Next Article in Journal
Enhancing Autonomous Vehicle Decision-Making at Intersections in Mixed-Autonomy Traffic: A Comparative Study Using an Explainable Classifier
Previous Article in Journal
FF3D: A Rapid and Accurate 3D Fruit Detector for Robotic Harvesting
Previous Article in Special Issue
Ensemble Learning Framework for DDoS Detection in SDN-Based SCADA Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Efficient Cyberattack Detection Methods in Industrial Control Systems

1
Institute of Control and Computation Engineering, Faculty of Electronics and Information Technology, Warsaw University of Technology, 00-665 Warsaw, Poland
2
Institute of Computer Science, Faculty of Electronics and Information Technology, Warsaw University of Technology, 00-665 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Sensors 2024, 24(12), 3860; https://doi.org/10.3390/s24123860
Submission received: 11 May 2024 / Revised: 5 June 2024 / Accepted: 12 June 2024 / Published: 14 June 2024
(This article belongs to the Special Issue Cyber Attacks in Industrial Control Systems)

Abstract

:
The article deals with the issue of detecting cyberattacks on control algorithms running in a real Programmable Logic Controller (PLC) and controlling a real laboratory control plant. The vulnerability of the widely used Proportional–Integral–Derivative (PID) controller is investigated. Four effective, easy-to-implement, and relatively robust methods for detecting attacks on the control signal, output variable, and parameters of the PID controller are researched. The first method verifies whether the value of the control signal sent to the control plant in the previous step is the actual value generated by the controller. The second method relies on detecting sudden, unusual changes in output variables, taking into account the inertial nature of dynamic plants. In the third method, a copy of the controller parameters is used to detect an attack on the controller’s parameters implemented in the PLC. The fourth method uses the golden run in attack detection.

1. Introduction

Modern, computerized production systems are equipped with network connectivity, taking advantage of industrial information and communication technologies. This provides the ability to communicate with other facilities and transmit information about the equipment. The interconnection of all systems leads to intelligent factories where production system components and people communicate via networks. In this way, production takes place almost autonomously.
Highly computerized and interconnected facilities open up digital access to equipment, and the benefits of information accessibility come at a price in terms of a higher risk of cyberattack, which poses an increasingly significant threat to industrial networks, e.g., in the energy, robotics, and automotive industries [1,2,3,4,5,6,7,8]. While there are obvious methods to protect against attacks, including the isolation of information technology (IT) and operational technology (OT) networks and the precise definition of network privileges [9], unauthorized access to critical elements of such networks constantly remains a real threat [10].

1.1. Related Works

Developing detection mechanisms of possible attacks is, therefore, a key issue. Many data-based detection mechanisms are using artificial intelligence approaches, e.g., support vector data description [11], neural autoencoders [12], k-nearest-neighbors, decision trees, support vector machines, naive Bayes and random forest methods [13], deep neural network models [14,15], genetic or evolutionary algorithms [16], Bayesian networks [17], or machine learning [18]. One can also use classifiers based on statistical properties [19,20,21]. The method described in [22] uses a fusion, adaptive, cubature Kalman filter. The concept presented in [23] relies on convolutional neural networks and Long-Short Term Memory (LSTM) models. A deep-neural-network-based approach is detailed in [24]. Thorough reviews of machine-learning- and artificial-intelligence-based cyberattack detection methods are given in [25,26,27]. On the other hand, simple threshold analysis [28] can also be applied. The relatively simple detection methods do not have the drawbacks of sophisticated data-based detection mechanisms, like complex preparation for application and tailoring to a specific application. Therefore, such relatively simple though effective methods were tested in this work.

1.2. Research Gap and Article Contribution

The aforementioned methods based on machine learning and artificial intelligence require precise models. Unfortunately, disturbances and changes in process parameters are inevitable in real-life control systems. As a result, sophisticated, model-based methods are likely to fail in such situations. Therefore, this work is concerned with alternative, model-free methods and one utilizing a golden-run approach, which relies on a non-parametric behavioral model of the entire control system, including the process itself and the control algorithm. The efficiency of all presented methods is evaluated and compared using laboratory software and a hardware environment mimicking industrial control systems. In particular, we confront the model-free methods and the golden-run approach in the case of nonstationary process behavior.

1.3. Article Structure

The article is structured as follows. Section 2 presents the laboratory control network, the investigated control plant, and the golden run used in our study. Then, Section 3 discusses efficient cyberattack detection methods and the results obtained when the control system is subjected to attacks; the attack scenarios are described, and the efficiency of the detection methods is investigated. Finally, Section 4 summarizes the article.

