Bond graph
A bond graph is a graphical modeling tool used to represent the dynamic behavior of physical systems by depicting the flow and conservation of energy through interconnected components, where power is defined as the product of effort and flow variables.[1] Developed by Henry M. Paynter at MIT and first published in 1961, it provides a unified, domain-independent framework applicable to diverse engineering domains such as mechanical, electrical, hydraulic, thermal, and chemical systems.[2][3] Bond graphs consist of bonds, which are directed lines symbolizing power exchange between elements, connected via junctions (0-junctions for equal effort and summing flows, 1-junctions for equal flow and summing efforts) and basic elements including resistors (R) for dissipation, inertias (I) and capacitors (C) for energy storage, sources (Se for effort, Sf for flow), and two-port modulators like transformers (TF) and gyrators (GY).[1][3] This structure enforces fundamental physical laws, such as energy conservation, without domain-specific analogies, enabling the systematic derivation of state-space equations for simulation and control.[4] Originally inspired by hydroelectric engineering and network theory, bond graphs emerged from Paynter's work on system dynamics in the 1950s, with the core concept documented in his seminal book Analysis and Design of Engineering Systems.[2][3] Their non-causal nature—where causality is assigned post-construction—facilitates modular, hierarchical modeling and reuse of submodels, making them particularly valuable for multidisciplinary systems like robotics, vehicles, and biochemical processes.[3][4] Over decades, the methodology has been extended through software tools and formal mathematical foundations, influencing fields from control theory to systems biology.[1]Introduction
Definition and Purpose
A bond graph is a graphical representation of a physical dynamic system, utilizing directed bonds to depict the exchange of power between interconnected components. Introduced by Henry M. Paynter, these diagrams abstract the functional structure of energetic systems by focusing on energy transactions rather than specific material properties.[5] Each bond connects multiport elements and represents a power flow defined by a conjugate pair of effort and flow variables, such as force and velocity in mechanical systems or voltage and current in electrical ones.[5] The core purpose of bond graphs is to enable the modeling of interdisciplinary dynamic systems—spanning mechanical, electrical, hydraulic, and thermal domains—within a common framework grounded in energy conservation and power continuity.[6] This unified approach facilitates the analysis and synthesis of complex systems by emphasizing the topological interconnections of energy ports, thereby avoiding the need for disparate equations tailored to individual domains.[7] Power in these models is conceptualized as the instantaneous product of effort and flow, providing a universal metric for energy exchange without delving into domain-specific derivations.[5] Bond graphs offer several key advantages, including modularity for reusable subsystem models and the explicit assignment of causality to determine computational directions, which improves simulation efficiency and model debugging.[6] By prioritizing energy balance at junctions and ports, they promote a deeper conceptual understanding of system behavior across disciplines, making them particularly valuable for engineering design and control applications.[8]History and Development
The bond graph methodology originated in 1959 when Henry M. Paynter, a professor at the Massachusetts Institute of Technology (MIT), introduced it as a unified framework for modeling dynamic systems across multiple engineering domains, drawing on concepts of energy flow and thermodynamic systems. Paynter presented the foundational ideas in a lecture titled "Ports, Energy and Thermodynamic Systems" on April 24, 1959, at MIT, aiming to bridge mechanical, electrical, hydraulic, and thermal systems through a graphical representation of power exchange. This innovation built on earlier work in analog computing and system analysis at MIT, where Paynter had been developing tools for engineering education and research since the 1940s. His seminal 1961 book, Analysis and Design of Engineering Systems, formalized the approach, establishing bond graphs as a tool for deriving state-space equations from physical principles without domain-specific reformulation.