Next Article in Journal
Investigation of Aeroelastic Energy Extraction from Cantilever Structures under Sustained Oscillations
Previous Article in Journal
Self-Sustaining Control Strategy for Proton-Exchange Membrane Electrolysis Devices Based on Gradient-Disturbance Observation Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Practical Online Characterization of the Properties of Hydrocracking Bottom Oil via Near-Infrared Spectroscopy

1
Key Laboratory of Smart Manufacturing in Energy Chemical Process, East China University of Science and Technology, Ministry of Education, Shanghai 200237, China
2
Engineering Research Center of Process System Engineering, East China University of Science and Technology, Ministry of Education, Shanghai 200237, China
*
Author to whom correspondence should be addressed.
Processes 2023, 11(3), 829; https://doi.org/10.3390/pr11030829
Submission received: 2 January 2023 / Revised: 28 February 2023 / Accepted: 7 March 2023 / Published: 10 March 2023

Abstract

:
Providing real-time information on the chemical properties of hydrocracking bottom oil (HBO) as the feedstock for ethylene cracker while minimizing processing time, is important to improve the real-time optimization of ethylene production. In this study, a novel approach for estimating the properties of HBO samples was developed on the basis of near-infrared (NIR) spectra. The main noise and extreme samples in the spectral data were removed by combining discrete wavelet transform with principal component analysis and Hotelling’s T2 test. Kernel partial least squares (KPLS) regression was utilized to account for the nonlinearities between NIR data and the chemical properties of HBO. Compared with the principal component regression, partial least squares regression, and artificial neural network, the KPLS model had a better performance of obtaining acceptable values of root mean square error of prediction (RMSEP) and mean absolute relative error (MARE). All RMSEP and MARE values of density, Bureau of Mines correlation index, paraffins, isoparaffins, and naphthenes were less than 1.0 and 3.0, respectively. The accuracy of the industrial NIR online measurement system during consecutive running periods in predicting the chemical properties of HBO was satisfactory. The yield of high value-added products increased by 0.26 percentage points and coil outlet temperature decreased by 0.25 °C, which promoted economic benefits of the ethylene cracking process and boosted industrial reform from automation to digitization and intelligence.

1. Introduction

1.1. Background

Ethylene is an important organic compound in the petrochemical industry, and its production substantially dwarfs that of other petrochemical products [1]. Ethylene production plays a pivotal role in petrochemical production, and the extent of this trade is often regarded as a significant indicator of a country’s level of industrialization [2]. Thermal cracking, as the primary force in the ethylene production process, utilizes cracking furnaces as major units to decompose feedstock into small molecules from a wide range of feedstock, such as ethane, propane, butane liquefied petroleum gas, naphtha, and hydrocracking bottom oil (HBO) [3,4].
HBO is obtained from the hydrocracking reaction unit (HCR), which is an essential process that converts heavy oil fractions with catalysts and excess hydrogen into a variety of valuable products [5]. Recently, the cost of feedstock accounts for over 90% of the total expense of ethylene production. The choice of ethylene feedstock largely depends on the availability of the raw material and profitability. Thus, HBO has become one of the preferred feedstocks due to its ability to produce high yields of ethylene, propylene, and butadiene.
Usually, the mass fraction of paraffins, isoparaffins, olefins, naphthenes, and aromatics (PIONA) is used to characterize the cracking performance of ethylene feedstock. In the cracking process, paraffins yield the highest amount of ethylene; isoparaffins readily form propylene in the cracking reaction; naphthenes tend to produce aromatic hydrocarbons; olefins generate acetylene and coke; and aromatics are prone to form heavy aromatic hydrocarbons [6,7]. Therefore, the properties of HBO, encompassing PIONA, density, boiling range, and the Bureau of Mines correlation indexes ( B M C I ), which change according to the properties of crude oil, processing technologies, and operating conditions, are crucial to the various operations involved in ethylene cracking, such as modeling, simulation, control, and optimization, especially in real-time monitoring and optimization [8].

1.2. Previous Research

These parameters can be assessed via standard procedures such as gas chromatograph-mass spectrometry; however, they tend to be time-consuming and costly. Nowadays, intelligent reformation of ethylene industries is of wide concern [9]. Fast and accurate property information of feedstock will improve aforementioned intelligence in the ethylene cracking process, leading to benefits for enterprises [10]. Therefore, real-time information on the chemical properties of HBO should be provided while minimizing the analysis and processing time to obtain a considerable amount of high-quality products.
Near infrared (NIR) spectroscopy has been proven to be highly valuable in predicting product quality as it can obtain information about multiple kinds of chemical bonds (e.g., C–H, O–H, or N–H) in molecules [11]. Thus, the application of this technique in the online detection of sample quality has received increasing attention due to its speed, low cost, and minimal sample preparation requirements [12,13,14]. It is extensively employed for predicting qualitative and quantitative properties of the products in various industries, such as agricultural, food, and pharmaceutical industries [15]. Optimal models on the basis of NIR spectroscopy have been developed to predict the quality of various types of materials, such as biorefinery feedstocks [16], tomatoes, and alternative fuels [17].
Recent studies have employed chemometric calibration modeling techniques combined with NIR spectroscopy to identify some of the most important properties of petroleum and its distillates [18], such as crude oil [19,20], gasoline [21], diesel, and naphtha [22]. By utilizing this technique, especially with a flexible NIR fiber probe [23,24], many critical parameters of petroleum and its products can be rapidly predicted. These predictions can be utilized to support rapid laboratory analysis of samples or online monitoring, controlling, and optimization of refinery processes, such as crude distillation, catalytic cracking, naphtha cracking, and gasoline and diesel blending [25,26,27].
The construction of high-quality quantitative models is challenging due to the large amounts of variables, noise, and overlapping bands [28]. To address this issue, some methods have been applied to accurately predict oil properties. For example, standard normal variate (SNV), wavelet transform (WT), and detrending have been employed to distinguish noise, improve resolution and sensitivity, and extract useful features in the NIR data preprocessing [29,30]. Methods that can be used to reduce the size of data sets used for training prediction models include correlation coefficient, two-dimensional correlation, principal component analysis (PCA), and Hotelling’s T2 [31,32,33,34,35]. Many modeling techniques are available for quantitative data analysis, including principal component regression (PCR), partial least squares (PLS), locally weighted regression, artificial neural networks (ANN), support vector regression, and the kernel partial least squares (KPLS), which fall under the category of multivariate or machine learning regression methods. They can be applied to analyze the correlation of NIR spectral data [36,37,38,39,40,41,42]. Additionally, bilinear approaches and the introduction of nonlinear kernels in support vector machine regression can assist in analyzing data with a strong nonlinear correlation. Recent studies indicate that novel chemometric approaches, including machine learning methods, can provide some advantage in the space nonlinearity and the time-varying issue of the process simultaneously [36,38].
However, unlike light oils or off-line measurements, HBO is a complicated hydrogenated oil with assorted internal molecular structures [43]. The NIR spectral peaks of the HBO samples overlap significantly. Their NIR spectral data fluctuate with the dynamic detection environment, such as the contamination degree of the NIR probe [44,45]. For a long-term online measurement of HBO properties in the dynamic ethylene cracking process, these factors pose a great challenge to the prediction accuracy and robustness of the HBO prediction model.
The aforementioned methods concentrate on addressing the nonlinearities of oil properties and NIR spectra data or ensuring the robustness of predictive models on an individual level. However, this approach may not yield a reliable predictive ability for HBO quality parameters. Generally, regression methods can provide a good solution to train the predictive model if the NIR spectra of the HBO samples can be denoised and abnormal oil samples can be removed. Meanwhile, there is interest in developing a model that incorporates the effects of oil properties on process control and profitability.

