Next Article in Journal
Empirical Study of Stability and Fairness of Schemes for Benefit Distribution in Local Energy Communities
Previous Article in Journal
A Survey on the Application of Machine Learning in Turbulent Flow Simulations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Research on Complex Lithology Identification Based on Well Logs: A Case Study of Lower 1st Member of the Shahejie Formation in Raoyang Sag

1
Shandong Provincial Key Laboratory of Deep Oil and Gas, Qingdao 266580, China
2
School of Geosciences, China University of Petroleum (East China), Qingdao 266580, China
3
Exploration and Development Research Institute of Daqing Oilfield Co., Ltd., Daqing 163712, China
*
Author to whom correspondence should be addressed.
Energies 2023, 16(4), 1748; https://doi.org/10.3390/en16041748
Submission received: 25 November 2022 / Revised: 4 February 2023 / Accepted: 8 February 2023 / Published: 9 February 2023

Abstract

:
Lithology identification is the basis for sweet spot evaluation, prediction, and precise exploratory deployment and has important guiding significance for areas with low exploration degrees. The lithology of the shale strata, which are composed of fine-grained sediments, is complex and varies regularly in the vertical direction. Identifying complex lithology is a typical nonlinear classification problem, and intelligent algorithms can effectively solve this problem, but different algorithms have advantages and disadvantages. Compared were the three typical algorithms of Fisher discriminant analysis, BP neural network, and classification and regression decision tree (C&RT) on the identification of seven lithologies of shale strata in the lower 1st member of the Shahejie Formation (Es1L) of Raoyang sag. Fisher discriminant analysis method is linear discriminant, the recognition effect is poor, the accuracy is 52.4%; the accuracy of the BP neural network to identify lithology is 82.3%, but it belongs to the black box and can not be visualized; C&RT can accurately identify the complex lithology of Es1L, the accuracy of this method is 85.7%, and it can effectively identify the interlayer and thin interlayer in shale strata.

1. Introduction

The majority of continental shale formations were generated by lake basin sedimentation and are characterized by a variety of rock types and frequent lithological shifts in the longitudinal direction. Lithology contains much geological information, which can reflect the geological characteristics of the reservoir [1]. The lithology of rocks can be accurately determined by analyzing the characteristics of rock samples. However, rock sample analysis is expensive and time-consuming. As a result, this method cannot be widely used in oil and gas exploration [2]. Well logs are an effective means of reservoir evaluation and an important way of lithology identification. It can quantitatively characterize depositional processes and reflect stratigraphic information at different depths based on precise vertical resolution [3].
In the traditional interpretation of well logs data, lithology definition is usually performed by experts with relevant professional knowledge, which greatly limits the development of the industry. Well logs are composed of a large amount of formation information, and data statistics can effectively sort out this information. Combining well logs and data statistics is an effective method to solve the problem. Density, neutron, and sonic logs are used to describe porosity. The lithology of conventional reservoirs can be identified based on using porosity logs to build cross plots separately [4]. With the change in exploration target, the lithology of the target interval becomes more complex, and the conventional mathematical statistics method can no longer meet the research needs. Principal component analysis, linear discriminant analysis, and Bayesian classifier were used to process multi-dimensional well logs data and identify lithology [5,6,7]. In unconventional oil and gas exploration, especially in the continental shale formations dominated by fine-grained sediments, the boundaries for dividing lithology are not obvious [8] and the well logs responses are similar. In addition, fluid type and saturation, salinity, fractures, and diagenesis can also produce log responses similar to those caused by lithology [9,10]. All the above are unfavorable factors for lithology identification.
Intelligent algorithms are more prominent in solving complex nonlinear recognition problems, and many scholars have also carried out related research. These algorithms include support vector machines (SVMs) [11,12,13], random forests (RFs) [7,14], boosted trees [15,16], decision trees (DTs) [17], K-means clustering algorithms [18], artificial neural networks (ANNs) [19,20], and fuzzy theory [21]. Some scholars have combined multiple algorithms to solve the identification problem of complex lithology [2,3,22,23]. Ren et al. (2022) processed logging data, including cluster analysis based on the K-means clustering algorithm and fuzzy processing based on triangular membership function, and then used the decision tree algorithm to build a model for lithology identification. Some scholars also believe that intelligent algorithms are black boxes that cannot be visualized to some extent [24]. Dong et al. [10,25] improved the linear discriminant analysis to maintain the characteristics of visualization and solve the complex lithology identification problem.
Fisher discriminant analysis (FDA) can identify the lithology of carbonate reservoirs with nonlinear characteristics [10]. The back propagation (BP) neural network has self-learning ability, particularly promotion, and generalization [26]. It can also be used for lithology identification. Lithology identification is a classification problem that can be efficiently solved by decision tree algorithms [23]. Most of the previous applications of the above methods are based on conventional reservoirs, and there is a lack of research on terrestrial shale formations where shale oil and gas resources are developed. The main rock type of the continental shale strata of fine-grained sedimentary rocks (FGSRs). Fine-grained sedimentary rocks have small particle size and complex mineral composition characteristics and are widely developed in the lower 1st member of the Shahejie Formation (Es1L) [26,27]. This study identifies the lithology of FGSR based on FDA, BP neural network, and decision tree algorithm. By comparing the identification accuracy of the three methods, the lithology identification method suitable for the Es1L is preferred. It is expected that accurately identifying the lithology of the Es1L can provide some guidance for oil and gas exploration and development. In addition, this research approach can also be applied to the lithology identification of other strata.

2. Geological Setting and Sample

The Raoyang Sag is located in the Bohai Bay Basin in eastern China [28]. The Es1L is the target interval of the study. It was formed during the sedimentary period of the extension of the lake basin. Shallow water deltas and deep lakes and semi-deep lakes are the main sedimentary microfacies in the Es1L [26,27] (Figure 1a). The shallow water delta is located in the lower part of the Es1L, and coarse-grained clastic rocks dominate its lithology. The lithology of deep and semi-deep lake microfacies includes dark gray mudstone, oil shale, biological limestone, argillaceous limestone, dolomite, calcareous mudstone, and thin sandstone [27] (Figure 1b). These lithologies frequently interstratify, making lithology identification much more challenging and restricting the development of shale oil and gas exploration. The lithology of 147 core samples from 13 wells was determined by core observation, thin section identification, and X-ray diffraction analysis. According to the needs of the production site, the lithology of the Es1L is divided into sandstone (37), sandy mudstone (6), mudstone (24), Calcareous mudstone (12), dolomite (17), limestone (3), and shale (48) (Figure 2). In addition, the logging curves used in this study include acoustic velocity wave (AC), spontaneous potential (SP), natural gamma logging (GR), caliper logging (CAL), resistivity logging, compensated neutron porosity logging (CNL), and density logging (DEN).

