Fact-checked by Grok 2 weeks ago

Signal-flow graph

A signal-flow graph (SFG) is a directed graphical representation of a set of linear algebraic equations describing a , where denote variables (such as signals or states) and directed branches (or arcs) represent the linear transformations, gains, or functional relationships between those variables, with incoming signals at a implicitly summed. Although the underlying ideas trace back to Claude Shannon's wartime research on modeling in the 1940s, the formalism was developed and named by Samuel J. Mason, who introduced it in 1953 as a for studying properties in electrical networks and control . Signal-flow graphs offer a compact alternative to traditional block diagrams, eliminating explicit summation and branching symbols by assuming addition at nodes and allowing multiple outgoing paths from a single node without pickoff points. They are widely applied in fields such as control systems engineering, where they facilitate the analysis of dynamic systems like servomechanisms and process control; in for circuit and network analysis; and in for modeling responses and loops. A key advantage is the ability to compute overall system transfer functions directly using , derived by Mason in 1956, which expresses the transmittance between input and output nodes as the ratio of the sum of forward path gains (weighted by their cofactors, or subgraphs excluding touched loops) to the graph determinant (accounting for all independent loops, nontouching loop pairs, and higher-order combinations). This topological approach simplifies the evaluation of complex interconnected systems compared to algebraic manipulation or stepwise block reduction, enabling efficient stability assessments and sensitivity analyses.

Fundamentals

Definition and Overview

A signal-flow graph (SFG) is a specialized type of that represents the flow of signals through a system as a of algebraic equations, where nodes correspond to system variables and directed branches indicate the gains or transfer functions relating those variables. These graphs model how signals propagate from inputs to outputs via operations such as by gains and at nodes, providing a visual framework for understanding without explicit procedural sequencing. The primary purpose of SFGs is to visualize and analyze linear systems in disciplines including , , and , where they illustrate the propagation of signals through networks of components like amplifiers, filters, and loops. Unlike general flowcharts, which depict algorithmic steps or decision processes in software or workflows, SFGs serve as concise mathematical representations focused solely on inter-variable relationships and signal transformations, eschewing any temporal or conditional logic. Originating in the mid-20th century, SFGs were formalized by Samuel J. Mason in 1953 during his work at , where they were developed to simplify the representation and analysis of complex systems by reducing reliance on lengthy equation manipulations. A key advantage of SFGs lies in their intuitive depiction of and paths, which streamlines the identification of signal interactions and facilitates subsequent computations of overall system gains compared to more cumbersome block diagrams.

Basic Components and Terminology

A signal-flow graph consists of nodes and directed branches that represent the flow of signals through a system. Nodes are points that denote variables or signals, such as voltages or wave amplitudes in electrical networks. There are three primary types of nodes: input nodes, also known as sources, which have only outgoing branches and represent independent signal inputs; output nodes, or sinks, which have only incoming branches and denote the final signal outputs; and mixed nodes, which receive multiple incoming signals that are summed and transmit outgoing signals. At mixed nodes, the value is the algebraic sum of all incoming signals. Branches are the directed edges connecting nodes, each associated with a transmittance that scales the signal passing through it. Transmittance, also called branch gain, is a multiplication factor—often a complex number like a scattering parameter or coefficient—indicating how the input signal at the originating node is modified before reaching the receiving node. Self-loops are a special type of branch that originates and terminates at the same node, representing feedback within that node. Key terminology in signal-flow graphs includes the forward , defined as a sequence of branches from a source to a sink without revisiting any . A is a closed of branches that starts and ends at the same , with no intermediate repeated. Loops can be classified as touching if they share at least one common , or non-touching if they do not share any . In graphical notation, nodes are typically depicted as circles or points, while branches are shown as arrows indicating signal direction, with values labeled along the arrows. Signal-flow graphs are generally constructed to model linear time-invariant systems, where relationships between signals do not vary with time unless otherwise specified.

Node and Branch Conventions

In signal-flow graphs, nodes represent system variables, such as signals or states, where the value at a is defined as the sum of all incoming branch signals. This summation convention implies that multiple incoming branches contribute additively to the 's output, with the 's value then serving as the input to all outgoing branches. The output from a along any branch is the 's value multiplied by the branch's transmittance, ensuring consistent signal propagation. Branches in signal-flow graphs are directed edges that indicate the flow of signals from one to another, with arrows specifying the unidirectional transmission path. Multiple branches originating from the same represent parallel signal paths, each scaled independently by their respective transmittances, which quantify the or (e.g., or delay) applied to the signal. Negative on branches account for signal inversion, commonly used to model phase reversals or subtractive operations in systems. Sign conventions in signal-flow graphs employ positive transmittances for direct signal propagation and negative values to denote shifts or inversions, facilitating the representation of both additive and subtractive interactions. Independent variables, typically associated with source nodes that have no incoming branches, initiate signal flows, while dependent variables at other nodes result from the of incoming signals modulated by transmittances. Scaling and normalization in signal-flow graphs are optional but recommended to align with physical domains, such as assigning units of voltage for electrical signals or for mechanical ones, ensuring dimensional consistency across nodes and branches. This practice aids in interpreting graph values relative to real-world quantities without altering the topological structure. Common pitfalls in applying these conventions include using undirected branches, which violate the directed flow requirement and can lead to ambiguous signal paths, or leaving nodes unlabeled, obscuring variable associations and complicating analysis. Adhering strictly to directed, labeled elements maintains the graph's validity as a linear algebraic .

Construction and Properties

Choosing Variables and Graph Construction

In constructing a signal-flow graph, the first step involves selecting appropriate variables that capture the essential dynamics of the , such as input signals, output signals, and variables derived from the 's mathematical description. These variables should be chosen to represent independent quantities, avoiding by eliminating variables that can be directly expressed as linear combinations of others, thereby minimizing the graph's while preserving all necessary relationships. For instance, in a governed by equations, state variables are selected based on their ability to describe the 's evolution over time, ensuring they form a minimal set that fully characterizes the behavior without forward dependencies. This selection is crucial, as different choices of output variables can lead to equivalent but structurally distinct graphs. The construction of the graph proceeds systematically from the system's equations. Begin by writing the algebraic or equations that model the , expressing each dependent as a of independent and other dependent variables; for example, an of the form y = a x + b z relates output y to inputs or states x and z. Next, assign a to each in the equations, where nodes symbolize the signal values at those points in the . Then, for each term in the equations, draw a directed branch from the node of the influencing (cause) to the node of the influenced (effect), labeling the branch with the corresponding coefficient or , such as a from the x-node to the y-node. This process directly translates the equations into a graphical form, with branches indicating signal and their gains representing multiplication factors. The foundational method for this construction was outlined by , who defined the graph by assigning nodes to variables and branches to direct transmissions between them, with branch gains equal to the transmission coefficients. Handling dependent variables requires expressing them iteratively in terms of inputs and prior states, starting from the most basic equations and substituting to resolve interdependencies. For complex systems, this may involve refining the graph through successive substitutions, where dependent nodes are eliminated by combining paths, ensuring all outputs are ultimately traceable to inputs without unresolved cycles in the initial setup. Such refinement maintains the graph's fidelity to the original equations while clarifying signal flows. Effective signal-flow graphs adhere to criteria that promote clarity and analyzability, including minimizing the number of nodes by selecting a parsimonious set of variables and enforcing strict through directional branches that reflect temporal or logical precedence (i.e., no branches pointing to past or simultaneous variables). Graphs should avoid unnecessary nodes or branches, focusing on direct cause-effect links to facilitate subsequent , as overly complex representations can obscure system insights. Node and branch conventions, such as labeling nodes with variable names and branches with gains, are applied consistently during this process to ensure standardization.