1.3. Motivation and Main Contributions

Real-time and accurate property information of HBO, such as PIONA, density, boiling range, and B M C I for a long run period, is essential and beneficial for intelligent operation and profitability in the ethylene cracking process. Aforementioned methods of NIR modeling are effective for offline prediction; however, the advantages of online prediction are not sufficient. The motivation of this study is to address the issue of using NIR data to measure the quality parameters including the physical and chemical properties of HBO in refineries, and to reveal the control and profitability improvement of this online measurement technique for real-time optimization of ethylene cracking.
To improve the accuracy and robustness of the property prediction model and realize the integration of online measurement and online optimization of ethylene cracking, a practical online characterization of HBO properties was proposed in this paper. An NIR absorption spectrometer that was equipped with a transmission probe was applied to implement the approach. The main noise of the spectral data was removed by applying discrete wavelet transform (DWT). Extreme samples were marked out to solve the robustness problem by combining principal component analysis (PCA) and Hotelling’s T2 test. The nonlinearities between the quality parameters of HBO and NIR data were constructed by KPLS algorithm to improve prediction accuracy. The method can improve the predictive ability of the models, thereby reducing the risk of performance degradation. Moreover, integrated applications of the aforementioned rapid measurement methods and real-time optimization of ethylene cracking unit were carried out.
The main contributions of the paper are as follows:
(1) A novel approach of DWT combining PCA with Hotelling’s T2 test was proposed to solve the robustness problem of NIR prediction model.
(2) NIR prediction models based on KPLS were constructed to predict HBO properties more accurately.
(3) An optimization framework integrating HBO online analysis and real-time optimization module was proposed.

2. Experimental Section

2.1. HBO Sample Collection

Over 130 HBO samples, which consisted of mixtures of bottom oils from a medium-pressure HCR unit and a high-pressure HCR unit, were collected from the feed pipeline of an ethylene cracking furnace located in an eastern Chinese refinery. An automatic sampler was utilized to collect HBO samples in the form of oil drops from the pipeline for a sampling period of 1 h. Approximately 100 mL of the HBO sample was continuously collected to ensure a greater number of HBO droplets were obtained from the sample. This step ensured that the as-obtained HBO samples were composed of many small HBO oil droplets at different time periods, which reflected the properties of the HBO sample as it flowed through the pipeline until relatively stable oil properties were achieved. The interval between collecting different samples, which lasted for 2–3 days, depended on changes in the properties of crude oil processed by the crude distillation unit and the operating conditions of the hydrocracking unit. Consequently, the different HBO samples had vastly different properties.

2.2. Laboratory Analysis of HBO Properties

The B M C I is widely used as an aromatic index for oil cracking, and the optimal BMCI value for hydrocracking of HBO is normally lower than 12. The B M C I was calculated as follows.
B M C I = 48,640 T + 273 + 473.7 × d 15.6 15.6 456.8
where T is the average boiling point (°C) and d 15.6 15.6 is the specific gravity at 15.6 °C (g/cm3).
The density, B M C I , and PIONA content of the liquid feedstock are crucial for the optimization and control of the ethylene cracking process.
Density and average boiling point ( T ) (such as 10% point, 30% point, 50% point, 70% point, and 90% point) were determined following the American Society for Testing and Materials (ASTM) D1298 and D86, respectively. The PIONA content of HBO samples was determined via the reproductive laboratory analysis method [46]. The chemical compositions of the HBO samples in terms of hydrocarbon families and carbon number distributions were analyzed via gas chromatograph [47] (Agilent 7890). Mass spectral data were obtained using an Agilent GC–MS/7890-5975C operating in the electron impact positive ion mode within the pyrolysis temperature range of 350–750 °C. The Agilent 7890 GC was equipped with a programmed temperature vaporization system, a flame ionization detector, a cryogenic cooled oven compartment, and a capillary column of fused silica, 30 m long and 0.25 μm wide. The mass range was 1 m/z–1050 amu m/z, the resolution was unit mass resolution, the sensitivity was 100 pg octafluoronaphthalene, S/N ≥ 10:1 at 272 m/z. The sample was extracted with a 0.4–1.0 μL microsyringe during the experiment. The working condition of the instrument is to adjust the repulsion pole of the ion source to maximize the molecular ion peak m/e 226 of n-cetane. The magnetic field scanning was done from m/e 65–410. The ionization voltage was 70 eV and the ionization current was 10–70 μA. PIONA content was calculated using software (Carbon Distribution of Disel Oil 2005) for calculating wax oil carbon number distribution and an MS software (Beijing Research Institute of Petroleum Processing of China Petroleum & Chemical Corporation, Beijing, China) for analyzing the hydrocarbon composition of vacuum gas oil.

2.3. Online NIR Detection System and HBO Spectra Collection

The HBO sampling site and the online method for measuring the NIR spectra of the HBO samples are shown in Figure 1. An online NIR Fourier transform spectrometer (Bruker, Germany), equipped with a tungsten halogen lamp, was used to online collect NIR spectra. A liquid immersion fiber optic probe with a 1.0 mm optical path length was installed on the feeding pipeline before the cracking furnace. The probe was connected to a spectrometer via a 50 m fiber optics cable. When the spectrometer and illumination are properly installed, this transmission mode of the NIR probe could collect more information about the HBO samples, as the light can penetrate through them. It can be used for a comprehensive assessment of the quality of HBO if satisfactory spectral information can be obtained and processed. The NIR spectra were scanned within the wavelength range of 4500 cm−1–12,000 cm−1. During a 1 h HBO sampling period, 30 NIR spectra of the HBO sample were recorded at 2 min intervals. Each spectrum consisted of 32 accumulation scans, resulting in over 4000 spectra collected in total. The NIR spectrum of a sample used for modeling was constructed by taking the average of the 30 spectra, ensuring spectral reproducibility.

2.4. Preprocessing and Analysis of the Spectral Data

2.4.1. WT for NIR Spectrum Preprocessing

A previous study had shown that WT can be used to denoise NIR data [48]. In the present study, the spectral data were denoised via DWT. The signals transformed by DWT were represented using Equation (2) below:
X ( t ) = j z k z C ( j , k ) 2 j 2 ψ ( 2 j 2 t k )
where X ( t ) is the function transferred by WT, ψ denotes the mother function, and j and k are dimensions, wherein k is an integer value.
The NIR spectra of the HBO samples were preprocessed by DWT to denoise and improve resolution. In this study, one-dimensional discrete wavelet denoising was employed. The wavelet discrete Meyer family has been shown to be relatively accurate in decomposing temporal wind signals and time-series data [49,50]. Moving average transform was utilized to smoothen the raw spectral data of different segment sizes. SNV, multiplicative scattering correction (MSC), SNV-minus a straight line (SNV-MASL), min-max normalization (MMN), and DWT were used for preprocessing spectral data.

