Fact-checked by Grok 2 weeks ago

Extent of reaction

In physical chemistry and chemical engineering, the extent of reaction, denoted by the symbol ξ (xi), is an extensive quantity that quantifies the progress of a chemical reaction by measuring the number of chemical transformations that have occurred, equivalent to the amount of substance transformed as indicated by the reaction equation on a molecular scale divided by the Avogadro constant. It describes the change in the amounts of reactants and products in a system, allowing for the tracking of reaction advancement in both closed and open systems. Mathematically, the change in the number of moles of a species j is given by Δnj = νj ξ, where νj is the stoichiometric coefficient of species j (negative for reactants and positive for products). This concept is distinct from the reaction rate, which measures the speed of change (dξ/dt), whereas ξ itself provides a cumulative measure of how far the reaction has proceeded. The extent of reaction is particularly useful in for analyzing spontaneous processes in closed systems, where it relates to the A = -∑νiμi (with μi as chemical potentials), driving the reaction until is reached when A = 0. In contexts, it facilitates calculations of (XA = ( NA0 - NA ) / NA0), selectivity, and in reactors, often normalized by initial amounts for intensive analysis, and is applied in processes like synthesis or production where reaction progress varies with conditions such as . For multiple simultaneous reactions, a separate extent ξr is defined for each, enabling comprehensive material balance assessments.

Fundamentals

Definition

The extent of reaction, commonly denoted by the symbol ξ, is an extensive quantity in and that describes the progress of a as the number of chemical transformations that have occurred, consistent with the of the reaction equation on a molecular scale. This measure quantifies the degree to which reactants have been converted into products in a , serving as a fundamental indicator of reaction advancement. Also known as the reaction advancement, ξ provides a stoichiometry-consistent way to track compositional changes during the reaction process. For an initial unreacted mixture, the extent of reaction is defined as zero, and it monotonically increases as the reaction proceeds under the influence of , continuing until is achieved or constrained by reactant availability. The maximum attainable value of ξ is governed by the limiting reactant, which is the species present in the smallest stoichiometric proportion relative to the reaction equation, thereby capping the total number of reaction events possible. In contrast to relative measures like percentage completion or fractional conversion, which normalize progress to a scale of 0 to 100% or 0 to 1 based on initial reactant amounts, the extent of reaction operates on an equivalent to moles of reaction occurrences, enabling consistent comparisons across systems regardless of feed composition. Conceptually, ξ accounts for stoichiometric multiplicities by representing the cumulative number of times the balanced has been fully executed, where each increment corresponds to the proportional of reactants and of products as dictated by their coefficients. This approach relates directly to the net changes in the amounts of reacting , forming the basis for stoichiometric analyses in reactive systems.

Historical Development

The concept of the extent of reaction traces its origins to the mid-19th century, when introduced foundational ideas in through in 1840. This law established that the total enthalpy change for a is independent of the pathway taken, providing an early framework for tracking the stoichiometric progress of reactions by quantifying energy changes associated with reaction advancement. In the 1870s, advanced this groundwork in his seminal papers on the equilibrium of heterogeneous substances, where he developed the concept of chemical potentials to describe the of reacting systems. Gibbs' formulation allowed for a more precise analysis of how reactions proceed toward , effectively laying the mathematical basis for quantifying reaction progress as a thermodynamic variable, though without the explicit notation later adopted. The specific concept of the extent of reaction was introduced by the Belgian physical chemist Théophile de Donder in the 1920s, who defined it as a measure of reaction progress in his work on and of irreversible processes. In the 1870s and later, applications in gained momentum in the early 20th century, particularly through applications of , first articulated in 1884 but widely applied in industrial contexts around the 1900s to predict shifts in reaction equilibria under varying conditions. This principle facilitated practical assessments of reaction extent in processes like ammonia synthesis, bridging thermodynamic theory with engineering design for optimizing reaction outcomes. A key milestone occurred in the 1980s with the International Union of Pure and Applied Chemistry (IUPAC) recommendations, which standardized the notation for extent of reaction using the symbol ξ in the first edition of the (Quantities, Units and Symbols in Physical Chemistry, 1988). This formalization ensured consistent usage across , emphasizing ξ as an extensive measuring the number of reaction events. Over time, the extent of reaction evolved from empirical methods of tracking—reliant on product formation or reactant —to a rigorous thermodynamic to modern and equilibrium analysis, enabling precise modeling of reaction pathways in complex systems.

Mathematical Formulation

Single Reaction Systems

In single reaction systems, the extent of reaction, denoted ξ, is an extensive quantity (in units of moles) that measures the progress of a single . For a general reaction written in the form \sum_i \nu_i A_i = 0, where \nu_i is the stoichiometric coefficient for species A_i (negative for reactants and positive for products), the change in the number of moles of species i from initial to current conditions is given by the fundamental equation: \Delta n_i = n_i - n_{i0} = \nu_i \xi This relation was originally formulated by Théophile de Donder as the "degree of advancement" to quantify displacement from equilibrium in thermodynamic systems. The derivation follows directly from stoichiometric proportionality: an infinitesimal advancement d\xi (in moles) produces a change dn_i = \nu_i d\xi in the moles of each species, assuming no other sources of species change. Integrating from the initial state (\xi = 0, n_i = n_{i0}) to the current state yields \Delta n_i = \nu_i \xi. To solve for \xi given initial moles n_{i0} and final composition n_i, rearrange the equation for any species i: \xi = \frac{n_i - n_{i0}}{\nu_i} This value must be consistent across all species involved in the reaction, serving as a check for data accuracy. For example, consider the combustion reaction \ce{C2H6 + 7/2 O2 -> 2 CO2 + 3 H2O}, with initial moles n_{\ce{C2H6},0} = 1, n_{\ce{O2},0} = 5, and no products initially. If final moles are n_{\ce{C2H6}} = 0.2, n_{\ce{O2}} = 2.2, then using : \xi = (0.2 - 1)/(-1) = 0.8; using oxygen: \xi = (2.2 - 5)/(-7/2) = 0.8, confirming the extent. The maximum possible extent, \xi_{\max}, is limited by the reactant that is depleted first and is calculated as \xi_{\max} = \min_i \left( -\frac{n_{i0}}{\nu_i} \right) over all reactants (\nu_i < 0). In the example above, for ethane: -1 / (-1) = 1; for oxygen: -5 / (-7/2) \approx 1.43; thus \xi_{\max} = 1, with oxygen in excess. To find \xi_{\max} step-by-step: (1) balance the reaction to obtain \nu_i; (2) determine initial moles n_{i0} for each reactant; (3) compute -n_{i0}/\nu_i for each; (4) select the minimum value as \xi_{\max}. This identifies the limiting reactant and sets the theoretical upper bound for progress. The fractional extent of reaction, defined as \alpha = \xi / \xi_{\max}, normalizes the progress to the range [0, 1], where \alpha = 0 indicates no reaction and \alpha = 1 complete consumption of the limiting reactant. This metric is particularly useful for comparing reaction efficiency across systems or scaling analyses. In the combustion example, if \xi = 0.8, then \alpha = 0.8 / 1 = 0.8. In the context of phase space, which encompasses the multidimensional space of system compositions and thermodynamic variables, \xi acts as a reaction coordinate parametrizing the stoichiometric manifold—the linear subspace traced by the reaction trajectory. Starting from initial conditions, increasing \xi moves the system state along this one-dimensional path toward equilibrium or complete conversion, with the direction dictated by the vector of stoichiometric coefficients \vec{\nu}. This geometric interpretation facilitates visualization of reaction paths in composition space for single reactions.