Linearity and Causality

In signal-flow graphs, linearity is characterized by branches with constant transmittances, or gains, that do not depend on signal amplitudes, resulting in linear algebraic equations relating node variables. This property ensures that the graph models systems where signals are scaled and summed without distortion, adhering to the : the response to a of inputs equals the corresponding combination of individual responses. Such graphs facilitate straightforward , including path reductions and computations for transfer functions. To verify linearity, all branches must be inspected for constant gains; any dependence on signal values violates this condition. A practical test involves constructing a signal-flow graph for a composite input, such as \alpha_1 x_1 + \alpha_2 x_2, and confirming that its output matches the scaled sum of outputs from separate graphs for x_1 and x_2. While standard signal-flow graphs assume to enable these manipulations, extensions to nonlinear cases introduce variable transmittances that depend on signal levels, complicating reduction techniques and requiring specialized methods for analysis. Causality requires that the directed branches of a signal-flow graph respect temporal order, with signals propagating from earlier to later s without backward links implying dependence on future values. Forward paths thereby represent causal chains from inputs to outputs, modeling systems where effects follow causes in time. In causal graphs, is possible to sequence evaluations, supporting iterative or recursive computations in simulations. Acausal elements, such as branches suggesting future-to-past flow, signal modeling inconsistencies and can lead to non-physical predictions, like non-zero responses before input application in transient analyses. Enforcing involves directing all branches forward and compensating for delays in distributed systems to ensure responses remain zero for negative times.

Non-Uniqueness of Representations

Signal-flow graphs provide a versatile graphical representation of linear systems, but they are inherently non-unique, meaning multiple distinct graphs can equivalently model the same underlying equations and system dynamics. This arises because the construction of a signal-flow graph depends on the choice of intermediate variables at nodes and the arrangement of branches, leading to isomorphic structures that preserve the overall system behavior despite differing topologies. Robichaud et al. emphasize that while the graph encodes the same information as the source equations, no one-to-one mapping exists between the equations and any particular graph form. Graphs are considered equivalent if they yield identical input-output relations, particularly the same from input sources to output sinks, regardless of internal labeling or configurations. Such holds for linear time-invariant systems where selections impact the graph's but not the computed response. Mason's foundational work establishes that topological variations in signal-flow graphs do not alter the fundamental properties or computations when is maintained. Specific transformations illustrate this non-uniqueness, such as merging series branches—where consecutive directed branches between nodes are replaced by a single branch representing their combined effect—or combining parallel branches incoming to the same node, which consolidates multiple paths without loss of . Another example involves rearranging loops, such as eliminating redundant self-loops or converting series-parallel configurations, to yield a structurally different yet equivalent that retains the overall structure. These operations, rooted in the algebraic of linear systems, allow isomorphic rearrangements that simplify while preserving . The absence of a unique representation has practical implications, enabling analysts to select graph forms optimized for simplification and computation, such as reducing complexity for manual or algorithmic processing. No canonical form exists, as the choice of representation can be tailored to the analysis needs without compromising accuracy. This flexibility underscores the graph's utility in system design but requires verification to ensure equivalence. Equivalence can be confirmed by applying to compute the for both graphs; identical results affirm that the representations are interchangeable despite structural differences.

Analysis Methods

Reduction to Sources and Sinks

One key approach to analyzing linear signal-flow graphs is the reduction to sources and sinks, which simplifies the graph by iteratively eliminating intermediate nodes to directly connect input sources (nodes with no incoming branches) to output sinks (nodes with no outgoing branches), thereby deriving the overall between them. This process preserves the graph's , equivalent to solving the associated , and was foundational in the topological analysis of systems as developed by Samuel J. Mason. The method applies to both acyclic and loop-containing graphs, provided holds, allowing for manual or computational simplification of complex structures in and communication systems. The reduction relies on three primary rules for manipulating branches and nodes:
  • Series reduction: Branches connected in series, where an intermediate node has exactly one incoming and one outgoing branch, are combined by multiplying their transmittances. For branches with gains G_1 from node A to intermediate node B and G_2 from B to node C, node B is eliminated, and a single branch with gain G_1 G_2 connects A directly to C. This rule simplifies cascaded paths without altering the signal propagation.
  • Parallel reduction: Multiple branches connecting the same pair of in the same direction are combined by summing their transmittances. If two branches from node A to node B have gains G_1 and G_2, they are replaced by a single branch with gain G_1 + G_2. This handles additive signal contributions at a node.
  • Self-loop elimination: A self-loop at a , representing , is removed by adjusting the incoming and outgoing branches. For a with incoming G from a prior , self-loop L, and outgoing H to a subsequent , the is eliminated, and the prior connects to the subsequent via G \cdot \frac{H}{1 - L}. This accounts for the infinite series of iterations in linear systems.
The step-by-step process begins by identifying eliminable nodes—those with simple structures like single input-output (for series), multiple parallel paths, or isolated self-loops—and applying the appropriate rule to update adjacent s. Nodes are eliminated one at a time, propagating changes through the ; for instance, after series , newly formed branches may enable further parallels or loops to be addressed. This continues until only sources and sinks remain, yielding the net as a of the original gains. Care must be taken to avoid eliminating sources or sinks prematurely, and the order of eliminations can vary due to the non-uniqueness of representations. Implementations range from manual application, suitable for small graphs in engineering design, to algorithmic approaches for larger systems. Manually, engineers visually scan for patterns and redraw the graph iteratively, often prioritizing series reductions to minimize early. Algorithmically, for acyclic graphs, topological ordering processes nodes from sources to sinks (or vice versa), applying series and parallel rules sequentially without loop handling; for graphs with loops, extensions incorporate self-loop resolution before or during elimination, potentially using methods or recursive to ensure convergence. These algorithms, developed for computational efficiency in tools, compute symbolic or numerical transmittances directly. Limitations include restriction to linear graphs, where branch transmittances are constants or functions of a complex variable (e.g., ); nonlinear elements disrupt the multiplicative and additive rules. The method preserves the only between specified sources and sinks, potentially requiring multiple reductions for multi-input/multi-output systems, and may become cumbersome for highly interconnected graphs with many loops, where alternative methods like path enumeration are preferable. As a generic example, consider a signal-flow graph with a connected via a to an intermediate node X, which has a self-loop of l and connects via b to another intermediate node Y; parallel paths from X to Y with p also exist, and Y connects via c to the . First, eliminate the self-loop at X by replacing the path through X with effective a \cdot \frac{b}{1 - l} to Y, then sum the parallel p to form a branch from the to Y with a \cdot \frac{b}{1 - l} + p, and finally apply series reduction to connect the directly to the via (a \cdot \frac{b}{1 - l} + p) \cdot c. This yields the overall without intermediate nodes.

