn)Ox materials. Therefore, the overall dataset consists of three elemental features (input variables) and one target OP (output variable). Figure 1 visualizes the original dataset by categorizing them based on the compositions of Fe, Co, Mn, and OP. As the compositions of Fe, Co, and Mn increase, the overall OP exhibits an initial decrease followed by an increase. This method clearly reveals the optimal composition for metal oxide catalysts in terms of overpotential. The comprehensive and complete dataset provides reliable evidence for predictive models and physical analysis.

2.2. Feature Construction

Feature selection plays a pivotal role in the performance of machine learning models. These features should not only be easily accessible but also accurately represent the geometrical and electronic characteristics of the metal atoms. This study considers eight independent features as follows: relative atomic mass (RAM), atomic number (AN), non-bonded atomic radius (RA), covalent radius (RC), valence electron number (VEN), first ionization energy (FIE), Pauling scale electronegativity (EP), and the number of electrons in the outermost d orbital (DE). The first four features primarily focus on the geometric structure of the catalysts, while the remaining four pertain to the electronic properties, indicating the active site’s ability to donate or accept electrons.
For Fe, Co, and Mn elements, the corresponding physical properties associated with these features were collected from the Royal Society of Chemistry’s interactive periodic table database [42]. Each catalyst sample was numerically represented using the composition (Ci) and relevant elemental properties. The feature transformation functions are given by Equations (1) and (2), with the aim of converting the original chemical element space into the primary physical feature space (Table 1). Specifically, for each of the 66 catalyst samples, X ¯ computes the weighted average of the element content for each physical feature, while δX calculates the variance of each physical feature. This variance reflects the physical diversity among the chemical elements. In Equations (1) and (2), Ci is the mole fraction of each element, and Pi corresponds to the properties of each element, respectively. Thus, X ¯ and δX construct 16 descriptive features that may be physically relevant to the OP. The original dataset was converted into a new dataset of shape 66 × 16 (Table S1).

2.3. Prediction Model

Herein, we used 11 algorithms for model construction, including random forest (RF), Adaptive Boosting (boosting), K-Nearest Neighbors (KNNs), Bagged Classification and Regression Trees (CART_bagged), Gaussian Kernel Support Vector Machine (SVM_radial), Cubist, Full Quadratic K-Nearest Neighbors (full_quad_KNNs), neural network, Multivariate Adaptive Regression Splines (MARSs), Classification and Regression Trees (CARTs), and Polynomial Kernel Support Vector Machine (SVM_poly). Among them, SVM_radial, SVM_radial, and KNNs are linear models, and multicollinearity can lead to unstable coefficient estimates and unreliable hypothesis testing in regression models. Prior to implementing machine learning techniques, it was necessary to compute the Pearson correlation coefficient in order to remove any linear correlation between two variables. Figure 2 shows a heat map of Pearson correlation coefficients between feature parameters. Features with absolute correlation coefficients greater than 0.95 are considered highly correlated, as shown by the dark blue and dark red boxes in Figure 2. In a pair of highly correlated features, one of them can be expressed linearly and replaced by the other. Multiple variables, including δAN and δDE with a correlation of 1 and EA and RA with a correlation of −1, show a significant correlation. The linear model prediction, after eliminating collinearity, incorporates the five variables VEN, FIE, δVEN, δFIE, and δDE.
During the model training, 80% of the instances in the database were randomly selected as the training set, while the remaining 20% were designated as the testing set. The model’s quality was assessed by calculating the mean squared error (MSE) and mean absolute error (MAE) due to the continuous nature of overpotential. To determine the optimal parameters, a ten-fold cross-validation was conducted on each machine learning model, followed by a grid search on the training set. This method involved dividing the dataset into ten subsets, using one as the validation set and the remaining nine as the training set, which was repeated ten times to ensure the model’s stability and generalization ability. As shown in Figure 3a, the variation in parameters in the neural network algorithms had the most significant impact on the model’s predictive ability, while the CART_bagged model remained relatively unaffected. CART_bagged is a decision tree model that utilizes bagging to improve stability and predictive performance by averaging the results of multiple decision trees. After selecting the optimal parameters for each model, their predictive ability is evaluated on new datasets using the validation set. This analysis reveals that random forests have the best predictive performance (MSE = 30.45, Figure 3b). This ensemble learning method builds multiple decision trees and averages their results to improve prediction accuracy and stability.
The RF model was retrained using the optimized parameters on the training set. The diagonal scatter plot in Figure 4a displays the predicted OP and ground truth values of the RF model during both training and testing, utilizing the transformed and selected datasets. The results indicate a consistent trend between the RF-predicted OP and ground truth values, with a clear linear relationship observed in both the training and testing sets. This demonstrates the high accuracy of the well-trained RF model in predicting OP.
Based on these aforementioned findings, the RF model was utilized to predict and screen the Fe-Co-Mn ternary catalyst system in order to design the optimal components for a highly active FeCoMn catalyst. The search space was designed as FexCoyMnz, where x, y, and z represent the mole fractions of Fe, Co, and Mn, respectively, ranging from 0 to 1 with a step of 0.01. To facilitate the description of the catalyst’s composition, a ternary phase diagram (Figure 4b) was created, with the mole fractions of Fe, Co, and Mn representing the three edges and the colors indicating the OER activity of the catalyst. It was observed that the FeCoMn ternary catalyst exhibited superior catalytic activity compared to binary compositions and single metal oxides. Among them, Fe12Co40Mn48Ox showed the best catalytic activity with the overpotential measured at 10 mA cm−2 of the current. Optimizing the components is an effective approach to enhance catalyst activity. Our work provides an efficient method for quickly determining the optimal component for catalyst optimization.