3. Method

3.1. Preprocessing

The quality of the training samples affects the accuracy of the prediction results [29]. The depth does not match the actual sampling point depth and the different correlations, splicing, and difference in the order of magnitude of the logging curves are all unfavorable factors that affect the prediction results. Core sample depth homing and logging curve splicing are preprocessing methods used in logging evaluation. Input curves have different units and degrees of variation, leading to large differences in predictions. To eliminate the influence of the dimension and the various size and magnitudes of the variable itself and to make the curves comparable, the input curves need to be standardized [23,30]. The log curve with approximate linear characteristics is processed by linear normalization, and the processing formula is expressed as:
X i = X i X m i n X m a x X m i n
Logarithmic normalization is used for logging curves of nonlinear characteristics such as resistivity, and the formula is expressed as:
X i = lg X i lg X m i n l g X m a x lg X m i n
where Xi is the normalized logging curve value; X i is the original logging value; X m a x and X m i n are the maximum and minimum values of the logging curve in the study interval.
Different types of logs have different responses to lithology, and selecting a log with a good response has a better effect on lithology identification [10]. The degree of correlation between categorical variables (lithology) and interval variables (logging curve) can be expressed by correlation ratios (E2). The formula for calculating E2 is expressed as:
E 2 = ( Y Y ¯ ) 2 ( Y Y k ¯ ) 2 ( Y Y ¯ ) 2
where Y represents the value of the interval variable, Y ¯ represents the mean of the interval variable, Yk represents the mean of the k-th category-spaced variable.
Usually, when E2 is less than 0.06, the two variables are weakly correlated, and when E2 is greater than 0.16, they are strongly correlated. From Figure 3, it can be found that the E2 of the different logging curves of the Es1L is quite different, and AC, SP, and CNL are strongly correlated with lithology. The E2 of the 2.5 m bottom gradient (R25) is 0.15, close to a strong correlation. The resistivity has reference significance for the identification of lithology. Therefore, AC, SP, CNL, and R25 are used as input parameters for lithology identification.

3.2. The FDA Principle

FDA, also known as linear discriminant analysis, is a classic classification discriminant method. It transforms data from multi-dimensional to low-dimensional through specific mathematical methods [25,31,32]. The core idea of the FDA is to minimize the distance between the same group and maximize the distance between different groups (Figure 4). Converting the idea into a mathematical formula can be expressed as:
J ( w ) = w T S b w w T S w w
where w is the projection vector, Sb is the between-class scatter matrix, Sw is the within-class scatter matrix, and their mathematical expressions are as follows:
S b = i p 1 k = i + 1 p N i N N k N ( m i m k ) ( m i m k ) T
S w = i = 1 p j = 1 n i 1 N ( x j i m i ) ( x j i m i ) T
where xj(i) represents the j-th sample in the i-th category; there are P elements in the vector x(i); ni is the number of samples in the i-th category and it satisfies i = 1 p n i = N ; and mi represents the centroid of the i-th sample, m i = 1 n i j = 1 n i x j ( i ) .
Previous research [25] has proved that the optimal solution of m is equivalent to finding the eigenvector of Sbu = λSwu. The expression J(w) can be obtained by bringing the eigenvector corresponding to the largest eigenvalue into the Formula (4).
There are seven lithologies to be identified in this study, and AC, SP, CNL, and R25 are used as input parameters. For each sample, x j ( i ) = ( A C , S P , R 25 , C N L ) T , i = 1, 2, …, 7; j = 1, 2, …, n147, p = 4. The eigenvalues and the corresponding canonical functions are determined, and the two canonical functions with the largest eigenvalues are selected as the eigenvectors for lithology discrimination. w 1 = ( w 11 , w 12 , w 13 , w 14 ) T , and w 2 = ( w 21 , w 22 , w 23 , w 24 ) T are the eigenvectors corresponding to the eigenvalues. Then, the projection vector is z j i = ( w 1 T x i i , w 2 T x i i ) T . Here
z 1 = A C × w 11 + S P × w 12 + R 25 × w 13 + C N L × w 14
z 2 = A C × w 21 + S P × w 22 + R 25 × w 23 + C N L × w 24
So far, the discriminant functions of seven kinds of lithology are established, and the logging curve is brought into the discriminant function. The one with the largest function value is this kind of lithology.

3.3. The BP Neural Network

The BP neural network is a multi-layer feedforward neural network trained by the error backpropagation algorithm, comprising an input layer, output layer, and hidden layer [33]. It has self-learning ability, particularly promotion, generalization, and self-adaptive ability, and is one of the most widely used neural network models [26]. The basic principle of the BP neural network for predicting mineral components is to find a mapping relationship by continuously modifying the network weights and correction threshold until a satisfactory accuracy is obtained [34,35]. The most basic algorithm of the BP neural network is the steepest descent method, including the forward propagation of the signal and the backpropagation of the error.
In the multi-layer forward propagation process, the output formula of each hidden layer unit is as follows:
Y j = f ( i = 1 n ( w i j x i ) + b j )
where f = 1 1 + e x is the activation function, wij represents the input weight, bi represents the offset, xi represents the output of the input layer, and Yj represents the output result of the j-th layer.
In the multi-layer forward propagation process, the output formula of each unit of the output layer is as follows:
Y k = f ( i = 1 n ( w i k x i ) + b k )
where Yk represents the output result of the output layer, wjk is the weight from the j-th hidden layer neuron to the k-th output layer neuron, xj represents the output of the hidden layer, and bk represents the neuron offset of the hidden layer.
In error backpropagation, the output error signal of each layer of neurons is calculated from the output layer to the input layer. Through the error gradient descent method, the weights and thresholds of each layer are adjusted to minimize the mean square error (MSE). MSE can be expressed as:
M S E = 1 2 n i n E i 2
where Ei is the error of the i-th input data, and n represents the amount of input data.
The adjustment formula of weight is as follows:
w i ( j + 1 ) = ϕ δ j Y k Y i + a Δ w i j
where wi(j+1) represents the weights from the i-th input layer to the j+1-th hidden layer, ϕ represents the learning step, δj represents the error, Yi represents the result of the input layer, and a represents the momentum factor. Δwij is the output result from the i-th input layer to the j-th hidden layer. The momentum factor a is to prevent partial error minimization, using momentum to slide past these minima.
The adjustment formula for the threshold is as follows:
b j + 1 = a b j + ϕ δ j + 1
where bj+1 is the threshold of the j+1-th hidden layer, and bj represents the threshold of the j-th hidden layer.
AC, SP, CNL, and R25 are used as input parameters. Seven lithologies are coded separately: sandstone is 1, sandy mudstone is 2, mudstone is 3, calcareous mudstone is 4, dolomite is 5, limestone is 6, and shale is 7. A total of 128 sets of data are randomly selected as training samples, and the weights are continuously updated for training to determine the network structure. In this study, the hidden layer contains ten neurons, and the value of the momentum factor is 0.95. The process of identifying lithology based on the BP neural network is shown in Figure 5. The remaining data are used as a test sample to test the model’s accuracy independently.