2.4.2. PCA and Hotelling’s T2 Test

PCA can be used to find combinations of variables that capture the main trends in the data set and to visualize the most important information [51]. Given a data matrix X, its covariance matrix is calculated as follows [52].
c o v ( X ) = X T X m 1
where m indicates samples. The result of the PCA procedure is given as follows:
X n × m = t 1 p 1 T + t 2 p 2 T + t i p i T + ... + t k p k T + E n × m
where t i is the score vector, p i T is the loading vector, and E is the residual matrix. The original dataset can be adequately described by several orthogonal principal components (PCs) without losing substantial information, and the relationships between samples can be easily detected when the PCS are relative to each other.
Extreme samples can affect the accuracy of the established model, so they should be identified. One way of doing this is by using the Hotelling’s T2 statistic. Hotelling’s T2 is commonly used to detect extreme sample values and has relatively high power in all unbiased tests [53]. Assuming that s 11 , s 12 , …, s 1 N 1 and s 21 , s 22 , …, s 2 N 2 are independent normally distributed samples with parameters λ 1 , λ 2 and σ 2 , i.e., N( λ 1 , σ 2 ) and N ( λ 2 , σ 2 ), Hotelling’s T2 test analyzes the null hypothesis H0: λ 1 ¯ = λ 2 ¯ against the alternative one, H1: λ 1 ¯ λ 2 ¯ . In performing this test, the mean ( s i ¯ ) and standard deviation ( s i ) of each sample are calculated as follows [54].
s j ¯ = 1 N j = 1 N i s i j   i = 1 , 2
s i 2 = 1 n j j = 1 N i ( x i j x i ¯ ) 2   i = 1 , 2
where N i is the sample size, and n i = N i 1 . The pooled variance is then represented by
S p 2 = n 1 S 1 2 + n 1 S 2 2 f
where f = n 1 + n 2 = N 1 + N 2 2 . The null hypothesis is rejected if and only if
[ ( N 1 N 2 N 1 + N 2 ) ( x ¯ 1 x ¯ 2 ) 2 S p 2 ] > t f , α / 2 2 = F 1 , f , α
where x i ¯ is the mean of each independent normally distributed sample, S p 2 is the pooled variance, t f , α / 2 is the critical value derived from the samples’ t-distribution, α is the significance level, and F 1 , f , α is the critical value derived from the Fisher distribution. The basic assumption of Hotelling’s T2 test is that the distribution of data is normal.
In this study, PCA and Hotelling’s T2 test were utilized to analyze the variation in spectral data and remove extreme samples.

2.4.3. Selection of Characteristic Variables

The NIR spectrum region extends from 12,000 cm–1 to 4000 cm–1, but the region generally used for rapid HBO analysis is between 10,000 cm–1 and 4500 cm–1. Due to the large amounts of redundant information unrelated to the prediction of density, B M C I , and PIONA content, spectral regions related to specific properties should be selected, and spectral regions that do not overlap with absorption peaks of other functional groups should be avoided. In spectral modeling, the spectral data obtained according to the spectral wavelength are input variables for the model, and several important wavelengths related to response information should be selected as far as possible to enhance the predictive ability of the constructed calibration model. Different wavelength ranges were chosen based on the correlation between HBO properties and NIR absorbance and chemical properties, which are discussed in detail in Section 3.2.

2.5. Development and Evaluation of Prediction Models

The NIR data and a corresponding estimation for the properties of the HBO are expressed as follows.
x = [ x 1 , ... , x n ] T
y ^ = [ y ^ 1 , ... , y ^ l ] T
where x i , i = 1 , ... , n denotes the NIR absorbance at the ith wavelength; y ^ , i = 1 , ... , l denotes the jth estimated property.
x k = [ x k 1 , ... , x k n ] T , k = 1 , ... , m
y k = [ x k 1 , ... , x k l ] T , k = 1 , ... , m
where x k and y k are oil samples. m is the number of oil samples.
Therefore, the NIR spectral peak height data matrix X and the corresponding property matrix Y are calculated as follows.
X = [ x 1 , x 2 , ... , x m ] T
Y = [ y 1 , y 2 , ... , y m ] T
Five prediction models were constructed by PCR [55], PLS [56], ANN, and KPLS,. Their results were compared and the best prediction approach was selected.

2.5.1. KPLS

The KPLS models can project the input X to a high-dimensional figure space, which are represented as follows [57]:
Φ ( X ) = T P T + E Y = U Q T + F
where Φ indicates a nonlinear transformation.
The relationship of score matrices T and U can be represented as Equation (16).
T = Φ ( X ) Φ ( X ) T U ( T T Φ ( X ) Φ ( X ) T U ) 1
The kernel function K is represented as Equation (17).
K x , y = < Φ ( x ) , Φ ( y ) > = Φ Φ T
where x and y are the inputs and outputs, respectively. < Φ ( x ) , Φ ( y ) > denotes the dot product. The deflation of it is calculated by Equations (18) and (19).
K ˜ = I 1 m l m l m T K l 1 m l m l m T
I = 1 0 0 1 m × m , l m = 1 1 1 1 m × m
In this research, the dot product calculated by Gaussian kernel function is represented as Equation (20).
K ( x , y ) = e x p ( x y 2 2 σ 2 )
where σ is the kernel parameter.

2.5.2. Significance Test on Model Accuracy

In building a prediction model, relevant samples from the dataset were randomly applied to construct a calibration model, and the remaining samples were applied to test the performance of the model. In the calibration subset, every method (i.e., PCR, PLS, and ANN) was employed to evaluate the methods by cross-validation [58]. This study used 10-fold cross-validation and averaged (or combined) the results of 10-fold results to produce a single estimation. The root mean square errors of cross validation ( R M S E C V ), the root mean squared errors of prediction set ( R M S E P ), and the mean absolute relative error ( M A R E ) were calculated using Equations (21)–(24), respectively, and they were applied to evaluate the constructed models. For prediction set samples, smaller R M S E P and M A R E values indicate higher accuracy. R measured the correlation between predicted and true values. The closer the value of R2 was to 1, the better the model performance was.
R M S E C V = i = 1 n ( y i , exp y i , p r e ) 2 n
R M S E P = i = 1 n ( y i , exp y i , p r e ) 2 n
M A R E = 1 n i = 1 n | y i . p r e y i . exp y i . exp × 100 % |
R 2 = 1 i = 1 n ( y i , exp y i , p r e ) 2 i = 1 n ( y i , exp y ¯ i , p r e ) 2
where n denotes the number of calibration/prediction set samples, y i , p r e denotes the predicted value for a given sample i , y i , exp denotes the experimental value, and y ¯ i , p r e denotes the average of the response variable.
The flow chart illustrating how the methods predict the properties of HBO rapidly through transmission NIR spectroscopy is shown in Figure 2. The same sample sets and the same characteristic wavelength ranges of NIR spectroscopy were used to construct NIR prediction modes via PCR, PLS, ANN, and KPLS.

2.6. Real-Time Optimization of Ethylene Cracking Process Integrated with Online NIR Measurement System

A real-time optimization system was constructed to optimize the 360,000 t/year HBO ethylene cracker. The system had several functions, including optimization mode selection and setting, monitoring of cracking feed data, selection of operating conditions and setting of constraint ranges, results display of cracking depth optimization and dilution ratio (stream to feed ratio), and system status verification. The real-time optimization framework of ethylene cracking process is shown in Figure 3. It was composed of an online NIR detection system for HBO and a real-time optimization module for the ethylene cracking process. Every module constructed a security policy to ensure the safety and stability of the system.
In this case study, the online property analysis module utilizes NIR spectroscopy to provide accurate property data for HBO, including density, distillation curve, B M C I , and PIONA. These properties are obtained through practical online characterization of the HBO, using the data interaction module. These data of properties are sent to the real-time optimization module. In this module, the yield of cracking products can be predicted by a process model of an ethylene cracking furnace with current operating conditions and feedstock properties.

3. Results and Discussions

3.1. Sample Statistics

The properties of over 130 HBO samples were investigated in this study using reference methods. These samples were randomly divided into two subsets with a ratio of 6:4, which represented the subset of validation and prediction, respectively. The range of properties and the mean of the samples were calculated. Results showed that the methods had a good range and distribution. Density was 821.5–839.8 kg/m3 with a mean of 832.2 kg/m3 for the full set; the mean of the calibration subset and prediction subset were 832.4 kg/m3 and 832.2 kg/m3, respectively. Moreover, the means of these properties for the calibration subset were close to those of the prediction subset used for external validation, which were 13.1, 26.2%, 26.8%, <0.1%, 46.1%, and 0.8% in the prediction subset. In addition, both subsets had distribution patterns that were similar and consistent with those of the full set. Owing to the low content of olefins in the HBO, these compounds were not considered in this modeling study. The characteristics of different types of samples provided are shown in the Table 1 below.