2.4. Experiment Validation

To validate the results predicted by machine learning, we prepared Fe12Co40Mn48Ox, Fe30Co40Mn30Ox, Fe30Co20Mn50Ox, Fe50Co50Ox, and Co50Mn50Ox at five different ratios of FeCoMn oxide catalysts using a hydrothermal method. Their morphologies are shown in Figure S1, and they exhibit similar nanosheet-like structures. The elemental compositions were confirmed by inductively coupled plasma-optical emission spectrometer (ICP-OES) measurements (Table S2). We then tested the OER activity of the prepared samples in N2 saturated in 1.0 M KOH (aq). Figure 5a shows the LSV curves of the samples, and the experimental overpotential and predicted overpotential 10 mA/cm−2 are shown in Table 2. The ground truth values of the experimental OP and the OP predicted by the RF model are presented in Figure 5b. The results show a close match between the experimental and predicted values for the five catalysts, with an MRE of 9.3%. While deviations may occur due to factors such as data bias, model complexity, and overfitting, the trend of RF-predicted OP remains consistent with the experimental value. This demonstrates the successful application of the RF model in catalyst screening and design optimization. Cdl was calculated by measuring the cyclic voltammetry (CV) curve (Figure S2). As depicted in Figure S3, compared with the other four catalysts, Fe12Co40Mn48Ox shows a higher Cdl value (3.23 mF cm−2), indicating that Fe12Co40Mn48Ox has a larger electrochemically active surface area. The Tafel slopes of the five catalysts are shown in Figure S4. Compared with the other four catalysts, the Tafel slope of Fe12Co40Mn48Ox is the smallest, which is 49.36 mV/dec, indicating that the OER kinetics of Fe12Co40Mn48Ox are the fastest.
In order to further understand the relationship between characteristic values and overpotentials, the Shapley Additive Explanation (SHAP) values of the characteristics in 66 sets of data instances were calculated. We evaluated the contribution of feature values to prediction results using the scatter plot of the feature impact from SHAP (Figure 6). In the figure, all features were sorted vertically from large to small by the sum of SHAP values, while the SHAP value distribution of each feature is displayed on the horizontal axis. Each dot represents a catalyst; the color of the dot represents the numerical value of the corresponding feature of the catalyst, and the position of the dot represents the SHAP value of the corresponding feature of the catalyst. Based on the results from SHAP plots, we can observe whether each feature has a positive or negative impact on the prediction outcome. Figure 6 displays the distribution of SHAP values for all feature values, arranged in descending order of feature importance. In the graph, δRC is at the top of the SHAP plot, indicating that it is the most important feature. δRC, δDE, δAN, δRAM, δRA, and δEP are widely dispersed in the SHAP plot. This indicates that they significantly influence the prediction of overpotential. When the values of these features are larger, their SHAP values become more negative. This implies that these features are negatively correlated with the predicted overpotential. The variance of the covalent radius and the variance of Pauling electronegativity reveal the stability of covalent radii data for different proportions of Fe, Co, and Mn metals. Smaller feature values indicate greater dataset stability and reduced fluctuations, implying minimal differences among the data points. Therefore, metal dopants with similar radii and small electronegativity differences can effectively modulate the geometric and electronic structures, thereby optimizing OER performance.
Moreover, the literature has shown that the number of electrons in the outermost d orbital has a direct impact on the d-band center energy of the metal active center, while the covalent radius is closely linked to the formation of covalent bonds between the metal atom and its coordination with oxygen atoms. This results in a lower d-band center energy for metals with more outermost d electrons, leading to a more stable energy-level structure. This, in turn, affects the interaction between metal sites and adsorbates, ultimately influencing the catalytic performance. Additionally, smaller covalent radii correspond to shorter bond distances and stronger covalent bonds. Metals with a higher number of d electrons form shorter metal–oxygen bonds, resulting in the release of additional active sites and stabilization of intermediate adsorption states, thereby enhancing catalyst activity. It is evident that these factors are closely interconnected, and understanding their interrelationships and effects is crucial to comprehend catalyst activity mechanisms and design efficient OER catalysts.

