Fact-checked by Grok 2 weeks ago

Process variable

A process variable (PV), also known as a process parameter or process value, is a measurable that characterizes the current state or behavior of a or industrial process, such as , , , liquid level, or composition. In process , the PV represents the controlled variable that is monitored by sensors and maintained at a desired setpoint through automated adjustments to ensure operational stability, safety, and product quality. Process variables are integral to loops, where the measured is compared to the setpoint, and any deviation triggers manipulation of an input variable—such as valve position or speed—to minimize the and regulate the process. Common examples include in piping systems, in heat exchangers, in reactors, and liquid levels in tanks, all of which are critical in industries like chemical manufacturing, oil refining, and pharmaceuticals. Effective management of these variables optimizes efficiency, prevents equipment damage, and complies with regulatory standards by enabling real-time monitoring and automated corrections.

Fundamentals

Definition

A process variable (PV), also known as a process value or process parameter, is the current measured or observed value of a specific in a physical, chemical, or that is subject to or . In , the PV represents a key output of the system, such as level, , or flowrate, which must be maintained within a desired range to ensure operational stability and product quality. The is distinct from related terms in : the set-point () is the desired or target value for the PV that the system aims to achieve, while the manipulated variable () is the adjustable input, such as a valve position or pump speed, used to influence the PV toward the SP. Basic examples of process variables include the temperature within a , where deviations could affect reaction rates, or the flow rate through a , which impacts material transport efficiency. Process variables are standardized using the () for consistency in measurement and communication across disciplines. Common examples include (K) for , Pascal (Pa) for , cubic meters per second (m³/s) for , and meters (m) for liquid level.

Key Concepts

Process variables are inherently dynamic, fluctuating over time in response to disturbances such as changes in feedstock quality, equipment malfunctions, or external environmental factors. In steady-state conditions, these variables maintain constant values, reflecting a balanced operation where inputs equal outputs and no net accumulation occurs. However, transient behavior dominates during process transitions, like startups, shutdowns, or grade changes, where variables deviate from set points until a new equilibrium is achieved, often requiring active control to minimize disruptions. For a quantity to qualify as a process variable, it must meet key criteria of measurability and relevance: it should be quantifiable through reliable and exert a significant influence on overall process performance or product quality. This ensures that the variable can be monitored and adjusted to maintain desired outcomes, distinguishing it from unobservable or insignificant factors in the . Controlled variables, in particular, are selected based on their direct impact on quality metrics, while disturbances—uncontrolled inputs—affect them and must also be measurable for effective management. Process variables exhibit interdependence within complex systems, where alterations in one can propagate effects to others, necessitating integrated modeling for accurate prediction and . For instance, in handling processes, variations in may alter flow rates, creating coupled dynamics that amplify or dampen responses across the network. This interconnectedness underscores the need for system-wide analysis rather than isolated variable treatment. The foundational ideas behind process variables trace back to early 20th-century advancements in industrial automation, particularly through pioneers like Elmer Sperry, who in the 1910s developed servo mechanisms and integrated control systems for naval applications, such as gyroscopic stabilizers and fire control directors that automated variable monitoring and adjustment. These innovations marked the shift from manual to feedback-based regulation, influencing the evolution of process control theory.

Types and Classification

Common Process Variables

In industrial and engineering contexts, the most frequently encountered process variables are temperature, pressure, flow rate, and level, often referred to as the "big four" parameters essential for monitoring and controlling chemical and physical processes. These variables dominate process measurements, accounting for the majority of instrumentation applications in sectors like , where they enable precise regulation of operations such as reactions, , and material transport. According to standards from the (ISA), these four variables represent the core of process control. Temperature is a fundamental process variable representing the average of molecules in a substance, arising from the degree of motion among particles. In processes like within chemical reactors, it dictates rates and changes, with typical ranges spanning from -50°C to 500°C, though extremes can reach -200°C in cryogenic applications or over 1000°C in high-temperature furnaces. For instance, maintaining reactor temperatures around 200–300°C ensures optimal catalytic activity in synthesis. Pressure measures the force per unit area exerted by a on a containing surface, fundamentally linked to the collisions of molecules against walls. It is critical in applications such as for transporting gases or liquids, where it influences flow dynamics and prevents or bursts; common ranges in chemical processes vary from levels below 0.1 to high pressures up to 100 in polymerization reactors. A representative example is monitoring pipeline pressures at 5–20 to sustain steady flow. Flow rate quantifies the volume or of passing through a cross-section per unit time, governed by principles of such as and Bernoulli's equation for velocity-pressure relationships. In systems, it ensures balanced material distribution, with typical ranges from 0.1 L/min in laboratory-scale mixing to 1000 m³/h in large-scale columns. For example, controlling flow rates at 50–200 L/min in reactor feeds maintains stoichiometric ratios during . Level indicates the height of a liquid or solid relative to a reference point in a , based on gravitational and displacement principles. It is vital for managing tank volumes to avoid overflows or runs, with ranges from 0 to 10 in storage vessels, adjustable via vessel geometry. In water treatment plants, level at 2–5 in sedimentation tanks prevents disruptions. Beyond industrial settings, variables like serve as common process parameters in environmental systems, where it represents the content in air—typically measured as relative humidity from 20% to 80%—to regulate comfort in HVAC applications such as data centers or greenhouses.