3.2. Spectral Features of HBO Samples and Spectroscopic Techniques

Noise elimination or reduction was performed to improve the model’s performance. The OPUS software comes with the MSC, SNV, SNV-MASL, MMN, and DWT methods, which are used to eliminate or reduce noise. The DWT of the software was mainly used to deduct the influence of the instrument background or drift on the signal. MSC and SNV were applied to eliminate the effect of scattering on the spectrum due to the uneven distribution of particles and different particle sizes. MMN could be employed to eliminate the adverse effects caused by large scale differences. Therefore, the different methods used to preprocess HBO parameters (such as paraffins) are shown as in Figure 4, which were SNV, MSC, SNV-MASL, MMN, and DWT.
As shown in Figure 4, the R M S E P and M A R E values of all pretreatment methods were lower than 0.8 and 0.6, respectively. The R M S E P value of SNV-MASL was the highest, while the M A R E value was relatively small. The RMSEP value of MMN was relatively small but the MARE value was particularly high. SNV and MSC had similar effects. Among the five methods, R M S E P and M A R E values of DWT were the smallest; considering R M S E P and M A R E comprehensively, the effect of DWT was best compared with other methods. Thus, DWT was chosen as the method for the pretreatment of NIR spectral data.
Over 4000 spectra were collected. The mean NIR spectra of the HBO samples with different properties are shown in Figure 5. The spectral curves of all samples clearly showed a similar tendency for both A and B orientations with no abnormal variation trend, suggesting accurately discriminating the properties of HBO was challenging. The line of different color represented different samples. In general, five obvious absorption peaks appeared at 8300 cm−1, 7350 cm−1, 5830 cm−1, 5765 cm−1, and 5658 cm−1. Specifically, the wavelength at around 5590 cm−1 and 6055 cm−1 was the maximum absorption region, which corresponded to the first overtones of the C–H stretching vibrations of –CH3, –CH2, and –HC=CH–. The peak ranging from 7200 cm−1 to 7500 cm−1 was the result of OH stretching overtones. The small peaks between 4800 cm−1 and 4550 cm−1 were attributed to the C=C and C–H stretching combination tones of unsaturated fatty acids [47].
In Figure 5a, the noise in the raw spectral data are obvious and the results of denoising absorbance data by DWT are given in Figure 5b. The spectral absorbance preprocessed by DWT with intensity and peak regions did not change, whereas the noises were removed and important data were retained. The pretreated spectra of HBO samples after eliminating baseline drift are given in Figure 5c. In the NIR spectra of HBO samples, the spectral features were enhanced and the baseline variations were weakened. The denoised spectral data were exported for further preprocessing and development of appropriate prediction models. The important information was mainly contained between 6780 cm−1 and 4800 cm−1. Therefore, this part of the spectra was considered as the characteristic variables for exploratory purposes to simplify the model.
Different types of uncertainty, such as the performance of the spectrophotometer, human uncertainty during the experimental process, and sample preparation, may generate extreme data points [59]. These uncertainties will increase the variance of the spectral data and/or quality parameters being measured [60]. Therefore, extreme points in the data were detected to determine the variance of the spectral data [61].
By combination PCA with the Hotelling’s T2 functions to pretreat the spectral data of HBO samples, extreme data were determined by the 95% confidence limits. Figure 5d clearly shows the explained variances of data (score plot) by PC-1 and PC-2. The first and second components explained 97% and 2% of the variance in the data, respectively. A total of six data points were located outside the 95% confidence interval, i.e., outside the ellipse, so they were determined to be extreme points, of which the Hotelling’s T2 values were higher than 7.07.

3.3. NIR Model Development for Predicting HBO Properties

Based on the data of different properties obtained from NIR spectroscopy, over 4000 spectral data points were divided into groups at a ratio of 6:4 into a verification set and a prediction set. The values obtained from the KPLS models were compared with the true values to continuously narrow the gap. Online monitoring was carried in three operating periods. After selecting the preprocessing transformations on the denoised data and trying various approaches, calibration models were developed for the density, B M C I , and PINA content of HBO samples. The RMSECV and MARE values obtained by PCR, PLS, ANN, and KPLS are shown in Figure 6.
As shown in Figure 6, R M S E C V and M A R E were basically consistent, indicating that the density, B M C I , and hydrocarbon composition of the HBO samples could be quickly analyzed by the four methods. KPLS consistently showed better performance and predicted results than PCR, PLS, and ANN. Its R M S E C V and M A R E values were smaller than those of the other models constructed by PCR, PLS, and ANN. In the KPLS model, the R M S E C V values of density, B M C I , and paraffins, isoparaffins, and naphthenes were 0.981, 0.340, 0.501, 0.328, and 0.723, respectively, whereas their MARE values were 0.136, 2.853, 0.218, 1.017, and 1.310, respectively. Therefore, the model developed by KPLS for predicting HBO properties was assessed using the prediction set. The KPLS method also showed good prediction performance in modeling the PIONA content in gasoline. This method was used to model the PIONA content in gasoline, and achieved good prediction results [25].
The results of the application of the KPLS model to the NIR spectral data of the HBO samples in the prediction subset are provided in Figure 7. The density, B M C I , PIONA content, 10% point, 30% point, 50% point, 70% point, and 90% point of the HBO samples measured by the standard test method were compared. Among them, the highest R2 value was for the 10% point, the second highest R2 value was for paraffins, and the smallest R2 value was for the 90% point, the corresponding value was 0.877. The smallest R M S E P was for isoparaffins, the next smallest was the B M C I , and the highest R M S E P value was for the 10% point, and the corresponding value was 2.150. The smallest M A R E value was for density, and the next smallest value was for paraffins; the highest M A R E value was for the 70% point, and the corresponding value was 2.854. The result of the three values of paraffins was relatively good compared to the results of other properties. The R M S E P and M A R E values of the models were less than 3 and their R 2 exceeded 0.87. The low R M S E P and M A R E values for these properties confirmed the models’ prediction performance.

3.4. Accuracy of Online Detection of HBO Properties

3.4.1. Implementation of Online NIR Detection

An online NIR system was established using an online NIR instrument for analyzing HBO properties under actual production conditions. The system is capable of reading spectra generated by spectral analysis software and using an encapsulated KPLS model for HBO NIR analysis to predict the properties of HBO samples. The model was implemented in C#. The predicted results can be transmitted to the OLE for process control (OPC) server through OPC communication and saved on the local server. A comprehensive program logic flow chart of the NIR analysis system formed herein is shown in Figure 8.
The program of the NIR analysis system determined whether the media access control (MAC) address of the current server meets the running permission. If yes, after the OPC connection, then the program would locate the latest spectrum in the folder generated by the spectroscopic analysis software based on time. If the file memory size was greater than a given threshold, then the spectrum file would be read, the present spectra pretreatment program and the NIR model parameters could be decoded and read, and the properties of the HBO samples were predicted. The predicted results were saved on the local server and transmitted to the OPC server through the OPC protocol. At this point, the whole program was completed and waited until the scan cycle finished the next round of prediction. In the long-term online prediction process, if the error was unacceptable, the spectrum pretreatment program and the quantitative model parameters were updated offline according to the method described in this paper. These updated spectroscopic preprocessing programs and NIR models were given into the online NIR system to maintain the accuracy of HBO properties online measurements.