3. Materials and Methods

The data flow of this article is based on analysis using the R language package of mr13. Firstly, the collected data were split 7:3 into training and validation sets. Secondly, 11 algorithms were used for model construction, including random forest (RF), Adaptive Boosting (boosting), K-Nearest Neighbors (KNNs), Bagged Classification and Regression Trees (CART_bagged), Gaussian Kernel Support Vector Machine (SVM_radial), Cubist, Full Quadratic K-Nearest Neighbors (full_quad_KNNs), neural network, Multivariate Adaptive Regression Splines (MARSs), Classification and Regression Trees (CARTs), and Polynomial Kernel Support Vector Machine (SVM_poly). For each machine learning model, parameter tuning of the model was performed using ten-fold cross-validation. Since the overpotential is a continuous variable, the goodness of the model is evaluated by calculating the mean squared error (MSE) and the mean absolute error (MAE), and the model parameter with the smallest MSE and MAE is selected as the best model. The experimental procedures for catalyst synthesis and electrochemical characterization are detailed in the Supplementary Information section.

4. Conclusions

A machine learning-based method was proposed to predict the activity of (Fe-Co-Mn)Ox catalysts in high-dimensional composition space. Physical features related to OER catalyst overpotential were constructed, including valence electron count, relative atomic mass, atomic number, non-bonding atomic radius, covalent radius, first ionization energy, Pauling electronegativity, and outermost d-orbital electron count. The random forest algorithm demonstrated the best-fitting performance and was used to predict overpotential. Experimental validation with five proportions of (Fe-Co-Mn)Ox catalysts resulted in a mean relative error (MRE) of 9.3% and a mean squared error (MSE) of 30.45. The most influential features were found to be the variance in covalent radius (RC) and outermost d orbital electron number (DE), which showed a linear decrease with overpotential (OP). This model was used to efficiently screen the Fe-Co-Mn ternary catalyst system and identify optimal composition ratios, reducing the number and time of experiments and overall development cost. This approach provides a valuable tool for predicting the activity of FeCoMn oxide OER electrocatalysts and has the potential to be extended to other electrocatalyst design applications.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/catal14080513/s1. Figure S1: SEM images of (a) Fe12Co40Mn48Ox, (b) Fe30Co20Mn50Ox, (c) Fe30Co40Mn30Ox, (d) Co50Mn50Ox,and (e) Fe50Co50Ox catalysts at different scan rates; Figure S2: CV curves of (a) Fe12Co40Mn48Ox, (b) Co50Mn50Ox, (c) Fe30Co20Mn50Ox, (d) Fe30Co40Mn30Ox, and (e) Fe50Co50Ox catalysts at different scan rates; Figure S3: Cdl values through CV curves in the non-faradic region of all catalysts; Figure S4. Tafel plots of metal oxides; Table S1: The transformed dataset with physical features with the shape of 66 × 18. Table S2. Composition of intended compositions with measured compositions by ICP-OES from the selected regions in the ternary composition space.

Author Contributions

Conceptualization, L.Z., J.Q. and X.S.; methodology, L.Z.; software, L.Z.; validation, L.Z., D.M., J.Q. and X.S.; formal analysis, L.Z.; investigation, J.H. and H.J.; resources, X.S.; data curation, L.Z.; writing—original draft preparation, J.H. and H.J.; writing—review and editing, L.Z., D.M. and J.Q.; visualization, L.Z.; supervision, X.S.; project administration, X.S.; funding acquisition, L.Z., D.M. and X.S. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are grateful for the support of the National Natural Science Foundation of China (No. 61973223), the Liao Ning Revitalization Talents Program (No. XLYC2007051), the Liaoning Educational Department Foundation (Nos. LJKMZ20220762, JYTMS20231510, LJKMZ 20220766), the Natural Science Foundation of Liaoning Province (Nos. 2023-MS-235, 2023-MSLH-270), and the Key Project in Science & Technology of SYUCT (No. 2023DB005).

