Fact-checked by Grok 2 weeks ago

Electroanalytical methods

Electroanalytical methods encompass a diverse set of techniques in that investigate chemical analytes by measuring electrical properties, such as potential, current, or charge, within an . These methods rely on the principles of reactions at the between an and an , where the applied or measured electrical signals correlate directly with the analyte's concentration, behavior, or reaction kinetics. By exploiting heterogeneous processes involving electrodes—typically composed of materials like , glassy carbon, or mercury—and supporting electrolytes, electroanalytical techniques enable precise control over reaction conditions through factors such as electrode surface and composition. The primary categories of electroanalytical methods include potentiometric, voltammetric, coulometric, and amperometric approaches, each distinguished by the electrical parameter emphasized and the mode of operation. Potentiometry measures the potential difference between electrodes at zero current flow, commonly used for ion-selective sensing, such as electrodes or detection of specific cations and anions. , the most versatile group, applies a varying potential while monitoring current, encompassing techniques like (CV) for studying mechanisms, (DPV), and square-wave voltammetry () for enhanced sensitivity in trace analysis. quantifies s by integrating current over time to determine total charge passed during , offering absolute measurements independent of cell constants. maintains a fixed potential to measure steady-state current proportional to , often integrated into flow systems like biosensors. Advanced variants, such as stripping voltammetry or electrochemical impedance spectroscopy, further improve selectivity through preconcentration steps or frequency-domain analysis. Electroanalytical methods are prized for their exceptional —often achieving detection limits in the nanomolar range—and versatility across diverse matrices, making them indispensable in fields like for pollutants, pharmaceutical for quantification, and clinical diagnostics for biomarkers such as glucose. Their non-destructive nature for many techniques, combined with the ability to provide mechanistic insights into processes, supports applications in , electrocatalysis, and development. Recent advancements, including electrode modifications with , have expanded their scope to real-time analysis, underscoring their ongoing evolution as robust tools in modern analytical science.

General Principles

Electrochemical Cells and Electrodes

In electroanalytical methods, electrochemical cells serve as the fundamental apparatus for studying processes through controlled interactions between s and an solution. Galvanic cells rely on spontaneous reactions to generate an electrical potential, whereas electrolytic cells apply an external voltage to drive non-spontaneous reactions, with the latter being predominant in techniques requiring precise potential control. The typical setup emphasizes three-electrode configurations to enhance measurement accuracy by isolating the 's response from solution resistance effects. In this system, the is where the target undergoes oxidation or , the maintains a stable potential for comparison, and the counter supplies or accepts electrons to balance the circuit, all immersed in a shared compartment. Early electroanalytical work in the utilized simple two-electrode systems, where the working and counter functions were combined, limiting precision due to uncompensated resistance. The shift to three-electrode systems occurred in the 1940s, driven by Archie Hickling's invention of the potentiostat, which enabled independent control of the potential relative to the reference, significantly improving reproducibility and accuracy in experiments. This configuration remains standard, as it minimizes drop—the voltage loss from current flow through the —and allows for reliable data in diverse analytical applications. Working electrodes are selected based on the need for inertness to avoid with reactions; electrodes, made from polished wire or foil, offer excellent conductivity and a broad potential range in aqueous media. Glassy carbon electrodes, formed by pyrolyzing a precursor into a non-porous, structure, provide chemical stability, low background currents, and resistance to adsorption, making them ideal for organic and biochemical analyses. Mercury-based electrodes, such as the dropping mercury electrode (DME) introduced by Jaroslav Heyrovský for , feature a vertical glass connected to a mercury , where drops form and detach at controlled intervals (typically 3-5 seconds) to renew the surface and suppress effects. Reference electrodes ensure a constant potential benchmark; the (SCE) consists of a containing a mercury pool covered by a paste of mercury and mercurous chloride (Hg/Hg₂Cl₂), filled with saturated , and connected to the external medium via a porous ceramic frit or fiber junction to allow ionic contact while preventing contamination. The (Ag/AgCl) electrode comprises a silver wire coated with (often as a paste or sintered layer) immersed in a (typically 3 M or saturated), housed in a similar tube with a for isolation. Counter electrodes are usually inert wires or meshes to facilitate efficient passage without altering the solution composition. Electrolyte solutions are essential for maintaining ionic conductivity and supporting charge transfer; they commonly include supporting electrolytes like (KNO₃), (NaCl), or tetramethylammonium salts at concentrations of 0.1-1 M in aqueous or non-aqueous solvents to reduce ohmic resistance, minimize electrostatic migration of analyte ions, and stabilize the double layer at the electrode interface. The choice of composition depends on the analyte's solubility and the method's requirements, ensuring uniform ion distribution without introducing side reactions.