3.4.2. Online Detection Accuracy of Three Run Periods

The predictive performance of the prediction model on the component data was evaluated using the performance indicators described above. For online NIR prediction of petroleum properties, the absolute errors between their predicted values and laboratory analysis values were always expected to be small and within a certain range. Moreover, the stability and accuracy of prediction models were normally required in the long-term applications of the refinery production processes. In this study, the normal operation of the online NIR detection system from the start to the cleaning of the fiber optic probe was recorded as an operating cycle (about 1 month). A small number of samples (one sample every three days) were collected in every period to verify the online predictive performance of the aforementioned NIR models using RMSEP and MARE values between the online predicted values and the experimental values. Over 30 samples were obtained in three run periods. The online predictive accuracy of density, B M C I , paraffins, isoparaffins, and naphthenes of the HBO samples in the three run periods is shown in Figure 9.
As shown in Figure 9, the ranges of the RMSEP values of density, B M C I , paraffins, isoparaffins, naphthenes, 10% point, 30% point, 50% point, 70% point, and 90% point of most samples were −0.99–1.16 kg/m3, 0.37–0.40 kg/m3, 0.54–0.58 kg/m3, 0.35–0.40 kg/m3, 0.77–0.80 kg/m3, 2.95–3.20 °C, 2.94–3.06 °C, 3.05–3.10 °C, 2.79–2.88 °C, and 1.85–1.90 °C, respectively. All RMSEP values of the samples’ properties were less than 4 in the three run periods. The MARE values of the density, B M C I , paraffins, isoparaffins, naphthenes, 10% point, 30% point, 50% point, 70% point, and 90% point of these samples were all less than 4.0%. It can see that NIR was very effective for chemical or physical parameters. Using the B M C I for example, the value of the relevant B M C I was calculated indirectly from Equation (1). The value of the B M C I is correlated with the mean boiling point ( T ) and density. HBO is rich in alkanes, a certain amount of cycloalkanes, and a small amount of aromatics, and its sulfur and nitrogen contents were low. Hydrocarbons contain C–H and C–C, whose molecular vibrations affect the yield of ethylene cracking. Thus, the online predictive effect of the NIR model did not significantly attenuate in the three run periods, which was acceptable for online prediction of HBO properties.
Overall, the results obtained were due to the combination of DWT with other preprocessing functions. DWT was used in denoising the raw spectral data before creating a regression model. The predictive performance of the KPLS model developed herein for the online estimation of the density; the B M C I ; and contents of paraffins, isoparaffins, naphthenes, 10% point, 30% point, 50% point, 70% point, and 90% point of the HBO samples had relatively acceptable regression parameters.

3.5. Effects of Online NIR Measurement System Assisted Online Cracking Depth Optimization

The cracking depth optimization control was implemented in a 0.36 Mt/year HBO cracker (SL-IV mode). The optimization results of the COT and the yield of key products, such as ethylene, propylene, and high value-added products, are shown in Figure 10. As can be seen, after optimization, the COT control value decreased by 2.5 °C from 802.4 °C to 799.9 °C, indicating a decrease in energy consumption of the cracking furnace decreases. Ethylene yield decreased by 0.22 percentage points from 30.18% to 29.97%, whereas the yield of propylene increased by 0.74 percentage points from 16.96% to 17.70%. The sum yield of ethylene and propylene increased from 47.14% to 47.67% by 0.53 percentage points. The yield of high value-added products (H2, ethylene, propylene, butadiene, benzene) increased by 0.26 percentage points from 61.42% to 61.68%. It can be seen that the depth optimization method integrated with online NIR measurement of HBO can effectively improve the yield of the target products and bring significant economic benefits.

4. Conclusions

In this study, a novel approach combining DWT, PCA, Hotelling’s T2 test, and the KPLS algorithm was proposed for online prediction of HBO samples characteristics using NIR spectroscopy. The characteristic distribution of the HBO samples investigated herein had a broad range. The noise in the spectral data was removed using DWT. The spectral data of the HBO samples were pretreated by combining PCA with the Hotelling’s T2 functions, and extreme data were identified by the 95% confidence limits. The prediction results of the HBO calibration set samples indicated that the KPLS method consistently performed better and had better results than PCR, PLS, and ANN, as demonstrated by the distinctly different values of R M S E C V and M A R E . The R M S E P , M A R E , and R 2 values for density, B M C I , paraffin, isoparaffin, naphthenes, and the 10% point, 30% point, 50% point, 70% point, and 90% point obtained from the KPLS model were in good agreement with the data obtained from the standard test methods, indicating that they were suitable for prediction subset and online measurement. The NIR method proposed for the analysis of HBO chemical composition was faster than GC-MS and other standard methods. The NIR models constructed were utilized for the online measurement of HBO properties in an industrial ethylene plant. The online measurement of quality parameters of HBO through NIR spectroscopy, along with the proposed practical online characterization of HBO properties, has enhanced the real-time optimization of the ethylene cracking process. The improvement in the yield of target products and the reduction in energy consumption had promoted the economic benefits of the ethylene cracking process. Moreover, applications of integration of real-time detection and real-time optimization in plants and enterprises can boost the industrial reform from automation to digitization and intelligence.

Author Contributions

Conceptualization, and funding acquisition, C.F.; writing—original draft and funding acquisition, J.L.; writing—review and editing, T.X. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by National Key Research and Development Program of China (2022YFB3305903), National Natural Science Foundation of China (Key Program: 62136003), National Natural Science Foundation of China (61973124), Shanghai Committee of Science and Technology (Grant No.22DZ1101500) and Fundamental Research Funds for the Central Universities.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

Acronyms and Parameters
ANNartificial neural network
ASTMAmerican Society for Testing and Materials
B M C I Bureau of Mines correlation index
DWTdiscrete wavelet transform
GCgas chromatograph
GC-MSgas chromatograph-mass spectrometry
HBOhydrocracking bottom oil
HCRhydrocracking unit
LVslatent variables
MAREthe mean absolute relative error
MLRmultiple linear regression
MMNmin-max normalization
MSmass spectrometry
MSCmultiplicative scattering correction
NIRnear-infrared
PCAprincipal component analysis
PCRprincipal component regression
PC-1the first PC
PC-2the second PC
PCsprincipal components
PLSpartial least squares regression
PIONAparaffins, isoparaffins, naphthenes, olefins, and aromatics
RMSECVroot mean squared error of cross validations
RMSEPthe root mean squared error of prediction set
SNVstandard normalized vector
SNV-MASLSNV-minus a straight line
WTwavelet transform
d 15.6 15.6 specific gravity at 15.6 °C (g/cm3)
Eresidual matrix
F1,f,αcritical value derived from the Fisher distribution
ia given sample
jdimension
K kernel function
K ˜ deflation of kernel function
kinteger value
mrow of samples
Nisample size
nnumber of calibration/prediction set samples
p i T loading vector
Sistandard deviation of each sample
sijindependent normally distributed samples
s ¯ i mean of each sample
S p 2 pooled variance
Taverage boiling point (°C)
Tscore matrix
tiscore vector
t f , α / 2 critical value derived from the samples’ t-distribution
Uscore matrix
Xspectral data matrix
X the NIR spectral peak height data matrix
X (t)the function transferred by WT
x k oil samples
x i NIR absorbance at the i-th wavelength
x ¯ imean of each sample
Y the corresponding property matrix
y k oil samples
y ^ i the j-th estimated property
y i , p r e predicted value
y i , exp experimental value
y ¯ i , p r e the average of the response variable
ψmother function
Φ a nonlinear transformation
σ kernel parameter