Multiple Reaction Systems

In multiple reaction systems, the extent of reaction is extended from a scalar quantity to a vector \boldsymbol{\xi} to account for the progress of each independent reaction simultaneously. This formulation allows for the analysis of complex networks where species may participate in several reactions, enabling the tracking of molar changes across all components without relying on individual conversion variables for each species. The change in the number of moles for each species j is given by the vector equation \Delta \mathbf{n} = \boldsymbol{\nu} \boldsymbol{\xi}, where \boldsymbol{\nu} is the and \boldsymbol{\xi} is the vector of . The \boldsymbol{\nu} has dimensions n_s \times n_r, with rows corresponding to species and columns to reactions; the entry \nu_{j i} is the stoichiometric coefficient of species j in reaction i (negative for reactants, positive for products, zero for non-participants). For a closed system, the initial moles \mathbf{n}_0 relate to final moles by \mathbf{n} = \mathbf{n}_0 + \boldsymbol{\nu} \boldsymbol{\xi}. To solve for the extent vector \boldsymbol{\xi}, the system \Delta \mathbf{n} = \boldsymbol{\nu} \boldsymbol{\xi} is inverted when possible, requiring knowledge of initial and final compositions or additional measurements. For independent reactions where the number of extents equals the rank of \boldsymbol{\nu}, direct inversion applies if the submatrix formed by measured species is square and invertible; otherwise, least squares minimization \boldsymbol{\xi} = (\boldsymbol{\nu}^T \boldsymbol{\nu})^{-1} \boldsymbol{\nu}^T \Delta \mathbf{n} is used for overdetermined cases with experimental error. Reaction independence is determined by the rank of the stoichiometric matrix \boldsymbol{\nu}, which equals the number of independent extents; if the rank is less than the number of proposed reactions n_r, the set is linearly dependent, as one reaction can be expressed as a combination of others. The rank can be computed via row reduction or singular value decomposition, ensuring only a basis of independent reactions is used to avoid redundancy in \boldsymbol{\xi}. Consider a system with two independent reactions sharing a reactant A:
Reaction 1: \ce{A + B -> C}
Reaction 2: \ce{A + D -> E}
The stoichiometric matrix is \boldsymbol{\nu} = \begin{bmatrix} -1 & -1 \\ -1 & 0 \\ 1 & 0 \\ 0 & -1 \\ 0 & 1 \end{bmatrix} with rows for species A, B, C, D, E. Assuming initial amounts n_{A0}, n_{B0}, n_{D0} and no initial C or E, the final moles \Delta n_B = - \xi_1 and \Delta n_D = - \xi_2, the extents as \xi_1 = n_{B0} - n_B and \xi_2 = n_{D0} - n_D. Then, \Delta n_A = -(\xi_1 + \xi_2) confirms consistency, and \Delta n_C = \xi_1, \Delta n_E = \xi_2 provide verification without coupling between \xi_1 and \xi_2. This derivation illustrates how shared species do not necessarily couple extents when unique species (B and D) allow direct solution. For dependent reactions, the is reduced to an independent set by selecting a basis of columns in \boldsymbol{\nu} that span the , discarding redundant reactions to match the . For instance, in the water-gas shift with reactions + H₂O ⇌ ₂ + H₂ and ₂ + H₂ ⇌ + H₂O, the second is the reverse of the first, yielding ; only one extent is needed. This reduction ensures the formulation remains parsimonious and solvable.

Applications and Examples

Batch Reactor Analysis

In batch reactors, a closed system where reactants are charged and the reaction proceeds without material addition or removal, the extent of reaction ξ serves as a fundamental parameter to quantify the reaction progress over time. The temporal evolution of ξ is described by the differential equation \frac{d\xi}{dt} = \frac{V}{\nu_i} r_i, where V is the reactor volume, \nu_i is the stoichiometric coefficient of species i, and r_i is the production rate of species i. This relation links the macroscopic advancement of the reaction (ξ, in moles) to the microscopic reaction rate, enabling the integration of kinetic models to predict system behavior. For constant-volume batch reactors, typically encountered in liquid-phase reactions, the concentration of any species i at time t is calculated as c_i = \frac{n_{i0} + \nu_i \xi}{V}, where n_{i0} is the initial moles of species i. This expression directly ties composition changes to the extent, facilitating the monitoring of reactant depletion and product formation without tracking individual mole balances. Under isothermal conditions, ξ(t) is obtained by integrating the rate law derived from \frac{d\xi}{dt}. For an irreversible first-order reaction A \to products with rate r = k c_A, the analytical solution is \xi(t) = \xi_{\max} (1 - e^{-kt}), where \xi_{\max} is the maximum achievable extent (limited by the initial amount of limiting reactant) and k is the rate constant. This integration highlights how kinetic parameters dictate the time required to reach a specified ξ. The application of ξ in design emphasizes its role in determining optimal holding times and maximizing yields. By solving for the time t to achieve a target ξ via numerical or analytical integration of the design equation t = \int_0^\xi \frac{d\xi'}{V r(\xi')}, engineers can size reactors and select operating conditions to balance productivity and selectivity, particularly for reactions with competing pathways. For instance, in pharmaceutical , ξ-based ensures high yields of desired intermediates by curtailing reaction time at the optimal extent. In gas-phase batch reactions, where the total number of moles may change, the volume V varies with ξ, requiring adjustments to the above formulations. Assuming ideal gas behavior at constant pressure, the volume expands as V = V_0 (1 + \epsilon_A X_A), where \epsilon_A is the fractional volume change upon complete conversion of limiting reactant A and X_A = \xi / \xi_{\max} is the conversion; this modifies the concentration expression to c_i = \frac{n_{i0} + \nu_i \xi}{V_0 (1 + \epsilon_A X_A)} and complicates the integration for ξ(t), often necessitating numerical methods for accurate prediction. Such adjustments are critical for processes like ammonia synthesis in variable-volume setups.