Mason's Gain Formula

Mason's gain formula, also known as Mason's rule, provides a systematic method for determining the of a linear signal-flow graph (SFG) by enumerating forward paths and feedback loops without requiring iterative simplification of the graph. Introduced by Samuel J. Mason in his foundational work on theory, the formula leverages the topological structure of the SFG to compute the overall gain T from input to output as the ratio of the sum of contributions from all forward paths to the graph's factor. This approach is particularly advantageous for complex graphs with multiple loops, as it avoids algebraic manipulation of the underlying system equations and directly incorporates loop interactions through nontouching combinations. The key elements of the formula include forward paths, loop gains, and determinants derived from the graph's topology. A forward path P_k is defined as a continuous directed from the input to the output that does not repeat any , with its gain being the product of the branch transmittances along that . A is a closed that starts and ends at the same without repeating in between, and its gain L_j is the product of the branch gains around the ; loops are considered nontouching if they share no common . The graph \Delta, also called the characteristic polynomial factor, accounts for all loop interactions and is given by: \Delta = 1 - \sum_j L_j + \sum_{i<j} L_i L_j - \sum_{i<j<k} L_i L_j L_k + \cdots where the summation terms alternate in sign based on the number of nontouching loops in each product, with positive signs for even numbers and negative for odd. For each forward path k, the cofactor \Delta_k is the value of \Delta for the subgraph obtained by removing all loops that touch or intersect the k-th forward path, ensuring that \Delta_k isolates the path's independent loop contributions. The overall transfer function T is then expressed as: T = \frac{\sum_k P_k \Delta_k}{\Delta} This formula arises from the expansion of the determinant associated with the SFG's node equations, where the numerator represents the cofactor expansion along forward paths from input to output, and the denominator is the full topological determinant excluding the input source. Mason derived this by generalizing properties of flow graph gains, treating the SFG as a weighted directed graph and applying combinatorial topology to sum over all relevant path and loop configurations, analogous to the permanent and determinant in matrix algebra but adapted for signal propagation. To apply Mason's gain formula, the following steps are followed: (1) Identify and compute the gains P_k for all forward paths from input to output; (2) Determine all individual loop gains L_j and the products of gains for sets of two or more nontouching loops to evaluate \Delta; (3) For each forward path k, compute \Delta_k by recalculating the determinant while excluding loops that share nodes with that path; (4) Substitute into the formula to obtain T. This process ensures a complete accounting of signal amplification and feedback effects without needing to solve the full set of nodal equations. Special cases simplify the computation significantly. If the SFG contains no loops, then \Delta = 1 and each \Delta_k = 1, yielding T = \sum_k P_k, the simple sum of all forward path gains. For a graph with a single forward path and no touching loops, T = P_1 \Delta_1 / \Delta, reducing to the path gain adjusted only by nontouching loops elsewhere in the graph. In systems with unity feedback loops, the loop gain is -1, directly contributing to the summation in \Delta with a negative sign. These cases highlight the formula's efficiency for feedforward or simple feedback structures, contrasting with more exhaustive enumerations in multi-loop scenarios.

Relation to Linear Equations

Signal-flow graphs provide a graphical representation of systems of linear algebraic equations, where each node corresponds to a variable and each directed branch represents a coefficient in the equation relating the variables. In this framework, the value of a node x_i is expressed as x_i = \sum_j g_{ji} x_j + b_i, where g_{ji} is the gain of the branch from node j to node i, and b_i represents any external input to node i. This direct mapping allows the graph's structure to encapsulate the cause-and-effect relationships inherent in the linear system, with branches denoting the multiplicative coefficients that propagate signals between variables. To align with standard analysis methods, such as , the equations are often rewritten in a form suitable for graph reduction: x_i - \sum_j g_{ji} x_j = b_i. This rearrangement highlights the self-dependency and feedback terms, transforming the system into the matrix equation \mathbf{x} = \mathbf{G} \mathbf{x} + \mathbf{b}, or equivalently, (\mathbf{I} - \mathbf{G}) \mathbf{x} = \mathbf{b}, where \mathbf{G} is the gain matrix with entries g_{ji}. The graph's branches thus embody the off-diagonal elements of \mathbf{G}, facilitating visual inspection of the system's interconnections. Solving for the variables involves applying Mason's gain formula to the graph, which computes the transfer function or node values as \mathbf{x} = (\mathbf{I} - \mathbf{G})^{-1} \mathbf{b}. Here, the graph determinant \Delta, defined as \Delta = 1 - \sum P_1 + \sum P_2 - \sum P_3 + \cdots (where P_k are products of gains from sets of k nontouching loops), equals the determinant of \mathbf{I} - \mathbf{G}. The cofactors and path gains in the formula correspond to elements in the inverse matrix, providing an algebraic solution through graphical enumeration rather than direct matrix inversion. The adjacency matrix of the signal-flow graph, weighted by branch gains, directly relates to the system matrix \mathbf{I} - \mathbf{G}, with the graph's topological features—such as loops and paths—mirroring the matrix's eigenvalues and structure. This equivalence ensures that operations like finding the system's overall gain align with matrix determinants and adjugates, confirming the graph as a visual analog to linear algebra. One key advantage of signal-flow graphs over pure matrix representations is their ability to visually identify feedback loops and signal paths, enabling engineers to intuitively assess system stability and dependencies without algebraic manipulation. This graphical insight simplifies the analysis of complex linear systems, particularly in feedback configurations, by highlighting nontouching loops and forward paths that might be obscured in matrix form.

