Fact-checked by Grok 2 weeks ago

Process simulation

Process simulation is a computational technique in chemical engineering that involves creating mathematical models to represent and predict the behavior of physical, chemical, and biological processes, typically by decomposing complex systems into interconnected unit operations such as reactors, separators, and heat exchangers. These models solve material and energy balances, along with thermodynamic and kinetic equations, to forecast key variables like flow rates, compositions, temperatures, pressures, and equipment sizing under various operating conditions. By simulating real-world scenarios virtually, process simulation reduces the need for costly physical experiments and enables engineers to test "what-if" analyses for process design, troubleshooting, and optimization. The roots of process simulation trace back to the early , when chemical engineers manually performed design calculations for unit operations as part of the emerging profession. Significant advancements occurred in the with the development of computer-based tools, including the SPEED-UP simulator in 1964, which introduced systematic flowsheet computation, and the 1972 GEMCS system, an early sequential modular approach. The 1980s marked a commercial breakthrough with Aspen Plus in 1982, a widely adopted steady-state simulator that integrated rigorous thermodynamic models for applications. Subsequent innovations, such as equation-oriented simulators like gPROMS in 1994, expanded capabilities to handle dynamic and nonlinear optimizations. In modern practice, process simulation plays a central role in process , supporting industries including , pharmaceuticals, and production. It facilitates steady-state analyses for initial design and economic evaluations, as well as dynamic simulations to study transient behaviors like startups, shutdowns, and responses. Key applications encompass optimizing , ensuring through hazard identification, and integrating metrics like life-cycle assessments. Popular software tools, such as and CHEMCAD, rely on extensive databases of physical properties and reaction kinetics to deliver accurate predictions, though results are highly sensitive to input . Overall, process simulation has evolved into an indispensable tool for , enabling faster development cycles and more robust industrial operations.

Fundamentals

Definition and Principles

Process simulation is a model-based computational technique used to replicate real-world chemical, physical, biological, or through software implementations, often visualized via diagrams or block representations that connect operations with material and streams. This approach enables engineers to predict process behavior without physical experimentation, focusing on variables such as rates, temperatures, pressures, and compositions by solving interconnected mathematical models. At its core, process simulation operates on fundamental principles derived from conservation laws, including the iterative solution of , , and balances to achieve or dynamic states within the modeled . These balances account for inputs, outputs, accumulations, generations, and consumptions across process elements, allowing for the of steady or transient conditions. Mathematical models representing thermodynamic properties, kinetics, and are integrated to simulate how changes in one part of the propagate throughout. For biological processes, models may incorporate population balance equations to describe and interactions. The typical begins with specifying system inputs, such as feed properties including , flowrate, , and , alongside definitions of unit operations like reactors, separators, heat exchangers, and pumps. Convergence algorithms, such as Newton-Raphson or sequential modular methods, are then employed to iteratively solve the until balances are satisfied, yielding outputs like product and equipment performance metrics. For steady-state simulations, a basic equation exemplifies this principle: \sum (\text{inflows} - \text{outflows} + \text{generation} - \text{consumption}) = 0 This equation enforces mass conservation under non-accumulating conditions, forming the basis for broader and analyses. To handle real-world complexities, process simulation incorporates approximations, particularly through and of physical and thermodynamic for unmeasured or extreme conditions, enabling reliable predictions within defined limits while acknowledging model assumptions. These techniques extend databanks of known to broader operating ranges, though accuracy diminishes beyond validated domains.

Importance and Applications

Process simulation plays a pivotal role in by enabling the , , , and of complex processes without the need for costly physical experiments, thereby significantly reducing trial-and-error during development. This approach allows engineers to virtually test scenarios, identify potential issues early, and refine systems iteratively in a controlled digital environment. Key benefits of process simulation include its ability to predict process behavior under diverse operating conditions, such as varying temperatures, pressures, or feed compositions, which supports informed decision-making and enhances overall system reliability. It also aids in achieving by facilitating environmental impact assessments and ensuring adherence to standards, while enabling seamless from prototypes to full operations. These advantages collectively contribute to substantial reductions in development time and operational expenses across industries. In , process simulation is essential for refinery optimization, where it models unit operations to maximize throughput and product quality while minimizing waste. The pharmaceutical sector employs it to simulate drug manufacturing processes, ensuring consistent quality, batch reproducibility, and compliance with stringent regulatory requirements. benefits from simulations of modeling, which optimize microbial growth, substrate utilization, and product yields to improve consistency and . In the energy sector, it enhances by analyzing and mass transfers to reduce fuel consumption and emissions. leverages process simulation for designs, predicting removal rates and treatment efficacy to mitigate ecological impacts. A representative case is the simulation of distillation columns in oil refining, where rigorous modeling has achieved energy minimization and yield maximization; for example, optimization studies have demonstrated reductions in fired heat demand by up to 20% through integrated tray-by-tray s and heat recovery adjustments. Emerging applications of process simulation are increasingly focused on , particularly in designing carbon capture processes to meet global emission reduction targets, where simulations optimize absorber configurations and solvent performance to lower energy penalties and achieve capture rates such as the U.S. target of 90%.