2. Materials and Methods

The control plant used in the experiments is part of a more extensive laboratory test bed depicted in Figure 1 used by the authors during testing advanced attack methods [29,30]. Three control processes are available: high-speed magnetic levitation, the MPS FESTO workstation, and the thermal heating and cooling stand. The first and the third processes are continuous, while the second is binary. A typical OT network’s main task is process control. The workstation is enhanced with a local and master computer station with Supervisory Control and Data Acquisition (SCADA) MAPS software ver. 4.0.5.1 to collect and archive data. The central PLC (master) is responsible for managing data from slave controllers and controlling the high-speed magnetic levitation process. The slave controllers perform their local tasks: the FX5U controller, produced by Mitsubishi Electric, Japan, controls the heating and cooling station whereas the S7-1200 controller, produced by Siemens, Germany, controls the FESTO station. Monitoring and diagnostics of the facilities can be performed on local Human–Machine Interface (HMI) panels. Process data are also transferred to the iQ-R master controller. Various protocols have been used, i.e., SLMP, Profinet, Melsoft, Ethernet Simple, Modbus TCP/IP, Modbus RTU, and Siemens S7, marked with appropriate lines on the station diagram. Communication between different devices is achieved using all implemented protocols. Data are exchanged through registers or individual data bits.

2.1. Laboratory Test Bed

The experiments were performed in a laboratory heating and cooling thermal control plant, where the user influences the temperature distribution through controllable fans and heaters. The bench can be controlled manually or via an automation system using the Modbus communication protocol. A picture of the laboratory thermal stand is presented in Figure 2.
The control plant has six manipulated variable (MV) inputs:
  • FLU, FLB, FRU, and FRB fans, with values from 0 to 100;
  • HL and HR heaters, with values from 0 to 100.
And seven process variable (PV) outputs:
  • TL, TM, TR, and TF bench temperature, with values from −55.0 °C to +125.0 °C;
  • TA ambient temperature, with values from −55.0 °C to +125.0 °C;
  • C current measurement;
  • V voltage measurement.
Figure 2. Laboratory thermal control plant.
Figure 2. Laboratory thermal control plant.
Sensors 24 03860 g002
A Pulse-Width Modulation (PWM) signal controls the actuators. The temperature sensors communicate internally using the OneWire bus, while the current and voltage measurements are carried out using dedicated electronics. All input and output signals are available via the Modbus protocol.
If the control of the fans is constant, then the control plant is linear, i.e., the temperatures linearly depend on the heater control signals. Changes in fan load cause changes in the dynamics and statics (amplification) of the control plant. The load of the left fan was set to  z FLU = 50 %  during the research. It should be noted that the thermal process under consideration is slow, with temperatures settling after about 150 s.
The control plant is versatile and allows different controllers to be tested in different configurations of control systems. For current research, the number of signals used is limited to the following:
  • u HL  is the left heater control signal (manipulated variable in the tested control system);  u HL min = 0 % u HL max = 100 % ;
  • y TL  is the temperature measured on the left-hand-side of the laboratory stand (process variable in the tested control system);  y TL min = 20 °C,  y TL max = 90 °C;
  • z FLU  is the left fan control signal (disturbance variable in the tested control system);  z FLU min = 0 % z FLU max = 100 % .
The PID control algorithm was implemented on a PLC. The controlled variable is  y TL , and the manipulated variable is  u HL . The PID control algorithm derives the value of the control signal  u HL  using the measurement of the output variable  y TL . Communication with the laboratory control plant is performed using the Modbus RTU protocol.

2.2. Control System Operation in Nominal Conditions

The responses obtained in the tested control system are depicted in Figure 3. The PID controller was tuned in such a way that the control system was sufficiently fast, with a relatively short settling time and reasonable rise time; the overshoot was very small; and the control signal changed smoothly. Thus, the obtained control quality was good. The temperature plot is taken as a reference plot (golden run) used for comparisons in the implementation of Method #4.

2.3. Attack Scenarios and Detection Methods

The tested detection methods are designed in such a way that they:
  • Are easy to implement;
  • Are efficient in detecting a cyberattack;
  • Generate few false alarms.
All evaluations were carried out using the laboratory test bed described in Section 2.1, which mimics real industrial control systems, i.e., it relies on software and hardware solutions used in industrial practice.

2.3.1. Method #1: Verification of the Control Value

Attack Scenario

In the first attack scenario, the control signal is frozen on the value of  u hl = 70 %  for  T atack = 400  s, then the controller is allowed to operate for  T noatack = 400  s, then the attack is repeated, and so on. The attack repeats five times. The periods during which the attack is performed are marked with blue in Figure 4. During the attack, the value of the output variable tends to settle around 44. The controller tries to achieve the setpoint value every time the attack stops. It is able to achieve its goal, but every time the setpoint is reached, a new attack is performed.