Linear Examples

Simple Amplifier Circuits

Signal-flow graphs offer a clear visual method for modeling basic linear electronic amplifiers, where nodes represent key voltages and branches denote amplification factors, facilitating the computation of overall voltage gain. In a non-feedback voltage amplifier, the signal-flow graph is constructed with two nodes: one for the input voltage v_{in} and one for the output voltage v_{out}. A single directed branch connects v_{in} to v_{out} with transmittance (gain) A, representing the open-loop amplification of the device, such as an operational amplifier without feedback connections. There are no loops or multiple paths in this minimal graph. To verify the transfer function, Mason's gain formula is applied: the forward path gain is A, the loop gain sum is zero, and the determinant \Delta = 1, yielding T = \frac{v_{out}}{v_{in}} = A. For a numerical illustration, if the amplifier has A = 10, then T = 10. The construction emphasizes branches solely for the amplifier gain, bypassing detailed internal circuit elements, while Mason's formula confirms the straightforward path dominance in such open-loop setups. An inverting amplifier configuration incorporates a negative gain branch in its signal-flow graph to capture the phase inversion. The graph features an input node for v_{in} and an output node for v_{out}, linked by a branch with transmittance T = -\frac{R_f}{R_{in}}, where R_f is the feedback resistor and R_{in} is the input resistor, under the assumption of an ideal operational amplifier. This representation treats the effective gain as a direct negative path, applicable after simplification of the underlying circuit equations. Mason's gain formula verifies T = -\frac{R_f}{R_{in}} for the dominant forward path in the reduced graph. Signal-flow graphs in these simple amplifiers highlight gain paths concisely, enabling intuitive understanding of amplification and inversion without complex algebraic manipulation.

Feedback Systems

In signal-flow graphs (SFGs), feedback systems model the interaction between forward amplification and return paths, particularly in ideal negative feedback amplifiers like operational amplifiers (op-amps). A representative example is the series-shunt feedback configuration, where the input signal v_s connects to a node representing the error voltage v_e, which then feeds into the forward path with open-loop gain A, producing the output v_o. The feedback path samples v_o and returns a fraction \beta v_o to subtract from v_s at the error node, forming a single loop. The SFG for this setup consists of three nodes: the source node for v_s, the error node for v_e = v_s - \beta v_o, and the output node for v_o = A v_e. The sole loop has gain L = -A \beta, as the negative sign arises from the inverting feedback. Applying Mason's gain formula, the closed-loop transfer function from input to output is T = \frac{v_o}{v_s} = \frac{A}{1 + A \beta}, where the denominator \Delta = 1 - L = 1 + A \beta accounts for the loop. For large A \beta \gg 1, T \approx 1 / \beta, independent of A. This configuration achieves desensitization, reducing the sensitivity of the closed-loop gain to variations in the open-loop gain A. The relative change in T due to a change in A is \frac{dT / T}{dA / A} = \frac{1}{1 + A \beta}, which is small when A \beta \gg 1, making the system robust to amplifier imperfections. Stability in such SFG-based feedback systems requires the loop gain magnitude |A \beta| < 1 for convergence, ensuring the feedback does not cause oscillation in the ideal DC case; more generally, this condition must hold at frequencies where the phase shift approaches 180 degrees. For instance, with A = 1000 and \beta = 0.1, the closed-loop gain T \approx 1000 / (1 + 100) = 9.9 \approx 10, demonstrating precise control via feedback.

Multi-Loop Servomechanisms

Multi-loop servomechanisms employ cascaded feedback structures to achieve precise control in mechatronic systems, such as position regulation in robotic actuators or machine tools, where an outer position loop generates commands for an inner velocity loop. In signal-flow graph representations of these systems, nodes correspond to key signals including the reference position \theta_r, position error e_p, actual position \theta, velocity command v_c, velocity error e_v, and actual velocity v. Branches represent the transfer functions or gains, such as the position gain K_p from e_p to v_c, the velocity gain K_v from e_v to the plant input, and the plant dynamics G(s) modeling the conversion from velocity command to position output, often approximated as G(s) = 1/s for integrator-like behavior in rotational servos. The inner velocity loop forms a self-loop with gain L_1 = -K_v G_v(s), where G_v(s) captures the velocity dynamics of the plant, while the outer position loop has gain L_2 = -K_p G_p(s), with G_p(s) representing the open-loop transfer function from the velocity command (through the inner velocity gain) to the position output, typically G_p(s) = K_v / s for simple models. These loops touch, sharing common nodes such as the velocity signal. In Mason's gain formula, the denominator includes only the sum of individual loop gains, without products of nontouching loops. Applying Mason's gain formula to this graph yields the overall transfer function from reference to output position: T(s) = \frac{\theta(s)}{\theta_r(s)} = \frac{K_p K_v G(s)}{1 - L_1 - L_2}, where the denominator is the graph determinant \Delta = 1 - L_1 - L_2, accounting for the sum of the individual (touching) loop gains. Signal-flow graphs of multi-loop servos highlight interactions between loops, such as bandwidth limitations where the outer position loop is tuned to 20-40% of the inner velocity loop's crossover frequency to ensure stability and minimize phase lag propagation. This visualization aids in identifying how gain adjustments in the inner loop affect outer-loop performance, facilitating sequential tuning starting from the fastest (innermost) loop. For instance, increasing K_v enhances velocity tracking but may introduce resonances if not damped properly. In applications involving PID control, signal-flow graphs extend the basic proportional gains K_p and K_v by incorporating integral and derivative branches, such as an integrator node for zero steady-state error in velocity and a differentiator for damping in position. This graphical depiction reveals how PID terms interact across loops, enabling designers to visualize and mitigate issues like overshoot in position responses (typically targeted below 10%) or steady-state velocity errors, thereby supporting robust tuning in industrial servomechanisms.

Advanced and Nonlinear Graphs

Nonlinear Branch Functions

In signal-flow graphs, nonlinearity is introduced by allowing branch transmittances to vary as functions of the signal at the originating node, expressed as \tau_{ij} = f(x_j), where x_j represents the value at node j. This extension enables the representation of systems containing nonlinear elements, such as those with memoryless or state-dependent behaviors, while preserving the graphical structure for signal propagation. Common examples of such nonlinear branch functions include the saturation nonlinearity, defined as f(x) = \min(\max(x, -L), L) for some limit L > 0, which models devices like amplifiers or actuators that clip signals beyond a to prevent excessive output. Another is the ideal function, f(x) = x if x > 0 and f(x) = 0 otherwise, capturing one-way conduction in devices. The deadzone nonlinearity, given by f(x) = 0 if |x| < \delta and f(x) = x - \delta \cdot \operatorname{sign}(x) otherwise for a width \delta > 0, represents insensitivity to small inputs, as seen in mechanical relays or hydraulic valves. introduces path-dependent behavior, often modeled as a loop where the output depends on input history, such as in magnetic or ferromagnetic components, complicating the branch function to f(x, s) with s tracking prior values. The presence of nonlinear branches precludes the use of superposition, rendering algebraic methods like inapplicable for exact analysis. Instead, solutions typically require numerical techniques, such as iterative solvers or time-stepping algorithms, to trace signal evolution through the graph. For approximate analysis near a specific , linearization of the function f(x) via Taylor expansion—e.g., f(x) \approx f(x_0) + f'(x_0)(x - x_0)—transforms the branch into a constant , permitting the application of linear signal-flow graph tools.

