Fact-checked by Grok 2 weeks ago

Cottrell equation

The Cottrell equation is a foundational relation in electrochemistry that quantifies the time-dependent current decay in diffusion-controlled processes following a sudden change in electrode potential, such as in chronoamperometry experiments._Chronoamperometry) Derived by American chemist Frederick Gardner Cottrell in his 1903 doctoral dissertation, it models the current arising from the diffusion of electroactive species to an electrode surface under semi-infinite linear diffusion conditions, assuming no convection or other mass transport mechanisms. The equation emerges from solving Fick's second law of diffusion with boundary conditions that establish a concentration gradient at the electrode interface upon applying the potential step._Chronoamperometry) In its standard form for a reduction process at a planar electrode, the Cottrell equation is expressed as: i(t) = n F A C \sqrt{\frac{D}{\pi t}} where i(t) is the current at time t, n is the number of electrons transferred per molecule, F is Faraday's constant ($96{,}485 C/mol), A is the electrode area, C is the bulk concentration of the electroactive species, D is its diffusion coefficient, and t is the elapsed time since the potential step._Chronoamperometry) This inverse square-root dependence on time (i \propto 1/\sqrt{t}) reflects the expanding diffusion layer thickness over time, which progressively limits the flux of species to the electrode._Chronoamperometry) Cottrell's original analysis, published in German as Der Reststrom bei galvanischer Polarisation, betrachtet als ein Diffusionsproblem ("The residual current in galvanic polarization, considered as a diffusion problem"), addressed the residual current after polarization and laid the groundwork for understanding unsteady-state diffusion in electrolytic systems. The equation's significance extends to practical electrochemical analysis, enabling the determination of key parameters such as diffusion coefficients, electrode areas, and reaction stoichiometries from experimental current-time transients._Chronoamperometry) It underpins techniques like chronoamperometry and chronocoulometry, which are widely used in studying redox kinetics, sensor development, and material characterization, particularly for species exhibiting reversible electron transfer._Chronoamperometry) Extensions of the Cottrell model account for complications like spherical diffusion at microelectrodes or coupled chemical reactions, but the original form remains a cornerstone for validating diffusion-dominated behavior in voltammetric methods._Chronoamperometry)

History

Origins and Development

Frederick Gardner Cottrell (1877–1948) was an American physical chemist and inventor, best known for developing the electrostatic precipitator in 1906, a device that uses high-voltage electric fields to remove fine particles from industrial exhaust gases, significantly advancing air pollution control. Born in Oakland, California, Cottrell earned a B.S. in chemistry from the University of California, Berkeley in 1896 and completed his Ph.D. at the University of Leipzig in 1902, focusing on metallurgical chemistry before returning to Berkeley as an instructor in physical chemistry. His early career emphasized electrolytic processes and diffusion phenomena in solutions, blending theoretical analysis with practical experimentation to address industrial challenges in metallurgy and electrochemistry. In 1903, Cottrell derived the equation that now bears his name while investigating residual currents during galvanic polarization of electrodes, treating the phenomenon as a diffusion problem. His experiments involved applying a constant potential to electrodes in electrolyte solutions, observing the decay of current over time due to the depletion of electroactive species near the electrode surface and their replenishment by diffusion from the bulk solution. This work stemmed from his studies on electrolytic deposition and dissolution rates, particularly in systems like metal ions in acidic media, where he noted the characteristic inverse square root time dependence of the current, providing a quantitative link between electrochemical response and diffusion kinetics. The derivation appeared in his paper "Der Reststrom bei galvanischer Polarisation, betrachtet als ein Diffusionsproblem" ("The Residual Current in Galvanic Polarization, Considered as a Diffusion Problem"), published in Zeitschrift für physikalische Chemie, volume 42, pages 385–431. Cottrell's equation received early validation through subsequent experiments by other researchers. These confirmations helped establish the equation's reliability in describing transient electrochemical behavior under diffusion limitations.