Specialized and Derived Variables

In chemical processes, emerges as a specialized process due to its profound impact on mechanisms, where even minor deviations can alter dissociation equilibria and catalyst activity, thereby dictating product purity and yield. Similarly, functions as a niche in fluid handling operations, quantifying a fluid's resistance to and influencing sizing, rates, and energy consumption in sectors like and . In batch reactions, concentration of reactants or products stands out as a critical specialized , as it directly modulates according to rate laws, enabling precise endpoint determination and resource optimization in pharmaceutical synthesis. Derived process variables extend basic measurements by combining them into composite indicators that capture system performance more holistically. For instance, in gas or streams is commonly derived from and data via equations of state, providing essential insights for mass flow calculations and without direct sensing. In energy systems, efficiency ratios—such as the energy efficiency ratio defined as divided by electrical input— are computed from power consumption and output metrics, serving as derived variables to benchmark operational viability and guide retrofits in HVAC or power plants. Domain-specific applications highlight the tailored nature of these variables. In biotechnology, biomass levels represent a specialized variable pivotal for regulating nutrient feeds and harvest timing in microbial fermentations, where maintaining optimal cell densities ensures high-titer production of biologics like vaccines. Conversely, in aerospace, vibration amplitudes act as a key variable for real-time health monitoring of airframes and engines, as excessive oscillations can accelerate material fatigue and compromise flight safety under dynamic loads. These contexts demand specialized handling because standard control algorithms often fail to account for their context-dependent behaviors, such as growth phases in bioprocesses or modal resonances in structures. Challenges in managing specialized and derived variables include inherent non-linearities that amplify small input changes into large output swings—for example, pH's logarithmic response or viscosity's variation with in non-Newtonian fluids—necessitating advanced modeling for stable operation. Furthermore, many such variables rely on indirect inference from proxies like optical signals for or accelerometers for vibrations, which can propagate estimation errors and reduce reliability in closed-loop control.