3.4. The Classification and Regression Tree (C&RT)

C&RT, as a classification model, was proposed by Breiman et al. in 1984 [36], and it is an efficient method for solving classification problems. The branch criterion is the core criterion of decision tree classification, and the change in the branch criterion produces different types of decision trees. The branching criterion of C&RT is established based on the Gini coefficient. The Gini (G) coefficient represents the probability that a randomly selected sample in the sample set is misclassified. It can be expressed as:
G ( p ) = 1 k = 1 k p k 2
where k represents the number of categories, Pk is the frequency with which the k-th class appears in the classification results.
For a given sample set, D, its Gini coefficient is expressed as:
G ( D ) = 1 k = 1 k ( | C k | | D | ) 2
where Ck is the number of samples in D that belong to class k.
When feature A takes the value of a, D is divided into two parts, D1 and D2. Under the condition of feature A, the G of D is defined as:
G ( D , A ) = | D 1 | | D | G ( D 1 ) + | D 2 | | D | G ( D 2 )
Overfitting is an unavoidable problem of C&RT, and pruning is the primary solution [17,37]. Pruning includes pre-pruning and post-pruning. Pre-pruning means stopping the tree from growing before the decision tree is fully formed. Post-pruning generates an initial decision tree according to the largest scale and prunes layer by layer from bottom to top according to certain rules.
The schematic diagram of C&RT identifying lithology is shown in Figure 6. Log data and lithology data of 147 samples were used as input variables and target parameters, respectively. To prevent overfitting, stopping rule was set, where the minimum number of records in the parent branch is 2%, and the maximum risk difference is 1 and set the depth of the tree to 5. The samples were randomly divided into a training set (131) and a test set (16).

4. Results

The canonical function characteristics of the Es1L are shown in Table 1. The eigenvalues of function 1 and function 2 are high, and the cumulative contribution rate is 92.7%, which contains most of the lithological information. As a result, function 1 and function 2 are selected as the characteristic variables of lithology discrimination, and the discriminant function of lithology is constructed. The discriminant function is as follows:
{ y s a n d s t o n e = 198.445 × A C + 13.132 × S P + 149.108 × R 25 30.769 × C N L 89.227 y s a n d y , m u d s t o n e = 203.262 × A C + 3.529 × S P + 157.391 × R 25 24.92 × C N L 94.781 y m u d s t o n e = 211.363 × A C + 11.009 × S P + 155.451 × R 25 22.273 × C N L 100.783 y c a l c a r e o u , s m u d s t o n e = 222.459 × A C + 13.391 × S P + 159.739 × R 25 24.898 × C N L 110.158 y d o l o m i t e = 213.758 × A C + 12.752 × S P + 163.179 × R 25 25.559 × C N L 107.449 y l i m e s t o n e = 223.749 × A C + 6.023 × S P + 151.782 × R 25 30.619 × C N L 100.398 y s h a l e = 225.139 × A C + 12.258 × S P + 167.792 × R 25 25.159 × C N L 114.750
The lithology identification results of FDA are shown in the Table 2 and Figure 7a, the identification effect of shale and sandstone is good, and the accuracy is 83.3% and 81.1%, respectively. Sandy mudstone and limestone have lower accuracy, both below 50%. The identification of other lithologies is not suitable. Calcareous mudstone and dolomite were not correctly classified. A total of 66.7% of calcareous mudstone and 70.6% of dolomite were wrongly classified as shale. The FDA’s overall accuracy in identifying lithology was 52.4%.
The accuracy rate of the BP neural network in identifying the Es1L was 82.3%. The accuracy rate of 19 of these test samples was 89.4%. Limestone and sandstone can be accurately identified by BP neural network with an accuracy of 100% and 97.3%, respectively, and shale, mudstone, and dolomite have an accuracy of 85.4%, 79.2%, and 76.5%, respectively, which can be well-identified. The calcareous mudstone has a poor recognition effect. The calcareous mudstone is wrongly divided into sandstone, mudstone, and shale (Figure 7b).
In the C&RT model, sandstone, shale, mudstone, and dolomite can be well-identified, with an accuracy of more than 75%, and the highest accuracy of sandstone is 97.3% (Figure 7c). Only half of the sandy mudstone is correctly identified, and the wrong part is mainly classified as sandstone. A total of 16 test samples were used, including sandstone, sandy mudstone, mudstone, calcareous mudstone, and dolomite. The CR&T model’s prediction accuracy was 88%. The accuracy rate of the C&RT model in identifying lithology was 85.7%.

5. Discussion

5.1. Logging Characteristics of Lithology

Lithology describes the physical properties of rock, such as color, texture, grain size, and composition [38]. Differences in these physical properties produce different response characteristics in logging, and analyzing these characteristics can invert the distribution of various subsurface lithologies [39,40]. The following is a brief description of the response characteristics of the well logs used in this study:
(1) AC is the recorded time it takes for a sound wave to travel at a fixed interval between formations. The formation contains not only rock but also water, CO2, CH4, and hydrocarbon fluids (oil and bitumen) in the pores of the rock [41,42,43,44]. Sound waves travel faster in solids such as rocks and minerals than in other fluids such as water and oil. Even the difference in lithology in the rock affects the propagation velocity of sound waves [10]. The P-wave velocity of mudstone, sandstone, limestone, and dolomite is 1800 m/s, 2130~5180 m/s, 3950~4900 m, 4000~5650 m/s, and 4600~6100 m/s, respectively [45], among which the porosity of sandstone, limestone, and dolomite is between 5% and 20%.
(2) SP represents the natural potential difference between the trajectory of the wellbore and the ground [46]. SP has a good effect in identifying sandstone with developed pores and mudstone. Tight rocks such as calcareous mudstone and shale often have higher electrical resistance than mudstone. However, SP is also affected by the type of reservoir fluid, salinity, and saturation. Using SP to judge lithology often needs to be combined with other logging.
(3) GR is recording the total natural gamma radiation intensity in the rock with a gamma-ray detector [47]. Uranium, thorium, and potassium play a decisive role in natural gamma radioactivity in rocks. Minerals such as gypsum (CaSO4·H2O), anhydrite (CaSO4), quartz (SiO2), dolomite (CaMgCO3), and calcite (CaCO3) in sedimentary rocks have low radioactivity, and with the increase in clay content, the radioactivity becomes stronger. Among the clay minerals, montmorillonite has a large surface area, a strong ability to adsorb radioactive substances, and contains more uranium oxide, which is the main contributor to radioactivity; potassium in illite can also adsorb uranium oxide and is radioactive.
(4) CAL measures the change in borehole diameter with depth, which is related to lithology and the mud used for drilling [48]. The well diameter of sandstone with good permeability is smaller than that of the drill bit. On the contrary, the well diameter of easily collapsed mudstone and carbonate rock with developed pores is larger than the drill bit. Lithology is classified based on these features. As a result, the change in well diameter is also related to cementing quality and wellbore azimuth, and lithology identification needs to be combined with other well logs.
(5) In general, sandstone with high resistance is sandstone, mudstone with low resistance, and reservoirs containing fluids are also low resistivity. However, many factors affect resistivities, such as mud invasion, electrode spacing, and wellbore inclination [49]. Therefore, it is mainly used for qualitative analysis, and further analysis needs to be combined with other data.
(6) CNL measures the amount of hydrogen because hydrogen slows down high-energy and fast neutrons [46]. Sandstone, dolomite, limestone, and migmatites are commonly found in reservoirs. The minerals that make up these rocks have no hydrogen, and most of the hydrogen is in the fluids in the pores of the rocks. Among them, the deceleration capacity of saturated freshwater sandstone is less than that of limestone, and the deceleration capacity of saturated freshwater limestone is lower than that of dolomite [9]. The pores of mudstone and shale contain bound water, and clay minerals contain crystal water. The higher the clay content, the higher the hydrogen content.
(7) DEN irradiates the formation with gamma rays and measures the bulk density of the formation according to the Compton effect [47]. The density of a liquid is much lower than that of solid minerals. The densities of several common minerals are quartz (2.65 g/cm3), calcite (2.71 g/cm3), dolomite (2.87 g/cm3), gypsum (2.32 g/cm3), and anhydrite (2.96 g/cm3). The reference densities of the corresponding rocks for these minerals are sandstone (2.644 g/cm3), limestone (2.710 g/cm3), dolomite (2.877 g/cm3), gypsum (2.355 g/cm3), and anhydrite (2.960 g/cm3). The porosity of these rocks is close to zero.
The logging characteristics of the lithology described above are ideal, but the actual formation is often more complex. In particular, the continental lake basins are affected by severe sedimentation and diagenesis, with strong inorganic heterogeneity [26,50], mostly mixed rocks, and blurred lithologic boundaries. Mudstone and shale strata are widely developed in the Es1L, mainly mudstone and shale, with dolomite, limestone, sandstone, and migmatite interlayers. Therefore, the rock density, radioactive element content, and well diameter of the whole section have little change. Combined with mathematical analysis, AC, SP, CNL, and R25 were selected as input parameters. The direct relationship between the four logging curves is poor, and the lithology is not classified on the intersection of the four logging curves (Figure 8). Clay minerals are rich in crystalline water, adsorbed water, and oil in the pores of mudstone [51], and mudstone has a high CNL value. Shale, calcareous mudstone, and some mudstones have a high degree of overlap, and the mineral compositions of the three are similar and difficult to distinguish. Effective identification of complex lithologies is therefore required with the help of some other measures.