Impact on Electrochemistry

The Cottrell equation, derived in 1903, was swiftly adopted in early 20th-century electrochemistry, particularly influencing the theoretical underpinnings of polarography developed by Jaroslav Heyrovsky in the 1920s. Heyrovsky's invention of polarography in 1922 relied on empirical observations of diffusion-limited currents at the dropping mercury electrode, but subsequent theoretical advancements integrated the Cottrell model to explain these phenomena quantitatively. In 1934, Miroslav Ilkovic extended the Cottrell equation to derive the Ilkovic equation, adapting it to account for the growing spherical diffusion layer at the mercury drop, which provided a rigorous basis for interpreting polarographic waves and enabled precise quantitative analysis in voltammetry. This integration marked the Cottrell equation as a cornerstone of electrochemical theory, formalizing the time-dependent nature of diffusion-limited currents and driving a paradigm shift from purely empirical electroanalysis to predictive, mechanism-based models. By quantifying how current decays inversely with the square root of time due to mass transport limitations, it allowed researchers to distinguish kinetic from diffusional control, foundational for advancing voltammetric techniques beyond qualitative observations. This theoretical framework influenced the standardization of electroanalytical methods, enabling reproducible measurements of analyte concentrations and reaction rates in diverse systems. In the 1940s and 1950s, the equation fueled a revival in electrochemistry alongside modern instrumentation, such as automated polarographs and controlled-potential devices, which facilitated precise chronoamperometric experiments aligned with Cottrell predictions. Pioneers like James J. Lingane and Izaak M. Kolthoff leveraged it to refine polarographic applications and develop complementary techniques like chronopotentiometry, expanding electroanalysis into complex matrices including biological and industrial samples. Cottrell's broader contributions earned him significant recognition, including the founding of the Research Corporation in 1912, a nonprofit funded by royalties from his inventions like the electrostatic precipitator, which supported scientific advancement and indirectly amplified the impact of his electrochemical work. The long-term legacy of the Cottrell equation is evident in its citation across thousands of papers, reflecting its status as a bedrock for mass transport analysis in electrochemistry. Recent applications include 2023 studies on CO2 electroreduction, where researchers at Cornell University applied the equation to decouple mass transport from catalytic effects, enabling control over product selectivity toward valuable chemicals like ethylene and ethanol during pulsed electrolysis. This enduring utility underscores its role in addressing contemporary challenges in sustainable energy conversion.

Theoretical Background

Diffusion-Controlled Processes

In electrochemistry, diffusion-controlled processes occur when the rate of an electrochemical reaction is limited by the mass transport of reactants to the electrode surface primarily through diffusion, rather than by the intrinsic kinetics of electron transfer or by migration of charged species. This regime arises in unstirred solutions where convection is negligible and migration is minimized, such as through the use of supporting electrolytes, leaving diffusion as the dominant mechanism for delivering species to the reactive interface. Under these conditions, the current is proportional to the flux of diffusing species, reflecting the rate at which molecules can reach the electrode to participate in the reaction. A key distinction exists between diffusion-controlled and kinetically controlled processes. In kinetically controlled reactions, the electron transfer step is rate-limiting, with mass transport occurring faster than the reaction itself, leading to currents that depend on the activation energy and follow exponential relationships with potential. Conversely, diffusion control prevails when a sufficiently large overpotential is applied, rendering the electron transfer rate constant extremely high and effectively setting the reactant concentration at the electrode surface to zero; here, the current is independent of further increases in overpotential and is instead constrained by how quickly species diffuse from the bulk solution. The transport is governed by Fick's laws of diffusion, which describe the flux as proportional to the concentration gradient. Potential steps in techniques like chronoamperometry initiate these processes by suddenly altering the electrode potential, causing rapid depletion of the reactant near the surface and establishing a concentration gradient that penetrates into the solution. This gradient forms a diffusion layer—a region of varying concentration whose thickness increases with the square root of time, starting thin and growing as more reactant is consumed without replenishment from bulk diffusion. The semi-infinite linear diffusion geometry, assumed in Cottrell scenarios, models this for planar macroelectrodes, where diffusion occurs unidimensionally perpendicular to the flat surface, valid when the layer thickness remains much smaller than the electrode dimensions. In contrast, spherical diffusion applies to microelectrodes or curved geometries, involving radial transport that allows steady-state conditions due to enhanced flux from all directions.