Detection Mechanism

The first method under investigation concentrates on the analysis of the control signal. It verifies whether the value sent to the control plant in the previous step is the actual value generated by the controller. If the verification result is negative, then an alarm is reported.
Figure 4. Responses of the control system operating during the first attack; detection method: verification of the control value with its copy (red blocks signal attack detection).
Figure 4. Responses of the control system operating during the first attack; detection method: verification of the control value with its copy (red blocks signal attack detection).
Sensors 24 03860 g004

2.3.2. Method #2: Detection of Sudden Change in Output Variable

Attack Scenario

In the second attack scenario, the process variable (temperature) is passed to the controller as the measurement is frozen on the value of  y tl = 50 C  for  T atack = 400  s, then the proper value of temperature is allowed to be sent to the controller for  T noatack = 400  s, then the attack is repeated, and so on. The attack repeats five times. The value sent to the controller as the temperature measurement is indicated with the dashed, red line in the upper subplot of Figure 5, and the real value is marked with the solid, blue line. The faked process variable value is larger during the attack than the setpoint. Therefore, the control signal tends to settle on 0 % (achieves the minimal possible value), and the real process variable tends toward a small value below the setpoint (it is the value close to the ambient temperature). Once the attack is stopped, the controller tends to achieve the setpoint value, but a new attack causes the process variable to get away from the setpoint.

Detection Mechanism

In each sampling instant, the measured temperature value is compared with the value obtained by the controller. For the mechanism to work, there must be two channels measuring the temperature created on the PLC in order to implement this security measure. Thus, temperature values exist in two different registers, but only one is used by the controller, and the attacker also overwrites this copy.
Figure 5. Responses of the control system operating during the second attack; detection of a sudden change in output variable using the copy of the temperature measurement (red blocks signal attack detection).
Figure 5. Responses of the control system operating during the second attack; detection of a sudden change in output variable using the copy of the temperature measurement (red blocks signal attack detection).
Sensors 24 03860 g005

2.3.3. Method #3: Copy of Controller Parameters Used to Detect an Attack

Attack Scenario

Two attack scenarios were tested. In the first one, it is assumed that the attacker changes the parameters of the Proportional–Integral (PI) controller in such a way that control quality deteriorates visibly (see Figure 6). It is assumed that  K = 0.3 T I = 20  (the nominal values of the parameter are  K = 0.2 T I = 50 ). The oscillations in the obtained responses occur. Interestingly, as shown in Table 1, the value of Mean Squared Error (MSE) drops compared to the golden run (but the Mean Absolute Error (MAE) increases); this illustrates that sole analysis of the values of control quality indexes may be insufficient to detect the attack.
In the second attack scenario, it is assumed that the parameters are altered slightly so that the operation of the control system deteriorates in such a way that it is difficult to detect the attack just by visual analysis of the control system response by the operator (see Figure 7). A minor change in the controller parameters was made ( K = 0.2 T I = 60 ); thus, only the integration time was slightly changed. The obtained responses are (visually) very close to the golden run; MSE and MAE deteriorate only slightly.
Figure 6. Responses of the control system operating during the third attack (significant change in the PID controller parameters); copy of controller parameters used to detect an attack (red blocks signal attack detection).
Figure 6. Responses of the control system operating during the third attack (significant change in the PID controller parameters); copy of controller parameters used to detect an attack (red blocks signal attack detection).
Sensors 24 03860 g006

Detection Mechanism

In each sampling instant, the values of the controller parameters are compared with the values of their copy (gain K and integration time  T I ). For the mechanism to work, there must be a copy of the controller parameters located in a safe place, outside reach of the attacker, ideally, in a place the attacker does not know about or in such a place so that the attacker cannot alter the values of the parameter copies.

2.3.4. Method #4: Using Golden Run in an Attack Detection

Attack Scenario

The method was tested using the same attack scenarios as already described:
  • To conduct the experiments depicted in Figures 8 and 12,the attack scenario described in Section 2.3.1, also used to obtain the result shown in Figure 4, was applied.
  • To conduct the experiments depicted in Figures 9 and 13, the attack scenario described in Section 2.3.2, also used to obtain the result shown in Figure 5, was applied.
  • To conduct the experiments depicted in Figures 10, 11, 14 and 15, the attack scenarios described in Section 2.3.3, also used to obtain the result shown in Figures 6 and 7, were applied.

Detection Mechanism

The mechanism uses the golden run (in our case, the responses depicted in Figure 3). Then, the absolute value of the difference between the current temperature and the temperature from the reference response ( | T real T gr | ) is calculated at each time step. If the difference exceeds the assumed limit, an alarm is reported. In the case of the experiments shown in Figures 8–11, the limit equal to 0 was assumed; in the case of the experiments shown in Figures 12–15, the limit equal to 1 was assumed. However, the tests in the simulation environment were also performed for other values of this parameter, and the mechanism gave the same result (in terms of attack detection) up to the limit equal to 6.