Continuous Flow Reactor Analysis

In continuous flow reactors operating at steady state, such as the (CSTR) and plug flow reactor (PFR), the extent of reaction ξ quantifies the progress of reaction by relating inlet and outlet molar flow rates of species. For a single reaction, the material for species i yields F_i = F_{i0} + \nu_i \xi, where F_i and F_{i0} are the outlet and inlet molar flow rates, respectively, \nu_i is the stoichiometric coefficient, and ξ is the total molar extent of reaction per unit time occurring within the reactor volume V. This formulation holds because, at steady state, the net accumulation is zero, and the difference between inlet and outlet flows equals the extent of reaction adjusted by stoichiometry. The reaction rate r (in terms of extent per unit volume per unit time) relates to ξ via r = \xi / V, allowing solution for ξ once concentrations and kinetics are known from the rate law. For a CSTR, the uniform mixing implies a single value of ξ for the entire , corresponding to the outlet conditions. The steady-state balance simplifies to solving F_i = F_{i0} + \nu_i \xi for ξ, often expressed for a key reactant A as \xi = (F_{A0} - F_A) / (-\nu_A). Substituting the rate law into r = \xi / V enables direct computation of reactor volume or performance for given flows and . This yields a uniform extent profile, where reaction progress is constant throughout the volume, leading to operation at outlet composition conditions. In contrast, for a PFR, the extent ξ varies along the reactor length due to no axial mixing. The differential is dF_i / dV = \nu_i r, or equivalently d\xi / dV = r for a single reaction, where r depends on local . Integrating from (ξ=0 at V=0) to outlet gives the total extent: V = \int_0^\xi \frac{d\xi'}{r(\xi')} This gradual increase in ξ along the reactor results in higher average reaction rates compared to a CSTR for kinetics with positive reaction orders, as concentrations remain closer to values longer. The ξ profile is thus linear in V for zero-order kinetics but nonlinear otherwise, reflecting cumulative progress. The space time τ = V / v_0, where v_0 is the inlet , links reactor size to performance, with ξ expressed as a function of τ and . For example, in an isothermal PFR with first-order (A → products, \nu_A = -1), ξ = F_{A0} (1 - e^{-k τ}), where k is the rate constant; analogous algebraic forms exist for CSTR but require larger τ for equivalent ξ due to backmixing. This relation highlights how τ scales ξ, guiding design for desired . For systems with or , ξ adjustments account for mixed feeds altering effective flows. In a PFR, the ( flow / net product flow) modifies the to F_{i,\text{in}} = (F_{i0} + R F_{i,\text{out}}) / (1 + R), yielding an adjusted balance F_{i,\text{out}} = F_{i,\text{in}} + \nu_i \xi where ξ is computed for the alone; use ratios S to apportion flows similarly, ensuring overall network balances close via global ξ summation. These modifications enable higher conversions or selectivity in complex setups without altering core equations.

