Fact-checked by Grok 2 weeks ago

Industrial process control

Industrial process control is the automated regulation of industrial manufacturing and production processes to maintain key variables—such as , , flow rates, and levels—at desired set points, using loops, sensors, controllers, and actuators to ensure consistency, safety, and efficiency. This discipline integrates hardware, software, and methodologies to measure, manipulate, and visualize processes in sectors like chemicals, oil and gas, pharmaceuticals, and food production, often through systems such as , Distributed Control Systems (DCS), and Programmable Logic Controllers (PLCs). Key components include for , final elements like valves and pumps for adjustments, and algorithms—such as Proportional-Integral-Derivative () schemes—that respond to deviations from set points via or mechanisms. By minimizing , optimizing use, and preventing hazards, industrial process control supports economic viability, environmental , and operational reliability in both continuous and batch environments. Advancements in this field, including adaptive and intelligent control strategies, address challenges like nonlinear dynamics and process uncertainties, enabling integration with broader plant automation networks for enhanced performance.

Historical Development

Early Innovations

The origins of industrial process control can be traced to the , when mechanical devices began to automate regulation of machinery to improve efficiency and safety. A pivotal innovation was the flyball governor, invented by in 1788 to maintain constant speeds despite varying loads. This device employed a centrifugal mechanism where rotating balls connected to the engine's throttle via linkages; as engine speed increased, the balls spread outward due to , lifting a sleeve that partially closed the steam valve through mechanical feedback, thereby stabilizing output. The flyball governor exemplified early principles, influencing subsequent control designs in power generation and manufacturing. During the late 19th century, the limitations of purely mechanical systems led to the development of pneumatic control technologies, which used signals for more precise and remote operation in harsh industrial environments. A landmark advancement was the introduction of the first industrial pneumatic controller by the Foxboro Company in the 1920s, a device that automatically adjusted valves based on pressure differentials to regulate processes like and in chemical plants. These systems enabled safer in sectors such as oil refining and textiles, where manual intervention was risky, by transmitting control signals over distances without electrical hazards. Early also extended to operations in manufacturing and chemical industries during the , where devices like float valves and thermostatic controls maintained water levels to prevent explosions from dry firing. For instance, by the mid-19th century, mercury-sealed float regulators in steam used to open or close feedwater valves, ensuring consistent and . Key figures such as Elmer Sperry advanced these efforts in the with gyroscopic controls, initially developed for ship stabilization but adapted for industrial applications like maintaining steady speeds through inertial mechanisms. Sperry's work laid groundwork for reliable in heavy machinery, bridging mechanical ingenuity with emerging precision needs. This era of mechanical and pneumatic innovations set the stage for later transitions to electronic systems in the mid-20th century.

Modern Advancements

The transition from mechanical and pneumatic systems to electronic and digital technologies in industrial process control began in the 1940s with the introduction of electronic analog controllers, which utilized vacuum tubes and operational amplifiers to implement proportional-integral-derivative (PID) control algorithms more reliably than earlier methods. These controllers enabled precise regulation of variables like temperature and pressure in chemical processes, reducing human intervention and improving response times in continuous manufacturing environments. By the late 1950s, the limitations of analog systems—such as susceptibility to noise and difficulty in reconfiguration—paved the way for digital innovation, exemplified by Texaco's implementation of the first direct digital computer control system in 1959 using the RW-300 computer at its Port Arthur refinery for polymerization processes. This milestone demonstrated the feasibility of real-time computation for multivariable control, marking the onset of computer-integrated process automation. The 1960s saw the parallel rise of programmable logic controllers (PLCs), initially developed by at Bedford Associates for ' automotive lines to replace hardwired in . PLCs offered reprogrammability via , facilitating rapid adjustments to production sequences and enhancing flexibility in batch-oriented industries like and . By the 1970s, the evolution accelerated with the advent of distributed control systems (DCS), pioneered by Honeywell's TDC 2000 in 1975, which decentralized control functions across networked microprocessors to improve reliability and scalability in large-scale operations such as power generation. Concurrently, Yokogawa launched its CENTUM DCS in the same year, emphasizing modular architecture for refineries and enabling redundant, fault-tolerant supervision of continuous processes. In the , minicomputer-based systems further bridged centralized and distributed paradigms, with platforms like the PDP-11 series from supporting supervisory control and (SCADA) for monitoring in utilities and . These systems integrated minicomputers with local controllers, allowing for more sophisticated data logging and operator interfaces while reducing wiring complexity compared to earlier mainframe setups. The 2010s introduced cloud-based SCADA architectures, leveraging and remote access to enable scalable data analytics and remote diagnostics, as seen in migrations to platforms like those from and for and energy sectors. By the 2020s, the adoption of Industry 4.0 principles has transformed process control through integration of the (IIoT), for , and digital twins—virtual replicas of physical assets that simulate and optimize operations in . In oil refining, for instance, digital twins combined with sensors have enabled predictive control models that forecast equipment failures and adjust feedstocks dynamically, reducing downtime in major facilities. algorithms, such as variants, enhance predictive control by learning from historical process data to optimize multivariable systems. These advancements, supported by , foster cyber-physical systems that adapt autonomously, though challenges like cybersecurity and persist. As of 2025, further integrations of and have advanced energy-efficient and sustainable control strategies.

Fundamental Concepts

Control Loops

A control loop serves as the basic operational unit in industrial process control, forming a closed sequence that involves sensing the current state of a process, comparing it to a desired , computing an adjustment based on the difference, and actuating changes to maintain . This sequence ensures that processes operate within specified parameters despite variations. The primary components of a control loop are the process variable (PV), which represents the measured output of the process such as , , or ; the setpoint (SP), the target value established for the PV; and the error signal, calculated as e = SP - PV, which quantifies the deviation driving corrective action. The loop typically includes a to detect the PV, a controller to process the error and generate an output signal, and an to implement adjustments in the process. A simple control loop is often depicted in a illustrating the flow: the enters a where it subtracts the feedback to yield the e, which inputs to the controller producing an output that drives the ; the modifies the process, whose response is sensed to generate the signal looping back to the . This structure enables continuous monitoring and correction. In practice, control loops appear in applications like temperature regulation in a , where a measures levels as the , compares them to the , and adjusts fuel input via a to sustain optimal ; or flow control in pipelines, where a meter tracks rate as the and a controller opens or closes a to match the desired throughput. These loops often employ controllers such as for error minimization. Control loops manage disturbances—unanticipated external influences like raw material variations or environmental shifts that perturb the —by detecting changes and applying compensatory adjustments to restore the . During steady-state operation, the loop achieves equilibrium where the holds constant at the , with the error approaching zero and process inputs balancing outputs for sustained performance.