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zhang, X.; Zhang, L.; Zhu, Y.; Li, Z.; Wang, Y.; Wågberg, T.; Hu, G. Increasing Electrocatalytic Oxygen Evolution Efficiency through Cobalt-Induced Intrastructural Enhancement and Electronic Structure Modulation. ChemSusChem 2021, 14, 467–478. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, X.; Li, Q.; Hu, J.; Zhou, C.; Yang, S.; Huang, X. Coordination confinement pyrolysis to hollow sea urchin shaped composite with embedded ultrasmall Co/Ni alloy for overall water splitting. Int. J. Hydrogen Energy 2022, 47, 3699–3708. [Google Scholar] [CrossRef]
  3. Dai, Z.; Du, X.; Zhang, X. Controlled synthesis of NiCo2O4@ Ni-MOF on Ni foam as efficient electrocatalyst for urea oxidation reaction and oxygen evolution reaction. Int. J. Hydrogen Energy 2022, 47, 17252–17262. [Google Scholar] [CrossRef]
  4. Tran-Phu, T.; Daiyan, R.; Leverett, J.; Fusco, Z.; Tadich, A.; Di Bernardo, I.; Kiy, A.; Truong, T.N.; Zhang, Q.; Chen, H. Understanding the activity and stability of flame-made Co3O4 spinels: A route towards the scalable production of highly performing OER electrocatalysts. Chem. Eng. J. 2022, 429, 132180. [Google Scholar] [CrossRef]
  5. Niu, S.; Jiang, W.J.; Wei, Z.; Tang, T.; Ma, J.; Hu, J.S.; Wan, L.J. Se-doping activates FeOOH for cost-effective and efficient electrochemical water oxidation. J. Am. Chem. Soc. 2019, 141, 7005–7013. [Google Scholar] [CrossRef] [PubMed]
  6. Song, J.; Wei, C.; Huang, Z.F.; Liu, C.; Zeng, L.; Wang, X.; Xu, Z.J. A review on fundamentals for designing oxygen evolution electrocatalysts. Chem. Soc. Rev. 2020, 49, 2196–2214. [Google Scholar] [CrossRef] [PubMed]
  7. Kim, M.; Kim, J.; Qin, L.; Mathew, S.; Han, Y.; Li, O.L. Gas-Liquid Interfacial Plasma engineering under dilute nitric acid to improve hydrophilicity and OER performance of nickel foam. Prog. Nat. Sci. Mater. Int. 2022, 32, 608–616. [Google Scholar] [CrossRef]
  8. Zhao, D.; Zhuang, Z.; Cao, X.; Zhang, C.; Peng, Q.; Chen, C.; Li, Y. Atomic site electrocatalysts for water splitting, oxygen reduction and selective oxidation. Chem. Soc. Rev. 2020, 49, 2215–2264. [Google Scholar] [CrossRef] [PubMed]
  9. Lee, Y.; Suntivich, J.; May, K.J.; Perry, E.E.; Shao-Horn, Y. Synthesis and activities of rutile IrO2 and RuO2 nanoparticles for oxygen evolution in acid and alkaline solutions. J. Phys. Chem. Lett. 2012, 3, 399–404. [Google Scholar] [CrossRef]
  10. Wang, C.; Jin, L.; Shang, H.; Xu, H.; Shiraishi, Y.; Du, Y. Advances in engineering RuO2 electrocatalysts towards oxygen evolution reaction. Chin. Chem. Lett. 2021, 32, 2108–2116. [Google Scholar] [CrossRef]
  11. Yoon, H.; Ju, B.; Kim, D.W. Perspectives on the development of highly active, stable, and cost-effective OER electrocatalysts in acid. Battery Energy 2023, 2, 20230017. [Google Scholar] [CrossRef]