Illustrative Calculations

To demonstrate the computation of the extent of reaction for a single reaction system, consider the complete of :
\ce{CH4 + 2O2 -> CO2 + 2H2O}
This balanced equation indicates that 1 of reacts with 2 of oxygen to produce 1 of and 2 of . Suppose an initial mixture contains 1 of CH₄ and 2 of O₂ in a , with no products present initially. After reaction, analysis shows 0.2 of CH₄, 0.4 of O₂, 0.8 of CO₂, and 1.6 of H₂O remaining.
The stoichiometric coefficients are ν_CH₄ = -1, ν_O₂ = -2, ν_CO₂ = +1, and ν_H₂O = +2. Since the initial mixture is stoichiometric (neither reactant is limiting), the extent of reaction ξ can be calculated using any species. For example, using CH₄:
\xi = \frac{n_{\ce{CH4}} - n_{\ce{CH4,0}}}{\nu_{\ce{CH4}}} = \frac{0.2 - 1}{-1} = 0.8 \, \text{mol}
Verification with O₂ yields:
\xi = \frac{0.4 - 2}{-2} = 0.8 \, \text{mol}
and with CO₂:
\xi = \frac{0.8 - 0}{+1} = 0.8 \, \text{mol}
The consistency confirms the value. The change in moles for each species is Δn_i = ν_i ξ.
SpeciesInitial moles (n_{i,0})Final moles (n_i)Stoichiometric coefficient (ν_i)Δn_i = ν_i ξ
CH₄1.00.2-1-0.8
O₂2.00.4-2-1.6
CO₂0.00.8+1+0.8
H₂O0.01.6+2+1.6
For multiple reaction systems, the extent of reaction is computed for each independent reaction using the stoichiometric matrix ν, where columns represent species and rows represent reactions. Consider synthesis with a side reaction, such as the formation of as a minor pathway:
Reaction 1: \ce{N2 + 3H2 -> 2NH3} (ν_1 = [-1, -3, +2, 0])
Reaction 2: \ce{N2 + 2H2 -> N2H4} (ν_2 = [-1, -2, 0, +1])
Initial amounts are 1 mol N₂ and 3 mol H₂, with no products. Outlet analysis reveals 0.4 mol N₂, 1.4 mol H₂, 0.8 mol NH₃, and 0.2 mol N₂H₄. The changes are Δn_N₂ = -0.6 mol, Δn_H₂ = -1.6 mol, Δn_NH₃ = +0.8 mol, Δn_N₂H₄ = +0.2 mol.
The is: -\xi_1 - \xi_2 = -0.6 \\ -3\xi_1 - 2\xi_2 = -1.6 \\ 2\xi_1 = 0.8 \\ \xi_2 = 0.2 \end{cases}$$ From the last [equation](/page/Equation), ξ₂ = 0.2 [mol](/page/Mol). Substituting into the first: -ξ₁ - 0.2 = -0.6 ⇒ ξ₁ = 0.4 [mol](/page/Mol). Verification with the second [equation](/page/Equation): -3(0.4) - 2(0.2) = -1.2 - 0.4 = -1.6 [mol](/page/Mol), and with NH₃: 2(0.4) = 0.8 [mol](/page/Mol). The matrix inversion or solving [linear system](/page/Linear_system) gives the extents directly from Δn = ν^T ξ, where ξ = (ν^T)^{-1} Δn for independent reactions.[](https://cbe255.che.wisc.edu/stoichiometry.pdf) | Species | Initial moles | Final moles | Δn_i | ν_{1,i} | ν_{2,i} | Contribution from ξ_1 | Contribution from ξ_2 | |---------|---------------|-------------|------|---------|---------|------------------------|------------------------| | N₂ | 1.0 | 0.4 | -0.6 | -1 | -1 | -0.4 | -0.2 | | H₂ | 3.0 | 1.4 | -1.6 | -3 | -2 | -1.2 | -0.4 | | NH₃ | 0.0 | 0.8 | +0.8 | +2 | 0 | +0.8 | 0 | | N₂H₄ | 0.0 | 0.2 | +0.2 | 0 | +1 | 0 | +0.2 | The [precision](/page/Precision) of measured compositions directly impacts the accuracy of ξ. For instance, if mole measurements have a relative [uncertainty](/page/Uncertainty) of 1%, the propagated [error](/page/Error) in Δn_i ≈ 1.4% (for two measurements), leading to similar [uncertainty](/page/Uncertainty) in ξ since ξ ≈ Δn_i / |ν_i| for single reactions; in multiple systems, ill-conditioned ν matrices can amplify errors up to 5-10% or more. This underscores the need for high-[precision](/page/Precision) analytical techniques, such as [gas chromatography](/page/Gas_chromatography), in determining extents from experimental data. ## Relations to Other Concepts ### Conversion and Yield In chemical reaction engineering, the fractional conversion $X_i$ of a key reactant $i$ measures the proportion of that reactant that has been consumed relative to its initial amount, defined as $X_i = \frac{n_{i0} - n_i}{n_{i0}} = -\frac{\nu_i \xi}{n_{i0}}$, where $\nu_i$ is the stoichiometric coefficient (negative for reactants), $\xi$ is the extent of reaction, and $n_{i0}$ is the initial molar amount of $i$.[](https://classes.engineering.wustl.edu/2017/fall/eece403/Course%20Notes/Part%20I/Lecture%201%20-%20Introduction%20to%20Chemical%20Reaction%20Engineering.pdf) This formulation links the absolute progress tracked by $\xi$ (in moles) to a dimensionless, normalized [metric](/page/Metric) that focuses on reactant [efficiency](/page/Efficiency).[](https://personalpages.manchester.ac.uk/staff/tom.rodgers/documents/CRE_Notes.pdf) The yield $\phi$ for a desired product $p$ from reactant $r$ quantifies the amount of $p$ actually formed relative to the maximum possible based on the initial amount of $r$, expressed as $\phi = \frac{\nu_p \xi}{-\nu_r n_{r0}}$, where $\nu_p$ (positive for products) and $\nu_r$ (negative for reactants) are stoichiometric coefficients, and $n_{r0}$ is the initial molar amount of $r$.[](https://www.sciencedirect.com/science/article/pii/B9780444627001000085) This metric emphasizes process efficiency in producing the target product while accounting for stoichiometry.[](https://www.sciencedirect.com/topics/engineering/extent-of-reaction) Selectivity $S$ assesses the preference for the desired product over byproducts and is defined as the ratio of yield to the conversion of the reference reactant, $S = \frac{\phi}{X_r}$.[](https://www.sciencedirect.com/science/article/pii/B9780444627001000085) In parallel reactions, where competing pathways convert the same reactant to different products simultaneously, selectivity depends on the relative rate constants of those pathways, often favoring the thermodynamically or kinetically preferred route.