3. Results and Discussion

In the current section, the effectiveness of all discussed detection methods is studied. The first three model-free methods require the availability of process measurements, while the golden-run approach relies on the availability of the behavioral controlled process model.

3.1. Method #1: Verification of the Control Value

The detection mechanism was used to detect the attack and invoke the alarm. Thus, every time the last value generated by the controller is not the value sent to the control plant, an alarm is reported (red blocks in Figure 4). The efficiency of the method is excellent. Every time the attack occurs, it is detected at once and in each instant it is present. The method can fail only when the attacker generates the same control value as the controller. However, in such a case, the operation of the control system would not have deteriorated as a result of the attack.

3.2. Method #2: Detection of Sudden Change in Output Variable

The detection mechanism used to detect the attack invoked the alarm every time temperature values from two registers differ (red blocks in Figure 5). The method’s efficiency is excellent because every time the attack occurs, it is detected at once. The method can fail when the attacker generates a temperature value equal to the actual temperature. Still, the operation of the control system would not have deteriorated as a result of the attack. The other possibility is that the attacker alternates values in both registers, but they must know the detection mechanism and location of the measurement copy.

3.3. Method #3: Copy of Controller Parameters Used to Detect an Attack

The detection mechanism detects the attack every time values of the controller parameters differ from their copy (red blocks in Figure 6 and Figure 7), then the alarm is invoked. The effectiveness of this method is excellent, as the attack is detected as soon as it is launched. It does not matter whether the change in parameters is small or significant. The method is, therefore, more difficult to cheat than the operator. The method can fail if the attacker alternates values of the parameters in the copy stored for reference. Still, they must know about the detection mechanism and the location of the copy of the controller parameters.

3.4. Method #4: Using Golden Run in an Attack Detection

The detection mechanism detects the attack every time values of the reference response differ enough from the responses obtained from the real control plant (red blocks in Figure 8, Figure 9, Figure 10 and Figure 11), then the alarm is signaled. The method detects the beginning of the attack reliably, but unlike methods #1 and #2, it signals the attack constantly, even if currently the attack is not performed (because the registered golden run differs from the responses generated in the real control system).

Obtained Results: Laboratory Experiments

The experiments in the simulation environment suggested that the method, though having some drawbacks, should work well. However, during the laboratory tests, very interesting results were obtained. Due to the nonstationary characteristic of the process, the method had problems detecting the attacks (it generated false alarms in situations depicted in Figure 12, Figure 13, Figure 14 and Figure 15, and was late in detecting the attack in the situation depicted in Figure 14).
The drawbacks of model-based methods (though a simplified model, in the form of a waveform, is used) have made their presence known. If the process is nonstationary, its model must be updated if a significant change occurs. It means that if one wants to be sure the proper model is employed, it must be obtained (or at least updated) each time the method is used, and in some cases, it is unacceptable. In all cases, when this is necessary, it causes significant inconvenience.
Figure 12. Responses of the control system operating during the first attack; detection method: golden run; laboratory experiments.
Figure 12. Responses of the control system operating during the first attack; detection method: golden run; laboratory experiments.
Sensors 24 03860 g012
Figure 13. Responses of the control system operating during the second attack; detection method: golden run; laboratory experiments.
Figure 13. Responses of the control system operating during the second attack; detection method: golden run; laboratory experiments.
Sensors 24 03860 g013
Figure 14. Responses of the control system operating during the third attack (small change in PID parameters); detection method: golden run; laboratory experiments.
Figure 14. Responses of the control system operating during the third attack (small change in PID parameters); detection method: golden run; laboratory experiments.
Sensors 24 03860 g014
Figure 15. Responses of the control system operating during the third attack (large change in PID parameters); detection method: golden run; laboratory experiments.
Figure 15. Responses of the control system operating during the third attack (large change in PID parameters); detection method: golden run; laboratory experiments.
Sensors 24 03860 g015

3.5. Comparison of Performance Metrics

Table 2 compares performance metrics obtained for the laboratory stand for a given attack scenario and the detection method. It is clear that the detection rate of model-free methods is  100 %  whereas the false positive rate does not exceed  0.1 % . Similarly, precision is not less than  99.8 % . The golden-run-based method gives poor performance metrics, e.g., the precision rate for the first and the second attacks is approximately  55 % . In the case of the third attack, the detection rate is very poor, especially when a small change in the controller parameters is made; the detection rate amounts to only  3.55 % .
Finally, we evaluated the discussed simulation methods. Table 3 compares performance metrics obtained in the simulation experiment for a given attack scenario and the detection method. Even in the simulated environment, the performance of the golden-run-based method is worse than that of model-free methods working in the laboratory environment.