Measurement Techniques

Sensors and Transducers

Sensors and transducers are essential hardware components in process control systems, designed to detect physical process variables and convert them into measurable electrical signals for monitoring and analysis. These devices form the foundational layer of measurement in industrial applications, enabling precise detection of parameters such as , , level, and . The historical development of sensors traces back to the 19th century with mechanical gauges, which relied on physical deformation to indicate measurements. A pivotal advancement was the invention of the Bourdon tube pressure gauge in 1849 by French engineer Eugène Bourdon, a curved, flattened tube that uncoils under internal pressure to drive a pointer on a dial, providing a direct visual readout without electrical components. This mechanical innovation marked a shift from rudimentary manometers and barometers to more reliable instruments for industrial pressure monitoring. The transition to sensors accelerated post-1950s, driven by advancements in technology and the need for higher precision in complex processes; by the late 1950s, sensors began surpassing mechanical ones in accuracy and integration potential, though initially more expensive to produce and maintain. Key milestones included the commercialization of silicon strain gauges in 1959, enabling electrical transduction of mechanical strain for pressure and force measurements. Various types of sensors are employed to measure specific process variables, each tailored to the physical phenomenon involved. For temperature, thermocouples are widely used; they consist of two dissimilar metal wires joined at a sensing junction, where a temperature difference generates a voltage proportional to the Seebeck effect, allowing measurement over a broad range from -200°C to over 2300°C depending on the type. In pressure measurement, strain gauges detect deformation in a diaphragm or bourdon element under applied force; the gauge, typically a foil pattern bonded to a substrate, experiences a change in electrical resistance as it stretches or compresses, which is then amplified to indicate pressure levels. Ultrasonic sensors serve for non-contact level detection in liquids or solids; a transducer emits high-frequency sound waves that reflect off the surface, and the time-of-flight—calculated as the round-trip duration divided by the speed of sound in the medium—determines the distance to the surface, yielding level height. For flow, turbine meters feature a multi-bladed rotor inserted in the fluid stream; as liquid or gas passes through, it spins the rotor at a speed directly proportional to the volumetric flow rate, with embedded magnets generating pulses via a pickup coil for electronic readout. Transducers operate on principles that convert physical inputs into electrical outputs, facilitating with systems. A common mechanism is the piezoelectric effect, where certain s like deform under mechanical stress (e.g., from or ), producing a charge separation that generates a measurable voltage; textually, this can be visualized as a sandwiched between electrodes, with applied compressing the material to displace internal dipoles, yielding an output signal proportional to the stress magnitude. In resistive transducers like strain gauges, elongation alters the conductor's length and cross-section, changing resistance according to R = \rho \frac{L}{A}, where \rho is resistivity, L length, and A area, converting to a voltage in a circuit. For thermocouples, the thermoelectric principle produces an at the junction due to differing electron diffusion rates in the metals at varying temperatures, described as a closed loop where the net voltage opposes the temperature gradient. Selecting appropriate sensors and transducers involves evaluating several critical factors to ensure reliability in operational environments. Accuracy specifies the closeness of measured values to the , often expressed as a of (e.g., ±0.5% for industrial thermocouples), essential for requiring precise control. defines the minimum and maximum measurable values, such as 0-1000 for transducers, ensuring coverage of expected variations without . Response time indicates how quickly the sensor reacts to changes, typically in milliseconds for dynamic like metering, where delays could affect or . Environmental robustness assesses durability against conditions like temperature extremes, humidity, or corrosives, often quantified by Ingress Protection (IP) ratings—e.g., IP67 for dust-tight and temporary immersion resistance in harsh industrial settings. These criteria guide choices, balancing performance with cost for applications from chemical plants to .

Signal Processing and Calibration

Signal conditioning refines raw analog signals from sensors into a form suitable for further processing and transmission in industrial systems. This process typically involves to boost weak signals to levels compatible with analog-to-digital converters (ADCs), ensuring they fall within the of downstream electronics. For instance, instrumentation amplifiers are commonly used for their high and ability to handle differential signals from bridges like Wheatstone configurations in or measurements. Filtering is another core technique, aimed at removing noise and unwanted frequency components to enhance . Low-pass filters, for example, attenuate high-frequency noise while preserving the signal of interest, often implemented using operational amplifiers in active configurations or resistors and capacitors in passive ones. Analog-to-digital conversion follows, where the conditioned is sampled and quantized into values, enabling computational analysis and storage. Calibration ensures that processed signals accurately represent true process variables by aligning outputs with known standards. The procedure begins with ing, where the is adjusted to read under no-input conditions to eliminate errors, followed by spanning to the full against reference values. to standards, such as those maintained by NIST, is achieved through a of certified references, ensuring measurements align with units via documented chains of comparison. sources like —differences in output when input is approached from increasing versus decreasing directions—arise from mechanical friction in components and must be quantified during to assess reliability. Data handling involves selecting appropriate sampling rates to capture signal dynamics without distortion, guided by the Nyquist theorem, which requires sampling at least twice the highest frequency component to avoid . In practice, rates of five times the signal frequency are often used in industrial for robustness. Transmission of these signals commonly employs the 4-20 mA current loop protocol, where 4 mA represents the minimum (live zero) and 20 mA the maximum process variable value, providing noise immunity over long distances in process control networks. Quality assurance in signal processing relies on metrics such as accuracy (closeness to true value), precision (consistency of measurements), and (variation under identical conditions). Calibration curves, typically plotted as output versus input in slope-intercept form (Output = m·Input + b), reveal deviations like zero shifts (vertical offsets) or span errors (slope mismatches), with appearing as separated ascending and descending paths. These assessments ensure processed process variables meet required tolerances, often verified through as-found and as-left tests against standards.