5.2. Comparison of the Three Models

Table 2 shows the differences in the prediction results of the three methods. Shale and sandstone can be accurately identified by FDA, BP neural network, and C&RT, and the identification accuracy is above 80% (Figure 7). Sandy mudstone and calcareous mudstone are mixed rocks with poor recognition effects. C&RT has the highest accuracy in identifying sandy mudstone and calcareous mudstone, with 50.0% and 66.7% accuracy rates, respectively (Figure 7). BP neural network and C&RT have the same results in identifying dolomite, while FDA cannot identify dolomite. The lithology recognition accuracies of FDA, BP neural network, and C&RT models are 52.4%, 82.3%, and 85.7%, respectively (Figure 7). FDA cannot be effectively applied to the identification of complex lithology. The complicated lithology of continental shale formations can be more accurately identified by BP neural network and C&RT; however, BP neural network is a “black box” and cannot be seen. This issue can be solved more effectively with C&RT since each step of the identification process can be presented and adjusted to the demands. C&RT is a lithology identification method therefore suitable to the Es1L.

5.3. Application

Well B11x is located in the Xiliu area of Raoyang sag. It is a typical shale oil exploration well in Raoyang sag. Using the established BP neural network and C&RT model to identify lithology, the lithology distribution characteristics of the Es1L of well B11x were analyzed. The Es1L of well B11x is dominated by mudstone and shale, with interlayers of calcareous mudstone, sandstone, and dolomite. The core lithology in Figure 9 is obtained from mud logging data. These data are greatly influenced by artificial subjective, especially in the continental shale strata with heterogeneity and strength, and the accuracy of the data remains to be discussed. Nonetheless, it can also be used as a reference point to advise on some geological work. The lithological identification results of the BP neural network show that sandstone can be effectively identified. However, shale, mudstone, calcareous mudstone, and dolomite with similar mineral compositions cannot be distinguished. At 3512.91~3513.13 m, the rock core shows mudstone, while the prediction result of the BP neural network is shale. The bedding fractures of shale have larger storage space [52], which reduces the value of AC. The natural potential of this section does not change much, and the value of AC is high. Combined with the photos of the rock core (Figure 10), the rock of this section is mudstone. Figure 10b shows that the upper bedding at 3547.90~3548.06 m is developed as shale, and the lower part is dolomite. At 3564.06~3564.36 m, the rock has a blocky structure without bedding.
The recognition result of C&RT is better than that of the BP neural network, especially in the recognition of interlayers and thin interlayers, which are usually favorable intervals for the exploration and development of shale oil. The lithology identified by C&RT is highly similar to the lithology obtained from mud logging, but there are also differences. For example, at 3509.37~3509.63 m, the core lithology shows calcareous mudstone, while the thin section identification result is shale, and the recognition result of C&RT is the same. The acoustic transit time of this section is small, the resistivity is low, and the value of CNL is high, indicating that it has a larger storage space and more water. This response is usually due to interlayer fractures in the shale in tight reservoirs. The acoustic time difference of carbonate rocks is higher than that of tight shale. Carbonate rocks can form reservoir spaces containing fluids due to dissolution [53]. Fluids and primary spaces increase the hydrogen content and reduce resistivity. At 3547.90~3548.06m, the values of AC and CNL are high, and the resistivity is low. Combined with thin section identification, the lithology is dolomite.
The accuracy of the BP neural network identification lithology model is lower than that of the C&RT model in identifying mixed lithology. Although the accuracy of the two models differs by only 3.4%, it can be found in the process of model application that identifying each lithology by C&RT is reasonable. Therefore, it is practical to use C&RT to identify lithology, and it can provide a reference for the exploration and development of shale oil in shale intervals.

6. Conclusions

The identification of complex lithology is a typical nonlinear classification problem. Intelligent algorithms can effectively solve this problem. Different algorithms have different identification effects on different lithologies. Through FDA, BP neural network, and C&RT to identify the complex lithology of the mudstone and shale strata in the Es1L of Raoyang Sag, the following conclusions are drawn:
(1) Each logging has a different sensitivity to lithology. The four logs AC, SP, CNL, and R25 of the Es1L the Raoyang sag have the highest sensitivity to seven lithologies and are used as input parameters for lithology identification.
(2) The intersection graph method cannot distinguish complex lithology. FDA, BP neural network, and C&RT have an accuracy of 52.4%, 82.3%, and 85.7%, respectively, in identifying the Es1L. The accuracy of the BP neural network in identifying mixed lithology is lower than that of C&RT, and the accuracy of C&RT in identifying sandstone, shale, and mudstone is 97.3%, 89.6%, and 87.5%, respectively.
(3) C&RT can identify interlayers and thin interlayers in shale strata. The C&RT identification results of well B11x show that the lithology is dominated by shale, with mudstone, sandstone, calcareous mudstone, and dolomite interlayers.