[9][10] In the 1960s and 1970s, bond graphs gained early adoption through the contributions of researchers like Dean Karnopp, Ronald C. Rosenberg, and Peter Wellstead, who expanded its applications in vehicle dynamics, control systems, and multidisciplinary modeling. Karnopp and Rosenberg's 1968 textbook, Analysis and Simulation of Multiport Systems: The Bond Graph, provided a rigorous mathematical foundation, demonstrating how bond graphs could simulate complex interactions in mechanical and electromechanical systems. Their subsequent works, including applications to drive-line dynamics in 1970 and further textbooks in 1975, popularized the method in academic and industrial settings. Wellstead advanced its integration with system identification techniques in the late 1970s, emphasizing bond graphs' role in parameter estimation for control engineering.[11][12][13] The 1980s and 1990s saw standardization of bond graph techniques through influential publications and growing acceptance in engineering practice. Karnopp and Rosenberg's 1983 book, Introduction to Physical System Dynamics, synthesized the methodology into a comprehensive pedagogical resource, emphasizing its unified approach to mechatronic systems and influencing curricula in mechanical and systems engineering programs. By this time, bond graphs had earned recognition in university courses worldwide, including at MIT and other leading institutions, where they were taught as a core tool for modeling dynamic systems.[14][15] Entering the 2000s, bond graphs evolved with the rise of computer-aided design tools, enabling automated simulation and analysis. Software like SYMBOLS 2000 and 20-sim facilitated graphical model construction and integration with numerical solvers, reducing manual equation derivation. In parallel, libraries such as the Modelica Bond Graph Library, introduced around 2005, embedded bond graphs within the object-oriented Modelica language for acausal modeling of complex systems. By the 2020s, this progressed to seamless integration with standards like the Functional Mock-up Interface (FMI), allowing bond graph-derived models to be exported and co-simulated across tools for multidomain applications in automotive and aerospace engineering.[16][17][18]Fundamental Concepts
Effort and Flow Variables
In bond graph modeling, the fundamental variables are the effort e and flow f, which represent generalized quantities analogous to force and velocity across diverse physical domains. The effort variable is an intensive quantity that acts across a system port, such as voltage in electrical systems or force in mechanical systems.[5] The flow variable is an extensive quantity that flows through a system port, such as current in electrical systems or velocity in mechanical systems.[5] The product of effort and flow defines power as p = e \times f, where the units yield watts, ensuring conservation of power across interconnected domains in the bond graph framework.[5] This power conjugation allows unified modeling of multi-domain systems by mapping domain-specific variables to these generalized forms.[1] Domain-specific interpretations of effort and flow are summarized in the following table, illustrating their role in power transmission for common engineering domains:| Domain | Effort (e) | Flow (f) |
|---|---|---|
| Electrical | Voltage (V) | Current (A) |
| Mechanical Translational | Force (N) | Velocity (m/s) |
| Mechanical Rotational | Torque (N·m) | Angular velocity (rad/s) |
| Hydraulic | Pressure (Pa) | Volume flow rate (m³/s) |
| Thermal | Temperature (K) | Entropy flow rate (J/(K·s)) |
| Chemical | Chemical potential (J/mol) | Molar flow rate (mol/s) |
Bonds and Power Conjugation
In bond graphs, bonds serve as the fundamental directed edges that represent the flow of power between system components. Each bond connects the ports of physical elements, such as storage or dissipative components, and is depicted as a line with a half-arrow indicating the reference direction of power flow. This direction is arbitrary but consistent within the graph, ensuring that power is positive when flowing in the direction of the arrow. The bond structure also accommodates a causal stroke, represented by a full arrowhead, which specifies the direction of information or causality flow but is introduced here only as an overlay on the power direction for later computational purposes.