Modeling Approaches

Mathematical Foundations

Process simulation relies on mathematical modeling to represent physical, chemical, and biological processes through systems of equations derived primarily from conservation laws, including those for , , and . These laws form the basis for constructing models that describe how materials and energy flow within units such as reactors, heat exchangers, and pipelines, enabling predictions of system behavior under various conditions. The , often expressed as the , states that the rate of change of within a equals the net flow in minus the net flow out, plus any generation or consumption due to reactions: \frac{dM}{dt} = \sum \dot{m}_{in} - \sum \dot{m}_{out} + \dot{G} - \dot{C}, where M is the total , \dot{m} denotes flow rates, and \dot{G} and \dot{C} represent generation and consumption rates. For open systems, the is derived from of and typically takes the rate form \frac{dU}{dt} = \dot{Q} - \dot{W} + \sum \dot{m}_{in} h_{in} - \sum \dot{m}_{out} h_{out}, where U is the , \dot{Q} and \dot{W} are and work rates, and h is the specific of inlet and outlet streams (neglecting kinetic and potential changes for simplicity in many process applications). conservation, crucial for in process equipment like , is governed by the Navier-Stokes equations, which for in a can be simplified to \rho \left( \frac{\partial \mathbf{v}}{\partial t} + \mathbf{v} \cdot \nabla \mathbf{v} \right) = -\nabla p + \mu \nabla^2 \mathbf{v} + \mathbf{f}, where \rho is , \mathbf{v} is , p is , \mu is , and \mathbf{f} represents body forces. These equations capture the balance between inertial, , viscous, and external forces in fluid motion. Solving these systems, which are often nonlinear due to coupled phenomena like reaction kinetics and phase equilibria, requires numerical techniques. The Newton-Raphson method is widely used for iteratively solving sets of nonlinear algebraic equations by linearizing around an initial guess and updating via \mathbf{x}_{k+1} = \mathbf{x}_k - \mathbf{J}^{-1} \mathbf{F}(\mathbf{x}_k), where \mathbf{J} is the matrix and \mathbf{F} represents the equations from the balances. For linear systems arising in flows, such as mass or energy balances across interconnected units, matrix methods like or sparse solvers are employed to solve \mathbf{A} \mathbf{x} = \mathbf{b}, where \mathbf{A} encodes the topology and coefficients from laws. Process models vary in complexity, with lumped-parameter approaches assuming uniform conditions within a (e.g., average and ), leading to ordinary or algebraic equations suitable for large-scale simulations. In contrast, distributed-parameter models account for spatial variations, resulting in partial equations that better capture phenomena like concentration gradients in reactors or profiles in , though at higher computational cost. To handle uncertainties in such as reaction rates or transport coefficients, quantifies how variations in inputs propagate to outputs, often using local methods like partial derivatives \frac{\partial y}{\partial \theta} (where y is a model output and \theta a ) or techniques such as variance-based to apportion total output . This is essential for model validation and identifying critical influencing reliability.

Data and Property Modeling

In process simulation, accurate representation of thermophysical properties such as , , and is essential for reliable model predictions. These properties are typically sourced from established databases like the DIPPR 801, which provides critically evaluated data for over 2,000 pure components, including temperature-dependent correlations derived from experimental measurements. Experimental data from laboratory measurements complement these databases, particularly for proprietary or novel compounds where database coverage is limited. Modeling techniques for these properties include empirical correlations and group contribution methods. Empirical correlations, such as the Antoine equation for vapor pressure, express properties as functions of temperature using fitted parameters: \log P = A - \frac{B}{T + C} where P is vapor pressure, T is temperature, and A, B, C are substance-specific constants obtained from regression of experimental data. Group contribution methods, like UNIFAC for estimating activity coefficients in nonideal mixtures, decompose molecules into functional groups and sum their contributions based on interaction parameters, enabling predictions without direct experimental data for the mixture. Fitted models, which regress parameters directly against available data, offer high accuracy within narrow or ranges but require substance-specific measurements. In contrast, predictive models, such as group contribution approaches, estimate properties for untested compounds solely from molecular structure, providing broader applicability at the cost of slightly reduced precision for well-studied systems. Challenges arise in handling multicomponent mixtures and phase equilibria, where interactions between components complicate property estimation. Equations of state like the Peng-Robinson model address this by incorporating mixing rules for parameters a and b: P = \frac{RT}{V - b} - \frac{a \alpha}{V(V + b) + b(V - b)} yet they often require binary interaction parameters to improve accuracy for nonideal behaviors in complex mixtures, and convergence issues can occur in highly asymmetric systems. Validation of these models involves systematically comparing predictions against independent experimental data to quantify deviations, often using metrics like average absolute relative error to ensure reliability within specified conditions before integration into process simulations.