Fick's Laws in Electrochemistry

Fick's first law of diffusion states that the diffusive flux J of a species is proportional to the negative gradient of its concentration, expressed mathematically as J = -D \frac{\partial c}{\partial x}, where D is the diffusion coefficient, c is the concentration, and x is the spatial coordinate perpendicular to the diffusion direction. This law quantifies the rate at which molecules move from regions of higher to lower concentration due to random thermal motion, assuming no bulk convection or migration effects. In electrochemical systems, it directly relates the flux of electroactive species to the electrode surface, which determines the faradaic current when the reaction is diffusion-limited. Fick's second law extends this to unsteady-state diffusion, describing how concentration profiles evolve over time. It is derived from the continuity equation combined with the first law and takes the form \frac{\partial c}{\partial t} = D \frac{\partial^2 c}{\partial x^2} for one-dimensional diffusion in a dilute solution where D is constant. This partial differential equation governs transient mass transport processes, such as those occurring after a sudden change in electrode potential, where concentration gradients develop dynamically near the electrode interface. In electrochemistry, Fick's laws are applied within the framework of the Nernst diffusion layer model, which conceptualizes a thin layer adjacent to the electrode where diffusion dominates mass transport, with a characteristic thickness \delta that approximates the region of significant concentration gradients. For potential step experiments, typical boundary conditions include an initial uniform bulk concentration c(x,0) = c_\text{bulk} for x > 0, a surface concentration set to zero at the electrode (c(0,t) = 0 for t > 0) assuming rapid consumption of the electroactive species, and maintenance of the bulk concentration far from the electrode (c(\infty,t) = c_\text{bulk}). These conditions model semi-infinite linear diffusion, where the diffusion layer expands with time as \delta(t) \approx \sqrt{\pi D t}, reflecting the unsteady nature of the process. Solutions to Fick's second law under semi-infinite boundary conditions often involve the error function, which provides analytical expressions for concentration profiles in such systems. The complementary error function \text{erfc}(z) = 1 - \text{erf}(z), where z = x / (2 \sqrt{D t}), captures the diffusive penetration into the solution over time, enabling prediction of how concentration varies with distance and elapsed time without steady-state assumptions. This approach is essential for interpreting transient electrochemical responses, though numerical methods like finite differences may supplement it for more complex geometries.

Derivation

Key Assumptions

The derivation of the Cottrell equation relies on several fundamental assumptions that define the idealized conditions for diffusion-controlled electrochemical processes. Primarily, it assumes semi-infinite linear diffusion, where the electrode is planar and the diffusion layer thickness remains much smaller than the depth of the solution throughout the experiment, allowing the concentration gradient to extend indefinitely away from the electrode surface without reaching the bulk boundary._Chronoamperometry) This setup ensures that the mass transport is governed solely by Fick's laws of diffusion in one dimension, perpendicular to the electrode surface. A critical assumption is the absence of convection and migration effects, meaning the solution is unstirred and contains an excess of supporting electrolyte to suppress ion movement due to the electric field, thereby isolating pure diffusional control._Chronoamperometry) Additionally, the potential step is presumed to cause an instantaneous change in the reactant concentration at the electrode surface to zero, reflecting complete depletion without kinetic barriers._Chronoamperometry) The electron transfer is assumed to be reversible, with fast kinetics that maintain the surface concentration fixed according to the , ensuring the is limited entirely by rather than the itself. Finally, is strictly one-dimensional, confined to the normal to the planar , neglecting any or radial contributions that could arise from non-ideal geometries. These assumptions collectively enable the mathematical simplification leading to the Cottrell equation's time-dependent response.

Mathematical Steps

The derivation of the Cottrell equation begins with the setup of the initial boundary value problem for one-dimensional linear diffusion under the assumptions of semi-infinite diffusion, constant diffusion coefficient, and no convection. The concentration profile c(x, t) of the electroactive species satisfies Fick's second law: \frac{\partial c}{\partial t} = D \frac{\partial^2 c}{\partial x^2}, where D is the diffusion coefficient, x is the distance from the electrode surface (x = 0), and t is time. The initial condition is c(x, 0) = c_0 for x \geq 0, where c_0 is the bulk concentration. The boundary conditions are c(0, t) = 0 for t > 0 (due to rapid consumption at the electrode surface) and c(\infty, t) = c_0 (maintaining bulk concentration far from the electrode). This partial differential equation is solved using the Laplace transform method or similarity transformation, yielding the concentration profile: c(x, t) = c_0 \erf\left( \frac{x}{2 \sqrt{D t}} \right), where \erf is the error function. This solution satisfies the initial and boundary conditions, with c(0, t) = 0 since \erf(0) = 0, and c(\infty, t) = c_0 as \erf(\infty) = 1. The diffusive flux at the electrode surface is obtained from Fick's first law: J(0, t) = -D \left. \frac{\partial c}{\partial x} \right|_{x=0}. Differentiating the concentration profile gives: \left. \frac{\partial c}{\partial x} \right|_{x=0} = \frac{c_0}{\sqrt{\pi D t}}, so the flux magnitude is: |J(0, t)| = c_0 \sqrt{\frac{D}{\pi t}}. This follows from the derivative of the error function, \frac{d}{du} \erf(u) = \frac{2}{\sqrt{\pi}} e^{-u^2}, evaluated at u = 0. The observed current i(t) is related to the flux by Faraday's law for the reduction (or oxidation) involving n electrons: i(t) = n F A |J(0, t)| = n F A c_0 \sqrt{\frac{D}{\pi t}}, where F is the Faraday constant and A is the electrode area. This is the Cottrell equation, describing the time-dependent current under diffusion control. Dimensional analysis confirms the consistency of units: i in amperes (A), n dimensionless, F in coulombs per mole (C mol⁻¹), A in square meters (m²), c_0 in moles per cubic meter (mol m⁻³), D in square meters per second (m² s⁻¹), and t in seconds (s), yielding i = (mol s⁻¹) × (C mol⁻¹) = A.