Electrode Potentials and the Nernst Equation

The equilibrium potential at an electrode-solution interface arises from the balance between oxidation and reduction half-cell reactions, where the electrochemical potential of the oxidized and reduced species is equal. This potential reflects the thermodynamic driving force for the electron transfer process at the interface, governed by the activities of the species involved in the half-reaction. The quantifies this equilibrium potential for a general half-cell reaction of the form Oxidized + ne⁻ ⇌ Reduced. It is derived from the relationship between the change (ΔG) and the cell potential (E), where ΔG = -nFE, combined with the standard expression for ΔG = ΔG° + RT ln Q, leading to -nFE = -nFE° + RT ln Q. Rearranging yields E = E° - (RT/nF) ln Q, where E is the , E° is the , R is the (8.314 J mol⁻¹ K⁻¹), T is the absolute temperature in , n is the number of electrons transferred, F is Faraday's constant (96485 C mol⁻¹), and Q is the defined as the ratio of activities (or concentrations for dilute solutions) of products to reactants. At 25°C (298 K), this simplifies to the logarithmic form E = E° - (0.059/n) log Q, using common logarithms for practical calculations. Standard electrode potentials (E°) are referenced to the (SHE), defined by IUPAC as a in contact with a solution of unit activity H⁺ ions (1 M) and H₂ gas at 1 bar pressure, assigned an E° of 0 V for the 2H⁺ + 2e⁻ ⇌ H₂. For example, the standard reduction potential for the Fe³⁺/Fe²⁺ couple is +0.771 V versus SHE at 25°C and 0, indicating that Fe³⁺ is a stronger oxidant than H⁺ under standard conditions. The is sensitive to changes in solution conditions through the Q term. Concentration influences the potential logarithmically; for the Ag⁺ + e⁻ ⇌ Ag with E° = +0.799 , a tenfold decrease in [Ag⁺] from 1 M to 0.1 M shifts E to +0.740 at 25°C, as log(1/[Ag⁺]) = +1. affects the RT/nF factor, increasing the slope of the logarithmic term and thus amplifying concentration effects; for instance, at 50°C, the becomes approximately 0.070/n instead of 0.059/n. For involving H⁺, such as the / couple Q + 2H⁺ + 2e⁻ ⇌ QH₂ with E° ≈ +0.699 , the potential decreases by 0.059 per unit increase at 25°C, since Q includes [H⁺]², making E = E° - (0.059/2) log([QH₂]/[Q][H⁺]²). In metal ion systems without direct H⁺ involvement, like Cu²⁺/Cu⁺ (E° = +0.153 ), effects are indirect through or , but concentration changes dominate, with E shifting by -0.059 log([Cu⁺]/[Cu²⁺]) at 25°C.

Faradaic and Capacitive Currents

In electroanalytical methods, currents at the electrode-solution interface arise from two primary mechanisms: Faradaic processes involving electron transfer and non-Faradaic processes due to interfacial charging. Faradaic currents result from redox reactions where electrons are transferred between the electrode and solution species, directly linking the current to the rate of chemical transformation according to Faraday's laws. These currents are proportional to the reaction rate, with the magnitude determined by the number of electrons transferred (n), the Faraday constant (F ≈ 96,485 C/mol), and the electrode area (A), such that the rate in mol/s equals i / (n F). The kinetics of Faradaic currents are described by the Butler-Volmer equation, which relates the net current density (i) to the overpotential (η = E - E_eq), where E is the applied potential and E_eq is the equilibrium potential. The equation is: i = i_0 \left[ \exp\left(\frac{\alpha n F \eta}{RT}\right) - \exp\left(-\frac{(1-\alpha) n F \eta}{RT}\right) \right] Here, i_0 is the exchange current density, α is the transfer coefficient (typically 0.3–0.7), R is the gas constant, and T is the temperature in Kelvin. This form captures both anodic and cathodic contributions, with the exponential terms reflecting activation barriers for oxidation and reduction, respectively. For irreversible processes, large overpotentials make one exponential term negligible, simplifying to the Tafel equation and highlighting kinetic limitations beyond thermodynamic equilibrium. In contrast, capacitive currents, also known as non-Faradaic or charging currents, originate from the accumulation of charge in the electrical double layer at the surface, without net to species. This layer behaves like a , with double-layer per unit area typically 10–40 μF/cm² for aqueous s (total C = c_dl A), and the current is given by: i_c = C \frac{dE}{dt} where dE/dt is the rate of change of potential. These currents are particularly prominent during potential transients, such as in voltammetric scans, and model the interface as an where the double-layer is in parallel with Faradaic resistance. Several factors influence the relative contributions of Faradaic and capacitive currents. Capacitive currents scale linearly with scan rate (v = dE/dt), increasing background noise at faster sweeps and potentially obscuring Faradaic signals, whereas reversible Faradaic currents in techniques like exhibit peak currents proportional to v^{1/2} due to control. Irreversibility in Faradaic processes, characterized by slow rates (low i_0 or extreme α), shifts the current-potential response, broadening peaks and reducing sensitivity compared to reversible systems. Electrode surface area and solution resistance also amplify capacitive effects, as larger areas increase C and uncompensated resistance (R_u) lengthens the charging (τ = R_u C). Quantitative separation of these currents is essential for accurate analysis, particularly in where Faradaic peaks overlay a capacitive . One approach involves plotting peak against scan rate: capacitive components yield a linear relationship through the origin, while Faradaic contributions show linearity with v^{1/2} for diffusion-limited processes, allowing subtraction via correction or fitting. In practice, techniques like square-wave voltammetry minimize capacitive interference by sampling differential currents, enhancing Faradaic signal resolution.