Types of Simulation

Steady-State Simulation

Steady-state simulation in process engineering involves modeling chemical or industrial processes under conditions where all variables remain constant over time, focusing on solving material, energy, and component balances at equilibrium without accumulation terms. This approach assumes the system has reached a stable operating point, where inputs equal outputs, and no transient effects occur. Such models are particularly suited for continuous processes operating at steady conditions, enabling engineers to predict performance without considering time-dependent changes. Key features of steady-state simulation include its computational efficiency, as it requires solving algebraic equations rather than ones, often yielding results orders of magnitude faster than dynamic methods. It is primarily used for initial , equipment sizing, and optimization of continuous operations, such as determining flow rates, temperatures, and pressures in a flowsheet to minimize energy use or maximize . For instance, in design, steady-state models facilitate heat and material balance calculations to ensure economic viability and compliance with specifications. These simulations have evolved since the , transforming from empirical methods to a data-driven practice, with significant impacts on energy optimization and . The simulation process typically employs one of two main strategies: the sequential modular approach, which solves unit operations and recycle streams one at a time in a predefined sequence, converging iteratively for loops; or the equation-oriented approach, which formulates and solves the entire system of nonlinear algebraic equations simultaneously using numerical methods like Newton-Raphson. The sequential modular method, widely adopted in early simulators, offers intuitive flowsheet construction but can struggle with complex recycles, while the equation-oriented method excels in optimization and sensitivity analysis by treating all variables holistically. A seminal review highlights how these approaches addressed challenges in accurate mathematical modeling and thermodynamic correlations by the mid-1970s. An illustrative example is the steady-state simulation of a network, where the model balances temperatures and flows across multiple units to achieve target heating or cooling without transient fluctuations. Using matrix-based methods grounded in , such simulations evaluate coefficients and account for property variations, enabling fouling analysis or performance predictions under stable conditions. This application demonstrates how steady-state tools support retrofit designs or operational assessments in and industries. Despite its advantages, steady-state simulation has limitations, as it cannot capture behaviors during startups, shutdowns, or disturbances where time-varying dynamics are critical, such as in design or safety analyses. These models provide a snapshot of equilibrium but overlook path-dependent effects, necessitating complementary dynamic simulations for comprehensive process evaluation.

Dynamic Simulation

Dynamic simulation in involves modeling the time-dependent behavior of chemical processes, incorporating transient phenomena through differential equations that account for accumulation over time. At its core, it extends the fundamental , , and balances by including time derivatives, such as the accumulation term in the equation \frac{dM}{dt} = \dot{m}_{\text{in}} - \dot{m}_{\text{out}} + r , where M represents the in the , \dot{m} denotes rates, and r is the rate of generation or consumption. This approach captures how processes evolve from initial conditions toward or in response to disturbances, contrasting with steady-state methods that assume constant conditions. Key features of dynamic simulation include the numerical solution of ordinary differential equations (ODEs) or partial differential equations (PDEs) derived from these balances, typically using integration methods like Runge-Kutta algorithms to advance the system state over discrete time steps. These integrators adapt step sizes based on process dynamics—ranging from milliseconds for fast transients like responses to minutes for slower batch operations—ensuring and accuracy. Compared to steady-state simulation, dynamic models impose higher computational demands due to the iterative time-stepping and inclusion of dynamic elements such as holdups, lags, and loops, but they provide insights into system and response trajectories. Dynamic simulation finds essential applications in analyzing transient operations, including process startups and shutdowns, where it models evolving flows, temperatures, and pressures to optimize sequences and minimize risks. It supports operator training simulators by replicating plant behavior in virtual environments, allowing practice of normal, upset, and emergency scenarios without operational hazards. Additionally, it enables real-time (MPC) by forecasting process responses to disturbances and optimizing control actions for improved yield and efficiency. A representative example is the modeling of pressure buildup in a research reactor during a sudden feed or malfunction, such as a spurious opening of a flapper leading to a loss-of-flow . Using thermal-hydraulic codes like RELAP5, which couple kinetics and via ODEs, simulations predict rapid power excursions and pressure rises—potentially reaching critical levels within seconds—while evaluating features like scram systems to ensure and prevent fuel damage. One key development in dynamic simulation is the use of hybrid steady-dynamic modes (as of the 2010s), which initialize with steady-state solutions for before switching to dynamic for transient analysis, reducing computational overhead in online optimization and applications. These hybrids facilitate seamless with systems, enhancing predictive capabilities for complex processes. As of 2025, emerging trends include with and for surrogate modeling to accelerate simulations, as well as digital twins for real-time in industries such as and pharmaceuticals.