Formulation and Interpretation

The Equation

The Cottrell equation provides the mathematical expression for the diffusion-limited current in a potential step experiment, given by i(t) = \frac{n F A c_0 \sqrt{D}}{\sqrt{\pi t}} where i(t) represents the current at time t. This formulation is equivalently expressed in some references as i = \frac{n F A c_0 D^{1/2}}{\pi^{1/2} t^{1/2}}. The equation highlights the characteristic time dependence of the current, which decreases proportionally to t^{-1/2}, reflecting the progressive expansion of the diffusion layer and the resulting diminution of the concentration gradient at the electrode surface. Graphically, when the current i is plotted against t^{-1/2}—known as a Cottrell plot—the relationship yields a straight line whose slope is proportional to the bulk concentration c_0 and the square root of the diffusion coefficient \sqrt{D}.

Parameter Meanings

The Cottrell equation relates the diffusion-limited current to fundamental electrochemical parameters, each carrying specific physical significance in the context of a potential step experiment under semi-infinite linear diffusion conditions. The parameter n is the stoichiometric number of electrons transferred in the redox reaction per molecule of the electroactive species. For instance, the reversible one-electron oxidation of ferrocene (\ce{Fe(C5H5)2}) to ferrocenium ion (\ce{[Fe(C5H5)2]+}) has n = 1, making it a standard reference for validating electrochemical measurements. Faraday's constant F quantifies the electric charge associated with one mole of electrons, with a value of 96,485 C/mol; it converts the molar flux of species to electrical current by linking chemical reaction stoichiometry to charge transfer. The electrode area A, typically in cm², represents the effective geometric surface of the working electrode exposed to the electrolyte solution; the current scales linearly with A because more area provides proportionally more sites for the electrochemical reaction to occur. The bulk concentration c_0, expressed in mol/cm³, is the initial uniform concentration of the electroactive species far from the electrode; it directly influences the magnitude of the concentration gradient driving diffusion, with higher c_0 yielding larger initial currents proportional to the available reactant supply. The diffusion coefficient D, in units of cm²/s, measures the average mobility of the electroactive species through the solution via random thermal motion; it determines how quickly species can diffuse to the electrode surface, with larger D resulting in steeper concentration gradients and thus higher currents for a given time. The parameter t, the time in seconds since the potential step, governs the temporal decay of the current; as t increases, the current diminishes proportionally to t^{-1/2} due to the progressive depletion of species near the electrode and the resulting extension of the concentration profile into the solution. Physically, the combination \sqrt{D}/\sqrt{t} in the equation embodies the diffusive flux's dependence on the inverse of the diffusion layer thickness, which grows as \delta \approx \sqrt{\pi D t}; this thickness marks the approximate extent of the concentration perturbation from the bulk value, leading to a diminishing gradient and current over time as the layer expands.