Feedback Mechanisms

Feedback mechanisms form the core of closed-loop systems in , where the output is measured and fed back to adjust the input, enabling regulation and optimization. subtracts the output signal from the input to counteract deviations, promoting system stability by reducing to disturbances and variations. In contrast, adds the output to the input, amplifying deviations and often leading to or oscillations, though it can be useful for signal in specific applications like oscillators. These mechanisms are fundamental to achieving desired performance in processes such as temperature regulation in chemical reactors or speed in motors. The closed-loop transfer function for a unity feedback system, which relates the output to the reference input, is given by G_{cl}(s) = \frac{G(s)}{1 + G(s)H(s)}, where G(s) represents the forward path transfer function and H(s) the feedback path transfer function; this formulation highlights how feedback modifies the overall system dynamics. In non-unity feedback configurations, the denominator $1 + G(s)H(s) determines the characteristic equation, whose roots influence stability and transient response. Stability analysis of feedback systems relies on frequency-domain methods to ensure that closed-loop poles lie in the left half of the . The , introduced in 1932, assesses by examining the encirclements of the critical point (-1, 0) by the Nyquist plot of the open-loop G(s)H(s); for , the plot must encircle this point a number of times equal to the number of right-half-plane poles of the open-loop system, with counterclockwise direction indicating . This graphical test provides necessary and sufficient conditions for bounded-input bounded-output without solving the explicitly. Bode plots complement Nyquist analysis by plotting the magnitude and phase of G(j\omega)H(j\omega) versus frequency on logarithmic scales, allowing evaluation of gain and phase margins to quantify relative stability. A gain margin greater than 6 dB and phase margin exceeding 45 degrees typically ensure robust performance against variations. These margins are derived from the frequency where the phase reaches -180 degrees (for phase margin) or magnitude unity (for gain margin), providing practical insights into feedback loop robustness. The root locus method visualizes the migration of closed-loop poles as the gain varies from 0 to infinity, aiding in pole placement for desired and times. Developed by Evans in , it plots loci starting from open-loop poles and ending at open-loop zeros, with branches following rules such as symmetry about the real axis and asymptotes at angles determined by the number of poles minus zeros. This technique is particularly valuable for designing compensators to shift poles away from instability regions. Time delays in feedback loops, often arising from transport lags in processes like pipeline flows, introduce phase shifts that reduce stability margins and can cause oscillations or instability even if the delay-free system is stable. , typically from sensors or actuators, propagates through the loop, amplifying errors in high-gain configurations; mitigates this by attenuating disturbances, but excessive may necessitate filtering to prevent destabilization. Both effects underscore the need for robust design, such as using predictors for delays or Kalman filters for noise suppression, to maintain performance in industrial settings.

System Architecture

Hierarchical Structure

Industrial process control systems are typically organized in a hierarchical structure to manage complexity, ensure efficient data flow, and separate (OT) from (IT). This multi-level architecture facilitates real-time monitoring and at lower levels while enabling strategic decision-making at higher levels, promoting scalability and security in industrial environments. The foundational framework for this hierarchy is the (PERA), developed in the to standardize enterprise integration in and systems. The Purdue model delineates five primary levels, each with distinct functions and time horizons. Level 0 represents the physical process itself, encompassing the actual industrial operations such as chemical reactions or mechanical movements in a . Level 1 includes intelligent devices for sensing and actuation, like sensors measuring or and actuators adjusting valves or motors to influence the process. Level 2 focuses on supervisory , where systems monitor and regulate equipment through local , operating on timescales of seconds to minutes. Level 3 handles manufacturing operations management, coordinating production scheduling and over hours to days. An additional Level 3.5, known as the Industrial Demilitarized Zone (DMZ), serves as a secure buffer between Level 3 and Level 4, enabling controlled data exchange while preventing direct connections to mitigate cybersecurity risks. Finally, Level 4 addresses business planning and logistics, integrating (ERP) systems for long-term strategy, typically spanning days to years. This layered approach ensures that critical controls remain isolated from higher-level business networks, enhancing system reliability. Within this hierarchy, key technologies integrate seamlessly across levels to enable cohesive control. Distributed Control Systems (DCS) primarily operate at Level 2, providing decentralized, for continuous processes by distributing computing tasks across multiple controllers. Supervisory Control and (SCADA) systems also function at Level 2, offering wide-area supervision and logging, often bridging to Level 3 for operator interfaces and alarms. Manufacturing Execution Systems (MES) reside at Level 3, optimizing production workflows by aggregating from lower levels to track work-in-progress, manage inventory, and ensure compliance with manufacturing standards. These integrations allow for vertical exchange, such as real-time process flowing upward for optimization while commands cascade downward for execution. A representative example is found in operations, where the supports end-to-end control from to product distribution. At Level 0 and 1, field instruments like flow meters and thermocouples detect variables in units, feeding data to DCS at Level 2 for automated adjustments to maintain optimal temperatures and pressures. systems at Level 2 provide plant-wide visibility, allowing operators to monitor multiple units remotely. At Level 3, integrates this data to schedule maintenance, track yields, and align production with demand forecasts, ultimately linking to Level 4 for planning via the DMZ. This structure enables refinery-wide optimization, such as balancing energy use across units to minimize costs while ensuring safety. Despite its benefits, implementing hierarchical structures in large-scale settings presents challenges, particularly in managing volume and latency. As systems expand to encompass thousands of devices, the across levels can lead to bottlenecks in , where high-frequency overwhelms upper-level processing, delaying . Additionally, ensuring reliability in or distributed becomes complex, requiring advanced protocols to maintain without compromising security or performance. These issues often necessitate hybrid architectures or to distribute loads and enhance resilience in expansive facilities like plants.