[](https://personalpages.manchester.ac.uk/staff/tom.rodgers/documents/CRE_Notes.pdf) In series reactions, involving sequential transformations (e.g., reactant to [intermediate](/page/Intermediate) to final product), selectivity for the [intermediate](/page/Intermediate) is influenced by the rates of formation and subsequent consumption, typically maximized at optimal residence times.[](https://personalpages.manchester.ac.uk/staff/tom.rodgers/documents/CRE_Notes.pdf) The extent of reaction $\xi$ enables the decoupling of stoichiometric coefficients from these normalized metrics, as [conversion](/page/Conversion) and [yield](/page/Yield) remain [invariant](/page/Invariant) under rescaling of the balanced [equation](/page/Equation) (e.g., multiplying all $\nu$ by a constant $k$ scales $\xi$ by $1/k$ but leaves $X_i$ and $\phi$ unchanged).[](https://wwwresearch.sens.buffalo.edu/karetext/unit_01/learning/01_Info.pdf) This independence simplifies [analysis](/page/Analysis) across different stoichiometric representations. ### Stoichiometric Coefficients In [chemical reaction](/page/Chemical_reaction) stoichiometry, the stoichiometric coefficients, denoted as $\nu_i$ for species $i$, quantify the relative number of moles of each species involved in the reaction. These coefficients are signed: negative for reactants, positive for products, and zero for inert species or catalysts that do not undergo net change.[](https://www.sciencedirect.com/topics/engineering/stoichiometric-coefficient)[](https://eng.libretexts.org/Bookshelves/Chemical_Engineering/Foundations_of_Chemical_and_Biological_Engineering_I_%28Verret_Qiao_Barghout%29/02%253A_Reaction_Chemistry/2.01%253A_Definitions_of_Reaction_Rate_and_Extent_of_Reactions) The coefficients are determined by balancing the [chemical equation](/page/Chemical_equation) to ensure conservation of atoms, meaning that for each element, the weighted sum of $\nu_i$ multiplied by the [atomic](/page/Atomic) [composition](/page/Composition) of [species](/page/Species) $i$ [equals](/page/The_Equals) zero. For instance, in the combustion reaction $\ce{H2 + 1/2 O2 -> H2O}$, the balanced coefficients are $\nu_{\ce{H2}} = -1$, $\nu_{\ce{O2}} = -0.5$, and $\nu_{\ce{H2O}} = +1$, satisfying [hydrogen](/page/Hydrogen) and oxygen balances.[](https://www.sciencedirect.com/topics/engineering/stoichiometric-coefficient) Normalization of stoichiometric coefficients is chosen for convenience, often using the smallest possible integers by multiplying all $\nu_i$ by a common factor, or setting the coefficient of a key reactant or product to $-1$ or $+1$ per [mole](/page/Mole) of that [species](/page/Species) or per [unit](/page/Unit) extent. This flexibility aids in simplifying calculations while preserving the reaction ratios.[](https://www.sciencedirect.com/topics/engineering/stoichiometric-coefficient)[](https://cbe255.che.wisc.edu/stoichiometry.pdf) Species in different phases or non-participating catalysts are assigned $\nu_i = 0$, as they act as spectators without altering the stoichiometric matrix. Errors in determining or assigning $\nu_i$, such as failing to balance the equation properly, directly propagate to incorrect values of the extent of reaction $\xi$, since $\xi$ scales with changes in mole numbers divided by these coefficients.[](https://www.sciencedirect.com/topics/engineering/stoichiometric-coefficient)[](https://eng.libretexts.org/Bookshelves/Chemical_Engineering/Foundations_of_Chemical_and_Biological_Engineering_I_%28Verret_Qiao_Barghout%29/02%253A_Reaction_Chemistry/2.01%253A_Definitions_of_Reaction_Rate_and_Extent_of_Reactions) ### Thermodynamic Considerations In thermodynamic [analysis](/page/Analysis) of reacting systems, the extent of [reaction](/page/Reaction) $\xi$ plays a central role in determining the equilibrium state by minimizing the [Gibbs free energy](/page/Gibbs_free_energy) $G$ at constant temperature and pressure. The differential change in [Gibbs free energy](/page/Gibbs_free_energy) is given by $dG = \sum_i \mu_i \, dn_i$, where $\mu_i$ is the [chemical potential](/page/Chemical_potential) of [species](/page/Species) $i$ and $dn_i$ its change in moles. For a [reaction](/page/Reaction) with stoichiometric coefficients $\nu_i$, the mole changes are $dn_i = \nu_i \, d\xi$, leading to $dG = \left( \sum_i \nu_i \mu_i \right) d\xi$. At [equilibrium](/page/Equilibrium), the extremum condition requires $\sum_i \nu_i \mu_i = 0$, which maximizes $\xi$ (or sets it to the equilibrium value) by balancing the chemical potentials.[](https://sites.engineering.ucsb.edu/~jbraw/chemreacfun/ch3/slides-thermo.pdf) This [equilibrium](/page/Equilibrium) condition directly relates to the [equilibrium constant](/page/Equilibrium_constant) $K$, expressed as $K = \prod_i a_i^{\nu_i}$, where $a_i$ are the activities of [species](/page/Species) $i$. The activities depend on $\xi$ through the mole numbers $n_i = n_{i,0} + \nu_i \xi$, with $a_i = \gamma_i (n_i / n_{\text{total}})$ for [ideal](/page/Ideal) solutions (where $\gamma_i$ is the [activity coefficient](/page/Activity_coefficient)). Substituting the chemical potentials $\mu_i = \mu_i^\circ + [RT](/page/RT) \ln a_i$ into the [equilibrium](/page/Equilibrium) [criterion](/page/Criterion) yields $ \Delta G^\circ + [RT](/page/RT) \ln \left( \prod_i a_i^{\nu_i} \right) = 0 $, or equivalently $K = \exp(-\Delta G^\circ / [RT](/page/RT))$, allowing $\xi$ to be solved iteratively from the activity expressions.