Author Contributions

Conceptualization, Z.S.; methodology, Z.S. and D.X.; formal analysis, Z.S.; investigation, X.W. and J.T.; data curation, Y.W. and R.Z.; writing—original draft preparation, Z.S.; writing—review and editing, D.X. and Y.W.; supervision, R.Z. and J.T.; All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation (grant nos. 41972123, 41972139, and 41922015).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Selley, R.C. An Introduction to Sedimentology; Academic Press: London, UK, 1976. [Google Scholar]
  2. Tian, M.; More, H.; Xu, H. Inversion of well logs into lithology classes accounting for spatial dependencies by using hidden markov models and recurrent neural networks. J. Petrol. Sci. Eng. 2021, 196, 107598. [Google Scholar] [CrossRef]
  3. Li, Z.; Kang, Y.; Feng, D.; Wang, X.-M.; Lv, W.; Chang, J.; Zheng, W.X. Semi-supervised learning for lithology identification using Laplacian support vector machine. J. Petrol. Sci. Eng. 2020, 195, 107510. [Google Scholar] [CrossRef]
  4. Burke, J.A.; Campbell, R.L.; Schmidt, A.W. The litho-porosity cross plot a method of determining rock characteristics for computation of log data. In Proceedings of the SPE Illinois Basin Regional Meeting, Evansville, IN, USA, 30–31 October 1969. [Google Scholar]
  5. Busch, J.M.; Fortney, W.G.; Berry, L.N. Determination of lithology from well logs by statistical analysis. Spe Form. Eval. 1987, 2, 412–418. [Google Scholar] [CrossRef]
  6. Liu, A.; Zuo, L.; Li, J.; Li, R.; Zhang, R. Application of principal component analysis in carbon ate lithology identification: A case study of the Cambrian carbonate reservoir in YH field. Oil Gas Geol. 2013, 34, 192. [Google Scholar]
  7. Xie, Y.; Zhu, C.; Zhou, W.; Li, Z.; Liu, X.; Tu, M. Evaluation of machine learning methods for formation lithology identification: A comparison of tuning processes and model performances. J. Petrol. Sci. Eng. 2018, 160, 182–193. [Google Scholar] [CrossRef]
  8. Jiang, Z.; Liang, C.; Wu, J.; Zhang, J.; Zhang, W.; Wang, Y.; Liu, H.; Chen, X. Several issues in sedimentological studies on hydrocarbon-bearing fine-grained sedimentary rocks. Acta Petrol. Sin. 2013, 34, 1031–1039. [Google Scholar]
  9. Carrasquilla, A.; Lima, R. Basic and specialized geophysical well logs to characterize an offshore carbonate reservoir in the Campos Basin, southeast Brazil. J. S. Am. Earth. Sci. 2020, 98, 102436. [Google Scholar] [CrossRef]
  10. Dong, S.; Zeng, L.; Du, X.; He, J.; Sun, F. Lithofacies identification in carbonate reservoirs by multiple kernel Fisher discriminant analysis using conventional well logs: A case study in A oilfield, Zagros Basin, Iraq. J. Petrol. Sci. Eng. 2021, 210, 110081. [Google Scholar] [CrossRef]
  11. Sebtosheikh, M.A.; Salehi, A. Lithology prediction by support vector classifiers using inverted seismic attributes data and petrophysical logs as a new approach and investigation of training data set size effect on its performance in a heterogeneous carbonate reservoir. J. Petrol. Sci. Eng. 2015, 134, 143–149. [Google Scholar] [CrossRef]
  12. Bhattacharya, S.; Carr, T.R.; Pal, M. Comparison of supervised and unsupervised approaches for mudstone lithofacies classification: Case studies from the Bakken and Mahantango-Marcellus Shale, USA. J. Nat. Gas Sci. Eng. 2016, 33, 1119–1133. [Google Scholar] [CrossRef]
  13. Deng, C.; Pan, H.; Fang, S.; Konaté, A.A.; Qin, R. Support vector machine as an alternative method for lithology classification of crystalline rocks. J. Geophys. Eng. 2017, 14, 341–349. [Google Scholar] [CrossRef]
  14. Sun, J.; Li, Q.; Chen, M.; Ren, L.; Huang, G.; Li, C.; Zhang, Z. Optimization of models for a rapid identification of lithology while drilling—A win-win strategy based on machine learning. J. Petrol. Sci. Eng. 2019, 176, 321–341. [Google Scholar] [CrossRef]
  15. Merembayev, T.; Yunussov, R.; Yedilkhan, A. Machine learning algorithms for classification geology data from well logging. In Proceedings of the International Conference on Electronics Computer and Computation, Kaskelen, Kazakhstan, 29 November–1 December 2018. [Google Scholar]
  16. Dev, V.A.; Eden, M.R. Evaluating the Boosting Approach to Machine Learning for Formation Lithology Classification. Comput. Chem. Eng. 2018, 44, 1465–1470. [Google Scholar] [CrossRef]
  17. Saporetti, C.M.; da Fonseca, L.G.; Pereira, E.; de Oliveira, L.C. Machine learning approaches for petrographic classification of carbonate-siliciclastic rocks using well logs and textural information. J. Appl. Geophys. 2018, 155, 217–225. [Google Scholar] [CrossRef]
  18. Yang, H.; Pan, H.; Ma, H.; Konaté, A.A.; Yao, J.; Guo, B. Performance of the synergetic wavelet transform and modified K-means clustering in lithology classification using nuclear log. J. Petrol. Sci. Eng. 2016, 144, 1–9. [Google Scholar] [CrossRef]
  19. Raeesi, M.; Moradzadeh, A.; Doulati Ardejani, F.; Rahimi, M. Classification and identification of hydrocarbon reservoir lithofacies and their heterogeneity using seismic attributes, logs data and artificial neural networks. J. Petrol. Sci. Eng. 2012, 82–83, 151–165. [Google Scholar] [CrossRef]
  20. Zych, M.; Stachura, G.; Hanus, R.; Szabó, N.P. Application of Artificial Neural Networks in Identification of Geological Formations on the Basis of Well Logging Data—A Comparison of Computational Environments’ Efficiency. In International Seminar of Metrology Methods and Techniques of Signal Processing in Physical Measurements; Springer: Rzeszow, Poland, 2018. [Google Scholar]
  21. Maia Ramos Lopes, D.; Neves Andrade, A.J. Lithology identification on well logs by fuzzy inference. J. Petrol. Sci. Eng. 2019, 180, 357–368. [Google Scholar] [CrossRef]
  22. Li, S.; Zhou, K.; Zhao, L.; Xu, Q.; Liu, J. An improved lithology identification approach based on representation enhancement by logging feature decomposition, selection and transformation. J. Petrol. Sci. Eng. 2021, 209, 109842. [Google Scholar] [CrossRef]
  23. Ren, Q.; Zhang, H.; Zhang, D.; Zhao, X.; Yan, L.; Rui, J. A novel hybrid method of lithology identification based on k-means++ algorithm and fuzzy decision tree. J. Petrol. Sci. Eng. 2022, 208, 109681. [Google Scholar] [CrossRef]
  24. Gardner, M.W.; Dorling, S.R. Artificial neural networks (the multilayer perceptron)—A review of applications in the atmospheric sciences. Atmos. Environ. 1998, 32, 2627–2636. [Google Scholar] [CrossRef]