Potentiometric Methods

Direct Potentiometry

Direct potentiometry is an electroanalytical technique that quantifies the activity of an ion by measuring the open-circuit potential difference between an ion-selective indicator electrode and a under conditions of zero or negligible current flow. This zero-current method ensures minimal perturbation to the sample, allowing the potential to directly reflect the electrochemical equilibrium at the electrode-solution interface. The measured potential arises from the selective interaction of the target ion with the indicator electrode's sensing element, such as a or , which establishes a phase boundary potential proportional to the ion's activity in solution. The relationship between the measured potential and activity follows a logarithmic response, as described by the , where the changes by approximately 59 mV per decade change in activity for monovalent s at 25°C, known as the Nernstian slope. curves are constructed by plotting the potential against the logarithm of known activities, yielding a linear segment typically spanning several orders of magnitude, from which unknown concentrations can be determined via . This logarithmic dependence arises from the thermodynamic basis of ion partitioning at the interface, enabling sensitive detection down to trace levels in many cases, though the exact slope and linear range depend on the material and solution conditions. A classic example of direct potentiometry is pH measurement using the , where a thin, hydrated silicate glass membrane selectively responds to hydrogen s, generating a potential that varies linearly with over the range of 0 to 14. This electrode, developed in the early and widely adopted for its robustness and accuracy, pairs with a like or Ag/AgCl to form a complete cell for routine laboratory and industrial pH monitoring. Another prominent application is analysis with lanthanum (LaF₃) electrodes, which utilize a single-crystal membrane to detect activities as low as 10⁻⁶ M, as pioneered in the for assessment and dental product evaluation. These solid-state electrodes exhibit near-Nernstian slopes of about -59 mV/decade and high selectivity for F⁻ over common interferents. Selectivity in direct potentiometry is crucial for accurate measurements in complex matrices, where interfering ions can contribute to the measured potential through non-ideal responses described by the Nikolsky-Eisenman equation: E = E^0 + \frac{RT}{F} \ln \left( a_i + \sum k_{ij} a_j \right) Here, E is the cell potential, E^0 is the standard potential, a_i and a_j are the activities of the primary ion i and interfering ion j, respectively, k_{ij} are selectivity coefficients quantifying the relative response to interferents, and RT/F is the Nernst factor (approximately 59 mV at 25°C). These coefficients, determined experimentally via methods like the separate solution technique, indicate the electrode's discrimination ability; for instance, a LaF₃ electrode has k_{F^-, OH^-} \approx 0.05, meaning hydroxide interference is minimal at neutral pH. Interference effects become pronounced when interferent activities approach or exceed those of the analyte, necessitating ionic strength adjustment or masking agents to maintain accuracy. This equation, rooted in phase boundary potential theory, underpins the design of ion-selective electrodes for multianalyte environments.