[](https://sites.engineering.ucsb.edu/~jbraw/chemreacfun/ch3/slides-thermo.pdf) The progress of the reaction also influences the [enthalpy](/page/Enthalpy) balance, with the total [enthalpy](/page/Enthalpy) change $\Delta H = \xi \Delta H_r$, where $\Delta H_r = \sum_i \nu_i \Delta H_{f,i}^\circ$ is the [standard enthalpy of reaction](/page/Standard_enthalpy_of_reaction) based on formation enthalpies $\Delta H_{f,i}^\circ$. This relation quantifies the [heat](/page/Heat) released or absorbed as the [reaction](/page/Reaction) advances, essential for understanding [thermal](/page/Thermal) effects in isothermal or adiabatic processes. For instance, in exothermic reactions, increasing $\xi$ corresponds to greater [heat](/page/Heat) generation proportional to the extent.[](https://www.che.msstate.edu/pdfs/fuel_cell_curriculum/felder/Chapter%209%20-%20Student.pdf) In multiphase systems, the integration of reaction extents into the Gibbs [phase rule](/page/Phase_rule) accounts for chemical equilibria: the [degrees of freedom](/page/Degrees_of_freedom) are $f = c - r - p + 2$, where $c$ is the number of components, $r$ the number of independent reactions (each associated with an independent $\xi$), and $p$ the number of phases. This modification reduces the system's variability by the constraints imposed by the reaction equilibria, linking the possible values of $\xi$ to the independent intensive variables like [temperature](/page/Temperature) and [pressure](/page/Pressure).[](https://louis.uah.edu/cgi/viewcontent.cgi?article=1023&context=research-horizons) Beyond equilibrium, the extent of reaction $\xi$ in [non-equilibrium thermodynamics](/page/Non-equilibrium_thermodynamics) quantifies irreversibility through the reaction [affinity](/page/Affinity) $A = -\sum_i \nu_i \mu_i = -(\partial G / \partial \xi)_{T,P}$, which drives the [reaction rate](/page/Reaction_rate) $d\xi / dt$. The local [entropy production](/page/Entropy_production) rate due to the reaction is $\sigma = (A / T) (d\xi / dt) \geq 0$, reflecting the dissipation of [free energy](/page/Free_energy) into heat and ensuring the second law holds for irreversible processes far from [equilibrium](/page/Equilibrium). This framework applies to systems like coupled biochemical reactions, where overall [affinity](/page/Affinity) ensures positive [entropy](/page/Entropy) generation despite individual steps.[](https://www.sciencedirect.com/topics/engineering/nonequilibrium-thermodynamics) ## Limitations and Extensions ### Assumptions and Validity The extent of reaction, denoted as ξ, relies on several fundamental assumptions to simplify the analysis of chemical systems. Primarily, it assumes a [closed system](/page/Closed_system) where [mass balance](/page/Mass_balance) is maintained without inflows or outflows, enabling the tracking of [species](/page/Species) changes solely through stoichiometric relations. This closed-system premise is essential for [batch reactor](/page/Batch_reactor) scenarios but extends to open systems under steady-state conditions with appropriate [flow](/page/Flow) adjustments. Additionally, the model presumes [constant](/page/Constant) stoichiometric coefficients (ν_ij) throughout the [reaction](/page/Reaction), implying fixed reaction pathways without variation in molecular composition. It also assumes no unspecified side reactions, ensuring that the balanced [equation](/page/Equation) fully captures all transformations. These assumptions hold for ideal, single-reaction systems but require validation in practice.[](https://www.sciencedirect.com/topics/engineering/extent-of-reaction)[](https://pubs.aip.org/aip/cha/article/33/4/043141/2886699/Reaction-extent-or-advancement-of-reaction-A) The validity of the extent of reaction is limited in scenarios where these assumptions break down. For instance, in [polymerization](/page/Polymerization) processes, particularly step-growth mechanisms, the stoichiometric coefficients effectively vary as chain lengths increase, making ξ inadequate for describing the full progress toward high molecular weights without high conversion levels. Non-stoichiometric phases, such as in heterogeneous solid-state [reaction](/page/Reaction)s or multiphase systems with variable compositions, further compromise accuracy, as the model does not account for phase-specific deviations from ideal [stoichiometry](/page/Stoichiometry). Error sources can also distort apparent ξ values; [diffusion](/page/Diffusion) limitations in porous [catalyst](/page/The_Catalyst)s create internal concentration gradients, reducing the observed [reaction](/page/Reaction) progress compared to bulk conditions, while [catalyst](/page/The_Catalyst) deactivation alters effective rates over time, leading to discrepancies between predicted and measured extents. In such cases, alternatives like elemental mass balances are preferred for [complex networks](/page/Complex_Networks), as they rely on conserved atomic species rather than [reaction](/page/Reaction)-specific [stoichiometry](/page/Stoichiometry).