[5][6] Power conjugation refers to the pairing of effort and flow variables along each bond, where these variables are defined such that their product yields the instantaneous power transmitted. Specifically, the power P on a bond is given by P = e \cdot f where e is the effort variable (e.g., force or voltage) and f is the flow variable (e.g., velocity or current), which are conjugate in the sense that they jointly describe energy exchange across domains. The power flow direction along the bond—from the side providing effort to the side receiving flow, or vice versa—ensures overall energy conservation in the system, as bonds link elements that store potential or kinetic energy without loss in ideal representations. This conjugation allows bond graphs to unify modeling across mechanical, electrical, hydraulic, and other domains by treating power as a common currency.[5][1][6] At junctions, where multiple bonds converge, conjugation rules enforce the equality of efforts or flows depending on the junction type, with bond orientations determining the sign of variables. For a 0-junction, all connected efforts are equal (e_1 = e_2 = \dots), and the algebraic sum of flows is zero (\sum f_i = 0); for a 1-junction, all connected flows are equal (f_1 = f_2 = \dots), and the algebraic sum of efforts is zero (\sum e_i = 0). These rules, combined with bond directions, maintain power balance at the junction, expressed as the sum of powers equaling zero: \sum p = 0 This equation reflects the conservation of power without dissipation at ideal junctions, where incoming and outgoing powers balance. Diagrammatically, bonds link elements by emanating from or entering junctions, forming a network that visually captures how energy is distributed and stored in components like capacitors (potential energy) or inductors (kinetic energy).[5][22][6]Tetrahedron of State
The tetrahedron of state serves as a conceptual framework in bond graph theory for classifying the state variables of dynamic systems, illustrating the interconnections among effort, flow, and their time integrals.[23] This geometric representation, introduced by Henry Paynter, depicts a tetrahedron with four vertices corresponding to the primary state variables: effort (e), flow (f), momentum (p = \int e \, dt), and displacement (q = \int f \, dt).[24] The edges of the tetrahedron symbolize the differential relationships, such as \dot{p} = e and \dot{q} = f, which highlight the power-conjugate nature of these variables in energy-based modeling.[25] Energy storage elements are positioned at specific vertices within this structure to represent their constitutive behaviors. Inertias, or I-elements, are associated with the momentum vertex (p), where the flow is related to momentum via f = \frac{1}{I} p for linear cases, storing kinetic energy.[22] Compliances, or C-elements, align with the displacement vertex (q), with effort related to displacement by e = \frac{1}{C} q, and flow given by f = \frac{dq}{dt}, thereby storing potential energy.[23] These single-port storage elements exemplify how the tetrahedron encapsulates the fundamental dynamics of accumulation without dissipation.[25] Geometrically, the tetrahedron provides a four-dimensional state space interpretation for dynamic systems, where the vertices and edges facilitate visualization of state transitions and energy flows across domains like mechanical, electrical, and hydraulic systems.[22] This structure draws conceptual parallels to Lagrangian and Hamiltonian mechanics by emphasizing energy coordinates and their conjugates, offering a unified view of system states without relying on domain-specific formulations.[26]Components
Single-Port Elements
Single-port elements in bond graphs represent fundamental physical components that exchange power through a single bond, capturing dissipation, storage, and imposition of variables across engineering domains such as mechanical, electrical, hydraulic, and thermal systems. These elements adhere to power conjugation, where effort e and flow f satisfy P = e \cdot f, and are connected via a half-arrow bond indicating power flow direction from effort to flow. The primary single-port elements include resistors (R) for dissipation, inertias (I) and compliances (C) for storage, effort sources (Se), flow sources (Sf), and sinks (Re, Rf) for boundary conditions. Their behaviors are defined by constitutive relations that depend on causality assignment, with standard notation using rectangular or circular symbols attached to the bond.