  25. Dong, S.; Wang, Z.; Zeng, L. Lithology identification using kernel Fisher discriminant analysis with well logs. J. Petrol. Sci. Eng. 2016, 143, 95–102. [Google Scholar] [CrossRef]
  26. Song, Z.; Li, J.; Li, X.; Chen, K.; Wang, C.; Li, P.; Wei, Y.; Zhao, R.; Wang, X.; Zhang, S.; et al. Coupling Relationship between Lithofacies and Brittleness of the Shale Oil Reservoir: A Case Study of the Shahejie Formation in the Raoyang Sag. Geofluids 2022, 2022, 2729597. [Google Scholar] [CrossRef]
  27. Wei, Y.; Li, X.; Zhang, R.; Li, X.; Lu, S.; Qiu, Y.; Jiang, T.; Gao, Y.; Zhao, T.; Songm, Z.; et al. Influence of a Paleosedimentary Environment on Shale Oil Enrichment: A Case Study on the Shahejie Formation of Raoyang Sag, Bohai Bay Basin, China. Front. Earth Sci.-Prc. 2021, 9, 736054. [Google Scholar] [CrossRef]
  28. Li, X.; Chen, K.; Li, P.; Li, J.; Geng, H.; Li, B.; Li, X.; Wang, H.; Zhang, L.; Wei, Y.; et al. A New Evaluation Method of Shale Oil Sweet Spots in Chinese Lacustrine Basin and Its Application. Energies 2021, 14, 5519. [Google Scholar] [CrossRef]
  29. Aghli, G.; Soleimani, B.; Moussavi-Harami, R.; Mohammadian, R. Fractured zones detection using conventional petrophysical logs by differentiation method and its correlation with image logs. J. Petrol. Sci. Eng. 2016, 142, 152–162. [Google Scholar] [CrossRef]
  30. Asante-Okyere, S.; Shen, C.; Yao, Y.Z.; Rulegeya, M.M.; Zhu, X. A novel hybrid technique of integrating gradient-boosted machine and clustering algorithms for lithology classification. Nat. Resour. Res. 2020, 29, 2257–2273. [Google Scholar] [CrossRef]
  31. Duda, R.O.; Hart, P.E.; Stork, D.G. Pattern Classification; John Wiley & Sons: New York, NY, USA, 2001. [Google Scholar]
  32. Subasi, A.; Gursoy, M.I. EEG signal classification using PCA, ICA, LDA and support vector machines. Expert Syst. Appl. 2010, 37, 8659–8666. [Google Scholar] [CrossRef]
  33. Hornik, K.; Stinchcombe, M.; White, H. Multilayer feedforward networks are universal approximators. Neural Netw. 1989, 2, 359–366. [Google Scholar] [CrossRef]
  34. Wu, D.; Zhang, D.; Liu, S.; Jin, Z.; Chowwanonthapunya, T.; Gao, J.; Li, X. Prediction of polycarbonate degradation in natural atmospheric environment of China based on BP-ANN model with screened environmental factors. Chem. Eng. J. 2020, 399, 125878. [Google Scholar] [CrossRef]
  35. Jawad, J.; Hawari, A.H.; Zaidi, S.J. Artificial neural network modeling of wastewater treatment and desalination using membrane processes: A review. Chem. Eng. J. 2021, 419, 129540. [Google Scholar] [CrossRef]
  36. Lewis, R.J. An introduction to classification and regression tree (CART) analysis. In Proceedings of the Annual Meeting of the Society for Academic Emergency Medicine, San Francisco, CA, USA, 22–25 May 2000. [Google Scholar]
  37. Shi, N.; Li, H.; Luo, W. Data mining and well logging interpretation: Application to a conglomerate reservoir. Appl. Geophys. 2015, 12, 263–272. [Google Scholar] [CrossRef]
  38. Haldar, S.K. Introduction to Mineralogy and Petrology; Elsevier: Amsterdam, The Netherlands, 2020. [Google Scholar]
  39. Zhou, Z.; Wang, G.; Ran, Y.; Lai, J.; Cui, Y.; Zhao, X. A logging identification method of tight oil reservoir lithology and lithofacies: A case from Chang7 Member of Triassic Yanchang Formation in Heshui area, Ordos Basin, NW China. Petrol. Explor. Dev. 2016, 43, 65–73. [Google Scholar] [CrossRef]
  40. Liu, G.; Liu, B.; Huang, Z.; Chen, Z.; Jiang, Z.; Guo, X.; Li, T.; Chen, L. Hydrocarbon distribution pattern and logging identification in lacustrine fine-grained sedimentary rocks of the Permian Lucaogou Formation from the Santanghu basin. Fuel 2018, 222, 207–231. [Google Scholar] [CrossRef]
  41. Li, J.; Lu, S.; Xie, L.; Zhang, J.; Xue, H.; Zhang, P.; Tian, S. Modeling of hydrocarbon adsorption on continental oil shale: A case study on n-alkane. Fuel 2017, 206, 603–613. [Google Scholar] [CrossRef]
  42. Li, J.; Lu, S.; Cai, J.; Zhang, P.; Xue, H.; Zhao, X. Adsorbed and free oil in lacustrine nanoporous shale: A theoretical model and a case study. Energy Fuel 2018, 32, 12247–12258. [Google Scholar] [CrossRef]
  43. Li, J.; Lu, S.; Zhang, P.; Cai, J.; Li, W.; Wang, S.; Feng, W. Estimation of gas-in-place content in coal and shale reservoirs: A process analysis method and its preliminary application. Fuel 2020, 259, 116266. [Google Scholar] [CrossRef]
  44. Li, W.; Lu, S.; Li, J.; Zhang, P.; Wang, S.; Feng, W.; Wei, Y. Carbon isotope fractionation during shale gas transport: Mechanism, characterization and significance. Sci. China Earth Sci. 2020, 63, 674–689. [Google Scholar] [CrossRef]
  45. Cannon, S. Petrophysics: A Practical Guide; John Wiley & Sons: Hoboken, NJ, USA, 2015. [Google Scholar]
  46. Lyu, W.; Zeng, L.; Liu, Z.; Liu, G.; Zu, K. Fracture responses of conventional logs in tight-oil sandstones: A case study of the Upper Triassic Yanchang Formation in southwest Ordos Basin, China. AAPG Bulletin. 2016, 100, 1399–1417. [Google Scholar] [CrossRef]
  47. Tokhmchi, B.; Memarian, H.; Rezaee, M.R. Estimation of the fracture density in fractured zones using petrophysical logs. J. Petrol. Sci. Eng. 2010, 72, 206–213. [Google Scholar] [CrossRef]
  48. Zazoun, R.S. Fracture density estimation from core and conventional well logs data using artificial neural networks: The Cambro-Ordovician reservoir of Mesdar oil field, Algeria. J. Afr. Earth Sci. 2013, 83, 55–73. [Google Scholar] [CrossRef]
  49. Shazly, T.F.; Tarabees, E. Using of Dual Laterolog to detect fracture parameters for Nubia Sandstone Formation in Rudeis-Sidri area, Gulf of Suez, Egypt. Egypt. J. Pet. 2013, 22, 313–319. [Google Scholar] [CrossRef]
  50. Zhang, X.; Pang, X.; Jin, Z.; Hu, T.; Toyin, A.; Wang, K. Depositional model for mixed carbonate-clastic sediments in the Middle Cambrian Lower Zhangxia Formation, Xiaweidian, North China. Adv. Geo-Energy Res. 2020, 4, 29–42. [Google Scholar] [CrossRef] [Green Version]
  51. Li, J.; Wang, S.; Lu, S.; Zhang, P.; Cai, J.; Zhao, J.; Li, W. Microdistribution and mobility of water in gas shale: A theoretical and experimental study. Mar. Petrol. Geol. 2019, 102, 496–507. [Google Scholar] [CrossRef]
  52. Sun, Y.; Ju, Y.; Zhou, W.; Qiao, P.; Tao, L.; Xiao, L. Nanoscale pore and crack evolution in shear thin layers of shales and the shale gas reservoir effect. Adv. Geo-Energy Res. 2022, 6, 221–229. [Google Scholar] [CrossRef]
  53. Guan, M.; Wu, S.; Hou, L.; Jiang, X.; Ba, D.; Hua, G. Paleoenvironment and chemostratigraphy heterogenity of the Cretaceous organic-rich shales. Adv. Geo-Energy Res. 2021, 5, 444–455. [Google Scholar] [CrossRef]