Historical Development

Early History

The origins of process simulation trace back to the pre-computer era, where empirical equations formed the basis for predicting physical properties essential to chemical processes. In 1888, French chemist C. Antoine proposed a semi-empirical relating to for pure substances, providing a foundational tool for and evaporation calculations that remains in use today. This equation built on earlier models and enabled engineers to estimate phase behavior without direct experimentation. Complementing such developments, the 1923 publication Thermodynamics and the Free Energy of Chemical Substances by and Merle Randall established rigorous principles for phase equilibria, including activity coefficients and fugacities, which were critical for modeling multicomponent systems in separation processes. These works shifted process analysis from purely experimental trial-and-error toward quantitative predictions, laying the groundwork for systematic simulation. In the early , process simulation relied on manual calculations for unit operations, such as and balances, performed using slide rules, log tables, and nomograms. Warren K. Lewis, a professor at , played a pivotal role in advancing modeling through his theoretical analyses of columns and multicomponent separations, publishing key papers in 1909 and 1922 that quantified vapor-liquid equilibria and tray efficiencies. His co-authorship of Principles of in 1923 further formalized unit operations as modular components of processes, facilitating manual design of interconnected systems. Concurrently, process flow diagrams emerged in the 1920s, pioneered by industrial engineer Frank Gilbreth in 1921 as "flow process charts" to visualize material and energy flows, aiding in the planning of complex plants without computational aids. Following , analog computers in the 1940s and 1950s enabled rudimentary simulations of simple material and energy balances in , particularly for dynamic systems like reactor control and dynamics, by solving differential equations through electrical analogs. Companies such as and Exxon (formerly ) contributed to early industrial applications, adapting these tools for process optimization in refining and production. The transition to digital methods began in labs during the late 1950s, with the first batch-oriented simulations developed by oil companies using precursors to , such as early assembly languages on machines, to perform iterative calculations for steady-state flowsheets. This period saw the development of pioneering digital simulators, including the SPEED-UP system in 1964 by (ICI), which introduced systematic flowsheet computation, and the GEMCS system in 1972, an early sequential modular approach. 's release in 1956 marked a milestone, allowing more accessible programming of equilibrium-based models and paving the way for broader adoption in .

Modern Evolution

The advent of mainframe computers in the facilitated the development of the first commercial process simulation software, enabling more sophisticated modeling of chemical and beyond manual calculations. Pioneering tools like Aspen Plus, released in 1982 by , marked a significant milestone by providing steady-state simulation capabilities for and optimization in the . During the , the field advanced with additional developments, while packages such as PRO/II from Simulation Sciences, under development since the early 1980s and released in 1990, leveraged mainframe power to simulate complex unit operations and flowsheets with greater accuracy and efficiency. The integration of personal computers in the 1990s democratized access to process simulation, shifting from centralized mainframes to desktop environments and accelerating adoption in workflows. This era included the introduction of equation-oriented simulators like gPROMS in 1994, expanding capabilities for dynamic and nonlinear optimizations. Software like , introduced in 1996 by Hyprotech (later acquired by ), was specifically designed for Windows-based , allowing dynamic simulations of oil and gas processes with user-friendly interfaces and reduced computational costs. This era saw widespread implementation in industries, as enabled iterative modeling without reliance on expensive hardware. From the 2000s onward, (HPC) revolutionized process simulation by supporting larger-scale and more intricate models, particularly for multiphase and reactive systems in . HPC clusters allowed simulations to handle millions of grid points, improving resolution for phenomena like and that traditional methods struggled with. Concurrently, the incorporation of (CFD) into process simulators enhanced detailed flow analysis, enabling hybrid models that couple 1D process flowsheets with 3D CFD for equipment-level predictions in reactors and separators. Post-2010 developments have increasingly integrated (AI) and (ML) to create surrogate models, such as neural networks, that approximate expensive physics-based with while reducing . These surrogates, often trained on simulation data, facilitate rapid optimization in chemical processes like and . Digital twins, virtual replicas updated in with sensor data, have emerged as a key application, enabling and control in plants through ML-enhanced process models. Key trends shaping modern process simulation include the shift to cloud-based platforms for scalable, collaborative computing and the rise of open-source tools like , which promote accessibility and customization for academic and small-scale industrial use. A growing emphasis on drives simulations for green process optimization, such as minimizing energy use in carbon capture and production. In the , process simulation saw broad adoption in renewables, modeling solar thermal systems and biomass conversion to support efficient design and scale-up. By the 2020s, AI-hybrid approaches have achieved significant computation time reductions in surrogate-assisted workflows, accelerating innovation in transitions.