  12. Osgood, H.; Devaguptapu, S.V.; Xu, H.; Cho, J.; Wu, G. Transition metal (Fe, Co, Ni, and Mn) oxides for oxygen reduction and evolution bifunctional catalysts in alkaline media. Nano Today 2016, 11, 601–625. [Google Scholar] [CrossRef]
  13. You, N.; Cao, S.; Huang, M.; Fan, X.; Shi, K.; Huang, H.; Chen, Z.; Yang, Z.; Zhang, W. Constructing P-CoMoO4@ NiCoP heterostructure nanoarrays on Ni foam as efficient bifunctional electrocatalysts for overall water splitting. Nano Mater. Sci. 2023, 5, 278–286. [Google Scholar] [CrossRef]
  14. Mane, R.S.; Mane, S.; Somkuwar, V.; Thombre, N.V.; Patwardhan, A.V.; Jha, N. A novel hierarchically hybrid structure of MXene and bi-ligand ZIF-67 based trifunctional electrocatalyst for zinc-air battery and water splitting. Battery Energy 2023, 2, 20230019. [Google Scholar] [CrossRef]
  15. Zhao, L.; Jiang, J.; Xiao, S.; Li, Z.; Wang, J.; Wei, X.; Kong, Q.; Chen, J.S.; Wu, R. PtZn nanoparticles supported on porous nitrogen-doped carbon nanofibers as highly stable electrocatalysts for oxygen reduction reaction. Nano Mater. Sci. 2023, 5, 329–334. [Google Scholar] [CrossRef]
  16. Zhang, Y.; Teng, X.; Ma, Z.; Wang, R.; Lau, W.M.; Shan, A. Cu2O-templated fabrication of Ni (OH)2·0.75 H2O hollow tubes for electrocatalytic oxygen evolution reaction. Prog. Nat. Sci. Mater. Int. 2022, 32, 554–560. [Google Scholar] [CrossRef]
  17. Ciambriello, L.; Cavaliere, E.; Vassalini, I.; Alessandri, I.; Ferroni, M.; Leoncino, L.; Brescia, R.; Gavioli, L. Role of Electrode Thickness in NiFe Nanogranular Films for Oxygen Evolution Reaction. J. Phys. Chem. C 2022, 126, 21759–21770. [Google Scholar] [CrossRef]
  18. Ciambriello, L.; Alessandri, I.; Ferroni, M.; Gavioli, L.; Vassalini, I. Unexpected Resilience of NiFe Catalysts for the Alkaline Oxygen Evolution Reaction. ACS Appl. Energy Mater. 2024, 7, 3462–3472. [Google Scholar] [CrossRef]
  19. Ciambriello, L.; Alessandri, I.; Gavioli, L.; Vassalini, I. NiFe Catalysts for Oxygen Evolution Reaction: Is There an Optimal Thickness for Generating a Dynamically Stable Active Interface? ChemCatChem 2024, e202400286. [Google Scholar] [CrossRef]
  20. Fan, M.; Cui, J.; Wu, J.; Vajtai, R.; Sun, D.; Ajayan, P.M. Improving the catalytic activity of carbon-supported single atom catalysts by polynary metal or heteroatom doping. Small 2020, 16, 1906782. [Google Scholar] [CrossRef]
  21. Cheng, R.; Min, Y.; Li, H.; Fu, C. Electronic structure regulation in the design of low-cost efficient electrocatalysts: From theory to applications. Nano Energy 2023, 115, 108718. [Google Scholar] [CrossRef]