References

  1. Al-Jamimi, H.A.; BinMakhashen, G.M.; Deb, K.; Saleh, T.A. Multiobjective optimization and analysis of petroleum refinery catalytic processes, A review. Fuel 2021, 288, 119678. [Google Scholar] [CrossRef]
  2. Gu, H.; Li, J.; Mu, P.; Zhu, Q. Improving the Operational Efficiency of Ethylene Cracking Integrated with Refining by Molecular Management. Ind. Eng. Chem. Res. 2020, 59, 13160–13174. [Google Scholar] [CrossRef]
  3. Browning, B.; Alvarez, P.; Jansen, T.; Lacroix, M.; Geantet, C.; Tayakout-Fayolle, M. A Review of Thermal Cracking, Hydrocracking, and Slurry Phase Hydroconversion Kinetic Parameters in Lumped Models for Upgrading Heavy Oils. Energy Fuels 2021, 35, 15360–15380. [Google Scholar] [CrossRef]
  4. Liu, J.; Fang, X.; Yang, T. Novel Ebullated Bed Residue Hydrocracking Process. Energy Fuels 2017, 31, 6568–6579. [Google Scholar] [CrossRef]
  5. Alawani, N.A.; Muller, H.; Panda, S.K.; Al-Hajji, A.A.; Koseoglu, O.R. Evaluation of Polycyclic Aromatic Hydrocarbon Removal from Hydrocracking Recycle Streams. Energy Fuels 2020, 34, 179–187. [Google Scholar] [CrossRef]
  6. Upreti, N.; Pakala, H.; Surasani, V.K.; Dinda, S. Modeling of thermal cracking reaction of kerosene range hydrocarbons. Mater. Today Proc. 2022, 59, 430–436. [Google Scholar] [CrossRef]
  7. Bai, Y.; Zhang, G.; Liu, D.; Zhang, Y.; Zhao, L.; Gao, J.; Xu, C.; Meng, Q.; Gao, X. The advance in catalytic pyrolysis of naphtha technology using ZSM-5 as catalyst. Appl. Catal. A Gen. 2021, 628, 118399. [Google Scholar] [CrossRef]
  8. Fakhroleslam, M.; Sadrameli, S.M. Thermal/catalytic cracking of hydrocarbons for the production of olefins: A state-of-the-art review III, Process modeling and simulation. Fuel 2019, 252, 553–566. [Google Scholar] [CrossRef]
  9. Meng, D.; Shao, C.; Zhu, L. Two-level comprehensive energy-efficiency quantitative diagnosis scheme for ethylene-cracking furnace with multi-working-condition of fault and exception operation. Energy 2022, 239, 121835. [Google Scholar] [CrossRef]
  10. Moro, M.K.; dos Santos, F.D.; Folli, G.S.; Romão, W.; Filgueiras, P.R. A review of chemometrics models to predict crude oil properties from nuclear magnetic resonance and infrared spectroscopy. Fuel 2021, 303, 121283. [Google Scholar] [CrossRef]
  11. Jurinjak Tušek, A.; Jurina, T.; Čulo, I.; Valinger, D.; Gajdoš Kljusurić, J.; Benković, M. Application of NIRs coupled with PLS and ANN modelling to predict average droplet size in oil-in-water emulsions prepared with different microfluidic devices. Spectrochim. Acta Part A 2022, 270, 120860. [Google Scholar] [CrossRef] [PubMed]
  12. Wang, K.; Du, W.; Long, J. Near-Infrared Wavelength-Selection Method Based on Joint Mutual Information and Weighted Bootstrap Sampling. IEEE Trans. Ind. Inform. 2020, 16, 5884–5894. [Google Scholar] [CrossRef]
  13. He, K.; Zhong, M.; Du, W. Weighted incremental minimax probability machine-based method for quality prediction in gasoline blending process. Chemom. Intell. Lab. Syst. 2020, 196, 103909. [Google Scholar] [CrossRef]
  14. Zhao, N.F.; Xu, Q.S.; Tang, M.L.; Wang, H. Variable Screening for Near Infrared (NIR) Spectroscopy Data Based on Ridge Partial Least Squares Regression. Comb. Chem. High Throughput Screen. 2020, 23, 740–756. [Google Scholar] [CrossRef]
  15. Du, C.; Sun, L.; Bai, H.; Liu, Y.; Yang, J.; Wang, X. Quantitative detection of azodicarbonamide in wheat flour by near-infrared spectroscopy based on two-step feature selection. Chemom. Intell. Lab. Syst. 2021, 219, 104445. [Google Scholar] [CrossRef]
  16. Varghese, G.; Saeed, K.; Rutt, K.J. Determination of the oxidative stability of biodiesel fuels by near-infrared spectroscopy. Fuel 2021, 290, 120015. [Google Scholar] [CrossRef]
  17. Liu, S.; Wang, S.; Hu, C.; Zhan, S.; Kong, D.; Wang, J. Rapid and accurate determination of diesel multiple properties through NIR data analysis assisted by machine learning. Spectrochim. Acta Part A 2022, 277, 121261. [Google Scholar] [CrossRef] [PubMed]
  18. Santos, F.D.; Santos, L.P.; Cunha, P.H.P.; Borghi, F.T.; Romão, W.; de Castro, E.V.R.; de Oliveira, E.C.; Filgueiras, P.R. Discrimination of oils and fuels using a portable NIR spectrometer. Fuel 2021, 283, 118854. [Google Scholar] [CrossRef]
  19. Joshi, N.B.; Mullins, O.C.; Jamaluddin, A.; Creek, J.; McFadden, J. Asphaltene Precipitation from Live Crude Oil. Energy Fuels 2001, 15, 979–986. [Google Scholar] [CrossRef]
  20. Barreira, F.R.; Reis, L.G.; Nunes, R.d.C.P.; Filipakis, S.D.; Lucas, E.F. Asphaltenes Precipitation Onset, Influence of the Addition of a Second Crude Oil or Its Asphaltenes Fractions (C3I and C5I). Energy Fuels 2018, 32, 10391–10397. [Google Scholar] [CrossRef]
  21. Wang, K.; He, K.; Du, W.; Long, J. Novel adaptive sample space expansion approach of NIR model for in-situ measurement of gasoline octane number in online gasoline blending processes. Chem. Eng. Sci. 2021, 242, 116672. [Google Scholar] [CrossRef]
  22. Yu, H.; Du, W.; Lang, Z.Q.; Wang, K.; Long, J. A Novel Integrated Approach to Characterization of Petroleum Naphtha Properties From Near-Infrared Spectroscopy. IEEE Trans. Instrum. Meas. 2021, 70, 1–13. [Google Scholar] [CrossRef]
  23. Long, J.; Wang, K.; Yang, M.; Zhong, W. Rapid crude oil analysis using near-infrared reflectance spectroscopy. Pet. Sci. Technol. 2019, 37, 354–360. [Google Scholar] [CrossRef]
  24. Giovenzana, V.; Beghi, R.; Tugnolo, A.; Brancadoro, L.; Guidetti, R. Comparison of two immersion probes coupled with visible/near infrared spectroscopy to assess the must infection at the grape receiving area. Comput. Electron. Agric. 2018, 146, 86–92. [Google Scholar] [CrossRef]
  25. Li, J.; Chu, X. Rapid Determination of Physical and Chemical Parameters of Reformed Gasoline by Near-Infrared (NIR) Spectroscopy Combined with the Monte Carlo Virtual Spectrum Identification Method. Energy Fuels 2018, 32, 12013–12020. [Google Scholar] [CrossRef]
  26. Long, J.; Jiang, S.; He, R.; Zhao, L. Diesel blending under property uncertainty: A data-driven robust optimization approach. Fuel 2021, 306, 121647. [Google Scholar] [CrossRef]
  27. Leal, A.L.; Silva, A.M.S.; Ribeiro, J.C.; Martins, F.G. Using Spectroscopy and Support Vector Regression to Predict Gasoline Characteristics, A Comparison of 1H NMR and NIR. Energy Fuels 2020, 34, 12173–12181. [Google Scholar] [CrossRef]