Tools and Implementation

Software Packages

Process simulation software packages are essential tools for modeling and analyzing chemical, , pharmaceutical, and other . These packages generally fall into two categories: , which often provides robust support, extensive validation, and industry-specific features at a significant cost, and open-source alternatives, which offer flexibility and no licensing fees but may require more user expertise for customization and validation. tools dominate in settings due to their reliability and integration capabilities, while open-source options are popular in and for prototyping. Capabilities vary by package, supporting steady-state simulations for equilibrium-based designs, dynamic simulations for transient operations, and sometimes 1D or for spatial distributions in reactors or pipelines. Key commercial packages include Aspen Plus and from AspenTech, widely used for steady-state and dynamic simulations in oil and gas, refining, and pharmaceuticals. Aspen Plus excels in and optimization with over 37,000 built-in components and thermodynamic models like Peng-Robinson for systems, enabling flowsheet simulations for and units. extends this to dynamic modeling, particularly for upstream oil and gas operations such as reservoir-to-refinery integration, and includes safety analysis tools like flare system design. gPROMS from Process Systems Enterprise (now ) specializes in advanced dynamic modeling for process control, optimization, and nonlinear , supporting custom equation-oriented models for complex phenomena like or . integrates process simulation with finite element analysis for multiphysics problems, such as coupled with chemical reactions in 3D geometries, making it suitable for electrochemical or catalytic processes. Open-source packages provide accessible alternatives, with standing out as a free dynamic simulator based on the language for object-oriented modeling of physical systems. It supports equation-based simulations for processes like heat exchangers or control systems, with libraries for components, and is extensible via integration for data analysis. Other open-source tools like offer steady-state capabilities with thermodynamic property calculations, though they lack the comprehensive validation of commercial suites. These tools facilitate collaborative development and are often used in educational settings to teach process dynamics without financial barriers. Common features across these packages include built-in property databases for pure components and mixtures, such as NIST-based thermophysical data for accurate vapor-liquid predictions, and graphical user interfaces for drag-and-drop flowsheet construction. Many support export functionalities to optimization tools like or gAMS for advanced parameter estimation and economic analysis, enhancing workflow efficiency in design iterations. For instance, Aspen packages integrate with Excel for , while COMSOL allows coupling with external solvers for hybrid simulations. Selection of a software package depends on several factors, including industry-specific libraries—such as models in Aspen Plus for pharmaceutical applications involving ionic solutions—and for large-scale simulations handling thousands of units. is a major consideration, with commercial licenses like AspenTech's often exceeding $10,000 annually per user, plus maintenance fees, whereas open-source options like incur no direct costs but may require investment in training or hardware for computational demands. Users prioritize packages with validated models against experimental data, strong , and compatibility with standards like OPC for plant integration. In the oil and gas sector, for example, is frequently selected for its proven accuracy in debottlenecking operations, where simulations have identified significant throughput increases by optimizing crude units.

Best Practices and Challenges

In process simulation, effective implementation relies on rigorous best practices to ensure model reliability and usability. Model validation against experimental or operational data is essential, involving comparisons of simulated outputs with real-world measurements to quantify discrepancies and refine parameters, such as adjusting thermodynamic models until predictions align within acceptable error margins. approaches, where simulations begin with simple components and progressively incorporate complexity, facilitate easier and , as seen in structured workflows that start with basic unit operations before integrating full systems. Thorough documentation of assumptions, including parameter sources and boundary conditions, supports and team , reducing errors in iterative . Managing uncertainty is critical for robust simulations, particularly in handling variability from input parameters like physical properties or operating conditions. simulations propagate uncertainties by sampling random values from probability distributions of inputs, generating statistical distributions of outputs to assess , such as variability in reactor yields due to feed composition fluctuations. studies complement this by systematically varying individual parameters to identify influential factors, enabling prioritization of data refinement efforts, for instance, in column designs where feed temperature impacts significantly more than minor impurities. Despite these practices, process simulation faces notable challenges. Computational intensity is a primary hurdle, especially for dynamic models that solve equations over time, often requiring hours or days on standard for large-scale systems like refinery operations due to the need for fine time steps and iterative solvers. Trade-offs between model fidelity and simulation speed are inherent, as high-fidelity representations with detailed increase accuracy but escalate resource demands, necessitating simplified approximations for applications like operator training. Integration with systems poses additional difficulties, as older architectures may lack compatible interfaces, complicating exchange and leading to inconsistencies in simulations for processes. Ethical considerations underscore the responsibility to mitigate overlooked risks in simulations. Simulations must account for to avoid underestimating hazards, such as in processes where probabilistic modeling of low-frequency accidents like coolant failures is vital to prevent catastrophic oversights, ensuring public safety aligns with engineering codes. Looking ahead, addressing simulation gaps involves emerging technologies like , which promises to handle ultra-complex molecular interactions intractable for classical systems, potentially revolutionizing simulations of catalytic reactions or formations by solving equations at unprecedented scales. Additionally, as of 2025, AI-supported simulation is gaining traction, enabling faster optimization and predictive capabilities in tools like .