4. Conclusions

Three effective, easy-to-implement, and relatively robust methods for detecting attacks on the control signal, output variable, and the PID controller parameters have been implemented and tested; also, a method using a simplified model—a waveform recorded during the golden run—has been considered. The first three methods, namely, Method #1: verification of the control signal value sent to the control plant, Method #2: detection of sudden changes in output variables, and Method #3: using a copy of the controller parameters to detect an attack, can be applied online and are very effective. They detect an attack almost at once as soon as it is performed and are relatively hard to fool.
The Method #4 uses the golden run in attack detection. A feature of this method is that under favorable circumstances (for stationary processes), it detects the occurrence of the attack fast. Still, once the attack is detected, the method signals it constantly. Therefore, in the case of attack scenarios from Section 2.3.1 and Section 2.3.2 it invokes the alarm even if the attack is halted for a while. Unfortunately, in the case of a nonstationary process, its significant drawbacks manifested, showing the need for relatively frequent model updating. On the other hand, the method can also detect unusual behavior of the control system subjected to attack scenarios that have not been researched in the article.
In the future, we plan to design more sophisticated detection methods for nonstationary control plants, i.e., with time-varying parameters. Secondly, as a result of this work, we have found that model-based detection methods are very susceptible to the control plant’s parameter changes. Hence, this complex problem is hard to solve and very important in the case of nonstationary processes.

Author Contributions

Conceptualization, P.M., R.N., A.W., K.C., P.C., M.Ł., S.P., K.Z.; methodology, P.M., R.N., A.W., K.C., P.C., M.Ł., S.P., K.Z.; software, R.N., A.W., P.C.; validation, P.M., R.N., A.W., K.C., P.C., M.Ł., S.P., K.Z.; formal analysis, P.M., K.C., M.Ł., S.P.; investigation, R.N., A.W.; resources, P.C., M.Ł.; data curation, R.N.; writing—original draft preparation, P.M.; writing—review and editing, P.M., M.Ł.; visualization, R.N.; supervision, P.M., K.C., M.Ł., S.P.; project administration, M.Ł.; funding acquisition, M.Ł. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partially financed by research grant from The National Centre for Research and Development no. CYBERSECIDENT/488240/IV/NCBR/2021 Laboratory for Vulnerability Analysis (LaVA) of stationary and mobile IT devices and algorithms and software.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

HMIHuman–Machine Interface
ITInformation technology
MAEMean Absolute Error
MSEMean Squared Error
MVManipulated variable
OTOperational technology
PIProportional–Integral
PIDProportional–Integral–Derivative
PLCProgrammable Logic Controller
PVProcess variable
PWMPulse-Width Modulation
SCADASupervisory Control and Data Acquisition