Applications

Chronoamperometry

Chronoamperometry is a potentiostatic electrochemical technique that involves applying a sudden potential step to a working electrode to initiate a diffusion-controlled faradaic process, measuring the resulting current as a function of time. In this method, the electrode potential is abruptly changed from a value where no electrolysis occurs (non-faradaic region) to one that drives the redox reaction at a diffusion-limited rate, allowing the study of mass transport phenomena under controlled conditions. This potential step method, also known as constant potential bulk electrolysis, provides insights into the kinetics of electron transfer and diffusion without the complications of varying potential. The experimental setup typically employs a three-electrode cell configuration connected to a potentiostat, which precisely controls the potential and measures the current. The working electrode, often a planar disk such as platinum or glassy carbon, is where the faradaic reaction occurs; a reference electrode (e.g., Ag/AgCl or saturated calomel) maintains a stable potential reference; and a counter electrode completes the circuit to pass the necessary current. The potentiostat applies the potential step, typically from an initial value determined by cyclic voltammetry to ensure no reaction, to a final value sufficiently negative to the standard potential (e.g., 120-180 mV below E^0 for reversible reduction systems) to achieve diffusion control. Solutions are usually unstirred to promote linear diffusion, and the experiment duration is set to capture the transient current decay, often with sampling intervals to resolve early rapid changes. Upon applying the potential step, the observed current exhibits an initial sharp spike due to double-layer charging, followed by a rapid decay of the faradaic current that adheres to a t^{-1/2} time dependence, as predicted by the Cottrell equation under ideal diffusion-limited conditions. This decay reflects the growing diffusion layer thickness at the electrode surface, where the flux of electroactive species diminishes over time, confirming the semi-infinite linear diffusion model. In practice, the current-time (i-t) transient is plotted to visualize this behavior, with the faradaic component dominating after the capacitive spike subsides. A representative example is the cathodic reduction of ferricyanide (\ce{[Fe(CN)6]^3-}) to ferrocyanide in aqueous solution using a 3 mm diameter platinum disk working electrode. The potential is stepped from +0.60 V to 0.00 V vs. Ag/AgCl in a 1.0 mM ferricyanide solution with 0.1 M KCl as supporting electrolyte, resulting in an i-t curve that shows the characteristic initial peak and subsequent t^{-1/2} decay. Fitting the linear portion of the i vs. t^{-1/2} plot to the Cottrell equation validates the diffusion-controlled nature of the process, with the slope providing a measure of consistency with theoretical predictions.

Measuring Diffusion Coefficients

The Cottrell equation provides a linear relationship between the measured current i and the inverse square root of time t^{-1/2} under diffusion-controlled conditions, enabling the determination of the diffusion coefficient D through graphical analysis. To extract D, experimental data are plotted as i versus t^{-1/2}, yielding a straight line with slope equal to n F A c_0 \sqrt{D / \pi}, where n is the number of electrons transferred, F is Faraday's constant, A is the electrode area, and c_0 is the bulk concentration of the electroactive species. If n, A, and c_0 are known, the slope allows direct calculation of D by rearranging the equation: D = \pi \left( \frac{\text{slope}}{n F A c_0} \right)^2. In practice, chronoamperometry involves applying a potential step to the working electrode and recording the transient current response at multiple time points following the step. The raw current-time data are corrected for background currents, such as charging or faradaic contributions from impurities, by subtracting a baseline measured prior to the potential step. The corrected data are then fitted to the linear plot, typically using least-squares regression on the later portion of the response (e.g., after the initial non-ideal transients subside), to obtain the slope with high precision. Accurate determination of D requires precise knowledge of n, A, and c_0, as uncertainties in these parameters propagate into the calculated value; for instance, electrode area A must be calibrated independently, often via geometric measurement or alternative electrochemical methods. Additionally, the experiment assumes semi-infinite linear diffusion, which holds best for short times and planar electrodes. For typical electroactive ions in aqueous solutions, diffusion coefficients fall in the range of $10^{-5} cm²/s, reflecting the mobility of small species like metal complexes or inorganic ions under ambient conditions. The Cottrell equation also applies to chronocoulometry, where the integrated charge Q(t) = 2 n F A C \sqrt{\frac{D t}{\pi}} provides an alternative method to determine D from the slope of Q versus \sqrt{t}._Chronoamperometry)