Role in Control Systems

Feedback and Deviation

In closed-loop control systems, the process variable (PV) is continuously measured and compared to a desired setpoint (SP) to compute the error signal, defined as e(t) = SP - PV(t). This error drives corrective actions through a loop, where the controller adjusts the manipulated variable to minimize deviations and maintain system stability. Such loops are essential in process control to counteract external influences and ensure consistent performance. Deviations in the PV arise primarily from disturbances, such as changes in feed composition, ambient conditions, or equipment wear, which perturb the system away from the SP. These deviations are minimized using proportional-integral-derivative () control, a widely adopted method that combines three actions to address different aspects of error dynamics. The PID control law is given by: u(t) = K_p e(t) + K_i \int_0^t e(\tau) \, d\tau + K_d \frac{de(t)}{dt} where u(t) is the controller output, K_p is the proportional , K_i is the , and K_d is the . The derivation of this equation stems from early , particularly Nicolas Minorsky's analysis of ship steering, where he modeled as a second-order incorporating inertia, friction, and external disturbances: A \ddot{\alpha} + B \dot{\alpha} + k \rho = D, with \alpha as heading deviation, \rho as angle, and D as disturbance torque. Minorsky proposed a starting with a proportional term (\rho = p \alpha) to directly counter the deviation, ensuring basic responsiveness proportional to magnitude. To eliminate persistent steady-state offsets from constant disturbances, he added an term (n \int \alpha \, dt), which accumulates past errors to drive the PV back to the SP over time. For improved transient response and damping of oscillations, a term (m \dot{\alpha}) was incorporated to anticipate error changes by responding to the rate of deviation. Stability analysis via the Hurwitz criterion confirmed that appropriate gains (k_p, k_i, k_d > 0) maintain bounded responses. The proportional action provides immediate correction scaled by K_p, reducing large errors quickly but often leaving a residual offset; the action sums errors to nullify this offset, though excessive can cause overshoot; the action predicts future errors to dampen rapid changes, enhancing smoothness without standalone use. System stability in feedback loops is assessed through concepts like gain margin, which quantifies the factor by which the can increase before the closed-loop system becomes unstable, providing a buffer against parameter variations or unmodeled dynamics. A larger gain margin indicates greater robustness to disturbances, ensuring the remains near the without oscillations or . In a example, the is the measured , which may deviate below the due to a cold draft disturbance; the feedback loop computes the error and triggers the controller to adjust the fuel valve opening, proportionally increasing heat input while integrating past shortfalls to fully restore the and differentiating to prevent overshoot.

Integration with Controllers

Process variables (PVs) integrate with proportional-integral-derivative (PID) controllers by providing real-time feedback for tuning parameters that minimize error in linear systems. The Ziegler-Nichols method, a seminal closed-loop tuning approach, involves increasing the proportional gain until sustained oscillations occur, then applying rules to set the and terms based on the ultimate gain and period. This ensures stability and performance for PVs like or in single-loop control. In (MPC), PVs serve as inputs to dynamic models that forecast future behavior over a horizon, enabling optimization of manipulated variables while respecting constraints. Originating in the process industry during the late , MPC uses PV to compute actions that minimize a quadratic cost function, particularly effective for systems with delays and interactions. Fuzzy logic controllers handle nonlinear PVs by mapping imprecise inputs, such as error and its rate of change, to outputs via linguistic rules and membership functions, avoiding explicit mathematical models. This approach excels in uncertain environments where PVs exhibit nonlinearity, like chemical reactions, by defuzzifying aggregated rules to produce smooth actions. For multivariable systems, multiple PVs are managed through techniques that eliminate interactions between loops, often via state feedback or inverse models to achieve independent . Hierarchical control structures layer supervisory optimization over local loops, prioritizing PVs based on process priorities to maintain overall . The integration of PVs shifted from analog pneumatic systems, prevalent before the , to digital programmable logic controllers (PLCs) introduced in the late 1960s, which replaced relay-based wiring with for flexible PV monitoring and actuation. By the , PLCs enabled scalable handling of PV signals through standardized I/O modules, facilitating the transition to computerized control. In supervisory control and data acquisition () systems, PV data is aggregated via protocols like OPC for real-time visualization and alarming, integrating with PLCs to oversee distributed processes. PV trends inform by adjusting parameters online to track changes in , with enhancements post-2010 incorporating to optimize policies from historical PV data. These methods enable self-tuning in varying conditions, improving robustness over fixed-gain approaches.