Figure 1. (a) Map showing the structural characteristic within and surround the Raoyang sag and distribution of sampling wells. (b) Stratigraphic column of the Shahejie Formation (modified after Wei et al., 2021 [27]).
Figure 1. (a) Map showing the structural characteristic within and surround the Raoyang sag and distribution of sampling wells. (b) Stratigraphic column of the Shahejie Formation (modified after Wei et al., 2021 [27]).
Energies 16 01748 g001
Figure 2. (a,b): sandstone, well G103, 2479.80 m, the mineral composition is dominated by quartz and feldspar; (c,d): sandy mudstone, well N92x, 3276.00 m, silty quartz and feldspar particles are unevenly distributed; (e,f): mudstone, well G34, 2517.56 m, clay structure, the main component is fine scaly clay minerals; (g,h): calcareous mudstone, well Q46, 2860.10 m, the main component is dolomite, a small amount of terrestrial quartz and feldspar, bright crystal dolomite filling between sand debris; (i,j): dolomite, well N202, 3565.26 m, the main component is dolomite, and a small amount of terrigenous fine-grained quartz and feldspar are unevenly distributed; (k,l): limestone, well XL25x, 3550.36 m; (m,n): shale, well B11x, 3543.90 m, development of page bedding structure; (o,p): shale, well G24, 2617.00 m, fine scaly clay minerals, silty quartz, feldspar and iron bands are interbedded with unequal thickness.
Figure 2. (a,b): sandstone, well G103, 2479.80 m, the mineral composition is dominated by quartz and feldspar; (c,d): sandy mudstone, well N92x, 3276.00 m, silty quartz and feldspar particles are unevenly distributed; (e,f): mudstone, well G34, 2517.56 m, clay structure, the main component is fine scaly clay minerals; (g,h): calcareous mudstone, well Q46, 2860.10 m, the main component is dolomite, a small amount of terrestrial quartz and feldspar, bright crystal dolomite filling between sand debris; (i,j): dolomite, well N202, 3565.26 m, the main component is dolomite, and a small amount of terrigenous fine-grained quartz and feldspar are unevenly distributed; (k,l): limestone, well XL25x, 3550.36 m; (m,n): shale, well B11x, 3543.90 m, development of page bedding structure; (o,p): shale, well G24, 2617.00 m, fine scaly clay minerals, silty quartz, feldspar and iron bands are interbedded with unequal thickness.
Energies 16 01748 g002
Figure 3. Correlation characteristics of conventional logging curves.
Figure 3. Correlation characteristics of conventional logging curves.
Energies 16 01748 g003
Figure 4. Diagram of the FDA.
Figure 4. Diagram of the FDA.
Energies 16 01748 g004
Figure 5. Schematic diagram of the principle of lithology recognition by BP neural network.
Figure 5. Schematic diagram of the principle of lithology recognition by BP neural network.
Energies 16 01748 g005
Figure 6. Schematic diagram of C&RT to identify lithology.
Figure 6. Schematic diagram of C&RT to identify lithology.
Energies 16 01748 g006
Figure 7. Confusion matrix of three lithology identification models.
Figure 7. Confusion matrix of three lithology identification models.
Energies 16 01748 g007
Figure 8. Cross plots of well logs against lithology labels.
Figure 8. Cross plots of well logs against lithology labels.
Energies 16 01748 g008
Figure 9. Lithology identification of well B11x by BP neural network and C&RT.
Figure 9. Lithology identification of well B11x by BP neural network and C&RT.
Energies 16 01748 g009
Figure 10. Rock core of well B11x. (a) At 3512.91~3513.13 m, the rock core shows mudstone; (b) The upper bedding at 3547.90~3548.06 m is developed as shale, and the lower part is dolomite; (c) At 3564.06~3564.36 m, the rock has a blocky structure without bedding.
Figure 10. Rock core of well B11x. (a) At 3512.91~3513.13 m, the rock core shows mudstone; (b) The upper bedding at 3547.90~3548.06 m is developed as shale, and the lower part is dolomite; (c) At 3564.06~3564.36 m, the rock has a blocky structure without bedding.
Energies 16 01748 g010
Table 1. Canonical function eigenvalue contribution rate statistics table.
Table 1. Canonical function eigenvalue contribution rate statistics table.
FunctionEigenvaluesPercent Variance (%)Cumulative Percentage (%)Canonical Correlation
10.73174.674.60.65
20.17818.192.7 0.388
30.0545.598.20.226
40.0181.81000.131
Table 2. Lithology of samples and results of three methods to predict lithology.
Table 2. Lithology of samples and results of three methods to predict lithology.
No.LithologyFDABP Neural NetworkC&RT
Prediction of LithologyPrediction of LithologySample TypePrediction of LithologySample Type
1sandstonesandstonesandstonetrainingsandstonetraining
2sandstonesandy mudstonesandstonetestsandstonetraining
3sandstonesandy mudstonesandstonetrainingsandstonetraining
4sandstonesandstonesandstonetrainingsandstonetraining
5sandstonesandstonesandstonetestsandstonetraining
6sandstonesandstonesandstonetrainingsandstonetraining
7sandstonesandstonesandstonetrainingsandstonetraining
8sandstonesandstonesandstonetrainingsandstonetraining
9sandstonesandstonesandstonetrainingsandstonetraining
10sandstonesandstonesandstonetrainingsandstonetraining
11sandstonesandstonesandstonetrainingsandstonetraining
12sandstonesandstonesandstonetrainingsandstonetraining
13sandstonesandstonedolomitetestsandstonetraining
14sandstonesandstonesandstonetrainingsandstonetraining
15sandstonesandstonesandstonetestsandstonetraining
16sandstonesandstonesandstonetrainingsandstonetraining
17sandstonesandstonesandstonetestsandstonetraining
18sandstonesandstonesandstonetrainingmudstonetraining
19sandstoneshalesandstonetrainingsandstonetraining
20sandstonesandstonesandstonetrainingsandstonetraining
21sandstoneshalesandstonetrainingsandstonetest
22sandstonesandstonesandstonetrainingsandstonetraining
23sandstonesandstonesandstonetrainingsandstonetraining
24sandstonesandstonesandstonetrainingsandstonetraining
25sandstonesandstonesandstonetrainingsandstonetraining
26sandstonesandstonesandstonetrainingsandstonetraining
27sandstonesandstonesandstonetrainingsandstonetraining
28sandstonesandstonesandstonetrainingsandstonetraining
29sandstonemudstonesandstonetrainingsandstonetraining
30sandstonesandstonesandstonetrainingsandstonetraining
31sandstonesandstonesandstonetrainingsandstonetraining
32sandstoneshalesandstonetestsandstonetraining
33sandstonesandstonesandstonetrainingsandstonetraining
34sandstoneshalesandstonetrainingsandstonetraining
35sandstonesandstonesandstonetrainingsandstonetraining