Limitations and Extensions

Validity Conditions

The Cottrell equation accurately describes the current-time response in chronoamperometric experiments under conditions where diffusion is the dominant mass transport mechanism, requiring short experimental timescales to minimize convective contributions. Specifically, the diffusion layer thickness δ must remain much smaller than the solution depth (δ << L, where L is the unstirred layer thickness), which typically holds for times t ≲ 10–100 s before natural convection arises due to density gradients from concentration changes near the electrode. At longer times, hydrodynamic effects distort the linear diffusion profile assumed in the derivation, leading to deviations from the predicted t^{-1/2} decay. Regarding electrode geometry, the equation is valid for macroplanar electrodes where the electrode dimensions significantly exceed the diffusion layer thickness (r_electrode >> √(Dt)), ensuring one-dimensional semi-infinite diffusion. On microelectrodes (radii < 25 μm), enhanced edge diffusion introduces spherical or hemispherical contributions, causing the current to transition from Cottrell-like behavior at short times (t << r^2/D) to a steady-state regime at longer times, invalidating the time-dependent form. Solution conditions must include a supporting electrolyte at concentrations at least 100 times higher than the analyte to suppress electromigration, as the Cottrell equation assumes negligible migration and convection, relying solely on diffusion. In unsupported or low-ionic-strength electrolytes, charged species experience additional transport due to the electric field, enhancing or altering the current beyond the predicted values, particularly for uncompensated ohmic drop or ion-selective reactions. The model further presumes dilute solutions where the diffusion coefficient remains constant and activity effects are minimal. For electrochemical kinetics, the equation requires reversible electron transfer with the applied potential sufficiently anodic or cathodic relative to the standard potential (E - E^0 >> RT/nF, typically > 200 mV) to establish zero surface concentration and full diffusional depletion. Irreversible systems, characterized by slow charge transfer, exhibit deviations at early times due to kinetic limitations overriding diffusion control, resulting in lower currents than predicted. A key experimental diagnostic for confirming the validity of these conditions is the linearity of the plot of current i versus t^{-1/2}, which should yield a straight line through the origin with a slope proportional to nF A C √(D/π); non-linearity indicates influences from convection, migration, or kinetics.

Modified Forms

The Cottrell equation, originally derived for planar electrodes under semi-infinite linear diffusion, has been extended to spherical electrodes to account for radial diffusion contributions, which become significant at shorter times or smaller electrode sizes. For a spherical electrode of radius r, the current is given by i(t) = n F A c_0 \sqrt{\frac{D}{\pi t}} \left(1 + \frac{\sqrt{\pi D t}}{r}\right), where the additional term \frac{\sqrt{\pi D t}}{r} introduces a steady-state-like contribution as time progresses, deviating from the pure t^{-1/2} decay. This modification is particularly useful for microelectrodes, where spherical geometry enhances mass transport and allows steady-state currents at longer times. Uncompensated solution resistance introduces an ohmic (iR) drop that distorts the applied potential, leading to a modified apparent current in chronoamperometric experiments. The effective potential at the electrode surface becomes E_{\text{eff}} = E_{\text{applied}} - i R_u, where R_u is the uncompensated resistance, causing the observed current to deviate from the ideal Cottrell behavior, especially at high currents or low conductivities, and requiring corrections via extrapolation or compensation techniques. This effect preserves the t^{-1/2} dependence but alters the magnitude, necessitating modeling of the feedback loop between current and potential shift. A related extension appears in cyclic voltammetry through the Randles-Ševčík equation, which adapts the Cottrell framework to describe the peak current under linear sweep conditions: i_p = 0.446 \, n F A c_0 \sqrt{\frac{n F v D}{R T}}, where v is the scan rate, linking the diffusion-limited transient to dynamic voltammetric responses for reversible systems. This equation highlights the square-root dependence on scan rate, analogous to the time dependence in chronoamperometry. In modern adaptations, the Cottrell equation integrates with ultramicroelectrode behavior, where the spherical correction dominates at short times before transitioning to steady-state diffusion, enabling low detection limits in analytical applications. For systems involving convection, hybrid models combine the Cottrell transient with the Levich equation for rotating disk electrodes, describing initial diffusive decay followed by convective steady-state plateaus, as seen in rotating ring-disk studies. A recent application involves modeling electrochemical CO2 reduction on copper catalysts, where the Cottrell equation is fitted to chronoamperometric transients to dissect charge transfer and intermediate formation kinetics, aiding selectivity control toward C2 products like ethylene.