References

  1. Toker, O.; Ozdemir, O. Physical-layer Cyberattack Resilient OFDM Automotive Radars. In Proceedings of the 2020 IEEE Vehicular Networking Conference (VNC), New York, NY, USA, 16–18 December 2020. [Google Scholar]
  2. Hou, B.; Zeng, S.; Xi, B.; Jia, S.; Guo, Q.; Xu, L.; Sun, H. Performance of Neighborhood-Watch-Based Resilient Distributed Energy Management Algorithm Under Different Types of Cyberattacks. In Proceedings of the 2021 IEEE 4th International Electrical and Energy Conference (CIEEC), Wuhan, China, 28–30 May 2021; pp. 1–5. [Google Scholar]
  3. Zabetian-Hosseini, A.; Mehrizi-Sani, A.; Liu, C.C. Cyberattack to Cyber-Physical Model of Wind Farm SCADA. In Proceedings of the IECON 2018—44th Annual Conference of the IEEE Industrial Electronics Society, Washington, DC, USA, 21–23 October 2018; pp. 4929–4934. [Google Scholar]
  4. Colter, J.; Kinnison, M.; Henderson, A.; Schlager, S.M.; Bryan, S.; O’Grady, K.L.; Abballe, A.; Harbour, S. Testing the Resiliency of Consumer Off-the-Shelf Drones to a Variety of Cyberattack Methods. In Proceedings of the 2022 IEEE/AIAA 41st Digital Avionics Systems Conference (DASC), Portsmouth, VA, USA, 18–22 September 2022; pp. 1–5. [Google Scholar]
  5. Mu, L.; Zhao, E.; Wang, Y.; Zomaya, A.Y. Buoy Sensor Cyberattack Detection in Offshore Petroleum Cyber-Physical Systems. IEEE Trans. Serv. Comput. 2020, 13, 653–662. [Google Scholar] [CrossRef]
  6. Khaw, Y.M.; Abiri Jahromi, A.; Arani, M.F.M.; Sanner, S.; Kundur, D.; Kassouf, M. A Deep Learning-Based Cyberattack Detection System for Transmission Protective Relays. IEEE Trans. Smart Grid 2021, 12, 2554–2565. [Google Scholar] [CrossRef]
  7. Pazouki, S.; Bibek, K.C.; Alkhwaildi, H.A.; Asrari, A. Modelling of Smart Homes Affected by Cyberattacks. In Proceedings of the 2020 52nd North American Power Symposium (NAPS), Tempe, AZ, USA, 11–13 April 2021; pp. 1–6. [Google Scholar]
  8. Zhang, Y.; Jiang, Y.; Xu, A.; Hong, C.; Chen, J. Method to Evaluate the Impact of Cyberattacks against Charging Piles on Distribution Network. In Proceedings of the 2020 12th IEEE PES Asia-Pacific Power and Energy Engineering Conference (APPEEC), Nanjing, China, 20–23 September 2020; pp. 1–5. [Google Scholar]
  9. Kbar, G.; Alazab, A. A Comprehensive Protection Method for Securing the Organization’s Network against Cyberattacks. In Proceedings of the 2019 Cybersecurity and Cyberforensics Conference (CCC), Melbourne, VIC, Australia, 8–9 May 2019; pp. 118–122. [Google Scholar]
  10. Sudar, K.; Deepalakshmi, P.; Nagaraj, P.; Muneeswaran, V. Analysis of Cyberattacks and its Detection Mechanisms. In Proceedings of the 2020 5th International Conference on Research in Computational Intelligence and Communication Networks (ICRCICN), Bangalore, India, 26–27 November 2020; pp. 12–16. [Google Scholar]
  11. Housh, M.; Kadosh, N.; Haddad, J. Detecting and Localizing Cyber-Physical Attacks in Water Distribution Systems without Records of Labeled Attacks. Sensors 2022, 22, 6035. [Google Scholar] [CrossRef] [PubMed]
  12. Gaggero, G.B.; Caviglia, R.; Armellin, A.; Rossi, M.; Girdinio, P.; Marchese, M. Detecting Cyberattacks on Electrical Storage Systems through Neural Network Based Anomaly Detection Algorithm. Sensors 2022, 22, 3933. [Google Scholar] [CrossRef] [PubMed]
  13. Yin, X.C.; Liu, Z.G.; Nkenyereye, L.; Ndibanje, B. Toward an Applied Cyber Security Solution in IoT-Based Smart Grids: An Intrusion Detection System Approach. Sensors 2019, 19, 4952. [Google Scholar] [CrossRef]
  14. Villalba, D.A.M.; Varon, D.F.M.; Pórtela, F.G.; Triana, O.A.D. Intrusion Detection System (IDS) with anomaly-based detection and deep learning application. In Proceedings of the 2022 V Congreso Internacional en Inteligencia Ambiental, Ingeniería de Software y Salud Electrónica y Móvil (AmITIC), San Jose, Costa Rica, 14–16 September 2022; pp. 1–4. [Google Scholar]
  15. Wang, W.; Wang, Z.; Zhou, Z.; Deng, H.; Zhao, W.; Wang, C.; Guo, Y. Anomaly detection of industrial control systems based on transfer learning. Tsinghua Sci. Technol. 2021, 26, 821–832. [Google Scholar] [CrossRef]
  16. Lysenko, S.; Bobrovnikova, K.; Shchuka, R.; Savenko, O. A Cyberattacks Detection Technique Based on Evolutionary Algorithms. In Proceedings of the 2020 IEEE 11th International Conference on Dependable Systems, Services and Technologies (DESSERT), Kyiv, Ukraine, 14–18 May 2020; pp. 127–132. [Google Scholar]