36sandstonesandstonesandstonetrainingsandstonetraining
37sandstonesandstonesandstonetrainingsandstonetraining
38sandy mudstonelimestonesandy mudstonetrainingsandy mudstonetraining
39sandy mudstoneshaledolomitetrainingdolomitetraining
40sandy mudstonesandstonesandy mudstonetrainingsandy mudstonetraining
41sandy mudstonesandy mudstonesandy mudstonetrainingsandstonetraining
42sandy mudstonesandy mudstonesandstonetrainingsandstonetest
43sandy mudstoneshalesandy mudstonetrainingsandy mudstonetraining
44mudstoneshalemudstonetrainingmudstonetraining
45mudstonemudstonemudstonetrainingmudstonetraining
46mudstonemudstonemudstonetrainingmudstonetraining
47mudstonemudstonemudstonetrainingmudstonetraining
48mudstonemudstonemudstonetrainingmudstonetraining
49mudstoneshalemudstonetestmudstonetraining
50mudstoneshalemudstonetrainingmudstonetraining
51mudstonesandstonemudstonetrainingmudstonetraining
52mudstonesandstonemudstonetrainingsandstonetraining
53mudstonesandy mudstonemudstonetestshaletraining
54mudstonesandy mudstonemudstonetrainingmudstonetest
55mudstonesandy mudstonesandstonetrainingmudstonetraining
56mudstonesandy mudstonemudstonetrainingmudstonetraining
57mudstonesandstonesandstonetrainingmudstonetraining
58mudstonesandstonemudstonetrainingmudstonetraining
59mudstoneshaleshaletrainingmudstonetraining
60mudstoneshalemudstonetrainingmudstonetraining
61mudstoneshalemudstonetrainingmudstonetraining
62mudstonesandstonemudstonetrainingmudstonetraining
63mudstoneshalemudstonetrainingmudstonetraining
64mudstoneshaleCalcareous mudstonetrainingmudstonetraining
65mudstoneshalemudstonetrainingsandstonetraining
66mudstoneshalemudstonetrainingmudstonetraining
67mudstoneshaleshaletrainingmudstonetraining
68Calcareous mudstonemudstoneCalcareous mudstonetrainingmudstonetraining
69Calcareous mudstonemudstoneCalcareous mudstonetestCalcareous mudstonetraining
70Calcareous mudstonesandstoneCalcareous mudstonetrainingCalcareous mudstonetraining
71Calcareous mudstoneshaleshaletrainingshaletraining
72Calcareous mudstoneshaleCalcareous mudstonetrainingCalcareous mudstonetraining
73Calcareous mudstoneshaleshaletrainingCalcareous mudstonetest
74Calcareous mudstoneshalesandstonetrainingCalcareous mudstonetraining
75Calcareous mudstonesandstonesandstonetrainingsandstonetraining
76Calcareous mudstoneshaleshaletrainingmudstonetraining
77Calcareous mudstoneshalemudstonetestCalcareous mudstonetraining
78Calcareous mudstoneshalemudstonetrainingCalcareous mudstonetraining
79Calcareous mudstoneshaleCalcareous mudstonetrainingCalcareous mudstonetest
80dolomiteshaledolomitetrainingdolomitetest
81dolomiteshaleshaletrainingdolomitetraining
82dolomiteshaleshaletrainingdolomitetraining
83dolomiteshaledolomitetrainingdolomitetraining
84dolomitesandstonedolomitetrainingshaletraining
85dolomiteshaledolomitetrainingdolomitetraining
86dolomiteshaledolomitetrainingdolomitetraining
87dolomiteshaledolomitetestdolomitetraining
88dolomitesandstonedolomitetrainingsandstonetraining
89dolomitesandstonedolomitetrainingdolomitetraining
90dolomitesandstonedolomitetrainingdolomitetraining
91dolomiteshaledolomitetrainingdolomitetraining
92dolomiteshaleshaletrainingshaletest
93dolomiteshaledolomitetestdolomitetraining
94dolomiteshaledolomitetrainingdolomitetraining
95dolomiteshaleshaletrainingshaletraining
96dolomitemudstonedolomitetrainingdolomitetest
97limestonesandy mudstonelimestonetraininglimestonetraining
98limestonelimestonelimestonetrainingshaletraining
99limestoneshalelimestonetraininglimestonetraining
100shalesandstoneCalcareous mudstonetrainingshaletraining
101shaleshaleshaletrainingshaletraining
102shaleshaleshaletrainingshaletraining
103shaleshaledolomitetrainingshaletraining
104shaleshaleshaletestshaletraining
105shaleshaleshaletrainingshaletraining
106shaleshaleshaletrainingshaletraining
107shaleshaleshaletestshaletraining
108shaleshaledolomitetrainingshaletraining
109shaleshaledolomitetrainingdolomitetraining
110shaleshaleshaletestshaletraining
111shaleshaleshaletrainingshaletraining
112shaleshaleshaletrainingshaletraining
113shaleshaleshaletrainingshaletraining
114shaleshaleshaletrainingshaletraining
115shaleshaleshaletrainingshaletraining
116shaleshaledolomitetrainingshaletraining
117shaleshaleshaletrainingshaletraining
118shaleshaleshaletrainingshaletraining
119shaleshaleshaletrainingshaletraining
120shaleshaleshaletrainingshaletraining
121shaleshaleshaletrainingshaletraining
122shaleshaleshaletrainingshaletraining
123shaleshaleshaletrainingshaletraining
124shaleshaleshaletestshaletraining
125shaleshaleshaletrainingshaletraining
126shaleshaleshaletrainingshaletraining
127shaleshaleshaletestshaletraining
128shalesandstoneCalcareous mudstonetrainingshaletraining
129shaleshaleshaletrainingshaletraining
130shaleshaleshaletrainingshaletraining
131shalesandy mudstoneshaletrainingshaletraining
132shalelimestoneshaletrainingshaletraining
133shaleshaledolomitetrainingshaletraining
134shalesandy mudstoneshaletraininglimestonetraining
135shaleshaleshaletrainingmudstonetraining
136shaleshaleshaletrainingshaletraining
137shaleshaleshaletrainingshaletraining
138shaleshaleshaletrainingshaletraining
139shaleshaleshaletrainingshaletraining
140shaleshaleshaletrainingmudstonetraining
141shalesandstoneshaletrainingshaletraining
142shalesandstoneshaletrainingshaletraining
143shaleshaleshaletrainingshaletraining
144shaleshaleshaletestshaletraining
145shaleshaleshaletrainingshaletraining
146shaleshaleshaletrainingshaletraining
147shalesandstoneshaletestsandstonetraining
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

Song, Z.; Xiao, D.; Wei, Y.; Zhao, R.; Wang, X.; Tang, J. The Research on Complex Lithology Identification Based on Well Logs: A Case Study of Lower 1st Member of the Shahejie Formation in Raoyang Sag. Energies 2023, 16, 1748. https://doi.org/10.3390/en16041748

AMA Style

Song Z, Xiao D, Wei Y, Zhao R, Wang X, Tang J. The Research on Complex Lithology Identification Based on Well Logs: A Case Study of Lower 1st Member of the Shahejie Formation in Raoyang Sag. Energies. 2023; 16(4):1748. https://doi.org/10.3390/en16041748

Chicago/Turabian Style

Song, Zhaojing, Dianshi Xiao, Yongbo Wei, Rixin Zhao, Xiaocheng Wang, and Jiafan Tang. 2023. "The Research on Complex Lithology Identification Based on Well Logs: A Case Study of Lower 1st Member of the Shahejie Formation in Raoyang Sag" Energies 16, no. 4: 1748. https://doi.org/10.3390/en16041748

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