References

  1. [1]
  2. [2]
    Frederick Cottrell | College of Chemistry
    Mar 17, 2020 · A native of Oakland, California, Frederick Gardner Cottrell (1877-1948) early revealed the intense scientific curiosity and drive toward achievement.Missing: biography | Show results with:biography
  3. [3]
    [PDF] Jaroslav Heyrovsky - Nobel Lecture
    Basing on this law the equations for the current-voltage curves due to the reductions and oxidations of depolarizers were deduced, from which the rule of ...Missing: Cottrell 1920s<|separator|>
  4. [4]
    Jaroslav Heyrovsky and polarography - Electrochemistry Knowledge
    Polarography was created in 1922 when Jaroslav Heyrovsky used dropping mercury electrode for measuring polarization curves. Encouraged by the results, he ...Missing: Cottrell | Show results with:Cottrell
  5. [5]
    [PDF] Polarography - National Academic Digital Library of Ethiopia
    the Ilkovic equation by simply substituting the geometrical proper ties of a sphere into the Cottrell equation. We have seen already (1.3.3) that for the ...
  6. [6]
    Understanding the Cottrell Equation - Macias Sensors
    Feb 3, 2023 · The Cottrell equation is a fundamental equation in the field of electrochemistry that describes the relationships between current and time in depletion ...Missing: primary | Show results with:primary
  7. [7]
    Introduction to Electrocatalysts | ACS Symposium Series
    Dec 9, 2022 · Gardner Cottrell established the Cottrell equation ( 54, 55 ), which governs the relationship between the kinetics of electrodes and mass ...
  8. [8]
    Cottrell's equation revisited: an intuitive, but unreliable, novel ...
    Cottrell showed that the concentration profile that develops is given, in modern terminology, by c(z,t)=C erf z 4Dt , where z is the coordinate normal to, and ...Missing: primary | Show results with:primary
  9. [9]
    Electrochemistry Encyclopedia -- Electroanalytical chemistry
    In the past, instruments were rebuilt or re-adjusted for different electrochemical methods. This adjustment is now totally in the hands of software. Blood ...
  10. [10]
    [PDF] Analytical Electrochemistry: The Basic Concepts
    according to the Cottrell equation. This was explained in terms of the concentration profile extending from the electrode surface into the bulk of solution ...<|separator|>
  11. [11]
    Our History – Research Corporation for Science Advancement (RCSA)
    To that end, he established Research Corporation (America's second oldest foundation) in 1912. The original board of directors, made up of eminent scientists, ...
  12. [12]
    Archaic equation helps scientists control CO2 transformations
    Apr 5, 2023 · The calculation – named the Cottrell equation for chemist Frederick Gardner Cottrell, who developed it in 1903 – can help today's ...Missing: original | Show results with:original
  13. [13]
    Diffusion Controlled Reaction - an overview | ScienceDirect Topics
    During a diffusion-controlled reaction, matter may be transported around an interface, which separates the reactants and the product. The progress of the ...
  14. [14]
    Diffusional Voltammetry in Finite Spaces | ACS Electrochemistry
    Jun 2, 2025 · It described the coexistence of semi-infinite diffusion from the bulk solution to the electrode surface and thin-layer diffusion within the ...
  15. [15]
    [PDF] ELECTROCHEMICAL METHODS
    Apr 5, 2022 · diffusion-controlled rate. What current ... Since semi-infinite linear diffusion applies, the diffusion-limited current is the Cottrell.
  16. [16]
    [PDF] Fick's Laws of Diffusion and Electrochemistry
    If the surface electron transfer reaction is very fast, the current just depends on the flux of electroactive species to the surface: i(t) = − nFAJ.
  17. [17]
    An Introduction to Electrochemical Methods in Neuroscience - NCBI
    Diffusion is a mode of mass transport that occurs as a result of the tendency of molecules in fluids to undergo continuous random motion, roughly equivalent to ...
  18. [18]
    Diffusion layer - SubsTech
    Dec 14, 2013 · Nernst diffusion layer is a virtual layer, within which the gradient of the ion concentration is constant and equal to the true gradient at the electrode- ...
  19. [19]
    [PDF] Electrochemistry
    Diffusion is quantitatively described by Fick's law where the rate of diffusion is known as flux in moles m-2 s-1. Fick's first law:
  20. [20]
    [PDF] Lecture #13 of 18(?)
    The Cottrell Equation can be used to describe the behavior during a potential step experiment as long as the thickness of the Nernst diffusion layer is small ...Missing: primary source
  21. [21]