References

  1. [1]
    None
    ### Summary for Encyclopedia Introduction: Process Simulation
  2. [2]
    Leverage Dynamic Process Simulation - AIChE
    Dynamic simulations are able to replicate actual plant behavior, making them valuable for plant design, operation, optimization, and training.
  3. [3]
    CEP: Process Automation Corner - Process Simulation and ... - AIChE
    Chemical engineers have been practicing process simulation since the advent of the profession in the early 20th century. It was called process design and ...
  4. [4]
    None
    ### Summary for Encyclopedia Intro: Process Simulation in Chemical Engineering
  5. [5]
    Chemical Engineering Process Simulation - ScienceDirect.com
    Chemical Engineering Process Simulation is ideal for students, early career researchers, and practitioners, as it guides you through chemical processes and ...
  6. [6]
    [PDF] Principles of Momentum, Mass and Energy Balances
    The basic principle used in modeling of chemical engineering process is a concept of balance of momentum, mass and energy, which can be expressed in a general ...
  7. [7]
    Process Simulation - an overview | ScienceDirect Topics
    Process simulation always uses models which introduce approximations and assumptions; but allow the description of a property over a wide range of temperatures ...
  8. [8]
    The Expert's Guide to Process Simulation
    Apr 29, 2025 · Process simulation is an everyday engineering task for designing, troubleshooting, and optimizing chemical processes.
  9. [9]
    Process Simulation & Why It Is Important - Audubon Companies
    Process simulation is a powerful software tool that allows refinery owners, operators, and engineers to virtually model a process in extreme detail.
  10. [10]
    What is Process Simulation & How to achieve efficiency with it?
    Nov 11, 2024 · Learn about process simulation & how it empowers companies to test changes, cut costs, and improve operations in a risk-free environment.
  11. [11]
    Using chemical process simulation to design industrial ecosystems
    Chemical process simulation (CPS) software has been widely used by chemical (process) engineers to design, test, optimise, and integrate process plants.
  12. [12]
    [PDF] The Role of Process Simulation in Pharmaceutical Process ... - ISPE
    Other tasks that can be handled by process simulators include process scheduling, environmental impact assessment, debottlenecking, and throughput analysis.
  13. [13]
    [PDF] Simulation Evolves in Power Plants | Emerson
    Long a tool for training operators, power plant simulation platforms find a growing role in advancing financial and operational goals. inside energy. Page 2. 2 ...
  14. [14]
    A review of commercial process simulators applied to food processing
    Nov 30, 2022 · This review includes a wide variety of food and beverages process like evaporation, concentration, distillation, fermentative reactions, membrane separation ...
  15. [15]
    Optimization of wastewater treatment facilities using process ...
    Mar 15, 1998 · The role of process simulation in designing, evaluating, and optimizing wastewater treatment facilities is discussed.
  16. [16]
    Carbon Capture Simulation for Industry Impact (CCSI 2 )
    The Carbon Capture Simulation for Industry Impact (CCSI2) project is focused on developing a fundamental understanding of CO2 capture technology, which will ...
  17. [17]
    Fluid Flow: Conservation of Momentum, Mass, and Energy - COMSOL
    Jun 29, 2018 · A flow field is characterized by balance in mass, momentum, and total energy described by the continuity equation, the Navier-Stokes equations, and the total ...<|control11|><|separator|>
  18. [18]
    [PDF] Conservation Laws in Continuum Modeling. - MIT Mathematics
    The first step in the modeling process is to identify conserved quantities (e.g. mass) and define the appropriate densities and fluxes | as in the following ...
  19. [19]
    Navier-Stokes Equations
    These equations describe how the velocity, pressure, temperature, and density of a moving fluid are related. The equations were derived independently by G.G. ...
  20. [20]
    first-law-open-systems-introduction - LearnChemE
    This module uses screencasts and an interactive simulation to explain the application of the first law of thermodynamics (energy balance) to open systems.
  21. [21]
    Dynamic simulation of chemical processes described by distributed ...
    In this paper, a general procedure is presented which permits the units to be described by both distributed and lumped parameter models.
  22. [22]
    A direct linear systems solver for pipe networks - ScienceDirect.com
    In this paper, an efficient method for the solution of matrix equations resulting from the flow of water in a pipe network is presented. The method is found ...Missing: process simulation