Applications

Industrial Processes

In chemical plants, process variables play a pivotal role in optimizing separation and reaction processes. In distillation columns, serves as a key process variable, particularly the reflux temperature, which is manipulated to maintain stable composition profiles and product purity by adjusting reflux flow rates. For example, advanced regulatory strategies calculate net internal reflux based on heat balances involving overhead vapor and reflux temperatures to counteract disturbances like feed changes. Similarly, monitoring in chemical reactors is essential for controlling reaction rates, gas solubility, and safety, as deviations can lead to hazardous conditions such as explosions or incomplete reactions; this is achieved through pressure regulators and relief valves integrated into control loops. In the oil and gas sector, and are fundamental process variables in pipelines, where they ensure efficient transport while mitigating risks like surges or blockages; flow rates are measured using differential devices such as orifice plates, and is monitored via transmitters to maintain operational integrity across varying viscosities and pipe diameters. A notable from a hydrocracker unit in a illustrates upset recovery: during high-throughput operations, elevated diesel outlet temperatures (exceeding 50°C) and fluctuating rates (up to 117.8 m³/hr) triggered an upset, but installing an active redundant allowed real-time monitoring and adjustment of these variables, reducing temperatures to 48°C and stabilizing to 100 m³/hr, thereby preventing shutdowns and enhancing . Manufacturing processes rely on process variables like level and speed to achieve precision and throughput. In filling lines, level control monitors and adjusts product volumes in containers using checkweighers or vision systems to detect over- or under-filling in , enabling automatic corrections without halting and ensuring compliance with standards. For processes, speed control—often as in serial operations like or part placement—modulates machine trajectories to match demands, reducing defects from excessive or insufficient rates while integrating with systems for consistent output. Standards and regulations standardize process variable handling in industrial settings. The ISA-5.1 standard establishes uniform instrumentation symbols and identification codes for process variables, facilitating clear depiction of measurement and control systems across diagrams in sectors like chemical and . Additionally, OSHA's (PSM) standard mandates monitoring of process variables such as pressures, flows, and temperatures through hazard analyses and operating procedures to prevent catastrophic releases of hazardous chemicals.

Modern and Emerging Contexts

In the context of Industry 4.0, process variables (PVs) such as temperature, vibration, pressure, and flow rates are streamed in real-time through (IIoT) networks, enabling for . Edge devices process these PVs locally to reduce and demands, allowing models to forecast equipment failures before they occur. For instance, in waste-to-energy plants, IIoT platforms monitor syngas heating value and temperature as key PVs, achieving prediction accuracies with R² values exceeding 0.98 through neural network-based models integrated with 5G-enabled gateways. This approach minimizes downtime by detecting anomalies in operational data streams, contrasting with traditional in earlier industrial paradigms. In , PVs like density and dissolved oxygen levels are pivotal in operations for optimizing microbial and mammalian cultures. density, often reaching densities above 10^10 s per liter in systems, directly influences nutrient uptake and product yields, monitored via or soft sensors to maintain viability during fed-batch processes. Dissolved oxygen, controlled through and via controllers or model predictive algorithms, ensures metabolic efficiency, with oxygen transfer rates (OTR) measured dynamically to prevent hypoxic conditions that limit growth. These PVs are integrated into supervisory control and (SCADA) systems for real-time adjustments in production, enhancing scalability for therapeutic protein synthesis. Renewable energy systems leverage PVs to enhance performance and reliability in dynamic environments. For wind turbines, vibration and acceleration serve as critical PVs, captured by accelerometers and strain gauges in setups, enabling early detection of in floating offshore installations through operational . In solar photovoltaic arrays, efficiency-tracking PVs such as , panel temperature, and shading factors are monitored to optimize power output, often via interfaces that adjust for environmental variations. These measurements support predictive strategies that extend asset life, with data from supervisory control and data acquisition () systems providing 10-minute resolution insights for maintenance scheduling. Emerging trends since 2020 highlight the integration of (AI) with digital twins for advanced , addressing limitations in traditional control by simulating system behaviors. AI-enhanced digital twins employ techniques, such as (LSTM) networks with , to identify deviations in PVs like or , facilitating proactive interventions in and sectors. models within these twins enable through and fault diagnosis via , reducing false positives in anomaly alerts. For example, in additive , AI-driven twins achieve by analyzing PV streams, aligning with ISO 30186 standards for adaptive systems. This evolution supports interdisciplinary applications, from urban grids to biotech simulations, by bridging physical and virtual domains for enhanced .