Control Models

Control models in industrial process control provide mathematical frameworks to represent, analyze, and predict the dynamic behavior of processes, enabling engineers to design effective strategies. These models approximate the underlying physics or empirical relationships of systems such as chemical reactors, columns, or lines, facilitating , optimization, and assessment before implementation. By capturing input-output relationships, control models support the development of algorithms that maintain desired operating conditions despite disturbances or setpoint changes. Process modeling begins with techniques like transfer functions, which describe the linear relationship between system inputs and outputs in the using Laplace transforms. For a , the transfer function G(s) = \frac{Y(s)}{U(s)} relates the output Y(s) to the input U(s), assuming zero initial conditions and . This approach is widely used in industrial applications for its simplicity in analyzing steady-state gains and time constants, such as in of heat exchangers. Another foundational method is , which models multi-input multi-output systems through a set of differential equations. The standard form is given by: \dot{x} = Ax + Bu \\ y = Cx + Du where x is the state vector, u the input, y the output, and A, B, C, D are system matrices derived from physical principles or data. State-space models are particularly valuable for MIMO processes like multivariable chemical plants, allowing representation of internal dynamics not visible in transfer functions. Many industrial processes exhibit nonlinear behavior, such as reaction kinetics or , necessitating linearization around operating points for applying linear control techniques. Linearization involves Taylor series expansion of the nonlinear equations about a nominal steady-state point (x_0, u_0), retaining only the first-order terms to yield a linear approximation \dot{\delta x} = A \delta x + B \delta u, where \delta x = x - x_0 and \delta u = u - u_0. The Jacobian matrices A = \frac{\partial f}{\partial x} \big|_{x_0, u_0} and B = \frac{\partial f}{\partial u} \big|_{x_0, u_0} (from the nonlinear form \dot{x} = f(x, u)) define the linearized model, valid near the operating point but requiring gain scheduling for wider ranges in processes like pH control. Model predictive control (MPC) leverages these models to optimize future process behavior over a horizon by solving an online . In MPC, a dynamic model forecasts outputs based on current measurements and proposed control moves, minimizing a objective function subject to constraints on inputs, outputs, and rates, such as \min J = \sum (y_k - r_k)^2 Q + \sum \Delta u_k^2 R with bounds u_{\min} \leq u_k \leq u_{\max}. This constraint-handling capability makes MPC suitable for industrial applications like , where it outperforms traditional controllers by anticipating disturbances and respecting equipment limits. Simulation tools are essential for validating control models against real-world data or hypothetical scenarios. and enable graphical block-diagram modeling of transfer functions or state-space systems, allowing simulation of closed-loop responses to verify stability and performance metrics like . For instance, 's MPC Designer app facilitates tuning of predictive controllers by simulating nonlinear and comparing predictions to measured outputs, ensuring model fidelity before deployment in processes such as batch reactors. Control models differ fundamentally in their treatment of : deterministic models assume exact predictability based on conditions and , as in linear transfer functions for stable systems, while models incorporate random variations like or parameter fluctuations using probability distributions. In industrial contexts, deterministic approaches suffice for well-characterized environments, but models, often employing techniques like Kalman filtering, are critical for noisy processes such as turbulent mixing or sensor-limited monitoring, providing probabilistic forecasts rather than point predictions.

Types of Control Systems

Open-Loop Systems

Open-loop systems operate without from the output to adjust the input, relying solely on a predefined relationship between the input and the desired output. In these systems, the controller generates an output command based on the setpoint and a model of the process, without measuring or correcting for actual process deviations. This approach is particularly suited to where the are well-understood and predictable, eliminating the need for error detection. One key advantage of open-loop systems is their faster response time, as there is no delay introduced by feedback loops that require sensing and computation. They also tend to be lower in cost due to the absence of sensors and feedback circuitry, making them economical for simple tasks. Additionally, open-loop designs avoid potential issues associated with , such as oscillations or from improper . Despite these benefits, open-loop systems are highly sensitive to external disturbances, such as variations in environmental conditions or component wear, which can lead to output errors without any corrective mechanism. Their performance also degrades if the internal model of the process is inaccurate, as there is no provision for adapting to unmodeled dynamics or changes over time. In contrast to closed-loop systems, open-loop approaches do not inherently provide against such perturbations. In industrial applications, open-loop systems are commonly used for timing in , where the speed is set based on requirements without needing to monitor belt position continuously. Another example is batch dosing in pharmaceuticals, where precise volumes of ingredients are dispensed according to a fixed , assuming consistent equipment behavior. Mathematically, an open-loop can be represented by the equation y(t) = G(u(t)) where y(t) is the system output at time t, u(t) is the input, and G denotes the or model of the process, with no term involved.

Closed-Loop Systems

Closed-loop systems in industrial control incorporate mechanisms to continuously monitor and adjust variables, enabling automatic correction of deviations from desired setpoints. Unlike open-loop configurations, these systems use output measurements to compute and generate corrective actions, enhancing and precision in dynamic environments. The core of most closed-loop implementations is the structure, where sensors detect the process output, compare it to the setpoint, and feed the error signal to a controller that modulates actuators accordingly. The proportional-integral-derivative () controller remains the most widely adopted mechanism in industrial applications due to its simplicity, robustness, and effectiveness in handling a broad range of processes. It computes the output u(t) based on the e(t), which is the difference between the setpoint and the measured . The proportional term, governed by K_p, provides an output proportional to the current , offering immediate response but potentially leading to steady-state offsets if used alone. The integral term, with K_i, accumulates past errors over time via K_i \int_0^t e(\tau) \, d\tau, eliminating residual offsets by addressing accumulated discrepancies. The derivative term, controlled by K_d, anticipates future errors through the rate of change K_d \frac{de(t)}{dt}, damping oscillations and improving stability. The complete PID law is expressed as: u(t) = K_p e(t) + K_i \int_0^t e(\tau) \, d\tau + K_d \frac{de(t)}{dt} This formulation allows fine-tuned regulation across diverse systems, from simple temperature loops to complex multivariable processes./09:Proportional-Integral-Derivative(PID)_Control/9.02:_P_I_D_PI_PD_and_PID_control) Tuning the PID parameters is essential for optimal performance, with the Ziegler-Nichols method providing a foundational, oscillation-based approach developed in 1942. This closed-loop technique involves setting integral and derivative gains to zero, then incrementally increasing the proportional gain until the system sustains sustained oscillations at the ultimate gain K_u and period P_u. Controller parameters are then derived from these values using empirical rules: for a PID controller, K_p = 0.6 K_u, K_i = 1.2 K_u / P_u, and K_d = 0.075 K_u P_u. These settings aim for a quarter-amplitude decay response, balancing responsiveness and stability, though they may require refinement for nonlinear processes./09:Proportional-Integral-Derivative(PID)_Control/9.03:_PID_Tuning_via_Classical_Methods) For processes exhibiting nonlinearity or time-varying dynamics, advanced variants extend traditional capabilities. Fuzzy logic controllers integrate rule-based inference to dynamically adjust gains based on linguistic variables like "large error" or "rapid change," mimicking human expertise without precise mathematical models; this approach proves effective in handling uncertainties, as seen in fuzzy- hybrids for motor speed . Adaptive controllers, meanwhile, employ online parameter estimation algorithms, such as model reference , to continuously retune K_p, K_i, and K_d in response to changing characteristics, ensuring consistent performance in varying operating conditions like load disturbances. These enhancements maintain the framework's familiarity while improving robustness in challenging industrial scenarios. In practice, closed-loop PID systems excel in applications requiring precise regulation, such as maintaining in oil and gas , where the controller adjusts positions to counteract variations and ensure safe operating limits. Similarly, in , PID controllers regulate by dosing acids or bases, stabilizing neutralization processes to meet environmental discharge standards and prevent equipment corrosion. These implementations demonstrate the versatility of feedback control in sustaining . Performance of closed-loop systems is evaluated through key metrics that quantify response quality to setpoint changes or disturbances. measures the duration to reach 90% of the setpoint from 10%, indicating speed of response. Overshoot quantifies the peak exceedance beyond the setpoint as a , reflecting effectiveness. Steady-state assesses the residual deviation after stabilization, ideally approaching zero with action. Well-tuned controllers typically achieve rise times under 10 seconds, overshoots below 20%, and negligible steady-state errors in stable processes, underscoring their role in reliable industrial automation./09:Proportional-Integral-Derivative(PID)_Control/9.02:_P_I_D_PI_PD_and_PID_control)