  23. [23]
    Lumped and distributed‐parameter pipe model framework for ...
    Jul 9, 2022 · Compared with the lumped-parameter pipe model, the distributed-parameter pipe model has higher computational efficiency. But only the input and ...INTRODUCTION · FORMULATION OF... · SIMULATION AND VALIDATION
  24. [24]
    Sensitivity analysis in the context of uncertainty ... - ScienceDirect.com
    The aim of sensitivity analysis is to identify the main contributors to model output uncertainty. Therefore the term “Uncertainty Importance Analysis” is ...
  25. [25]
    Uncertainty Quantification and Sensitivity Analysis in Process ...
    This work presents a PSP simulation framework for laser powder bed fusion with a focus on uncertainty quantification through probabilistic calibration.
  26. [26]
    Validation of thermophysical data for scientific and engineering ...
    The present work describes reporting recommendations and property data validation methods ... Development of on-demand critically evaluated thermophysical ...
  27. [27]
    The Antoine Equation for Vapor-pressure Data. | Chemical Reviews
    Mahsa Gholami, Joey H. T. Wolbers, Tim Schuttevaer, Meik B. Franke, Boelo Schuur. Experimental Study on Vapor–Liquid and Solid–Liquid Equilibria Data for ...
  28. [28]
    Group‐contribution estimation of activity coefficients in nonideal ...
    The group-contribution method predicts activity coefficients in liquid mixtures using functional-group concepts and group-interaction parameters, often with ...
  29. [29]
    Group contribution-based property estimation methods
    This perspective paper gives a brief overview on the state of the art in group-contribution-based property estimation methods and their further development ...Missing: thermophysical | Show results with:thermophysical
  30. [30]
    (PDF) Thermodynamic Modeling with Equations of State: Present ...
    In this work, we review present challenges associated with established models, and give suggestions on how to overcome them in the future.
  31. [31]
    Steady State Simulation - an overview | ScienceDirect Topics
    Steady-state simulation is defined as a modeling approach that assumes the system operates in a perfectly stable manner, where variation with respect to ...Missing: seminal | Show results with:seminal
  32. [32]
    Steady State Chemical Process Simulation: A State-of-the-Art Review
    May 30, 1980 · Perspective. The use of a mathematical model on a computer to simulate a chemical process is now approximately two decades old.Missing: seminal | Show results with:seminal
  33. [33]
    Steady-State and Dynamic Simulation: What is the Difference
    Steady-state simulation maximizes productivity; you obtain the result directly and faster, by several orders of magnitude, and it simplifies post-processing of ...
  34. [34]
    Equation oriented approach to process flowsheeting - ScienceDirect
    Steady State Simulation Review. Paper presented at 178th ACS National ... Efficient solution of design problems using a sequential-modular flowsheeting programme.
  35. [35]
    A matrix approach for steady-state simulation of heat exchanger networks
    ### Summary of Abstract on Steady-State Simulation of Heat Exchanger Networks
  36. [36]
  37. [37]
    Historical development of the vapor pressure equation from dalton to ...
    The vapor pressure of a pure liquid or of a solution is a property that has been observed since antiquity. Here, we trace the different equations that have.
  38. [38]
    Thermodynamics - Gilbert Newton Lewis, Merle Randall
    Since its first publication in 1923, this volume has been considered one of the great books in the literature of chemistry. In the early 1960s, ...
  39. [39]
    Warren Kendall Lewis | Biographical Memoirs: Volume 70
    Lewis published thirteen papers on distillation and nine on evaporation; nineteen of his eighty-one patents were on distillation. The movement of underground ...
  40. [40]
    What is a Process Flow Diagram - Lucidchart
    This type of diagram has its roots in the 1920s. In 1921, industrial engineer and efficiency expert Frank Gilbreth, Sr. introduced the “flow process chart” to ...
  41. [41]
    Analog Computer - an overview | ScienceDirect Topics
    In the early 1940s and 1950s, analog computers were commonly used to simulate continuous dynamic systems, mainly, automatic control systems, mechanical systems ...
  42. [42]
    Fortran - Wikipedia
    Fortran was originally developed by IBM with a reference manual being released in 1956; however, the first compilers only began to produce accurate code two ...Intel Fortran Compiler · Fortran 95 language features · GNU Fortran · Speedcoding
  43. [43]
    [PDF] Introduction - JUST