[27][28] The resistor (R) models irreversible energy dissipation, such as friction in mechanics, ohmic losses in electricity, or viscous drag in fluids, converting mechanical or electrical power into heat. Its constitutive relation is e = R f under flow causality (flow input, effort output) or f = \frac{1}{R} e under effort causality (effort input, flow output), where R > 0 is the resistance parameter with units ensuring dimensional consistency (e.g., ohms in electrical systems). In bond graph notation, it is depicted as a rectangle labeled "R" with the parameter value inside, and modulated variants (MR) allow R to depend on external signals. This element ensures entropy production in irreversible processes and is essential for realistic modeling of damping effects.[27][28] Inertias (I) capture kinetic energy storage through momentum accumulation, analogous to mass in translational mechanics or inductance in electrical circuits. The core constitutive relation is p = I f, where p is the momentum state variable and I > 0 is the inertia coefficient (e.g., kg for mass, H for inductance). Under preferred integral causality (effort input to the element), p = \int e \, dt and f = \frac{p}{I} = \frac{1}{I} \int e \, dt; alternatively, under derivative causality (flow input), e = I \frac{df}{dt}. The element is represented as a rectangle labeled "I", with modulated forms (MI) for variable inertia. Integral causality is favored in simulations to avoid numerical stiffness from differentiating flow.[27][28] Compliances (C) represent potential energy storage via displacement or charge buildup, such as in springs (mechanical), capacitors (electrical), or hydraulic accumulators. The constitutive relation is q = C e, where q is the displacement state variable and C > 0 is the compliance (e.g., F for capacitance, m/N for spring compliance). With preferred integral causality (flow input), q = \int f \, dt and e = \frac{q}{C} = \frac{1}{C} \int f \, dt; in derivative causality (effort input), f = C \frac{de}{dt}. Notation uses a rectangle labeled "C", and modulated compliances (MC) accommodate varying C. This form promotes stable integration in computational models by integrating flow to update the state.[27][28] Effort sources (Se) impose a prescribed effort value, modeling ideal actuators or drivers like constant voltage sources or pressure pumps, independent of the conjugate flow. The constitutive relation is simply e = u(t), where u(t) is a specified time-varying function, with flow f determined by the connected system. It is symbolized by a circle containing "Se" and the effort specification. Modulated effort sources (MSe) incorporate external modulation for controlled inputs.[27][28] Flow sources (Sf) dictate a fixed flow, representing devices such as constant current sources or imposed velocities, irrespective of effort. The relation is f = u(t), with effort e set by the system response. Denoted by a circle with "Sf" and the flow function, modulated versions (MSf) enable dynamic control. These sources are key for defining input boundaries in dynamic simulations.[27][28] Sinks provide termination conditions at system boundaries, effectively setting one variable to zero while allowing the conjugate to vary freely. An effort sink (Re) enforces e = 0, modeled as a resistor with effort causality and infinite resistance (f = \frac{1}{R} e with R \to \infty, yielding f arbitrary), such as a grounded electrical terminal or zero-pressure reference. It uses R notation with "Re" to indicate causality. A flow sink (Rf) sets f = 0, as a resistor with flow causality and zero resistance (e = R f with R = 0), like an open circuit or free-floating end, denoted "Rf". These are crucial for open-system modeling and variable measurement without power exchange.[27][28]Two-Port Elements
Two-port elements in bond graphs represent ideal transducers that modulate power between exactly two ports without energy storage or dissipation, ensuring power conservation across the ports.[1] These elements transmit energy from one domain or subsystem to another, facilitating the modeling of mechanical, electrical, or fluidic couplings where variables are scaled by a fixed or variable modulus.[22] The transformer (TF) is a two-port element that scales both effort and flow variables proportionally between its ports using a modulus n, often representing mechanical leverage or electrical turns ratios.[22] Its constitutive relations are given by: \begin{align} e_1 &= n e_2, \\ f_2 &= n f_1, \end{align} where e_1, f_1 are the effort and flow at port 1, and e_2, f_2 at port 2.[22] This structure ensures power conservation, as e_1 f_1 = e_2 f_2, making the TF lossless and reversible.[1] Physical analogies include ideal gears or levers in mechanical systems, where n corresponds to a gear ratio or lever arm length ratio, scaling force and velocity inversely to maintain power balance.