  28. Zhao, N.; Ma, L.; Wang, K.; Zhang, F.; Li, M.; Liu, X.; Zhu, M.; Lu, Y.; Song, X.; Yan, H.; et al. NIR robustness model of variable selection investigation of critical quality attributes coupled with different simulate noises by prediction capability and reproducibility. Spectrochim. Acta Part A 2022, 267, 120522. [Google Scholar] [CrossRef]
  29. Shi, H.; Yu, P. Comparison of grating-based near-infrared (NIR) and Fourier transform mid-infrared (ATR-FT/MIR) spectroscopy based on spectral preprocessing and wavelength selection for the determination of crude protein and moisture content in wheat. Food Control 2017, 82, 57–65. [Google Scholar] [CrossRef]
  30. Dotto, A.C.; Dalmolin, R.S.D.; Grunwald, S.; ten Caten, A.; Pereira Filho, W. Two preprocessing techniques to reduce model covariables in soil property predictions by Vis-NIR spectroscopy. Soil Tillage Res. 2017, 172, 59–68. [Google Scholar] [CrossRef]
  31. Chu, X.-L.; Xu, Y.-P.; Tian, S.-B.; Wang, J.; Lu, W.-Z. Rapid identification and assay of crude oils based on moving-window correlation coefficient and near infrared spectral library. Chemom. Intell. Lab. Syst. 2011, 107, 44–49. [Google Scholar] [CrossRef]
  32. Cai, C.-B.; Xu, L.; Han, Q.-J.; Wu, H.-L.; Nie, J.-F.; Fu, H.-Y.; Yu, R.-Q. Combining the least correlation design, wavelet packet transform and correlation coefficient test to reduce the size of calibration set for NIR quantitative analysis in multi-component systems. Talanta 2010, 81, 799–804. [Google Scholar] [CrossRef] [PubMed]
  33. Liu, H.; Zhou, W.; Ying, D.; Xiang, B.; Li, P. Generalized two-dimensional correlation NIR spectroscopy analysis of the structures on n-propanol and n-butanol. J. Mol. Struct 2019, 1181, 436–443. [Google Scholar] [CrossRef]
  34. Cruz-Tirado, J.P.; Oliveira, M.; de Jesus Filho, M.; Godoy, H.T.; Amigo, J.M.; Barbin, D.F. Shelf life estimation and kinetic degradation modeling of chia seeds (Salvia hispanica) using principal component analysis based on NIR-hyperspectral imaging. Food Control 2021, 123, 107777. [Google Scholar] [CrossRef]
  35. Bobba, S.; Zinfollino, N.; Fissore, D. Application of Near-Infrared Spectroscopy to statistical control in freeze-drying processes. Eur. J. Pharm. Biopharm. 2021, 168, 26–37. [Google Scholar] [CrossRef] [PubMed]
  36. Cui, C.; Fearn, T. Hierarchical mixture of linear regressions for multivariate spectroscopic calibration: An application for NIR calibration. Chemom. Intell. Lab. Syst. 2018, 174, 1–14. [Google Scholar] [CrossRef]
  37. Reda, R.; Saffaj, T.; Derrouz, H.; Itqiq, S.E.; Bouzida, I.; Saidi, O.; Lakssir, B.; El Hadrami, E.M. Comparing CalReg performance with other multivariate methods for estimating selected soil properties from Moroccan agricultural regions using NIR spectroscopy. Chemom. Intell. Lab. Syst. 2021, 211, 104277. [Google Scholar] [CrossRef]
  38. Franzoi, M.; Ghetti, M.; Di Monte, L.; De Marchi, M. Investigation of weight loss in mozzarella cheese using NIR predicted chemical composition and multivariate analysis. J. Food Compos. Anal. 2021, 102, 104002. [Google Scholar] [CrossRef]
  39. Lei, T.; Sun, D.-W. A novel NIR spectral calibration method: Sparse coefficients wavelength selection and regression (SCWR). Anal. Chim. Acta 2020, 1110, 169–180. [Google Scholar] [CrossRef]
  40. Fung Shek, C.; Betenbaugh, M. Taking the pulse of bioprocesses: At-line and in-line monitoring of mammalian cell cultures. Curr. Opin. Biotechnol. 2021, 71, 191–197. [Google Scholar] [CrossRef]
  41. Labbé, N.; Lee, S.-H.; Cho, H.-W.; Jeong, M.K.; André, N. Enhanced discrimination and calibration of biomass NIR spectral data using non-linear kernel methods. Bioresour. Technol. 2008, 99, 8445–8452. [Google Scholar] [CrossRef] [PubMed]
  42. Binetti, G.; Del Coco, L.; Ragone, R.; Zelasco, S.; Perri, E.; Montemurro, C.; Valentini, R.; Naso, D.; Fanizzi, F.P.; Schena, F.P. Cultivar classification of Apulian olive oils: Use of artificial neural networks for comparing NMR, NIR and merceological data. Food Chem. 2017, 219, 131–138. [Google Scholar] [CrossRef] [PubMed]
  43. Pan, H.; Li, X.; Shen, Y.; Wu, X.; Ju, F.; Li, Y.; Wu, G.; Ai, B.; Xu, B.; Ling, H. Novel design of lubricant-type vacuum distillation process for lube base oils production from hydrocracking tail oil. Chin. J. Chem. Eng. 2022, 45, 121–132. [Google Scholar] [CrossRef]
  44. Campos, M.I.; Antolin, G.; Debán, L.; Pardo, R. Assessing the influence of temperature on NIRS prediction models for the determination of sodium content in dry-cured ham slices. Food Chem. 2018, 257, 237–242. [Google Scholar] [CrossRef] [PubMed]
  45. Zhong, Y.; Gu, J.; Su, Y.; Zhao, L.; Zhou, Y.; Peng, J. Real-time screening of hepatotoxins in natural medicine by peroxynitrite responsive lanthanide-based NIR-II luminescent probes. Chem. Eng. J. 2022, 433, 133263. [Google Scholar] [CrossRef]
  46. Standard Test Method for Hydrocarbon Types Analysis of Gas-Oil Saturates Fractions by High Ionizing Voltage Mass Spectrometry. Available online: https//docin.com/p-1271811601.html (accessed on 27 August 2015).
  47. Zhang, K.; Tan, Z.; Chen, C.; Sun, X.S.; Wang, D. Rapid Prediction of Camelina Seed Oil Content Using Near-Infrared Spectroscopy. Energy Fuels 2017, 31, 5629–5634. [Google Scholar] [CrossRef]
  48. Abasi, S.; Minaei, S.; Jamshidi, B.; Fathi, D.; Khoshtaghaza, M.H. Rapid measurement of apple quality parameters using wavelet de-noising transform with Vis/NIR analysis. Sci. Hortic. 2019, 252, 7–13. [Google Scholar] [CrossRef]
  49. Jamei, M.; Ahmadianfar, I.; Chu, X.; Yaseen, Z.M. Prediction of surface water total dissolved solids using hybridized wavelet-multigene genetic programming: New approach. J. Hydrol. 2020, 589, 125335. [Google Scholar] [CrossRef]
  50. Zucatelli, P.J.; Nascimento, E.G.S.; Santos, A.Á.B.; Arce, A.M.G.; Moreira, D.M. An investigation on deep learning and wavelet transform to nowcast wind power and wind power ramp: A case study in Brazil and Uruguay. Energy 2021, 230, 120842. [Google Scholar] [CrossRef]
  51. Long, J.; Li, T.; Yang, M.; Hu, G.; Zhong, W. Hybrid Strategy Integrating Variable Selection and a Neural Network for Fluid Catalytic Cracking Modeling. Ind. Eng. Chem. Res. 2019, 58, 247–258. [Google Scholar] [CrossRef]
  52. Hotelling, H. Analysis of a complex of statistical variables into principal components. J. Educ. Psychol. 1933, 24, 417. [Google Scholar] [CrossRef]