Implementation Components

Sensors and Actuators

Sensors and actuators form the foundational field devices in industrial process control systems, where sensors detect and quantify process variables such as , , and , while actuators apply corrective actions to maintain desired operating conditions. These devices operate at the lowest level of the control hierarchy, directly with the physical process to enable monitoring and manipulation. Sensors convert physical phenomena into measurable electrical signals, with common types including thermocouples for temperature measurement, which operate on the Seebeck effect and are suitable for high-temperature environments up to 1700°C with accuracies typically around ±1.5°C or better in industrial settings. Resistance temperature detectors (RTDs), often platinum-based, offer higher accuracy (±0.1°C) and stability over narrower ranges like -200°C to 850°C, making them ideal for precise applications. Flow meters, such as orifice plates, measure differential pressure to infer flow rates with accuracies of ±1-2% of full scale across ranges from 0.1 to 10 m/s for liquids and gases. Ultrasonic flow meters provide non-intrusive measurement using transit-time differences, achieving ±1% accuracy for clean fluids over velocities up to 30 m/s without pressure drop. Industrial pressure transducers, typically strain-gauge or piezoresistive based, measure pressures from near 0 Pa (gauge) to over 100 MPa with typical accuracies of ±0.075% to ±0.5% of full scale. Calibration of these sensors involves comparison against traceable standards, often using environmental chambers or flow rigs, to ensure accuracy within specified tolerances, with intervals typically every 6-12 months depending on process criticality. Actuators translate control signals into mechanical actions to adjust variables, including valves like types for linear control with tight shutoff capabilities and response times under 1 second for small sizes, or butterfly valves for quick quarter-turn operation in large-diameter with requirements up to several hundred . Electric serve as actuators in positioning systems, delivering torques from 0.1 to 1000 with response times of milliseconds to seconds based on drive electronics. Pumps, often centrifugal or positive displacement, actuate fluid movement with rates up to thousands of liters per minute and pressure heads exceeding 100 , where response time is influenced by and integration. Signal conditioning processes raw sensor outputs to make them suitable for control systems, involving to boost weak signals, filtering to remove noise (e.g., low-pass filters for high-frequency interference), and analog-to-digital conversion using successive approximation or sigma-delta methods to achieve 12-24 bit resolution with effective noise reduction through . These steps ensure , minimizing errors from common in industrial environments. In chemical reactors, RTDs monitor temperature gradients to prevent runaway reactions, providing stable readings essential for safe operation. valves, with fast response times under 50 ms, pneumatic actuators in assembly lines for precise sequencing of material handling. Emerging technologies include wireless sensors that enable mesh networks for distributed monitoring, reducing cabling costs while maintaining accuracies comparable to wired systems through protocols like , with advancements in 2024-2025 focusing on low-power for real-time data processing. Smart actuators incorporate embedded diagnostics and self-tuning capabilities, such as via vibration analysis, integrated into cyber-physical systems for enhanced autonomy by 2025.

Controllers and Interfaces

Controllers in industrial process control serve as the core computational units that receive input signals from sensors, process control algorithms, and issue commands to actuators to maintain desired process conditions. These devices interpret data, execute logic, and ensure real-time responsiveness in environments like manufacturing and chemical processing. Programmable Logic Controllers (PLCs) are rugged, modular hardware devices designed for discrete automation tasks, featuring input/output modules, central processing units, and power supplies to handle binary signals and sequential operations reliably in harsh industrial settings. Distributed Control Systems (DCS) consist of interconnected modules distributed across a plant, providing centralized monitoring and decentralized control for continuous processes such as oil refining, with redundant architectures to enhance fault tolerance. Embedded microcontrollers, smaller and more integrated than PLCs, are used in compact applications like machine tools, offering customizable processing via microprocessors with integrated memory and I/O interfaces for cost-effective, low-power control. Control software for these hardware platforms includes graphical and textual languages standardized under IEC 61131-3. Ladder logic programming mimics relay circuits with rung-based diagrams, enabling electricians to implement logic for tasks like motor sequencing without deep coding knowledge. Function block diagrams (FBD) represent control as interconnected blocks for modular, data-flow-oriented programming, facilitating complex functions like control and reusable code in process industries. Interfaces bridge controllers with operators and other systems for seamless interaction. Human-Machine Interfaces (HMIs) provide graphical touchscreens or web-based dashboards for real-time visualization, alarming, and manual overrides, allowing operators to monitor variables and adjust setpoints intuitively. OPC UA (Open Platform Communications Unified Architecture) is a secure, platform-independent protocol for standardized data exchange between controllers, HMIs, and enterprise systems, supporting semantic modeling and publish-subscribe mechanisms to integrate heterogeneous devices. Cybersecurity in controllers has become critical by 2025, with protocols like —widely used for simple, master-slave communication in legacy systems—vulnerable to eavesdropping and spoofing due to its lack of built-in . To counter threats, intrusion detection systems (IDS) employ AI-driven on Modbus traffic and controller logs, identifying deviations like unauthorized commands in real-time while complying with standards such as for resilience in . A representative example is the series, deployed in automotive plants for synchronized assembly lines, where PLCs and motion controllers process sensor signals to regulate robotic welding and part handling with sub-millisecond precision in facilities like those of .