[22] The gyrator (GY) is another two-port element that couples effort at one port to flow at the other via a modulus r, commonly modeling transducers like electric motors or fluid-mechanical interfaces. Its equations are: \begin{align} e_1 &= r f_2, \\ e_2 &= r f_1, \end{align} with power conservation holding as e_1 f_1 = e_2 f_2.[22] Examples include electromagnetic devices, such as a DC motor where electrical voltage relates to mechanical torque and current to angular velocity through a motor constant r.[1] Gyrators are distinct from transformers in that they represent non-reciprocal power flow directions in certain physical realizations, though both are ideal and energy-preserving.[22] Modulated versions of these elements, denoted as mTF and mGY, allow the moduli n or r to vary as functions of external signals or system states, enabling the representation of nonlinear or controlled transducers.[3] In bond graph notation, the modulating signal is indicated by an arrow pointing to the element, with the variable modulus computed from other graph variables.[3] Causality assignment for two-port elements follows preferences that support efficient computational integration during simulation.[22] For the TF, the preferred causality has one port as effort-causal (stroke on the bond indicating effort direction) and the other as flow-causal, allowing direct propagation of variables without integration loops.[22] The GY prefers both ports to be either effort-causal or flow-causal, which determines whether efforts or flows are solved algebraically first in the system's state equations.[22] These conventions minimize derivative causality and ensure numerical stability in bond graph-based modeling tools.[1]Multi-Port Junctions
Multi-port junctions in bond graphs serve as essential power-conserving nodes that interconnect multiple bonds, enabling the representation of complex interactions where multiple energy pathways converge. These junctions enforce specific constraints on effort and flow variables across the connected bonds, ensuring that power balance is maintained without storage or dissipation. There are two fundamental types: 0-junctions and 1-junctions, which are dual to each other and correspond to parallel and series configurations in physical systems, respectively.[1][3] A 0-junction enforces a common effort across all connected bonds, such that the effort e is equal for every bond (e_1 = e_2 = \dots = e_n), while the algebraic sum of the flows is zero (\sum_{i=1}^n f_i = 0). This structure represents parallel power flow, analogous to parallel electrical circuits where voltages are equal and currents sum to zero (Kirchhoff's current law), or mechanical systems with common force and summing velocities. For instance, in a three-port 0-junction, the flows satisfy f_1 + f_2 + f_3 = 0, with the common effort e shared among all.[1][3] In contrast, a 1-junction imposes a common flow across all bonds (f_1 = f_2 = \dots = f_n), with the efforts summing to zero (\sum_{i=1}^n e_i = 0). This models series power flow, similar to series electrical connections where currents are equal and voltages sum to zero (Kirchhoff's voltage law), or mechanical systems with shared velocity and additive forces. For a three-port 1-junction, the relation is e_1 + e_2 + e_3 = 0, with identical flow f on each bond.[1][3] Bond orientation plays a critical role in handling signs for the summation rules, with bonds classified as attached (directly connected to an element or source) or detached (intermediate, between junctions). The half-arrow on each bond indicates the direction relative to the junction: flows entering the junction are positive, while those leaving are negative, ensuring the sum-to-zero condition accounts for directionality. Detached bonds, often used in cascades, do not affect the overall constraints but simplify graph reduction.[1][3] Power conservation in multi-port junctions arises automatically from the equality and summation rules. For a 0-junction, the total power is e \sum f_i = e \cdot 0 = 0, confirming no net power accumulation. Similarly, for a 1-junction, it is f \sum e_i = f \cdot 0 = 0. This inherent property aligns with the bond graph's focus on power conjugation, where each bond carries power e \cdot f.[1][3] Notation for junctions uses a circular symbol labeled with "0" or "1" inside, often annotated with the number of ports in parentheses, such as 0(3) for a three-port 0-junction or 1(4) for a four-port 1-junction. This compact representation highlights the junction type and connectivity, facilitating clear visualization in bond graph diagrams.[1][3]Modeling Techniques