    Commercial process simulators first appeared in the 1970's. Currently ... commercial (and one open source) software applications for process simulation.Missing: history 1990s
  44. [44]
    [PDF] The Development of Chemical Process Simulation Software ... - Aidic
    In history Borland C++ and Visual Basic 6.0 are the first language for GUI designing, then Visual C++ MFC becomes the tendency, and now days C# and JAVA ...
  45. [45]
    Aspen Plus | Leading Process Simulation Software - AspenTech
    Applications · Adsorption & Chromatography Improvement. Optimize separation processes. · Batch Process Improvement. Simplify batch process development. · Energy ...
  46. [46]
    Aspen HYSYS - Wikipedia
    Aspen HYSYS (or simply HYSYS) is a chemical process simulator currently developed by AspenTech used to mathematically model chemical processes.
  47. [47]
    What is Computational Fluid Dynamics (CFD)? - Ansys
    Apr 24, 2024 · High-Performance Computing (HPC) (2000s-Present):. With advancements in HPC, running larger, more complex CFD models in less time is possible.
  48. [48]
    (PDF) High-Performance Computing to Accelerate Large-Scale ...
    This study paper presents an in-depth review of the function of high-performance computing (HPC) in enhancing computational fluid dynamics (CFD) simulations on ...
  49. [49]
    Computational fluid dynamics - Wikipedia
    Recently CFD methods have gained traction for modeling the flow behavior of granular materials within various chemical processes in engineering. This ...<|separator|>
  50. [50]
    A machine learning approach for the surrogate modeling of ...
    This paper explores the capability of different machine learning techniques for modeling different chemical process with different non-linear behavior.Missing: post- | Show results with:post-
  51. [51]
    Digital twins in process engineering: An overview on computational ...
    A digital twin (DT) is an automation strategy that combines a physical plant with an adaptive real-time simulation environment.Missing: post- | Show results with:post-
  52. [52]
    Computational Processes → Term - Sustainability Directory
    Mar 14, 2025 · Term → Computational processes aid sustainability through data-driven analysis, optimization, and the modeling of complex environmental ...
  53. [53]
    Impacts of digitalization on smart grids, renewable energy, and ...
    Oct 26, 2024 · Common applications of AI include load forecasting, renewable energy production forecasting, detecting power quality issues, and identifying ...
  54. [54]
    AI-Based Surrogate Models for the Food and Drink Manufacturing ...
    Surrogate models provide virtual representations that mirror physical objects or processes, serving distinct purposes in simulations and digital ...3. Surrogate Modelling · 3.2. Surrogate Modelling... · 4.3. 2. Gaussian Process...<|separator|>
  55. [55]
    Seven Golden Rules of Process Modelling - SysCAD
    Dec 19, 2023 · 1. Clearly Define your Scope and Objectives. Before starting any simulation, it is essential to define the scope and objectives of the model.
  56. [56]
    Simultaneous Process and Economic Uncertainty Analysis - AIChE
    In the early stages of technology development, process simulators such as Aspen Plus and Aspen HYSYS enable engineers to quickly evaluate new processes. However ...
  57. [57]
    The Monte Carlo driven and machine learning enhanced process ...
    Jun 9, 2019 · This paper presents a methodology that integrates property uncertainty analysis and sensitivity analysis in commercial process simulators.
  58. [58]
    What Is Dynamic Simulation? Engineering Applications & Methods
    Dynamic simulation evaluates stability, control performance, and emergency response. A hybrid approach combines steady-state optimization with dynamic analysis, ...
  59. [59]
    Dynamic Scheduling: A Comparison of High-Fidelity Models with ...
    Oct 29, 2025 · This work investigates the trade-off between model fidelity and optimization complexity in dynamic scheduling under integrated dynamic ...
  60. [60]
    The role of process engineering in the digital transformation
    This paper examines several challenges specific to how process simulation limits the transformation of process engineering.
  61. [61]
    Duck, Duck, Black Swan: How the Brain Can Simplify Rare, High ...
    The human brain tends to simplify high-consequence, low-frequency events, which can be particularly dangerous in process hazard analysis (PHA) development.<|control11|><|separator|>
  62. [62]
    Quantum Computing: An Emerging Tool for Chemical Engineers
    May 18, 2021 · Quantum computing technology has been shown to be especially useful when solving problems involving chemistry, artificial intelligence, simulation tasks.