References

  1. [1]
    What is a Process and Process Variable - Instrumentation Academy
    A process variable is a measurable quantity or parameter that characterizes the state or behaviour of a physical system or process.
  2. [2]
    5.4: Process Control - Engineering LibreTexts
    May 22, 2024 · Process Variable: the variable in the system or process that we desire to control. Controlled Variable: the output process variable we compare ...
  3. [3]
    AutoQuiz: What Is an Industrial Process Variable? - ISA Interchange
    A process variable, which is to be maintained at some desired value (temperature, pressure, level, flow) by means of manipulating another process variable, is ...<|control11|><|separator|>
  4. [4]
    Process controls - processdesign
    Feb 21, 2016 · Process control is an important part in maintaining the output of a system within a desired range by manipulating various inputs.
  5. [5]
    SI Units
    ### SI Units for Process Variables in Engineering
  6. [6]
    [PDF] Chapter 1
    Process Dynamics a) Refers to unsteady-state or transient behavior. b) Steady-state vs. unsteady-state behavior i. Steady state: variables do not change with ...
  7. [7]
    [PDF] Introduction to Process Control Topic I - Madar
    Remarks: All important variables to be controlled (CV) must be identified and measurable. (CV's are usually direct or indirect quality variables).
  8. [8]
    Process Control – Understanding the Basics - Newell Automation
    In manufacturing, a wide number of variables from temperature to flow to pressure can be measured simultaneously. All of these can be interdependent variables ...
  9. [9]
    [PDF] sperry.pdf - MIT
    Elmer Ambrose Sperry (I 860-1930) and the company he founded, the Sperry Gyroscope Company, led the engineering of control systems between 1910 and 1940. ...<|control11|><|separator|>
  10. [10]
    Temperature Measurement and Control Fundamentals
    Aug 31, 2023 · The “big four” process control parameters are temperature, pressure, flow, and level. Other parameters, such as pH, conductivity, and ...Missing: prevalence statistics
  11. [11]
    The Four Key Parameters of Process Control: Flow, Level, Pressure ...
    Among the many factors that can be monitored and adjusted, four parameters stand out as fundamental: flow, level, pressure, and temperature. These four are ...
  12. [12]
    Temperature: The Most Misunderstood Process Variable
    Jan 1, 2023 · (Page 1) Measuring temperature is required in nearly all chemical processes. Here are suggestions to avoid common pitfalls.Missing: basis | Show results with:basis
  13. [13]
    [PDF] The Engineer's Guide to Industrial Temperature Measurement
    Figure 4-39 – Typical Process Flow Diagram. 4.4.3.1.3 – Piping and Instrument ... including pressure, temperature, flow rate , fluid viscosity and ...
  14. [14]
    Understanding the Four Major Process Variables in Process ...
    Pressure: Perhaps the most ubiquitous of all process variables, pressure is the force exerted perpendicular to the surface of an object per unit area. In ...
  15. [15]
    Pressure Ranges and Units for Fluid System Monitoring
    Jun 27, 2023 · Low Pressure: Low-pressure systems typically operate at pressures below 100 psi (pounds per square inch) or 7 bar. · Medium Pressure: Medium- ...
  16. [16]
    Mass Flow vs Volumetric Flow - Alicat Scientific
    Volumetric flow measures the physical space a fluid occupies as it moves through a system, while mass flow quantifies the number of molecules passing through ...
  17. [17]
    [PDF] The Engineer's Guide to DP Flow Measurement | Emerson
    Chapter 3 – Theory of DP Flow focuses on derivations and basis for the conservation of energy and mass behind the principals of DP flow.
  18. [18]
    Main Process Variable - an overview | ScienceDirect Topics
    Main process variables are defined as the key parameters in upstream production facilities, including pressure, liquid level, temperature, and flow, ...
  19. [19]
  20. [20]
    Humidity Control - an overview | ScienceDirect Topics
    Humidity control is very important to guarantee the thermal comfort and health of occupants in buildings, which is particularly true in hot and humid climates.
  21. [21]
    What are the Essentials of pH in Industrial Applications?
    The actual solution pH changes despite a constant hydrogen ion concentration because of changes in dissociation constants with process temperature, and activity ...
  22. [22]
    Viscosity Management | PCI Magazine
    Nov 1, 2006 · Because viscosity can be directly affected by temperature, shear rate and other variables that can be very different off-line from what they are ...
  23. [23]