Graph Construction and Associations
Bond graphs are assembled by connecting storage, dissipation, and source elements through junctions that enforce power-conserving relationships, enabling the representation of complex systems in a modular fashion. The primary associations used in construction are series and parallel configurations, which correspond directly to the structure of 1-junctions and 0-junctions, respectively. These associations allow for the systematic buildup of models from basic components without specifying causality at the outset, preserving the multi-domain applicability of the formalism.[3] In a series association, equivalent to a 1-junction, all connected elements share a common flow variable while the efforts across them sum to balance the junction equation (∑ e = 0). This setup models scenarios where elements are connected end-to-end, such as inductors in series within an electrical circuit or masses connected by springs in a mechanical chain, ensuring that the total effort drop equals the sum of individual efforts. Conversely, a parallel association, represented by a 0-junction, imposes a common effort across all bonds while the flows sum to zero (∑ f = 0), suitable for elements sharing the same potential, like capacitors in parallel or parallel mechanical dampers. These junction-based associations maintain power conjugation, as the product of effort and flow remains consistent in magnitude but opposite in sign across bonds entering and leaving the junction.[3][29] The construction of a bond graph follows a systematic procedure to ensure completeness and consistency. First, identify the physical domains involved and list all energy storage (I, C), dissipation (R), and source (Se, Sf) elements. Next, define reference efforts and flows for the system boundaries, then introduce internal efforts and flows at interconnection points using modulated sources (Me, Mf) if necessary. Connect these via 0- and 1-junctions to represent parallel and series associations, respectively, and finally incorporate any transformers (TF) or gyrators (GY) for variable scaling. This step-by-step approach, applied to examples like electromechanical actuators, facilitates the translation from physical schematics to a unified graph.[3] Hierarchical modeling enhances scalability by treating subsystems as super-elements, where internal bonds are encapsulated, and only external bonds interface with the larger graph. For instance, a motor subsystem can be abstracted as a single multi-port element with input effort/flow bonds, allowing reuse across models without exposing details. Simplification rules streamline the graph post-construction: redundant junctions connecting only two bonds in the same power direction can be eliminated by direct connection; identical adjacent 0- or 1-junctions merge into one; and fused effort or flow differences (from modulated sources) reduce clutter while preserving equations. These techniques, rooted in the energy-flow principles, enable efficient representation of large-scale multidomain systems.[3][29]Causality Assignment
Causality assignment in bond graphs determines the direction of computational cause-and-effect relationships along each bond, specifying whether effort or flow is the independent (input) variable for connected elements. This process transforms the acausal power-oriented structure into a directed signal-flow model suitable for simulation and analysis. The causality stroke, a perpendicular bar placed on the bond near one end, indicates the input variable: if the stroke is adjacent to an element, effort is imposed on it (flow is computed as output); if the stroke is at the opposite end, flow is imposed (effort is computed as output).[30][3] Element-specific rules guide preferred causality to favor integral forms over derivatives, promoting numerical stability. Resistors (R) prefer effort causality, where effort is input and flow is output, following the relation e = R f. Inertias (I) prefer flow causality, integrating effort to yield flow as f = \frac{1}{I} \int e \, dt. Capacitors (C) prefer effort causality, integrating flow to yield effort as e = \frac{1}{C} \int f \, dt. Sources have fixed causality: effort sources (Se) output effort (stroke away), while flow sources (Sf) output flow (stroke near).