State Transition and Closed Graphs

State transition signal-flow graphs (SFGs) extend conventional SFGs to represent the of discrete-time systems, where nodes correspond to variables and directed branches depict transition functions of the form x_{k+1} = f(x_k, u_k), with f potentially nonlinear to capture complex behaviors in autonomous or led systems. This graphical structure facilitates the visualization of evolution, enabling the study of interactions among variables through adjacency matrices that encode branch gains as functional dependencies. In nonlinear contexts, such as systems, the graph reveals isomorphic substructures and recurrent patterns, aiding in the identification of system invariants and strategies. Closed flowgraphs represent a specialized variant of SFGs characterized by fully recurrent topologies, lacking sources or sinks, where every node participates in cycles, making them ideal for modeling closed-loop processes without external inputs or outputs. These structures are particularly useful in settings, such as s, where branches denote transition probabilities forming directed cycles, and analysis focuses on stationary distributions derived from graph trees and paths. For instance, in a model, the flowgraph allows computation of long-term state probabilities by enumerating spanning trees, analogous to formulas in deterministic SFGs but adapted for probabilistic gains. Analysis of state transition and closed SFGs often employs eigenvalue methods on the linearized Jacobian for stability assessment in deterministic cases, while simulation traces trajectories in nonlinear regimes to evaluate qualitative behaviors like chaos or convergence. In stochastic closed graphs, spectral properties of the transition matrix, visualized via the flowgraph, determine recurrence and ergodicity. Transitions in these graphs are classified as deterministic, relying on fixed functional mappings, or stochastic, incorporating probability distributions for uncertain evolutions, as seen in Markovian cyclic processes. A representative example is a nonlinear oscillator modeled as a single-node closed SFG with a self-loop branch gain of \sin(x), yielding the discrete update x_{k+1} = \sin(x_k), which exhibits bounded oscillatory trajectories and can be analyzed for fixed points via the graph's recurrent structure. Shannon's formula provides an analytic expression for the of a signal-flow featuring a single input and a single output that touches all loops within the . Under these conditions, the overall gain T simplifies to T = \frac{P}{\Delta}, where P denotes the direct forward path gain from input to output, and the \Delta = 1 - \sum L_k + \sum_{i < j} L_i L_j - \sum_{i < j < m} L_i L_j L_m + \cdots, with L_k representing individual loop gains and higher summations accounting for products of gains from combinations of two or more non-touching loops (with alternating signs). This structure ensures that all feedback paths interact with the primary signal path (making the path cofactor 1), though higher-order terms are included unless no such combinations of three or more non-touching loops exist. The derivation arises as a topological simplification of the general determinant in , where the single-path condition touching all loops eliminates untouched loops from the cofactor. Discovered by Claude E. Shannon in during analysis of in analog solvers, the formula remained unpublished initially but influenced subsequent developments in ; it refers to the general topological gain formula later formalized by . The Happ extension builds on Shannon's approach by addressing graphs with disjoint (non-intersecting) loop sets, enabling more efficient computation for modular structures. Here, the determinant \Delta factors as the product \Delta = \prod_k \Delta_k, with each \Delta_k = 1 - L_k where L_k encapsulates the signed sum of gains (including higher-order terms) within a non-intersecting cluster k. This multiplicative form exploits the independence of clusters, avoiding expansive summations over all possible interactions. Formulated by W.W. Happ in 1966, the extension derives from converting open flowgraphs to closed forms via auxiliary branches and solving topology equations H = \sum L(N) (summing loops of order N) to isolate functions and sensitivities, such as S_Q = \frac{H(Q, P') - H(Q, P)}{H} for parameter Q. It applies to networks with modular loops, such as multi-loop servomechanisms or designs (e.g., Chebyshev or Butterworth types using operational amplifiers), where disjoint enables independent subsystem optimization. Despite their efficiency, both formulas are limited to specific linear topologies and do not apply universally; for arbitrary graphs lacking a touching or disjoint clusters, reverts to the full to account for all and loop interactions. They assume small-signal linearity and ideal components, with performance degrading under nonlinearities or non-ideal elements.

Representations and Applications

Comparison to Block Diagrams

Block diagrams represent linear systems using rectangular blocks to denote operations or subsystems with functions, connected by directed arrows indicating signal , and explicit summing s—typically depicted as circles or points—to add or subtract signals. In contrast, signal-flow graphs (SFGs) employ s to represent signal variables and directed branches with values to connect them, where occurs implicitly at each without dedicated symbols. This fundamental difference means block diagrams require explicit adders for combining signals, whereas SFG s inherently perform this function by aggregating incoming branch contributions. To convert a to an SFG, signals are labeled and assigned to nodes, blocks are replaced by branches carrying their functions, and summing junctions are eliminated by connecting incoming branches directly to the appropriate output node; , for instance, is handled by incorporating a negative sign into the relevant branch . Both representations maintain equivalence in deriving system functions, as block diagram reduction techniques—such as series-parallel simplifications and feedback loop algebra—parallel the path-gain summation and loop elimination in for SFGs, ultimately yielding identical results for linear time-invariant systems. SFGs offer advantages over block diagrams in compactness, particularly for systems with multiple feedback loops, where they facilitate quicker sketching and clearer identification of loops and paths without the clutter of explicit junctions. However, SFGs can be less intuitive for those unfamiliar with , as the implicit nature of nodes may obscure signal combinations compared to the visual explicitness of elements. Historically, SFGs evolved from block diagrams in the mid-20th century as a simplification tool, with Samuel J. Mason's 1953 formulation providing a graphical method to streamline the analysis of interconnected systems previously handled through more cumbersome block reductions.

Use in System Design and Synthesis

Signal-flow graphs (SFGs) play a pivotal role in the of dynamic systems, particularly in , where designers begin by defining primary forward paths that represent the desired signal propagation from input to output. Feedback loops are then incorporated iteratively to enhance , adjust margins, and meet performance criteria such as specified or . This graphical approach allows for modular construction, enabling engineers to evaluate and refine system by adding or modifying branches without algebraic reformulation from scratch. For instance, in multivariable systems, SFGs support systematic of controllers by decomposing complex interactions into manageable subgraphs, facilitating the of gains to achieve responses. In dynamic analysis, SFGs are extensively applied to Laplace-domain representations of linear time-invariant systems, where branch transmittance functions incorporate the complex variable s to model time-domain behavior, such as G(s) = \frac{K}{s + a} for a first-order . Mason's gain formula then computes the overall by summing forward path gains while accounting for loop interactions via the determinant \Delta = 1 - \sum L_i + \sum L_i L_j - \cdots, where L_i are individual loop gains. This enables precise evaluation of system and zeros, essential for assessing transient and steady-state responses. SFGs also integrate seamlessly with classical : loop gains extracted from the graph directly inform root locus plots, tracing closed-loop locations as a parameter (e.g., controller gain) varies from 0 to \infty, aiding in margin optimization. Similarly, forward path and loop contributions facilitate construction by providing asymptotic approximations of and versus , highlighting crossover frequencies and phase shifts for robustness . A practical synthesis example involves augmenting an existing SFG with a branch to meet design specifications. Consider a first-order system with G(s) = \frac{1}{s + 2}; to achieve closed-loop poles at s = -9.2 \pm j 9.13 for less than 5% overshoot and 600 ms , a D(s) = \frac{118}{s + 16.4} is inserted in the forward path, reshaping the locus and as verified through the updated graph's formula. The visual structure of SFGs offers key benefits in pole-zero placement, allowing designers to intuitively observe how added branches shift dominant poles toward desired regions in the s-plane, reducing trial-and-error in cycles compared to purely algebraic methods. This is particularly evident in multi-loop servomechanisms, where path-loop interactions are for coordinated .