  53. Glazko, G.V.; Emmert-Streib, F. Unite and conquer: Univariate and multivariate approaches for finding differentially expressed gene sets. Bioinformatics 2009, 25, 2348–2354. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Sadeghi, Z.; Mashhadi-Hossainali, M. Application of the T2-Hotelling test for investigating ionospheric anomalies before large earthquakes. J. Atmos. Sol. Terr. Phys. 2019, 185, 7–21. [Google Scholar] [CrossRef]
  55. Hosani, E.A.; Meribout, M.; Al-Durra, A.; Al-Wahedi, K.; Teniou, S. A New Optical-Based Device for Online Black Powder Detection in Gas Pipelines. IEEE Trans. Instrum. Meas. 2014, 63, 2238–2252. [Google Scholar] [CrossRef]
  56. Haghi, R.K.; Yang, J.; Tohidi, B. Fourier Transform Near-Infrared (FTNIR) Spectroscopy and Partial Least-Squares (PLS) Algorithm for Monitoring Compositional Changes in Hydrocarbon Gases under In Situ Pressure. Energy Fuels 2017, 31, 10245–10259. [Google Scholar] [CrossRef]
  57. Zhu, Q.; Zhao, Z.; Liu, F. Developing new products with kernel partial least squares model inversion. Comput. Chem. Eng. 2021, 155, 107537. [Google Scholar] [CrossRef]
  58. Rohani, A.; Taki, M.; Abdollahpour, M. A novel soft computing model (Gaussian process regression with K-fold cross validation) for daily and monthly solar radiation forecasting (Part I). Renew. Energy 2018, 115, 411–422. [Google Scholar] [CrossRef]
  59. Toscano, G.; Rinnan, Å.; Pizzi, A.; Mancini, M. The Use of Near-Infrared (NIR) Spectroscopy and Principal Component Analysis (PCA) To Discriminate Bark and Wood of the Most Common Species of the Pellet Sector. Energy Fuels 2017, 31, 2814–2821. [Google Scholar] [CrossRef]
  60. Heidari, P.; Rezaei, M.; Sahebi, M.; Khadivi, A. Phenotypic variability of Pyrus boissieriana Buhse: Implications for conservation and breeding. Sci. Hortic. 2019, 247, 1–8. [Google Scholar] [CrossRef]
  61. Jamshidi, B.; Mohajerani, E.; Farazmand, H.; Mahmoudi, A.; Hemmati, A. Pattern recognition-based optical technique for non-destructive detection of Ectomyelois ceratoniae infestation in pomegranates during hidden activity of the larvae. Spectrochim. Acta Part A 2019, 206, 552–557. [Google Scholar] [CrossRef]
Figure 1. Sampling site and online method for measuring the NIR spectra of the HBO samples.
Figure 1. Sampling site and online method for measuring the NIR spectra of the HBO samples.
Processes 11 00829 g001
Figure 2. Flow chart of the proposed integrative approach for the rapid determination of HBO.
Figure 2. Flow chart of the proposed integrative approach for the rapid determination of HBO.
Processes 11 00829 g002
Figure 3. Real-time optimization framework of ethylene cracking process integrated with online analysis property of HBO.
Figure 3. Real-time optimization framework of ethylene cracking process integrated with online analysis property of HBO.
Processes 11 00829 g003
Figure 4. Effects of different preprocessing methods for paraffins.
Figure 4. Effects of different preprocessing methods for paraffins.
Processes 11 00829 g004
Figure 5. NIR spectra of HBO samples within the range of 12,000 cm−1–4500 cm−1. (a) Raw NIR spectral data. (b) Spectra denoised by DWT. (c) Spectra after baseline correction. (d) Score plot from the PCA of the NIR spectra of the HBO samples.
Figure 5. NIR spectra of HBO samples within the range of 12,000 cm−1–4500 cm−1. (a) Raw NIR spectral data. (b) Spectra denoised by DWT. (c) Spectra after baseline correction. (d) Score plot from the PCA of the NIR spectra of the HBO samples.
Processes 11 00829 g005
Figure 6. RMSECV and MARE of different methods (PCR, PLS, ANN, and KPLS) in the calibration subset.
Figure 6. RMSECV and MARE of different methods (PCR, PLS, ANN, and KPLS) in the calibration subset.
Processes 11 00829 g006
Figure 7. Estimation of the values of density, B M C I , paraffins, isoparaffins, naphthenes, 10% point, 30% point, 50% point, 70% point, and 90% point by the KPLS model according to the NIR spectra compared with the values determined by reference methods.
Figure 7. Estimation of the values of density, B M C I , paraffins, isoparaffins, naphthenes, 10% point, 30% point, 50% point, 70% point, and 90% point by the KPLS model according to the NIR spectra compared with the values determined by reference methods.
Processes 11 00829 g007
Figure 8. Program logic flow chart of the online NIR analysis system.
Figure 8. Program logic flow chart of the online NIR analysis system.
Processes 11 00829 g008
Figure 9. Online prediction accuracy of HBO samples in three run periods. (a) RMSEP of density, B M C I , paraffins, isoparaffins, naphthenes, 10% point, 30% point, 50% point, 70% point, and 90% point. (b) MARE of 10% point, 30% point, 50% point, 70% point, and 90% point.
Figure 9. Online prediction accuracy of HBO samples in three run periods. (a) RMSEP of density, B M C I , paraffins, isoparaffins, naphthenes, 10% point, 30% point, 50% point, 70% point, and 90% point. (b) MARE of 10% point, 30% point, 50% point, 70% point, and 90% point.
Processes 11 00829 g009
Figure 10. Effects of real-time optimization of cracking depth by combination of online analysis of ethylene cracking feed.
Figure 10. Effects of real-time optimization of cracking depth by combination of online analysis of ethylene cracking feed.
Processes 11 00829 g010
Table 1. Measured properties of HBO.
Table 1. Measured properties of HBO.
PropertiesRange of Full SetMean
Full SetCalibration on SubsetPrediction on Subset
Density (kg/m3)821.5–839.8832.2832.4832.2
B M C I 9.0–16.713.113.313.1
PIONA (Volume%)Paraffins21.9–30.2826.226.426.2
Isoparaffins22.6–28.4526.826.526.8
Olefins<0.5<0.1<0.1<0.1
Naphthenes41.2–53.946.146.346.1
Aromatics0–2.230.80.80.8
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

Long, J.; Xu, T.; Fan, C. Practical Online Characterization of the Properties of Hydrocracking Bottom Oil via Near-Infrared Spectroscopy. Processes 2023, 11, 829. https://doi.org/10.3390/pr11030829

AMA Style

Long J, Xu T, Fan C. Practical Online Characterization of the Properties of Hydrocracking Bottom Oil via Near-Infrared Spectroscopy. Processes. 2023; 11(3):829. https://doi.org/10.3390/pr11030829

Chicago/Turabian Style

Long, Jian, Tiantian Xu, and Chen Fan. 2023. "Practical Online Characterization of the Properties of Hydrocracking Bottom Oil via Near-Infrared Spectroscopy" Processes 11, no. 3: 829. https://doi.org/10.3390/pr11030829

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