  22. Yan, Y.; Wang, P.; Lin, J.; Cao, J.; Qi, J. Modification strategies on transition metal-based electrocatalysts for efficient water splitting. J. Energy Chem. 2021, 58, 446–462. [Google Scholar] [CrossRef]
  23. Zhang, A.; Liang, Y.; Zhang, H.; Geng, Z.; Zeng, J. Doping regulation in transition metal compounds for electrocatalysis. Chem. Soc. Rev. 2021, 50, 9817–9844. [Google Scholar] [CrossRef] [PubMed]
  24. Huang, J.; Jiang, Y.; An, T.; Cao, M. Increasing the active sites and intrinsic activity of transition metal chalcogenide electrocatalysts for enhanced water splitting. J. Mater. Chem. A 2020, 8, 25465–25498. [Google Scholar] [CrossRef]
  25. Strmcnik, D.; Uchimura, M.; Wang, C.; Subbaraman, R.; Danilovic, N.; Van Der Vliet, D.; Paulikas, A.P.; Stamenkovic, V.R.; Markovic, N.M. Improving the hydrogen oxidation reaction rate by promotion of hydroxyl adsorption. Nat. Chem. 2013, 5, 300–306. [Google Scholar] [CrossRef] [PubMed]
  26. Du, H.; Pu, W.; Yang, C. Morphology control of Co3O4 with nickel incorporation for highly efficient oxygen evolution reaction. Appl. Surf. Sci. 2021, 541, 148221. [Google Scholar] [CrossRef]
  27. Yan, K.L.; Shang, X.; Li, Z.; Dong, B.; Li, X.; Gao, W.K.; Chi, J.Q.; Chai, Y.M.; Liu, C.G. Ternary mixed metal Fe-doped NiCo2O4 nanowires as efficient electrocatalysts for oxygen evolution reaction. Appl. Surf. Sci. 2017, 416, 371–378. [Google Scholar] [CrossRef]
  28. Huang, Y.; Han, J.; Wang, H.; Liu, L.; Liang, H. Multiple metallic dopants in nickel nanoparticles for electrocatalytic oxygen evolution. Prog. Nat. Sci. Mater. Int. 2023, 33, 67–73. [Google Scholar] [CrossRef]
  29. Kou, C.; Han, J.; Wang, H.; Han, M.; Liang, H. Unveiling the role of Zn dopants in NiFe phosphide nanosheet for oxygen evolution reaction. Prog. Nat. Sci. Mater. Int. 2023, 33, 74–82. [Google Scholar] [CrossRef]
  30. Xiao, Z.; Zhou, W.; Yang, B.; Liao, C.; Kang, Q.; Chen, G.; Liu, M.; Liu, X.; Ma, R.; Zhang, N. Tuned d-band states over lanthanum doped nickel oxide for efficient oxygen evolution reaction. Nano Mater. Sci. 2023, 5, 228–236. [Google Scholar] [CrossRef]
  31. Raccuglia, P.; Elbert, K.C.; Adler, P.D.; Falk, C.; Wenny, M.B.; Mollo, A.; Zeller, M.; Friedler, S.A.; Schrier, J.; Norquist, A.J. Machine-learning-assisted materials discovery using failed experiments. Nature 2016, 533, 73–76. [Google Scholar] [CrossRef] [PubMed]
  32. Xue, D.; Balachandran, P.V.; Hogden, J.; Theiler, J.; Xue, D.; Lookman, T. Accelerated search for materials with targeted properties by adaptive design. Nat. Commun. 2016, 7, 11241. [Google Scholar] [CrossRef] [PubMed]
  33. Xue, D.; Xue, D.; Yuan, R.; Zhou, Y.; Balachandran, P.V.; Ding, X.; Sun, J.; Lookman, T. An informatics approach to transformation temperatures of NiTi-based shape memory alloys. Acta Mater. 2017, 125, 532–541. [Google Scholar] [CrossRef]