[](https://sites.chemengr.ucsb.edu/~ceweb/faculty/doherty/pdfs/molebal.pdf)[](https://learncheme.com/quiz-yourself/interactive-self-study-modules/diffusion-and-reaction-in-porous-catalysts/diffusion-and-reaction-in-porous-catalysts-example-problems/) Experimentally, ξ is validated by inferring it from direct measurements of species concentrations, often using spectroscopic methods like UV-Vis or [infrared spectroscopy](/page/Infrared_spectroscopy) to track [absorbance](/page/Absorbance) changes indicative of reactant depletion or product formation. [Titration](/page/Titration) techniques, such as acid-base or [redox](/page/Redox) titrations, provide quantitative data for reactions involving titratable groups, allowing [calculation](/page/Calculation) of ξ from initial and final concentrations. These methods confirm the model's applicability in simple systems but highlight deviations in non-ideal conditions, where extensions to account for transport or deactivation may be necessary. ### Extensions to Non-Ideal Systems In systems where stoichiometry varies, such as chain reactions or enzyme kinetics, the standard extent of reaction $\xi$ must be modified to account for changing stoichiometric coefficients $\nu_{ij}$. For polymerization reactions, $\xi$ is often redefined as the fractional conversion $p$, representing the proportion of functional groups that have reacted, allowing tracking of molecular weight distribution despite evolving chain lengths and branching. This approach is particularly useful in step-growth polymerization, where the degree of polymerization $P_n$ relates to $p$ via $P_n = \frac{1}{1 - p}$ for the stoichiometric case ($r = 1$), enabling control of polydispersity through staged reaction extents.[](https://chem.libretexts.org/Bookshelves/Organic_Chemistry/Polymer_Chemistry_(Schaller)/03%3A_Kinetics_and_Thermodynamics_of_Polymerization/3.02%3A_Kinetics_of_Step-Growth_Polymerization) In enzyme kinetics, such as Michaelis-Menten mechanisms, $\xi$ can be extended to variable stoichiometry by incorporating substrate inhibition or product activation, where rate constants adjust dynamically, though this requires numerical integration of differential extents over time-dependent $\nu$.[](https://www.sciencedirect.com/science/article/abs/pii/S0040603102000424) For non-ideal reactors, the dispersion model extends $\xi$ profiles in [plug flow](/page/Plug_flow) reactors (PFRs) by incorporating axial mixing via a dispersion [coefficient](/page/Coefficient) $D_{ax}$, which smooths concentration gradients and alters [reaction](/page/Reaction) progress along the reactor length. The axial dispersion number $D/uL$ (where $D$ is dispersion [coefficient](/page/Coefficient), $u$ velocity, $L$ length) quantifies back-mixing; low values ($<0.01$) approximate ideal PFR behavior, while higher values require solving the dispersed [plug flow](/page/Plug_flow) equation $\frac{\partial C_i}{\partial t} + u \frac{\partial C_i}{\partial z} = D_{ax} \frac{\partial^2 C_i}{\partial z^2} + \sum \nu_{ij} r_j$ to compute spatially varying $\xi(z) = \int_0^z \frac{r_j}{\nu_{ij}} dz'$.[](https://classes.engineering.wustl.edu/eece503/Lecture_Notes/Module_3.pdf) This adjustment is critical for tubular reactors with turbulent flow or wall effects, where dispersion reduces selectivity in consecutive reactions compared to ideal models.[](http://umich.edu/~elements/5e/18chap/Fogler_Web_Ch18_final.pdf) In multiphase systems, separate extents $\xi_k$ are defined for each phase $k$, coupled through mass transfer rates to capture interphase dynamics beyond ideal assumptions. For gas-liquid reactions, the moles vector decomposes into extents of reaction $\xi_r$, mass transfer $\xi_{mt}$, inlet/outlet flows, and invariants via linear transformation $T$, yielding $\mathbf{n} = T [\xi_r, \xi_{mt}, \mathbf{n}_{in}, \mathbf{n}_{out}, \mathbf{i}]^T$, where $\xi_{mt}$ links phases with transfer coefficients up to $p_{lg} + p_{gl} \leq N_C$ (number of components).[](https://www.researchgate.net/publication/40741296_Extents_of_Reaction_Mass_Transfer_and_Flow_for_Gas-Liquid_Reaction_Systems) This framework applies to absorption or extraction, as in chlorination of butanoic acid, where mass transfer extents ensure conservation across phases. For distributed multiphase systems like packed beds, extents generalize to include diffusion and advection, transforming concentrations via stoichiometry: $c(\mathbf{x},t) = \nu \xi(\mathbf{x},t) + c_0(\mathbf{x})$, optimizing reactive separation by isolating rate effects.[](https://doi.org/10.1016/j.ces.2017.05.051) Stochastic extensions treat $\xi$ as the cumulative count of reaction firings in microscopic simulations, suitable for small systems where fluctuations dominate. In Monte Carlo methods, such as Gillespie's [stochastic simulation](/page/Stochastic_simulation) [algorithm](/page/Algorithm), $\xi$ emerges as the expected number of reaction [events](/page/2000_in_anime) $\langle \xi \rangle = \int_0^t r(\mathbf{n}(t')) dt'$, with variance from [Poisson](/page/Poisson) statistics, applied to reaction-diffusion in cellular environments.[](https://www.dna.caltech.edu/courses/cs191/paperscs191/gillespie1.pdf) This is [essential](/page/Essential) for non-ideal nanoscale reactors, where deterministic $\xi$ fails [due](/page/A_due) to low [molecule](/page/Molecule) counts, predicting bimodal distributions in chain reactions with up to 50% deviation from mean-field limits.[](https://pubs.aip.org/aip/jcp/article/138/17/170901/1061609/Perspective-Stochastic-algorithms-for-chemical) Software implementations like Aspen Plus and [COMSOL Multiphysics](/page/COMSOL_Multiphysics) incorporate $\xi$ for non-ideal optimization. In Aspen Plus, molar extent is directly specified for stoichiometric reactors (RStoic) or computed via [kinetics](/page/Kinetics) in RPLUG for dispersed PFRs, enabling [sensitivity analysis](/page/Sensitivity_analysis) on $\xi$ to maximize [yield](/page/Yield) in multiphase flows.[](https://www.just.edu.jo/~yahussain/files/reactors.pdf) COMSOL's Chemical Reaction Engineering Module uses $\xi$ implicitly in mass balances for porous multiphase reactors, supporting axial dispersion via Transport of Diluted Species and parameter estimation to optimize $\xi$ profiles against experimental data, as in non-isothermal gas-liquid simulations.[](https://doc.comsol.com/6.0/doc/com.comsol.help.chem/ChemicalReactionEngineeringModuleUsersGuide.pdf)