[3][7][31] The standard assignment follows the sequential causal assignment procedure (SCAP), beginning with elements of fixed or preferred causality and propagating constraints through the graph. First, assign causality to sources and propagate to connected bonds via junctions (0-junctions require one effort input, enforcing equal efforts; 1-junctions require one flow input, enforcing equal flows) and transducers (transformers preserve causality direction; gyrators reverse it). Next, assign preferred integral causality to storage elements (I and C), propagating as before. Finally, assign arbitrary causality to resistors, ensuring no bond has conflicting inputs. This iterative propagation continues until all bonds are assigned or conflicts arise.[30][7][3] Conflicts occur when an element receives multiple inputs (overconstrained) or when storage elements receive derivative causality (input opposite to preferred, yielding differentiation instead of integration). Resolution involves model refinement, such as adding small parasitic storage elements to break loops or accepting derivative causality in implicit solvers, while avoiding mixed causality that leads to ill-posed systems. The procedure ensures computational solvability by revealing algebraic loops and producing differential-algebraic equations (DAEs) from acausal models, facilitating efficient simulation without manual equation derivation.[3][7][31]Domain-Specific Conversions
Bond graphs facilitate the translation of classical equations from various physical domains into a unified modeling framework by mapping domain-specific effort and flow variables to standard bond graph elements such as resistors (R), inductors (I), and capacitors (C). This conversion process leverages the power conjugate nature of effort (e) and flow (f), where power p = e × f remains consistent across domains, enabling multidomain system integration without loss of physical insight.[3][32] In the electromagnetic domain, effort is represented by voltage (V) and flow by current (I). Inductors (L) map to I-elements storing magnetic flux linkage, with the relation \dot{\lambda} = V where \lambda = L I; capacitors (C) map to C-elements storing charge, following I = C \dot{V}; and resistors (R) map to R-elements dissipating power via V = R I. Mutual inductance between coils is modeled using a transformer (TF) element with modulus n equal to the turns ratio N1/N2, combined with I-elements representing the self-inductances and a shared magnetizing inductance to capture the coupling effect.[3][33] For linear mechanical systems, translational motion uses force (F) as effort and velocity (v) as flow. Mass (m) corresponds to an I-element, with momentum p = m v and \dot{p} = F; springs (k) map to C-elements storing potential energy, where displacement x satisfies F = k x; and dampers (b) act as R-elements with F = b v. Rotational systems employ torque (T) as effort and angular velocity (\omega) as flow, with analogous mappings: rotational inertia (J) to I, torsional springs to C, and friction to R. Couplings between translational and rotational motion, such as in rack-and-pinion mechanisms, are captured by gyrator (GY) elements with gyration ratio r relating force to torque and velocity to angular velocity, e.g., T = r F and v = r \omega.[3][34][35] In hydraulic systems, effort is pressure (P) and flow is volumetric flow rate (Q). Fluid inertia, arising from mass in pipes, maps to I-elements with I = \rho L / A^2 where \rho is density, L length, and A cross-section, relating \dot{p} = P for momentum p. Compressibility of the fluid or pipe elasticity corresponds to C-elements, with C = V / \beta for bulk modulus \beta and volume V, following Q = C \dot{P}. Thermal systems treat temperature (T) as effort and entropy flow (\dot{S}) or heat flow (\dot{Q}) as flow in pseudo-bond graphs, where thermal capacity maps to C-elements via T = (1/C) \int \dot{Q} dt, and thermal resistance to R-elements with \dot{Q} = (1/R) (T_1 - T_2).[35][33][32] A general mapping of effort-flow pairs and core elements across domains is summarized below:| Domain | Effort (e) | Flow (f) | I-Element Example | C-Element Example | R-Element Example |
|---|---|---|---|---|---|
| Electrical | Voltage (V) | Current (I) | Inductor (L) | Capacitor (C) | Resistor (R) |
| Mechanical (Translational) | Force (F) | Velocity (v) | Mass (m) | Spring (k) | Damper (b) |
| Mechanical (Rotational) | Torque (T) | Angular velocity (\omega) | Inertia (J) | Torsional spring | Friction |
| Hydraulic | Pressure (P) | Volumetric flow (Q) | Fluid inertia (\rho L / A^2) | Compressibility (V / \beta) | Pipe resistance |
| Thermal | Temperature (T) | Heat flow (\dot{Q}) | - | Thermal capacity | Thermal conductance |
| Acoustic | Pressure (P) | Volume velocity (U) | Acoustic mass | Acoustic compliance | Acoustic resistance |