  22. [22]
    [PDF] Analytical Electrochemistry, Second Edition. Joseph Wang
    FIGURE 4-29 Cottrell plot of the chronoamperometric response for 1 10 3. M. Ru(NH3)6. 3 З at a Kel-F=gold composite electrode. Points are experimental data ...
  23. [23]
    Faraday constant - CODATA Value
    Numerical value, 96 485.332 12... C mol ; Standard uncertainty, (exact) ; Relative standard uncertainty, (exact).
  24. [24]
    Data Analysis - Chronoamperometry/chronocoulometry - BASi
    The analysis of chronoamperometry (CA) data is based on the Cottrell equation, which defines the current-time dependence for linear diffusion control.
  25. [25]
    Chronoamperometry (CA) | Pine Research Instrumentation
    Sep 24, 2024 · Cottrell, F. G. Residual current in galvanic polarization regarded as a diffusiona problem. Z. Phys. Chem., 1903, 42, 385–431. Tennyson ...Missing: title | Show results with:title
  26. [26]
    Theory
    ### Summary of Chronoamperometry for Ferricyanide Reduction
  27. [27]
    Diffusion Coefficient - an overview | ScienceDirect Topics
    As order of magnitude, they are 10− 1 to 10− 2 cm2/s for gases and 10− 4 to 10− 5 cm2/s for liquids.
  28. [28]
    Rapid and Direct Determination of Diffusion Coefficients Using ...
    May 27, 2016 · The diffusion coefficient for ferricyanide was determined to be 7.3 ± 0.7 × 10─6 cm2/s (n = 15) using the slope of the line in Figure 7, right, ...Abstract · Experimental · Results and Discussion · Conclusions
  29. [29]
    Calibration-Free Analysis with Chronoamperometry at Microelectrodes
    Sep 3, 2024 · Current values calculated using the Cottrell equation (blue dots), and the Mahon and Oldham equation for short times (black dots) and for long ...
  30. [30]
    [PDF] Lecture #13 of 18(?)
    Nov 28, 2023 · The Cottrell Equation can be used to describe the behavior during a potential step experiment as long as the thickness of the Nernst diffusion.Missing: assumptions | Show results with:assumptions<|separator|>
  31. [31]
    Cyclic voltammetry at microelectrodes. Influence of natural ...
    In chronoamperometry, this results in a progressive deviation of the currents from the Cottrell equation towards steady-state values [9], [10].
  32. [32]
    [PDF] CHEM465/865, 2004-3, Lecture 11, 4th Oct
    Oct 4, 2025 · ❑ Short times: spherical diffusion resembles the solution for planar diffusion (Cottrell equation). ❑ At times. 2. 0 r t. D π. >> 0 r δ >> for ...
  33. [33]
    The Cottrell Experiment and Diffusion Limitation 3/3 - PalmSens
    The layer of ions and the electrode act like a capacitor and this has impact on most electrochemical techniques. This is only a rudimentary description of the ...
  34. [34]
    Limiting currents in potentiostatic voltammetry without supporting ...
    The imposition of a large enough potential step on a macroelectrode in such a solution results in a limiting current that obeys Cottrell's equation, declining ...
  35. [35]
    Accounting for the concentration dependence of electrolyte diffusion ...
    The applicability of these equations is limited because they are deduced under the assumption of a constant diffusion coefficient.Missing: key | Show results with:key
  36. [36]
    [PDF] Direct Determination of Diffusion Coefficients by ... - Allen J. Bard
    kinetic limitations), semi-infinite linear diffusion to and from a microdisk takes ... place, and that the diffusion controlled current obeys the Cottrell ...
  37. [37]
    Chronoamperometry (CA) | Pine Research Instrumentation
    Jan 4, 2016 · ... Cottrell equation. Examining the Cottrell Current plot in this region reveals that the current is linear with respect to t^{-1/2} (see black ...Missing: original paper
  38. [38]
    The Cottrell Experiment and Diffusion Limitation 2/3 - PalmSens
    During the Cottrell experiment a potential step across a redox potential is applied and the current caused by the electrochemical reaction is recorded. The ...
  39. [39]
    corrections to the Cottrell equation
    The Cottrell equation predicts that, under conditions of semi-infinite linear diffusion, the current decays as t-'I2 aiid the quantity It 'IZ should be ...
  40. [40]
    Uncompensated Resistance. 1. The Effect of Cell Geometry
    Aug 6, 2025 · ... Cottrell's equation, declining as t−1/2. In the present article it is demonstrated that t−1/2 dependence is preserved even when supporting ...
  41. [41]
    Halate electroreduction via autocatalytic mechanism for rotating disk ...
    Sep 20, 2021 · Within this initial time range the current decreases according to the Cottrell equation for X2 species. Then, if the convection intensity is ...
  42. [42]