References

  1. [1]
    extent of reaction (E02283) - IUPAC Gold Book
    Extensive quantity describing the progress of a chemical reaction equal to the number of chemical transformations.
  2. [2]
    Definitions of Reaction Rate and Extent of Reactions
    Extent of Reaction · d n j = change in the number of moles of a certain substance · ν j = the stoichiometric coefficient · d ξ = the extent of reaction.
  3. [3]
    1.14.5: Extent of Reaction
    ### Summary of Extent of Reaction
  4. [4]
    Extent of Reaction - an overview | ScienceDirect Topics
    The extent of reaction is defined as the amount of material processed by a specific reaction at a given position and time, represented mathematically by the ...
  5. [5]
    1.14.5: Extent of Reaction - Chemistry LibreTexts
    May 20, 2024 · The extent of chemical reaction is measured by a quantity d ⁢ ξ , where the chemical composition is described by the symbol ξ . An example makes ...
  6. [6]
    Chemical Equilibrium - David Keffer
    Feb 19, 2001 · You cannot react more than the limiting reactant allows. The maximum value of the extent of reaction is determined by the limiting reactant.<|control11|><|separator|>
  7. [7]
    [PDF] ChE 471 – LECTURE 1 1 Introduction to Chemical Reaction ...
    Limiting reactant is the reactant present in amount less than or equal to ... Mj – molecular weight of j, (kg/ kmol). The relationship between extent of reaction ...
  8. [8]
    Germain Henri Hess | Law of Conservation of Energy, Heat Transfer ...
    Hess's law heralded the formulation of the more complete law of conservation of energy and prepared the way for the development of chemical thermodynamics in ...
  9. [9]
    Josiah Willard Gibbs Issues The Principia of Thermodynamics
    It introduced concepts such as chemical potential, phase rule, and others which form the basis for modern physical chemistry. “In this monumental, densely woven ...
  10. [10]
    Le Chatelier's Principle Fundamentals - Chemistry LibreTexts
    Jan 29, 2023 · Hence, Le Châtelier's principle states that any change to a system at equilibrium will adjust to compensate for that change. In 1884 the French ...
  11. [11]
    [PDF] Quantities, Units and Symbols in Physical Chemistry - iupac
    The first IUPAC Manual of Symbols and Terminology for Physicochemical Quantities and. Units (the Green Book) of which this is the direct successor, was ...
  12. [12]
    Reaction extent or advancement of reaction: A definition for complex ...
    Apr 25, 2023 · The concept of reaction extent (the progress of a reaction, advancement of the reaction, conversion, etc.) was introduced around 100 years ago.
  13. [13]
    [PDF] Celebrating the birth of De Donder's chemical affinity (1922-2022)
    He invented the concept of the degree of advancement of a reaction, also known as extent of reaction, which is a mathematical tool for describing the velocity ...
  14. [14]
    [PDF] 1 Lecture 14 Limiting and Excess Reactants The reactant that would ...
    The reactant with the smallest maximum extent of reaction is the limiting reactant. For example, for the chemical reaction equation: C7H16 + 11O2. 7CO2 + 8H2O.<|control11|><|separator|>
  15. [15]
    Reversible Reactions: Extent of Reaction and Theoretical Yield
    Dec 24, 2019 · Nearly 100 years ago de Donder introduced the term "extent of reaction", ξ. ... extent of reaction is that which has been recommended by IUPAC.
  16. [16]
    [PDF] Stoichioimetry of Chemical Reactions∗
    Sep 1, 2014 · The stoichiometric matrix for a single reaction is a row vector, in this case, ν = h. −2 −1 2 i. 2 Independent Reactions. To motivate the ...
  17. [17]
    [PDF] CHAPTER 2:The Stoichiometry of Reactions—Introduction Chemical ...
    The ith row of the stoichiometric matrix contains the stoichiometric numbers of all species in the ith reaction. Since there is no reason to prefer one ordering ...
  18. [18]
    [PDF] Chemical Reaction Engineering
    ... νi, for the reactant is the same in each case, and this ratio is called the extent of reaction, ξ, ξ = dNi νi. = Ni − Ni,0 νi. (1.3.1). Extent of Reaction. The ...
  19. [19]
    Reactor Design - an overview | ScienceDirect Topics
    2.b Variable volume batch reactors. In general, the equations developed previously assumed constant volume or constant density. For gas-phase reaction such as ...
  20. [20]
    Extended Definition of Conversion and Reaction Extent for a ... - MDPI
    Jan 1, 2024 · The aim of this work is to present in a systematic way a novel general methodology to develop the design equations (heat and mass balances) for networks of ...
  21. [21]
    [PDF] 10.37 HW 3 Spring 2007 Problem 1. A CSTR of volume 0.602 liters ...
    Plugging in the rate expression in the extent of reaction, ξx = rV = k[A][B]V , the steady state A balance becomes: 0 = [A]0 q − [A]q −ξx = ([A]0 − [A])q ...
  22. [22]
    [PDF] Stoichiometry of Chemical Reactions
    Consider as an example the reaction between one methane molecule (CH4) and two diatomic oxygen molecules (O2) to produce one carbon dioxide molecule (CO2) and ...Missing: 2H2O | Show results with:2H2O
  23. [23]
    Recommendations for performing measurements of apparent ... - NIH
    Mar 15, 2023 · ... measure the extent of reaction. If there are any side reactions, one must apply a correction to the measured enthalpy change ΔH for the ΔH ...
  24. [24]
    None
    ### Summary of Definitions and Formulas from Lecture 1
  25. [25]
  26. [26]
    [PDF] A First Course on Kinetics and Reaction Engineering Unit 1 ...
    The extent of reaction is a common extensive reaction progress variable. ... quotient has the smallest value is the limiting reactant. ... the fractional conversion ...
  27. [27]
    Stoichiometric Coefficient - an overview | ScienceDirect Topics
    ... stoichiometric coefficients is a constant, and ξ is called the extent of reaction. Eq. (1.2) can be used for any reactions: (1.3) n i − n i 0 = ν i ξ. We ...<|control11|><|separator|>
  28. [28]
    2.1: Definitions of Reaction Rate and Extent of Reactions
    May 22, 2024 · ... stoichiometric coefficients are positive for products and negative for reactants: ... Extent of Reaction. We use the extent of reaction ( ξ ) to ...
  29. [29]
    [PDF] CHAPTER 3:Review of Chemical Equilibrium — Introduction
    The Gibbs energy is the convenient function for solving reaction equilibrium problems ... reaction extent for reaction i. µj chemical potential of species j.<|control11|><|separator|>
  30. [30]
    None
    ### Summary on Enthalpy of Reaction and Extent of Reaction ξ, and ΔH Relation to ξ
  31. [31]
    [PDF] The Application of the Gibbs Phase Rule and Critical Point ... - LOUIS
    Gibbs Phase Rule Theory​​ where 𝐶 is the total number of chemical species, 𝑃 is the number of coexisting phases, 𝑅 is the number of linear independent chemical ...
  32. [32]
    Nonequilibrium Thermodynamics - an overview | ScienceDirect Topics
    The central function of nonequilibrium thermodynamics is the rate of “irreversible” production of entropy, or the dissipation function which gives the ...
  33. [33]
    [PDF] A systematic method for reaction invariants and mole balances for ...
    feed rate of the limiting reactant. The recycle flow rates for all the ... extent of reaction for reaction r can be defined as: mr = (ni −ni. 0)r wi,r.
  34. [34]
    diffusion-and-reaction-in-porous-catalysts-example-problems
    The first-order reaction A → B takes place in a spherical catalyst pellet. The effectiveness factor due to diffusion limitations in the catalyst pores is 0.80.
  35. [35]
    A reaction-diffusion framework for hydrolytic degradation of ...
    Sep 1, 2023 · The short chain concentration C o ( X , t ) and the extent of reaction ξ ( X , t ) together specify the state of the degrading polymer as a ...
  36. [36]
    Kinetic evaluation of DSC curves for reacting systems with variable ...
    Kinetic evaluation of DSC curves for reacting systems with variable stoichiometric compositions ... extent of reaction, k1 and k2 the rate constants for the ...
  37. [37]
    [PDF] Mixing Effects in Chemical Reactors-III -Dispersion Model
    2. Dispersion occurs in the axial direction. The extent of the dispersion is sufficient to account for the combined effects of all dispersive phenomena ( ...
  38. [38]
    [PDF] Models for Nonideal Reactors
    Oct 9, 2017 · Both of these assumptions are false to some extent in every tubular reactor; frequently, they are sufficiently false to warrant some ...
  39. [39]
    Extents of Reaction, Mass Transfer and Flow for Gas-Liquid ...
    Aug 5, 2025 · For gas-liquid reaction systems with inlet and outlet streams, this paper proposes a linear transformation to decompose the numbers of moles ...
  40. [40]
  41. [41]
    [PDF] Stochastic Simulation of Chemical Kinetics - Caltech
    Abstract. Stochastic chemical kinetics describes the time evolution of a well- stirred chemically reacting system in a way that takes into account.
  42. [42]
    Perspective: Stochastic algorithms for chemical kinetics
    May 1, 2013 · We outline our perspective on stochastic chemical kinetics, paying particular attention to numerical simulation algorithms.
  43. [43]
    [PDF] Reactors - JUST
    The extent of reaction ( ) is used to define the output from the reaction using the following expression: (32). Reactor. nAo. nBo. nCo. nDo. nIo. nA. nB. nC. nD.<|control11|><|separator|>
  44. [44]
    [PDF] The Chemical Reaction Engineering Module User's Guide
    ... Reaction Engineering Module, an optional package that extends the COMSOL Multiphysics. ® modeling environment with customized physics interfaces and ...