    Modeling of pH for Control | Industrial & Engineering Chemistry ...
    Indirect adaptive backstepping control of a pH neutralization process based on recursive prediction error method for combined state and parameter estimation.
  24. [24]
    Liquids - Densities vs. Pressure and Temperature Change
    Liquid density changes with temperature and pressure. Temperature increases cause most liquids to expand, and pressure increases decrease volume.
  25. [25]
    Energy Efficiency Ratio - an overview | ScienceDirect Topics
    Energy Efficiency Ratio (EER) is the ratio of cooling capacity to electrical energy consumed, calculated as EER = Cooling capacity / Electrical energy input.
  26. [26]
    Biomass Monitoring in Microbial Fermentation
    Biomass monitoring tracks microbial cell concentration over time, measuring it to understand, optimize, and control the bioprocess and product production.
  27. [27]
    Vibration Analysis: Fundamentals, Types, FEA - SimScale
    Aug 28, 2024 · Displacement, velocity, and acceleration are the three primary vibration measurement variables. These are measured in terms of their magnitudes ...
  28. [28]
    Estimation of biomass concentrations in fermentation processes for ...
    As the biomass concentration cannot be measured online during the production to sufficient accuracy, indirect measurement techniques are required.
  29. [29]
    Sensors and Transducers in Instrumentation & Control
    Mar 12, 2021 · The different types of sensors & transducers; their operating principles as employed in measurement and process control.
  30. [30]
    History and Innovation | Bourdon Instruments
    In 1849, the French engineer Eugène Bourdon patented the pressure-measuring device commonly known today as the Bourdon tube.
  31. [31]
    Electronic Sensing - an overview | ScienceDirect Topics
    Initially—that is, in the 1950s—the cost of electronic sensors greatly exceeded that of the traditional mechanical measuring tools, and their servicing ...Missing: post- | Show results with:post-
  32. [32]
    [PDF] HISTORY OF MICROELECTOMECHANICAL SYSTEMS (MEMS)
    Strain gauges began to be developed commercially in 1958. Kulite was founded in 1959 as the first commercial source of silicon strain gages .
  33. [33]
  34. [34]
  35. [35]
    Functionality and technology of ultrasonic sensors - Baumer
    Most ultrasonic sensors are based on the principle of measuring the propagation time of sound between send and receive (proximity sensor). The barrier principle ...
  36. [36]
    Turbine Flow Meter Explained | Operation & Calibration - RealPars
    Mar 1, 2021 · A Turbine Flow Meter is inserted in a pipe directly in the flow path and has a turbine rotor placed in the path of a flowing stream.
  37. [37]
    Piezoelectric Transducer: Applications & Working Principle
    May 7, 2024 · Working Principle: The piezoelectric effect allows these transducers to generate voltage when mechanical stress is applied, which is then used ...
  38. [38]
    What is a Strain Gauge and How Does it Work?
    Aug 13, 2020 · Strain gauges convert the applied force, pressure, torque, ect., into an electrical signal which can be measured.
  39. [39]
    How Do Thermocouples Work? A Quick Tutorial - WIKA blog
    Thermocouples are electrical devices used to measure temperature. · The working principle of a thermocouple follows the Seebeck effect, or thermoelectric effect, ...
  40. [40]
  41. [41]
    How to Choose the Right Sensor for My Industrial Application
    May 27, 2025 · Range and Span: Ensure the sensor's measurement range covers the minimum and maximum expected values. Response Time: Faster response is crucial ...
  42. [42]
  43. [43]
    [PDF] "Signal Conditioning and Preprocessing". In
    This operation is performed by means of interface circuits. Second, the electrical signal undergoes analog conditioning (e.g., filtering) to enhance its ...Missing: industrial | Show results with:industrial
  44. [44]
    [PDF] INSTRUMENTATION CALIBRATION Dick A. Mack ... - OSTI.GOV
    Nov 9, 1976 · Calibration is usually carried out by comparing the value of a known standard with the readings of the instrument being calibrated in a series.
  45. [45]
    [PDF] Calibration Procedures for Weights and Measures Laboratories
    This publication contains standard operating procedures for calibrations that were not previously published in other NIST Office of Weights and Measures ...
  46. [46]
    Acquiring an Analog Signal: Bandwidth, Nyquist Sampling Theorem, and Aliasing
    ### Nyquist Theorem in Context of Data Acquisition for Industrial Signals
  47. [47]
    What Is a 4-20 mA Current Loop? | Fluke
    ### Summary of 4-20 mA Protocol in Process Control