  34. Zhang, Y.; Wen, C.; Wang, C.; Antonov, S.; Xue, D.; Bai, Y.; Su, Y. Phase prediction in high entropy alloys with a rational selection of materials descriptors and machine learning models. Acta Mater. 2020, 185, 528–539. [Google Scholar] [CrossRef]
  35. Wu, L.; Guo, T.; Li, T. Rational design of transition metal single-atom electrocatalysts: A simulation-based, machine learning-accelerated study. J. Mater. Chem. A 2020, 8, 19290–19299. [Google Scholar] [CrossRef]
  36. Wen, C.; Wang, C.; Zhang, Y.; Antonov, S.; Xue, D.; Lookman, T.; Su, Y. Modeling solid solution strengthening in high entropy alloys using machine learning. Acta Mater. 2021, 212, 116917. [Google Scholar] [CrossRef]
  37. Li, S.; Zhang, Y.; Hu, Y.; Wang, B.; Sun, S.; Yang, X.; He, H. Predicting metal-organic frameworks as catalysts to fix carbon dioxide to cyclic carbonate by machine learning. J. Mater. 2021, 7, 1029–1038. [Google Scholar] [CrossRef]
  38. Wang, X.; Xiao, B.; Li, Y.; Tang, Y.; Liu, F.; Chen, J.; Liu, Y. First-principles based machine learning study of oxygen evolution reactions of perovskite oxides using a surface center-environment feature model. Appl. Surf. Sci. 2020, 531, 147323. [Google Scholar] [CrossRef]
  39. Chen, A.; Zhang, X.; Zhou, Z. Machine learning: Accelerating materials development for energy storage and conversion. InfoMat 2020, 2, 553–576. [Google Scholar] [CrossRef]
  40. Jiang, X.; Wang, Y.; Jia, B.; Qu, X.; Qin, M. Prediction of oxygen evolution activity for NiCoFe oxide catalysts via machine learning. ACS Omega 2022, 7, 14160–14164. [Google Scholar] [CrossRef]
  41. Ahmed, M.G.; Tay, Y.F.; Chi, X.; Zhang, M.; Tan, J.M.R.; Chiam, S.Y.; Rusydi, A.; Wong, L.H. Efficient Ternary Mn-Based Spinel Oxide with Multiple Active Sites for Oxygen Evolution Reaction Discovered via High-Throughput Screening Methods. Small 2023, 19, 2204520. [Google Scholar] [CrossRef] [PubMed]
  42. The Royal Society of Chemistry’s Interactive Periodic Table Database. Available online: https://www.rsc.org/periodic-table/ (accessed on 11 January 2024).
Figure 1. Original dataset visualization by categorical plots of Fe, Co, Mn, and overpotential.
Figure 1. Original dataset visualization by categorical plots of Fe, Co, Mn, and overpotential.
Catalysts 14 00513 g001
Figure 2. Heat map of Pearson correlation coefficients between feature parameters and OP.
Figure 2. Heat map of Pearson correlation coefficients between feature parameters and OP.
Catalysts 14 00513 g002
Figure 3. (a) Mean absolute errors for different models during model selection. (b) Mean squared errors for different models during model selection.
Figure 3. (a) Mean absolute errors for different models during model selection. (b) Mean squared errors for different models during model selection.