Applications and Benefits

Industrial Applications

Industrial process control is extensively applied across diverse sectors to optimize operations, ensure safety, and maintain product quality. In chemical processing, columns are critical for separating liquid mixtures based on differences in , with control systems regulating variables such as reflux ratio, heat input, and feed flow to achieve desired separation efficiency. Advanced control strategies, including multivariable predictive control, address interactions between temperature, pressure, and composition to minimize energy consumption while maximizing purity, as demonstrated in models accounting for complexities. For instance, inferential composition control using soft sensors integrates active disturbance rejection to maintain product specifications under varying feed conditions. In the oil and gas industry, pipeline flow control systems monitor and adjust pressure, flow rates, and temperature to prevent surges and ensure efficient transport, often employing supervisory control and (SCADA) architectures for oversight. Safety shutdown systems, mandated by regulations, utilize automatic shut-off valves and emergency shutdown (ESD) protocols to isolate segments within 30 minutes of detecting anomalies like ruptures, thereby mitigating environmental and operational risks. These controls incorporate sensors for and remote valve actuation, with operators trained to respond to alarms for rapid intervention. Manufacturing processes leverage process control in computer numerical control (CNC) machines and robotic assembly lines to achieve precision and repeatability. CNC systems use feedback loops to regulate spindle speed, tool position, and feed rates, enabling that compensates for and material variations. In robotic assembly, coordinated control integrates vision systems and programmable logic controllers (PLCs) to synchronize movements, ensuring accurate part placement and reducing cycle times in automotive and production. Such implementations enhance by automating machine tending and quality inspection tasks. Power generation in thermal relies on level to maintain stable production and prevent damage from carryover or dry-out. Three-element strategies combine level, flow, and feed flow measurements to counteract shrink-and-swell effects caused by load changes, using proportional-integral-derivative () or for precise regulation. In coal-fired , statistical models based on mass and energy balances predict level dynamics, allowing operators to adjust and inputs for efficient operation under varying demands. In the food and pharmaceutical industries, process control ensures sterilization and batch consistency to meet stringent safety standards. For pharmaceuticals, FDA-compliant controls under (CGMP) regulate temperature, pressure, and time in to achieve parametric release without end-product sterility testing, while batch records document deviations for validation. Production controls require formulation to provide at least 100% of active ingredients, with in-process adjustments to maintain uniformity across batches. In , similar controls oversee thermal sterilization in or , using sensors to monitor microbial inactivation while preserving nutritional quality, aligned with FDA guidelines for and critical control points (HACCP). Emerging applications in renewables include turbine control systems, which optimize power output through , yaw, and regulation to maximize energy capture while minimizing structural loads. As of 2025, advanced controls incorporate real-time wind forecasting and for farm-level coordination, enhancing overall efficiency in installations. In (EV) battery production lines, process control manages coating, cell assembly, and formation stages to ensure uniformity and safety, addressing challenges like material sensitivity through automated inspection and feedback systems. University-scale pilot lines, such as those expanded in 2025, demonstrate scalable controls for high-capacity , focusing on defect detection and yield optimization.

Economic and Operational Advantages

Industrial process control systems deliver substantial economic benefits by minimizing operational costs through reduced waste and enhanced . For instance, (MPC) implementations in refineries and chemical plants have achieved energy savings of 15-18%, such as an 18% reduction in in a para-xylene production unit at Chemical's facility and a 15% decrease in steam usage in PVC plant columns, translating to annual savings of approximately $220,000 (in 1978 dollars) for the latter. These improvements stem from MPC's ability to optimize multivariable processes in , directly lowering fuel and utility expenses while maintaining production levels. Return on investment for (DCS) installations is often realized within 1-3 years in the manufacturing industry, driven by cumulative savings from and reduced manual interventions. Case studies highlight rapid financial recovery through deferred capital expenditures and optimized asset utilization. On the operational front, these systems enhance safety by incorporating fail-safe mechanisms like (SIS) and emergency shutdown (ESD) protocols, which could have mitigated disasters such as the 1984 incident. In , the absence of an operational SIS/ESD allowed a runaway reaction in a to proceed unchecked, leading to a catastrophic release; post-incident analyses show that such automated safeguards, when integrated into process control, provide independent protection layers with probability of failure on demand () as low as 0.01, significantly reducing the risk of similar events. Regulatory advancements following Bhopal, including the U.S. Process Safety Management standard, have mandated these controls, resulting in fewer major incidents across global chemical operations. Process control also boosts key operational metrics, including increased throughput by up to 10-20% through optimized utilization and reduced bottlenecks, greater product consistency via stable variables that minimize variability, and improved for expanding operations without proportional cost increases. These gains arise from closed-loop mechanisms that maintain setpoints, ensuring reliable output in sectors like and . In modern smart factories, AI-driven control extensions further amplify these advantages, reducing unplanned downtime by 35-45% through algorithms that forecast failures using . This integration supports by lowering emissions; for example, precise control of processes in industrial plants can cut outputs by 10-20% via optimized fuel use and recovery, aligning with net-zero goals as accounts for about 40% of global emissions.