  48. [48]
    Calibration Errors and Testing | Basic Principles of Instrument ...
    Hysteresis errors are almost always caused by mechanical friction on some moving element (and/or a loose coupling between mechanical elements) such as bourdon ...
  49. [49]
    [PDF] PID Theory Explained - Experimentation Lab
    Mar 29, 2011 · This is called a closed loop control system ... Steady-State error is the final difference between the process variable and set point.
  50. [50]
    [PDF] introduction to - process control
    1.4 CONTROL SYSTEM EVALUATION. A process-control system is used to regulate the value of some process variable.Missing: engineering | Show results with:engineering
  51. [51]
    Process Control Introduction — Dynamics and Control - APMonitor
    Aug 17, 2021 · A feedback control system consists of a sensor, actuator and controller that are connected with information flowing in a loop.
  52. [52]
  53. [53]
    [PDF] Nicolas Minorsky and the Automatic Steering of Ships - Robotics
    his main interests have been computer control and the history of control engineering. He is the author of a book on the history of control engineering and ...
  54. [54]
    Introduction: Frequency Domain Methods for Controller Design
    The gain margin is defined as the change in open-loop gain required to make the closed-loop system unstable. Systems with greater gain margins can withstand ...
  55. [55]
    [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 ...
  56. [56]
    [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.Missing: seminal | Show results with:seminal
  57. [57]
    Fuzzy control theory: A nonlinear case - ScienceDirect.com
    We analyze the performance of a simple fuzzy controller with linear and nonlinear defuzzification algorithms.
  58. [58]
    [PDF] DECOUPLING IN THE DESIGN AND SYNTHESIS OF ...
    Necessary and sufficient conditions for the "decoupling" of an m-input, m-output time-invariant linear system using state variable feedback are determined.
  59. [59]
    [PDF] History of Control History of PLC and DCS
    Jun 15, 2012 · The early history of the PLC goes back to the 1960's when control systems were still handled using relay control. During this time the control ...
  60. [60]
    Sensors Data Analysis in Supervisory Control and Data Acquisition ...
    Apr 14, 2021 · This paper presents the design and implementation of a supervisory control and data acquisition (SCADA) system for automatic fault detection.
  61. [61]
    [PDF] A Historical Perspective of Adaptive Control and Learning - arXiv
    Feb 22, 2022 · Adaptive control is real-time control of uncertain systems through adaptation and learning, with its history spanning the field of control ...
  62. [62]
    Use Model Predictive Control to Improve Distillation Process
    Typical process variables (PVs) are temperatures, pressures, and compositions at different points in the column. The main controlled variables (CVs) are feed ...
  63. [63]
    The Dos and Don'ts of Distillation Column Control - ScienceDirect
    In two-product columns it is common to do this by manipulate e.g. reflux to maintain a certain temperature somewhere between the feed stage and product stage. ...
  64. [64]
    Process Pressure - an overview | ScienceDirect Topics
    Pressure is another key process variable since its level is critical for boiling, chemical reaction, distillation, extrusion, vacuuming, and air conditioning.
  65. [65]
    Process Control Automation in Oil & Gas (Downstream) - EPCM
    In oil & gas downstream, the critical process variables which are very important to control include pressure, level, flow, temperature, and density.
  66. [66]
    Process Safety and Performance Improvement in Oil Refineries ...
    May 11, 2020 · The case study presents the excerpts of process safety and performance improvement of a HCU cooling system by installing an additional ...
  67. [67]
    Fill Level Inspection and Control | METTLER TOLEDO
    Fill level inspection with checkweighing, x-ray inspection, or vision inspection helps manufacturers to control against over-filling and underfilling ...
  68. [68]
    [PDF] Manufacturing Processes and Process Control - MIT
    To help define internal variables in the process as well as the inputs and outputs, the basic output causality of the process model is shown in Fig. 3 using a ...
  69. [69]
    ISA5.1, Instrumentation Symbols and Identification
    The purpose of this standard is to establish a uniform means of designating instruments and instrumentation systems used for measurement and control.ANSI/ISA-5.1-2024 · ISA-TR5.1.03-2024 · ISA-TR5.1.02-2024
  70. [70]
  71. [71]
  72. [72]
  73. [73]
  74. [74]
  75. [75]