  17. Naik Sapavath, N.; Muhati, E.; Rawat, D.B. Prediction and Detection of Cyberattacks using AI Model in Virtualized Wireless Networks. In Proceedings of the 2021 8th IEEE International Conference on Cyber Security and Cloud Computing (CSCloud)/2021 7th IEEE International Conference on Edge Computing and Scalable Cloud (EdgeCom), Washington, DC, USA, 26–28 June 2021; pp. 97–102. [Google Scholar]
  18. Tang, Z.; Jiao, J.; Zhang, P.; Yue, M.; Chen, C.; Yan, J. Enabling Cyberattack-Resilient Load Forecasting through Adversarial Machine Learning. In Proceedings of the 2019 IEEE Power & Energy Society General Meeting (PESGM), Atlanta, GA, USA, 4–8 August 2019; pp. 1–5. [Google Scholar]
  19. Kao, J.B.; Jiang, J.R. Anomaly Detection for Univariate Time Series with Statistics and Deep Learning. In Proceedings of the 2019 IEEE Eurasia Conference on IOT, Communication and Engineering (ECICE), Yunlin, Taiwan, 3–6 October 2019; pp. 404–407. [Google Scholar]
  20. Wang, H.; Tang, M.; Park, Y.; Priebe, C.E. Locality Statistics for Anomaly Detection in Time Series of Graphs. IEEE Trans. Signal Process. 2014, 62, 703–717. [Google Scholar] [CrossRef]
  21. Heymann, S.; Latapy, M.; Magnien, C. Outskewer: Using Skewness to Spot Outliers in Samples and Time Series. In Proceedings of the 2012 IEEE/ACM International Conference on Advances in Social Networks Analysis and Mining, Istanbul, Turkey, 26–29 August 2012; pp. 527–534. [Google Scholar]
  22. Wu, P.; Zhang, J.; Luo, S.; Song, Y.; Zhang, J.; Wang, Y. A Fusion Adaptive Cubature Kalman Filter Approach for False Data Injection Attack Detection of DC Microgrids. Electronics 2024, 13, 1612. [Google Scholar] [CrossRef]
  23. Söğüt, E.; Erdem, O.A. A Multi-Model Proposal for Classification and Detection of DDoS Attacks on SCADA Systems. Appl. Sci. 2023, 13, 5993. [Google Scholar] [CrossRef]
  24. Siva Kumar, C.; Kolla, H.; Sravya, B.; Sri, D.L.; Nikitha, G. Obtrusion Unmasking of Machine Learning-Based Analysis of Imbalanced Network Traffic. In Proceedings of the 2023 International Conference on Computer Communication and Informatics (ICCCI), Coimbatore, India, 23–25 January 2023. [Google Scholar]
  25. Liu, H.; Lang, B. Machine Learning and Deep Learning Methods for Intrusion Detection Systems: A Survey. Appl. Sci. 2019, 9, 4396. [Google Scholar] [CrossRef]
  26. Ahsan, M.; Nygard, K.E.; Gomes, R.; Chowdhury, M.M.; Rifat, N.; Connolly, J.F. Cybersecurity Threats and Their Mitigation Approaches Using Machine Learning—A Review. J. Cybersecur. Priv. 2022, 2, 527–555. [Google Scholar] [CrossRef]
  27. Govea, J.; Gaibor-Naranjo, W.; Villegas-Ch, W. Transforming Cybersecurity into Critical Energy Infrastructure: A Study on the Effectiveness of Artificial Intelligence. Systems 2024, 12, 165. [Google Scholar] [CrossRef]
  28. Chae, Y.; Katenka, N.; DiPippo, L. An Adaptive Threshold Method for Anomaly-based Intrusion Detection Systems. In Proceedings of the 2019 IEEE 18th International Symposium on Network Computing and Applications (NCA), Cambridge, MA, USA, 26–28 September 2019; pp. 1–4. [Google Scholar]
  29. Zarzycki, K.; Chaber, P.; Cabaj, K.; Ławryńczuk, M.; Marusak, P.; Nebeluk, R.; Plamowski, S. Wojtulewicz, A. GAN Neural Networks Architectures for Testing Process Control Industrial Network Against Cyber-Attacks. IEEE Access. 2023, 11, 49587–49600. [Google Scholar] [CrossRef]
  30. Zarzycki, K.; Chaber, P.; Cabaj, K.; Ławryńczuk, M.; Marusak, P.; Nebeluk, R.; Plamowski, S.; Wojtulewicz, A. Forgery Cyber-Attack Supported by LSTM Neural Network: An Experimental Case Study. Sensors 2023, 23, 6778. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Diagram of the laboratory test bed.
Figure 1. Diagram of the laboratory test bed.
Sensors 24 03860 g001
Figure 3. Responses of the control system operating in nominal conditions.
Figure 3. Responses of the control system operating in nominal conditions.
Sensors 24 03860 g003
Figure 7. Responses of the control system operating during the third attack (small change in the PID controller parameters); copy of controller parameters used to detect an attack (red blocks signal attack detection).
Figure 7. Responses of the control system operating during the third attack (small change in the PID controller parameters); copy of controller parameters used to detect an attack (red blocks signal attack detection).
Sensors 24 03860 g007
Figure 8. Responses of the control system operating during the first attack; detection method: golden run; simulation experiments (red blocks signal attack detection).
Figure 8. Responses of the control system operating during the first attack; detection method: golden run; simulation experiments (red blocks signal attack detection).
Sensors 24 03860 g008