Applications Across Disciplines

Signal-flow graphs find extensive application in electronics for circuit analysis, where they graphically represent the linear relationships among voltages, currents, and impedances, enabling the computation of transfer functions via Mason's rule without explicit matrix inversion. This approach is particularly advantageous for complex networks, as it simplifies the identification of loops and paths, reducing computational complexity in analog and digital circuit design. For instance, in operational amplifier circuits and transistor networks, SFGs facilitate rapid evaluation of gain and stability margins. In analysis, signal-flow graphs model interactions between input and output ports using parameters such as (S-parameters) or (ABCD-parameters), allowing straightforward conversions between representations and assessment of network performance under mismatched conditions. This is especially useful in , where SFGs visualize power flow and reflections in transmission lines and amplifiers, aiding in the of high-frequency components like filters and antennas. Control systems leverage signal-flow graphs for design, providing a compact visualization of that highlights criteria through loop gains and nontouching loops. They integrate seamlessly with state-space representations, where variables serve as nodes, enabling the derivation of and matrices for modern controller synthesis, such as in and applications. In , signal-flow graphs depict the architecture of digital filters and algorithms, such as infinite impulse response (IIR) structures, by mapping delay elements, multipliers, and adders to branches and nodes. This representation supports efficient realization of algorithms on hardware like FPGAs, optimizing latency and resource usage in applications including audio processing and communications. For example, transposed forms of SFGs minimize quantization errors in fixed-point implementations. Beyond engineering, signal-flow graphs extend to interdisciplinary domains. In , they model input-output systems akin to Leontief matrices, with sectors as nodes and production flows as branches, facilitating analysis of economic interdependencies and policy impacts. In , linear approximations of signaling pathways use SFGs to trace in cellular networks, predicting response under perturbations. In physics, brief adaptations appear in processing to diagram interactions and measurement flows in quantum circuits. Modern extensions include with simulation software like for virtual prototyping of multidisciplinary systems, and emerging trends in where SFGs visualize forward and backward propagation in neural networks to interpret model decisions.

Historical Development

Origins and Key Contributors

Signal-flow graphs originated in the 1930s and 1940s within the domain of theory, emerging as a tool to graphically depict and interactions in linear systems. This was driven by the need to analyze mechanisms in communication networks and early devices, particularly amid advances in where solving large sets of simultaneous equations manually was cumbersome. The approach built on foundational ideas from circuit analysis, allowing engineers to visualize how signals transformed through networks without relying solely on algebraic manipulation. A key precursor was Claude Shannon's 1937 master's thesis, "A Symbolic Analysis of Relay and Switching Circuits," which introduced methods for representing logical operations in relay-based systems using . Shannon further advanced the concept during his time at Bell Laboratories, inventing the signal-flow graph in a 1942 internal report to model complicated communication systems, motivated by the challenges of wartime and switching networks. This innovation provided a structure where nodes signified variables and branches indicated signal transmissions with associated gains, simplifying the study of multi-loop in telephony. Preceding signal-flow graphs were block diagrams, popularized in the 1940s for analysis; notably, Rudolf Oldenbourg and Hans Sartorius employed them in their 1948 book Die Grundlagen der Automatischen Regeltung (English edition: The Dynamics of Automatic Control, 1949) to represent transfer functions in control loops for . Samuel J. , working at from the mid-1940s, became a pivotal figure by extending Shannon's ideas into a comprehensive . In his 1956 , Mason introduced the topological gain formula for computing overall system gains from flow graphs, addressing the limitations of equation-based methods in large-scale network design. A major milestone came with Mason's 1956 paper, which formalized signal-flow graphs as a standard tool for analysis, emphasizing their utility in by detailing properties like path independence and rules. These efforts collectively shifted focus from rigid algebraic solving to intuitive graphical , enabling efficient handling of in telephony and emerging servomechanisms.

Evolution and Standards

Following the initial development of linear signal-flow graphs in the mid-20th century, significant extensions occurred in the to accommodate nonlinear systems, enabling the representation of more complex dynamics through nonlinear branch functions. In the , signal-flow graphs were increasingly linked to state-space representations, facilitating the formulation of systems in terms and improving computational of dynamic systems. By the , with software tools emerged, allowing automated and simulation of signal-flow graphs for design iteration and system modeling in engineering applications. Standardization efforts formalized signal-flow graph notation within broader diagramming practices. Complementary international standards, like those in ISO/IEC 24765 for systems and vocabulary, support consistent modeling of signal flows in complex systems, though they emphasize conceptual frameworks over specific graph notations. Terminology for signal-flow graphs evolved to classify them as linear versus nonlinear; additionally, graphs are distinguished as open (acyclic, without loops) or closed (with cycles enabling recurrent signal paths). In the and , modern extensions include signal-flow graphs for cyber-physical systems, combining and continuous to model interactions between computational and physical components. Quantum signal-flow graphs have also advanced in , representing non-Markovian feedback networks for quantum optical devices and enabling analysis of coherent signal propagation in quantum feedback loops. Despite these developments, gaps persist in standardization for nonlinear signal-flow graphs, particularly in handling multiple equilibria and elements, which recent literature addresses through nonlinear model techniques to enhance and consistency.