References

  1. [1]
    Process Automation and Control - ISA
    Process automation and control encompasses hardware, software, systems and methods that measure, regulate, manipulate and visualize industrial processes.
  2. [2]
    industrial control system (ICS) - Glossary | CSRC
    An ICS is an information system controlling industrial processes, including SCADA, DCS, and PLC, using control components to achieve industrial objectives.
  3. [3]
  4. [4]
    Process controls - processdesign
    Feb 21, 2016 · Process control maintains a system's output within a desired range by manipulating inputs, modeling system behavior when a perturbation occurs.
  5. [5]
    Unit Manufacturing Processes: Issues and Opportunities in Research
    Process control can be viewed as the executive portion of a unit process. It provides the means to direct a process so that it produces the desired results.
  6. [6]
    [PDF] Advanced - PID Control - International Society of Automation (ISA)
    Analog PID controllers were used at the base level and the computer supplied set point to the controllers. As the systems developed, the attention focused ...
  7. [7]
    Computers in Control - CHM Revolution
    Close cooperation between Texaco and Ramo Wooldridge achieved “closed loop control” of an industrial polymerization process using a RW-300 computer in 1959.
  8. [8]
    Digital control of industrial processes | ACM Computing Surveys
    Describes the results of a joint venture by Monsanto and Bunker-Ramo in computer control. ... Discusses the use of a digital computer for first:-tevel plant ...
  9. [9]
    [PDF] History of Control History of PLC and DCS
    Jun 15, 2012 · The need for improved industrial processes in the 50s and 60s led to the development of PLCs and DCSs, with early PLCs in the 1960s as relay ...<|separator|>
  10. [10]
    Distributed Control System - Honeywell Process Solutions
    The first Distributed Control System (DCS) was developed by Honeywell in 1975, with their TDC 2000 system. This system introduced a new approach to industrial ...
  11. [11]
    CENTUM-History | Yokogawa America
    To achieve this philosophy in the control industry, Yokogawa launched CENTUM, the world's first distributed control system (DCS), in 1975.Missing: rise Honeywell
  12. [12]
    Rise and Fall of Minicomputers
    Oct 24, 2019 · Minis were designed for process control and data transmission and switching, whereas mainframes emphasized data storage, processing, and ...
  13. [13]
    [PDF] EVOLUTION OF SCADA SYSTEMS
    This paper presents the evolution, the architecture and the main components of a SCADA system. Key words: supervisory control, data acquisition, process control ...
  14. [14]
    Digital Twin-Based Predictive Maintenance of a Polish Filter in Palm ...
    Aug 29, 2025 · This paper applies an Industry 4.0 model using a digital twin for a filter in palm oil refining, using machine learning to predict sleeve ...
  15. [15]
    Integrating industry 4.0 for enhanced sustainability: Pathways and ...
    This paper provides a comprehensive analysis of how Industry 4.0 technologies, such as artificial intelligence, the Industrial Internet of Things, blockchain, ...
  16. [16]
    (PDF) Advancing Industrial IoT and Industry 4.0 through Digital Twin ...
    Aug 9, 2025 · This paper presents a comprehensive review of self-tuning databases using machine learning techniques. We explore various approaches to ...
  17. [17]
    Single-loop control- Still the mainstay of advanced process control
    Single-loop control (i.e., a control loop with one input and one output) as implemented in various versions of the proportional, integral, derivative control ...
  18. [18]
    [PDF] Process Control Fundamentals
    Feedback loops are commonly used in the process control industry. The advantage of a feedback loop is that it directly controls the desired process variable.
  19. [19]
    [PDF] Chapter 15 of: - Fundamentals of Process Control
    In discussing control loops, each controlled variable in a process must be paired with a specific manipulated variable at any given instant of time. However, ...
  20. [20]
    The Components of a Control Loop - Control Guru
    Mar 31, 2015 · Block diagrams help us visualize the components of a loop and see how the pieces are connected. A home heating system is simple on/off control ...
  21. [21]
    8. FEEDBACK CONTROL SYSTEMS
    As a rule negative feedback systems are more stable than positive feedback systems. Negative feedback also makes systems more immune to random variations in ...
  22. [22]
    Positive and Negative Feedback - SpringerLink
    Positive feedback reinforces a given tendency of a system and can lead a system away from equilibrium states, possibly causing quite unexpected results.
  23. [23]
    [PDF] Regeneration Theory - By H. NYQUIST
    Regeneration Theory. By H. NYQUIST. Regeneration or feed-back is of considerable importance in many appli- cations of vacuum tubes. The most obvious example ...
  24. [24]
    Control Design and Stability Analysis of Power Converters
    Mar 1, 2021 · This method is based on the Nyquist criterion but developed from the open-loop Bode diagram, evaluated also at 0 Hz and at the Nyquist frequency ...Missing: plots | Show results with:plots
  25. [25]
  26. [26]
  27. [27]
  28. [28]
    [PDF] PURDUE MODEL FRAMEWORK FOR INDUSTRIAL CONTROL ...
    The Purdue model provides a framework for segmenting industrial control system networks from corporate enterprise networks and the internet.
  29. [29]
    The Purdue enterprise reference architecture - ScienceDirect.com
    This paper presents the basic concepts which comprise the Purdue Enterprise Reference Architecture along with a description of its development and use.Missing: original | Show results with:original
  30. [30]
  31. [31]
    [PDF] Networking and Security in Industrial Automation Environments - Cisco
    Recommended models would be the Purdue Model of Control Hierarchy, International Society of Automation 95. (ISA95) and IEC 62443, NIST 800-82 and NERC CIP ...
  32. [32]
    Oil and Gas Pipeline Security Reference Document - Cisco
    Feb 11, 2019 · The Purdue Model for Control Hierarchy also describes a hierarchical data flow model, where sensors and other field devices are connected to the ...
  33. [33]
    [PDF] More General Transfer Function Models
    Most industrial process control applications involved a number of input (manipulated) and output (controlled) variables. • These applications often are referred ...
  34. [34]
    [PDF] introduction to - process control
    Process control uses methods to force parameters to specific values, regulating a quantity to a desired value, maintaining it regardless of external influences.
  35. [35]
    Introduction: System Modeling
    The state-space representation, also referred to as the time-domain representation, can easily handle multi-input/multi-output (MIMO) systems, systems with non ...
  36. [36]
    State Space Representations of Linear Physical Systems
    The state space representation of a system replaces an n th order differential equation with a single first order matrix differential equation.Simple Example · Developing a state space... · Developing state space model...