Figure 9. Responses of the control system operating during the second attack; detection method: golden run; simulation experiments (red blocks signal attack detection).
Figure 9. Responses of the control system operating during the second attack; detection method: golden run; simulation experiments (red blocks signal attack detection).
Sensors 24 03860 g009
Figure 10. Responses of the control system operating during the third attack (small change in PID parameters); detection method: golden run; simulation experiments (red blocks signal attack detection).
Figure 10. Responses of the control system operating during the third attack (small change in PID parameters); detection method: golden run; simulation experiments (red blocks signal attack detection).
Sensors 24 03860 g010
Figure 11. Responses of the control system operating during the third attack (large change in PID parameters); detection method: golden run; simulation experiments (red blocks signal attack detection).
Figure 11. Responses of the control system operating during the third attack (large change in PID parameters); detection method: golden run; simulation experiments (red blocks signal attack detection).
Sensors 24 03860 g011
Table 1. Comparison of control quality indicators obtained for the laboratory stand for a given attack scenario and the detection method.
Table 1. Comparison of control quality indicators obtained for the laboratory stand for a given attack scenario and the detection method.
Attack ScenarioDetection MethodMSEMAE
No attack   1.7362 × 10 1   1.9370 × 10 0
First attackMV copy   4.0905 × 10 1   4.2441 × 10 0
First attackGolden run   3.8889 × 10 1   4.1876 × 10 0
Second attackPV copy   5.3678 × 10 1   4.8352 × 10 0
Second attackGolden run   3.9102 × 10 1   4.1863 × 10 0
Third attack v. 1Copy of parameters   1.8223 × 10 1   2.0957 × 10 0
Third attack v. 1Golden run   1.7591 × 10 1   2.0220 × 10 0
Third attack v. 2Copy of parameters   1.7877 × 10 1   2.4675 × 10 0
Third attack v. 2Golden run   1.6701 × 10 1   2.2712 × 10 0
Table 2. Comparison of performance metrics obtained for the laboratory stand for a given attack scenario and the detection method.
Table 2. Comparison of performance metrics obtained for the laboratory stand for a given attack scenario and the detection method.
Attack ScenarioDetection MethodDetection RateFalse Positive RatePrecisionRecallTime to Detect (First Attack/Mean)
First attackMV copy   100.0 %   0.05 %   99.9 %   100.0 %   0.0 / 0.0
First attackGolden run   0 77.5 %   30.2 %   56.2 %   0 77.5 %   38.0 / 37.2
Second attackPV copy   100.0 %   0.1 %   99.8 %   100.0 %   0.0 / 0.0
Second attackGolden run   100.0 %   41.0 %   54.95 %   100.0 %   0.0 / 0.0
Third attack v. 1Copy of parameters   100.0 %   0.0 %   100.0 %   100.0 %   0.0 / n . a .
Third attack v. 1Golden run   3.55 %   0.0 %   100.0 %   3.55 %   527.0 / n . a .
Third attack v. 2Copy of parameters   100.0 %   0.0 %   100.0 %   100.0 %   0.0 / n . a .
Third attack v. 2Golden run   51.3 %   5.5 %   94.91 %   51.3 %   48.0 / n . a .
Table 3. Comparison of performance metrics obtained in the simulation experiment for a given attack scenario and the detection method.
Table 3. Comparison of performance metrics obtained in the simulation experiment for a given attack scenario and the detection method.
Attack ScenarioDetection MethodDetection RateFalse Positive RatePrecisionRecallTime to Detect (First Attack/Mean)
First attackGolden run   84.8 %   28.5 %   59.8 %   84.8 %   49.0 / 45.2
Second attackGolden run   83.8 %   23.38 %   64.19 %   83.8 %   38.0 / 44.0
Third attack v. 1Golden run   3.28 %   0.0 %   100.0 %   3.28 %   545.0 / n . a .
Third attack v. 2Golden run   55.85 %   0.0 %   100.0 %   55.85 %   29.0 / n . a .
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

Marusak, P.; Nebeluk, R.; Wojtulewicz, A.; Cabaj, K.; Chaber, P.; Ławryńczuk, M.; Plamowski, S.; Zarzycki, K. Efficient Cyberattack Detection Methods in Industrial Control Systems. Sensors 2024, 24, 3860. https://doi.org/10.3390/s24123860

AMA Style

Marusak P, Nebeluk R, Wojtulewicz A, Cabaj K, Chaber P, Ławryńczuk M, Plamowski S, Zarzycki K. Efficient Cyberattack Detection Methods in Industrial Control Systems. Sensors. 2024; 24(12):3860. https://doi.org/10.3390/s24123860

Chicago/Turabian Style

Marusak, Piotr, Robert Nebeluk, Andrzej Wojtulewicz, Krzysztof Cabaj, Patryk Chaber, Maciej Ławryńczuk, Sebastian Plamowski, and Krzysztof Zarzycki. 2024. "Efficient Cyberattack Detection Methods in Industrial Control Systems" Sensors 24, no. 12: 3860. https://doi.org/10.3390/s24123860

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