Modern Extensions

In the early , software tools have enabled automated generation, , and of signal-flow graphs, enhancing their utility in engineering . The File Exchange provides "Signal_Flow_Graphz," a for symbolic reduction of diagrams and signal-flow graphs, facilitating efficient computation of s and system responses since its release in 2008. Similarly, supports implementation of signal-flow graphs for and , as demonstrated in educational labs and design workflows post-2000, allowing seamless integration with modeling. In , the open-source "sfg" library offers functions for constructing, calculating gains, and plotting signal-flow graphs using for directed graph , promoting accessibility in computational environments. Interdisciplinary applications have extended signal-flow graph concepts to and bioinformatics. In graph neural networks, flow-attentional mechanisms inspired by signal propagation in s improve and performance on and tasks, as shown in models developed in the 2020s. For gene regulatory networks in bioinformatics, structures analogous to signal-flow graphs model regulatory interactions, with methods like convolutional networks inferring dependencies from data to reconstruct network topologies. Emerging developments include extensions for probabilistic systems and adaptations for applications. Stochastic flow diagrams generalize signal-flow graphs by incorporating weighted directed edges to represent uncertainties in dynamic systems, enabling probabilistic analysis of complex processes as introduced in 2014. signal-flow graph implementations in systems leverage for synthesizing algorithms, supporting deployment on hardware with timing constraints for applications like digital filters. Advances address challenges in large-scale graphs through AI-driven optimization, while quantum extensions explore . AI integration with graph enhances scalability and interpretability for large networks, such as in systems where explainable models optimize . In quantum , recent 2023–2024 research on fusing photonic graph states develops flow-based protocols for reliable entanglement generation, extending signal-flow principles to architectures. Looking ahead, AI-driven system holds potential for automating signal-flow graph-based designs in networks and structures, leveraging algebraic representations for conceptual innovation.