Catalysts 14 00513 g003
Figure 4. Machine-learning model by RF: (a) diagonal scatter plot for the predicted OP and the ground truth by RF; (b) contour map of the predicted overpotential of the RF model under different compositions.
Figure 4. Machine-learning model by RF: (a) diagonal scatter plot for the predicted OP and the ground truth by RF; (b) contour map of the predicted overpotential of the RF model under different compositions.
Catalysts 14 00513 g004
Figure 5. (a) Current–potential curves of Fe12Co40Mn48Ox, Fe30Co40Mn30Ox, Fe30Co20Mn50Ox, Fe50Co50Ox and Co50Mn50Ox after the iR correction measured in 1.0 mol L−1 KOH. (b) Bar plot of the experimental and predicted overpotentials by the RF model.
Figure 5. (a) Current–potential curves of Fe12Co40Mn48Ox, Fe30Co40Mn30Ox, Fe30Co20Mn50Ox, Fe50Co50Ox and Co50Mn50Ox after the iR correction measured in 1.0 mol L−1 KOH. (b) Bar plot of the experimental and predicted overpotentials by the RF model.
Catalysts 14 00513 g005
Figure 6. Distributions of SHAP values with feature values.
Figure 6. Distributions of SHAP values with feature values.
Catalysts 14 00513 g006
Table 1. Material physical features, abbreviations, units and transformation formulas.
Table 1. Material physical features, abbreviations, units and transformation formulas.
FeaturesAbbreviationUnitFormula
valence electron numberVEN (1) X ¯ = i F e , C o , M n C i P i (2) δ X = i F e , C o , M n P i X ¯ 2 C i
relative atomic massRAM
atomic numberAN
non-boned atomic radiusRAÅ
covalent radiusRCÅ
first ionization energyFIEkJ mol−1
Pauling scale electronegativityEP
the number of electrons in the outermost d orbitalDE
Table 2. Composition and overpotentials of different ratios of FeCoMn oxides.
Table 2. Composition and overpotentials of different ratios of FeCoMn oxides.
SampleFeCoMnExperimental (mV)Predicted (mV)
exp-10.120.40.48344369.3
exp-20.30.40.3367389.1
exp-30.30.20.5372415.6
exp-40.50.50391442.2
exp-500.50.5356385.6
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, L.; Hou, J.; Ji, H.; Meng, D.; Qi, J.; San, X. Prediction of Oxygen Evolution Activity for FeCoMn Oxide Catalysts via Machine Learning. Catalysts 2024, 14, 513. https://doi.org/10.3390/catal14080513

AMA Style

Zhang L, Hou J, Ji H, Meng D, Qi J, San X. Prediction of Oxygen Evolution Activity for FeCoMn Oxide Catalysts via Machine Learning. Catalysts. 2024; 14(8):513. https://doi.org/10.3390/catal14080513

Chicago/Turabian Style

Zhang, Lei, Jinfei Hou, Honglin Ji, Dan Meng, Jian Qi, and Xiaoguang San. 2024. "Prediction of Oxygen Evolution Activity for FeCoMn Oxide Catalysts via Machine Learning" Catalysts 14, no. 8: 513. https://doi.org/10.3390/catal14080513

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