  37. [37]
    [PDF] 8.6 Linearization of Nonlinear Systems
    Using the state space model, the linearization procedure for the multi-input multi-output case is simplified. Consider now the general nonlinear dynamic control ...
  38. [38]
    [PDF] Nonlinear Control Using Linearizing Transformations - DSpace@MIT
    The usual approach to controlling nonlinear systems is to linearize, about an operating point, the nonlinear dynamics and apply proven linear control design ...
  39. [39]
    [PDF] Tutorial overview of model predictive control - IEEE Control Systems ...
    Model predictive control (MPC) optimizes forecasts of process behavior using a process model, and is aimed at control practitioners.
  40. [40]
    [PDF] A survey of industrial model predictive control technology - CEPAC
    Model predictive control (MPC) uses a process model to predict future plant response and optimize future behavior by computing adjustments.
  41. [41]
    Simulink - Simulation and Model-Based Design - MATLAB
    Simulink is a block diagram environment used to design systems with multidomain models, simulate before moving to hardware, and deploy without writing code.Simulink Online · Model-Based Design · For Students · Getting Started
  42. [42]
    Simulation - MATLAB & Simulink - MathWorks
    Simulate controllers against linear or nonlinear plants in MATLAB and Simulink. The toolbox provides tools for simulating your controller from the command line ...
  43. [43]
    Stochastic systems - Industrial & Operations Engineering
    Stochastic systems involve uncertainty, where output is not always the same for a given input, unlike deterministic systems.
  44. [44]
    [PDF] STOCHASTIC MODELING OF CHEMICAL, PROCESSES
    Abstract-Conventional simulation models and process simulators treat the design variables in process calculations deterministically.
  45. [45]
  46. [46]
  47. [47]
    [PDF] Optimum Settings for Automatic Controllers
    By J. G. ZIEGLER' and N. B. NICHOLS,' ROCHESTER, N. Y.. In this paper, the three principal control effects found in present controllers are examined and ...
  48. [48]
  49. [49]
    A Fuzzy Logic Based PID Self-Tuning Scheme For Process Control
    A new fuzzy logic based PID self-tuning regulator is proposed in this paper. The basic concept behind the new scheme is to combine the classical PID control ...
  50. [50]
  51. [51]
    [PDF] Design and Simulation of PID Controller for PH Neutralization Process
    Jan 30, 2020 · Abstract: PH control system plays an important role in a wide range of industrial applications particularly in wastewater treatment management. ...
  52. [52]
    [PDF] Process Control Basics - International Society of Automation (ISA)
    Some actuators include a horsepower rating, which is basi- cally used to determine how quickly the valve can move with the maximum torque applied.
  53. [53]
    [PDF] A Guide to the Automation Body of Knowledge
    The valve stem of linear motion valves moves toward or away from the orifice driven by reciprocating actuators. Ball valves and butterfly valves are both rotary ...
  54. [54]
    [PDF] Advanced Temperature Measurement and Control
    In industrial environments, high process temperatures, pressures, and vibration make it necessary to have a robust temperature sensor. Fast response time, ...
  55. [55]
    [PDF] Guide to secondary Thermometry: Industrial Platinum Resistance ...
    Nov 22, 2021 · the best temperature range and accuracy that can be expected from the main types of PRT. Page 18. Guide on Secondary Thermometry: IPRTs. 18.Missing: transducers | Show results with:transducers
  56. [56]
    [PDF] Introduction to liquid flow metering and calibration of liquid flowmeters
    Usually the temperature of liquids is measured with immersion sensors such as thermo- meters, thermocouples and thermistors. In general, the accurate ...<|control11|><|separator|>
  57. [57]
    [PDF] Flow Measurement
    Flow measurement includes differential pressure, critical, magnetic, mass, open channel, and ultrasonic flowmeters, among others.Missing: transducers | Show results with:transducers
  58. [58]
    [PDF] PRESSURE AND VACUUM MEASUREMENTS - Charles R. Tilford
    Oct 7, 2010 · It is confined to the industrial pressure range, 10-8- 10 Pa (10-1º torr to 150,000 psi). In all cases the accuracy obtained depends not only ...
  59. [59]
    [PDF] On-site calibration of flow metering systems installed in buildings
    On-site calibration of flow metering systems in buildings is about measuring flow of fluids like air, water, and steam, which is often the least accurate ...
  60. [60]
    [PDF] Essentials of Modern Measurements and Final Elements in the ...
    The most common example of a measurement with a primary element is a differential head flowmeter with a capacitance type differential pressure transmitter (DP).
  61. [61]
    [PDF] Flow measurement in open channels and closed conduits Volume 2
    This document, 'Flow Measurement in Open Channels and Closed Conduits', is Volume 2, published by the National Bureau of Standards.
  62. [62]
    Trade-Offs in Sensor Systems Design: A Tutorial - IEEE Xplore
    May 31, 2022 · Analog to Digital Converter (ADC). The conversion of the analog signal into digital is essential for further processing and storage of the ...
  63. [63]
    [PDF] Deconvolution of time domain waveforms in the presence of noise
    Frequency and time domain methods were studied along with the synthesis of the filters required to obtain stable and smooth results.
  64. [64]
    Monitoring Reactor Temperatures in Chemical Plants
    Jun 9, 2023 · RTD sensors are used for monitoring reactor temperatures in chemical plants, ensuring safety, efficiency, and optimal performance. They are ...Missing: solenoid valves lines
  65. [65]
    Solenoid Valves & Discrete Valve Controllers
    Available are ASCO™ HV216, HV226, Numatics Series 520, 521, L01, L1, L2, 015, and 025 solenoid valves, including gas shutoff, mini-spool, and manual reset ...
  66. [66]
    Wireless Sensor Networks: From Fundamentals and Applications to ...
    Jun 9, 2025 · WSNs have emerged as a transformative technology for distributed sensing, data collection, and communication in diverse applications ranging ...
  67. [67]
    The New Paradigm Toward Autonomous Industrial Cyber-Physical ...
    Sep 9, 2025 · Abstract—Autonomous Industrial Cyber-Physical Systems. (ICPS) represent a future vision where industrial systems achieve.
  68. [68]
    PLC/PAC and Controllers | Schneider Electric USA
    A PLC (Programmable Logic Controller) is a computing device, like your PC or smart phone, that is used to monitor and control the operation of machines and ...
  69. [69]
    Types of Industrial Control Systems: Examples and Applications
    Nov 25, 2024 · From PLC and DCS to SCADA, each ICS type enables precision, scalability, and adaptability, thereby enhancing operational efficiencies across global industries.Missing: microcontrollers | Show results with:microcontrollers
  70. [70]
    PLC vs Microcontroller Difference: Choosing the Ideal Controller for ...