References

  1. [1]
    Feedback Theory-Some Properties of Signal Flow Graphs
    **Summary of Signal-Flow Graph Construction from IEEE Paper (DOI: 10.1109/TCT.1956.1086302)**
  2. [2]
    [PDF] Signal Flow Graphs: Props, Presentations, and Proofs
    of signal flow graphs, due to Claude Shannon in the 1940s. Here is an ... our simplified signal flow graphs from Definition 5.45 because now we can also use the.Missing: origin | Show results with:origin
  3. [3]
    [PDF] Feedback Theory-Further Properties of Signal Flow Graphs
    Mason: Further Properties of Signal Flow Graphs closed loops and which also do not contain any paths from node 1 to ground or from node 2 to ground. Write ...
  4. [4]
    3.2 Signal Flow Graphs – Introduction to Control Systems
    Signal flow graphs are much more succinct and only include nodes for signals and branches for gains. Any signals entering a node are assumed to be added or ...
  5. [5]
    3.2: Signal Flow Graph - Engineering LibreTexts
    May 22, 2022 · SFGs are used in many disciplines, but they are particularly useful with RF and microwave circuits.
  6. [6]
    Analysis of Electronic Circuits with the Signal Flow Graph Method
    A signal flow graph (SFG) describes a system's signal flow using directed, weighted graphs, where signals are applied to nodes and functions on edges. It helps ...
  7. [7]
  8. [8]
    [PDF] 4.5 – Signal Flow Graphs
    Mar 19, 2009 · Signal flow graphs represent microwave networks, consisting of nodes connected by branches with values, and help understand network behavior.
  9. [9]
    Signal Flow Graphs - an overview | ScienceDirect Topics
    A signal flow graph (SFG) is defined as a representation of the interacting relationships within linear time invariant discrete time systems, illustrating how ...
  10. [10]
    [PDF] Signal Flow Graphs IIR Filter Structures - MIT OpenCourseWare
    In a signal flow graph, the value carried by a specific branch is equal to the value of its originating node. Nodes in signal flow graphs represent variables.
  11. [11]
    None
    ### Summary of Node and Branch Conventions in Signal Flow Graphs
  12. [12]
    Chapter 2 Construction and Analysis of Signal Flow Graphs
    This chapter focuses on the constructions and analyses of signal-flow graphs (SFGs) and discusses SFG reduction and solution of equations.
  13. [13]
    Signal Flow Graphs and Applications - Google Books
    Author, Louis P. A. Robichaud ; Publisher, Prentice-Hall, 1962 ; Original from, the University of Wisconsin - Madison ; Digitized, Jul 30, 2008 ; Length, 214 pages.
  14. [14]
    None
    ### Four Important Rules for Decomposing Signal Flow Graphs
  15. [15]
    An algorithm for automatic reduction of complex signal flow graphs
    Dec 1, 1976 · A computer algorithm is developed that provides efficient means to compute transmittances directly from a signal flow graph or a block ...
  16. [16]
    [PDF] MASON'S GAIN FORMULA EE371 TECHNICAL NOTE
    Signal-flow diagrams are composed of nodes and arcs. Summation junctions and pickoff points become nodes in the graph. Signals are branches connecting the ...
  17. [17]
    Analog circuit analysis simplified by K9 using a signal flow graph
    May 7, 2014 · Author's note: This article shows how to derive circuit gain equations via a signal flow graph. With practice you may be able to write a ...
  18. [18]
    [PDF] ECE145a / 218a Signal Flow Graphs g p - UCSB ECE
    represents visually and compactly graph flow signal. The system. the describing equations many the. Page 5. class notes, M. Rodwell, copyrighted 2009.
  19. [19]
    [PDF] Feedback Amplifiers - Marshall Leach
    ... signal-flow graph. A signal-flow ... Because the current fed back through RF to the input node combines in parallel with the source current, the circuit is.
  20. [20]
    [PDF] Inversion of Nonlinear and Time-Varying Systems
    We present a corollary to Theorem 2 that addresses this concern, facilitating the inversion of systems realized as signal flow graphs, without requiring repre-.
  21. [21]
    [PDF] Nonlinear Systems
    Figure 2.15: Hysteresis characteristics of the tunnel-diode circuit. ... limiter or saturation nonlinearity (a typical nonlinearity in feedback control systems ...
  22. [22]
    Signal flow graph and control of realizable autonomous nonlinear ...
    A signal flow graph and an electronic circuit implementation are designed, analyzed, and constructed to realize the novel Chen system. The control problem of ...<|control11|><|separator|>
  23. [23]
  24. [24]
    [PDF] 19700004922.pdf - NASA Technical Reports Server (NTRS)
    W.W. Happ generalized the. Shannon. formula to include flowgraphs and went on further to develop flowgraph sensitivity algorithms. Flowgraphs are perhaps best ...
  25. [25]
    [PDF] Reduction of Multiple Subsystems - Aerostudents
    Mason, S. J. Feedback Theory—Some Properties of Signal-Flow Graphs. Proc. IRE,. September 1953, pp. 1144-1156. Milhorn, H. T, Jr. The Application of Control ...Missing: construction | Show results with:construction
  26. [26]
    [PDF] On Teaching the Simplification of Block Diagrams* - IJEE
    Ogata, Modern Control Engineering, 3rd edition, Prentice-Hall, New Jersey ... The signal flow graph (SFG) developed by Mason. [5±6], which is based on a ...
  27. [27]
    Multivariable controller design by signal-flow graph techniques
    Insufficient relevant content. The provided URL (https://ieeexplore.ieee.org/document/1105616/) does not display accessible content for extraction and summarization due to access restrictions or lack of visible text in the given snippet. No method, key steps, or benefits for multivariable controller synthesis using signal-flow graphs can be extracted.
  28. [28]
    [PDF] ROOT-LOCUS METHOD FOR COMPENSATING AND ...
    The signal flow graphs can readily be utilized in control theory, in this case a transfer function can be assigned to every hranch. The relation between any two ...
  29. [29]
    [PDF] Signal Flow Graph In Loop Gain Analysis Of DC-DC PWM CCM ...
    Mason, “Feedback theory—Some properties of signal flow graphs,”. Proc. IRE, vol. 41, pp. 1144–1156, 1953. [11] R. Middlebrook, “Measurement of loop gain in ...
  30. [30]
    Analysis of Electronic Circuits with the Signal Flow Graph Method
    Aug 10, 2025 · This article shows how to use network flow diagrams to arrive at the equations which must be solved to perform a full 1-port VNA calibration.
  31. [31]
    [PDF] Analysis of Electronic Circuits with the Signal Flow Graph Method
    There are various methods in the circuit technology to calculate trans- fer functions of electrical circuits such as Kirchhoff's laws, two-port net- work theory ...
  32. [32]
    The Application of Signal Flow Graph in Conversion of Two-port ...
    The relationship of the four voltages and currents in two-pore networks can be described using six different kinds of parameters and equations using signal ...
  33. [33]
    Signal Flow Graphs and Data Flow Graphs - ResearchGate
    Signal flow graphs and data flow graphs are powerful representations of signal processing algorithms ... DSP systems in image processing and radar, sonar ...
  34. [34]
    SIGNAL-FLOW GRAPHS | McGraw-Hill Education
    BLOCK DIAGRAM REDUCTION USING SIGNAL FLOW GRAPHS AND THE GENERAL INPUT-OUTPUT GAIN FORMULA · Cover of Schaum's Outline of Feedback and Control Systems. Chapter ...Missing: Nise | Show results with:Nise<|control11|><|separator|>
  35. [35]
    The linear framework: using graph theory to reveal the algebra and ...
    The linear framework uses finite, directed graphs with labelled edges to model biomolecular systems. Graph vertices represent biochemical species or molecular ...Missing: flow signaling
  36. [36]
    AI for Signal Processing - MATLAB & Simulink - MathWorks
    Use deep learning for signal processing applications with MATLAB such as extracting features, transforming signals, building models, and more.
  37. [37]
    [PDF] Control: A perspective - Maths Homepage
    Oldenburger and Sartorius of Siemens showed that concepts and methods from servomechanism theory were useful for process control (Oldenbourg & Sartorius, 1944).
  38. [38]
    [PDF] Abstractions of Graph Models - BYU ScholarsArchive
    Jun 4, 2020 · sink or source node missing in its abstraction. Table 1. A table ... assume that we abstract away the last input node. So WR(s) has the ...
  39. [39]
    The Calculus of Signal Flow Diagrams I: Linear relations on streams
    A certain class of diagrams captures the orthodox notion of signal flow graph used in control theory; we show that any diagram of our syntax can be realised ...
  40. [40]
    [PDF] PROCEEDINGS OF THE IRE® - World Radio History
    ... H. D. Crane 7 and. W. F. Kosonocky worked in the later stages and con ... Signal Flow Graph and. How to Avoid Them. S. J. MASON, Massachusetts. Institute ...
  41. [41]
    Feedback Theory-Further Properties of Signal Flow Graphs
    Feedback Theory-Further Properties of Signal Flow Graphs · S. J. Mason · Published in Proceedings of the IRE 1 July 1956 · Engineering, Computer Science.
  42. [42]
    The use of matrix signal-flow graph in state-space formulation of ...
    Presents a unified summary on the formulation of the general feedback theory in terms of the coefficient matrices of the state equations.
  43. [43]
    [PDF] Generalised Models of Design Iteration Using Signal Flow Graphs
    In this paper, signal flow graphs are presented as a flexible tool for design process modeling, and illustrated using an industrial example. Analysis of the.
  44. [44]
    [PDF] Analysis and Design of Cyber-Physical Systems: A Hybrid Control ...
    A model of the system performing the conversion of analog signals into digital signals is first proposed. This model uses a timer to trigger the sampling events.Missing: 2020s | Show results with:2020s
  45. [45]
    [PDF] Signal Flows in Non-Markovian Linear Quantum Feedback Networks
    Dec 17, 2014 · By contrast, a signal flow graph uses nodes to represent signals, while the system is represented by a directed arc. The directed arcs indicate ...
  46. [46]
    (PDF) Nonlinear model standardization for the analysis and design ...
    Apr 22, 2021 · This transformation is, in the present study, referred to as nonlinear model standardization, which is the extension of the well-known concept ...Missing: 2020s | Show results with:2020s<|control11|><|separator|>
  47. [47]
  48. [48]
    [PDF] ECE-205 Lab 6 Transfer Functions, Signal Flow Graphs, and ...
    In this lab we will first learn how to implement transfer function in both Matlab and in Simulink. Next we show how to write block diagrams as signal flow ...
  49. [49]
    sfg - Signal flow graph calculation and plotting in Python - GitHub
    Create a plot of the SFG by calling the plot() function. The function will return the graph as a graphviz.Digraph object, which you can save or display.
  50. [50]
    [2506.06127] Flow-Attentional Graph Neural Networks - arXiv
    Jun 6, 2025 · We demonstrate that flow attention enhances the performance of attention-based GNNs on both graph-level classification and regression tasks.
  51. [51]
    Inference of gene regulatory networks based on directed graph ...
    Jun 27, 2024 · In this study, we propose Directed Graph Convolutional neural network-based method for GRN inference (DGCGRN).
  52. [52]
    (PDF) Stochastic flow diagrams - ResearchGate
    Aug 9, 2025 · We introduce Stochastic Flow Diagrams (SFDs), a new mathematical approach to represent complex dynamic systems into a single weighted ...
  53. [53]
    [PDF] Designing and Implementing Real-Time Signal Processing Systems
    This hands-on, three-day course focuses on developing models in the. Simulink® environment to deploy on embedded systems. The course is designed for Simulink ...
  54. [54]
    Enhancing explainable AI with graph signal processing
    This paper presents a novel framework integrating Explainable AI (XAI) with graph signal processing to enhance the interpretability of AI models applied to WDS.Missing: standardization 2020s
  55. [55]
    Fusion of deterministically generated photonic graph states - Nature
    May 8, 2024 · Ring 25 and tree 26 graph states with up to eight qubits, with the names reflecting the entanglement topology, are efficiently fused from the photonic states ...Missing: flow | Show results with:flow
  56. [56]
    [PDF] A NEW APPROACH TO CONCEPTUAL DESIGN SYNTHESIS OF ...
    Jan 14, 2015 · It incorporates signal flow graph algebra to support quantification. The model has been used to capture physical effects and populate a database ...