    PLCs are industrial controllers explicitly developed for high-performance manufacturing and process automation applications. They are purpose-built for ...
  71. [71]
    What Is the Difference Between Ladder Logic and Function Block ...
    Jan 28, 2019 · Ladder logic is read from top to bottom and left to right. Function block diagrams can essentially convert several lines of ladder logic into boxes.
  72. [72]
    Function Block Diagram (FBD) PLC Programming Tutorial for ...
    Mar 13, 2018 · FBD, just like ladder logic and structured text, follows the IEC 61131-3 standard from PLCOpen. It's common in PLC programming, often used ...What is Function Block Diagram? · Function Blocks in FBD
  73. [73]
    Automatic HMI tag creation through OPC UA servers - AVEVA
    Apr 30, 2025 · InTouch now enables you to drag and drop items from the OPC UA server to the Model-Tagname dictionary, bypassing the need for manual tag creation.<|separator|>
  74. [74]
    What is OPC Unified Architecture (OPC UA)? - PTC
    OPC UA helps to enable good security practices for process control and data acquisition for SCADA/HMI and Industry 4.0/Digital Transformation within modern ...
  75. [75]
    Enhancing Modbus/TCP-Based Industrial Automation and Control ...
    In this paper, we propose the use of a misuse-based intrusion detection system (IDS) to detect out-of-bound process values.
  76. [76]
  77. [77]
    2025 State of Industrial Cybersecurity Solutions
    Sep 22, 2025 · AI-driven threat detection is cited as a driving force along with use of “zero-trust” security architectures and advanced network segmentation.
  78. [78]
    Siemens unveils breakthrough in automation technology with new ...
    May 8, 2024 · The initial deployment of Simatic Workstations will be deployed across a number of Ford Motor Company's manufacturing facilities. The Simatic ...
  79. [79]
    Distillation Columns
    Apr 5, 2022 · The column operation dictates the separation of reactants and products. The reactants are more volatile than the products in this example, and ...Conventional Distillation · Packed Beds · Advanced Distillation... · Vacuum
  80. [80]
    [PDF] ITP Chemicals: Distillation Column Modeling Tools - eere.energy.gov
    The multiphase gas/liquid flow patterns in distillation columns are complex, making them difficult to predict and control.
  81. [81]
    Inferential Composition Control of a Distillation Column Using Active ...
    Jan 16, 2023 · This paper presents the integration of active disturbance rejection control (ADRC) with soft sensors for enhancing the composition control ...<|separator|>
  82. [82]
    [PDF] PIPELINE SAFETY Better Data and Guidance Needed to Improve ...
    Jan 23, 2013 · Six minutes later, the control room operator shut down the pipeline and isolated the affected segment with remote-control valves. About two ...
  83. [83]
    PHMSA Announces Requirements for Pipeline Shut-off Valves to ...
    Mar 31, 2022 · New rules require remotely controlled shut-off valves on pipelines 6 inches or greater, with a 30-minute max shut-down time after rupture, and ...Missing: industry | Show results with:industry
  84. [84]
    CNC and Flexible Automation - Yoram Koren - University of Michigan
    Dr. Koren has developed innovative control methods for machine tools aimed at enhancing their precision and increasing their productivity.
  85. [85]
    [PDF] Robotic Automation of a CNC Machine - ucf stars
    [8] Discusses the capability of industrial robots to automate applications such as machine tending, painting and assembly. Further it discusses controls and ...
  86. [86]
    Scalable production of large components by industrial robots and ...
    Dec 14, 2022 · This paper presents a novel approach for manufacturing large components by industrial robots and machine tools through segmented manufacturing.
  87. [87]
    Drum-boiler control system employing shrink and swell effect ...
    Abstract: This paper develops a control system for a drum-boiler unit, which employs a three element cascade control (water level, steam flow, feed-water) ...
  88. [88]
    Statistical modeling of an integrated boiler for coal fired thermal ...
    Jun 13, 2017 · Modeling of the boiler drum of the coal-fired thermal power plant through the mass balance and energy balance was initiated by Astrom and Bell ...
  89. [89]
    Q7A Good Manufacturing Practice Guidance for Active ... - FDA
    This document is intended to provide guidance regarding good manufacturing practice (GMP) for the manufacturing of active pharmaceutical ingredients (APIs)
  90. [90]
    21 CFR Part 211 Subpart F -- Production and Process Controls - eCFR
    (a) The batch shall be formulated with the intent to provide not less than 100 percent of the labeled or established amount of active ingredient. (b) Components ...
  91. [91]
    Wind Turbine Control Systems | Wind Research - NREL
    Feb 21, 2025 · Researchers design, implement, and test advanced wind turbine controls to maximize energy extraction and reduce structural dynamic loads.
  92. [92]
    Floating Offshore Wind Farm Control via Turbine Repositioning
    Sep 19, 2024 · Situated at sea, offshore turbines can capitalize on stronger and steadier wind, resulting in higher energy production compared to their onshore ...
  93. [93]
    Challenges and opportunities for high-quality battery production at ...
    Jan 12, 2025 · The rise in battery production faces challenges from manufacturing complexity and sensitivity, causing safety and reliability issues.
  94. [94]
    U-Michigan 'Battery Lab 2.0' expansion open for innovation
    Aug 27, 2025 · And its three-megawatt-hour lithium-ion battery production line is believed to be the highest-capacity pilot-scale line based at a university.Missing: process | Show results with:process
  95. [95]
    [PDF] Process Control and Energy Efficiency - OSTI.GOV
    Mar 23, 2020 · Abstract. We review the impact of control systems and strategies on the energy ef- ficiency of chemical processes.
  96. [96]
    [PDF] Modernization justification - Rockwell Automation
    Some companies use payback period as a financial metric for new investments, which in this case would simply calculate the number of years it would take to ...<|separator|>
  97. [97]
    [PDF] Consider the Role of Safety Layers in the Bhopal Disaster - AIChE
    The process vent header (PVH) also fed the scrubber. As the situation unfolded, numerous layers of protection failed. Take a look at the safety layers that ...
  98. [98]
    Process Safety Since Bhopal - C&EN - American Chemical Society
    Jan 24, 2005 · And Bhopal gave new impetus to a concept called "inherently safer design," in which chemical plants are designed in a manner to reduce risk ...<|separator|>
  99. [99]
  100. [100]
    Predictive Maintenance Machine Learning: A Practical Guide
    35–45% reduction in downtime; 70–75% elimination of unexpected breakdowns; 25–30% reduction in maintenance costs. These findings underscore the substantial ...
  101. [101]
    Many paths to industrial sustainability go through process control
    Apr 22, 2022 · "Approximately 40% of greenhouse gas emissions today are produced by industry, so we need to help them decarbonize to achieve net-zero emissions ...