Potentiometric Titrations

Potentiometric titrations involve the measurement of the potential difference between an and a as a function of the volume of titrant added to an , allowing for the determination of the through changes in . This method relies on the Nernstian response of the indicator electrode to the activity of involved in the reaction. The technique was first introduced in the late , with Robert Behrend performing the initial in 1893 at Ostwald's Institute in , where he titrated mercurous nitrate with halides using a mercury . Early applications focused on reactions, marking the beginning of instrumental detection in volumetric analysis. Endpoint identification in potentiometric titrations is achieved by analyzing the potential-volume curve, which typically exhibits a gradual change before and after the , with a sharp inflection at the for systems with suitable responses. One widely used approach is the Gran plot method, developed by Gunnar Gran in the early 1950s, which transforms the nonlinear potential data into linear segments by plotting functions proportional to the or titrant concentration against volume. For acid-base titrations, the Gran function before the (e.g., 10^{pH} times volume) extrapolates to zero at the equivalence volume, providing precise determination even in dilute solutions where inflection points are shallow. This method enhances accuracy by avoiding direct reliance on the inflection and is particularly effective for systems with conditional stability constants. Alternative endpoint detection employs first- and second- methods applied to the potential-volume . The first , calculated as the change in potential per unit volume (ΔE/ΔV), plotted against volume, reaches a maximum at the , corresponding to the steepest slope of the curve. The second , the change in the first per unit volume (Δ(ΔE/ΔV)/ΔV), shows a sharp peak or sign change precisely at the , offering higher sensitivity for detecting subtle inflections in weak systems. These techniques are computationally straightforward and improve in automated , though they require smooth to minimize noise effects. Potentiometric titrations are classified by reaction type, each exhibiting characteristic potential changes at the due to shifts in the predominant species sensed by the . In acid-base titrations, such as the neutralization of with using a glass , the potential () remains low in the acidic region and jumps abruptly to the basic region at the , reflecting the rapid change from H⁺ dominance to OH⁻ dominance. This jump, often exceeding 6-8 units in strong acid-strong base systems, enables accurate detection without visual indicators. titrations, exemplified by the determination of with using a silver , show a constant potential before the governed by excess Ag⁺, followed by a sharp decrease after as sparingly soluble AgCl forms, reducing [Ag⁺] dramatically and shifting control to Cl⁻ activity. The potential change can span 100-200 mV or more, depending on . titrations, such as the oxidation of ferrous ion with cerium(IV) using a , exhibit a low potential before the (controlled by Fe²⁺/Fe³⁺ ) that rises steeply to a high value after, dominated by the Ce³⁺/Ce⁴⁺ , with jumps often around 400-600 mV due to differing standard potentials. The primary advantages of potentiometric titrations include the objective determination of the without reliance on color-changing indicators, making them suitable for , turbid, or opaque solutions where visual methods fail. They also provide high precision and versatility across reaction types, with automation enabling reproducible results in routine analyses. However, limitations arise from the slow response time of electrodes near the , particularly in systems with gradual potential changes, which can prolong titration duration and introduce errors if not equilibrated properly. Additionally, the method requires stable and suitable electrodes responsive to the , limiting applicability in highly irreversible systems.

Amperometric and Voltammetric Methods

Amperometry

is an electroanalytical technique that involves applying a constant potential to a and measuring the resulting , which is proportional to the concentration of an electroactive diffusing to the surface. At sufficiently positive or negative potentials, the becomes diffusion-limited, governed by the rate at which the reaches the . In steady-state conditions, such as those achieved with or thin-layer cells, the limiting i is described by i = n F A D C / \delta, where n is the number of electrons transferred, F is the , A is the area, D is the coefficient, C is the concentration, and \delta is the diffusion layer thickness. This steady-state provides a direct measure of concentration, making suitable for in flowing systems or sensors. Capacitive currents may contribute to the background signal but are typically minimized at longer times. A key variant is chronoamperometry, where a potential step is applied, and the transient current is monitored over time. The current decays as t^{-1/2} due to the expanding diffusion layer, following the : i(t) = n F A C \sqrt{D / \pi t}. Plotting i versus t^{-1/2} yields a straight line (Cottrell plot), from which the diffusion coefficient D can be determined using the slope n F A C \sqrt{D / \pi}. This method is valuable for studying mass transport and reaction kinetics, particularly for reversible systems, and is often performed at microelectrodes to approach steady-state conditions more rapidly. Amperometry finds widespread application in biosensors, such as those for detection, where catalyzes the oxidation of glucose, consuming oxygen that is subsequently reduced at the to generate a measurable . In amperometric titrations, the remains constant or decreases until the , after which excess titrant causes a sharp increase or "jump" in , enabling precise detection for reactions involving electroactive species like halides or metals. These applications highlight amperometry's sensitivity and simplicity for real-time monitoring in clinical and environmental analysis. Despite its advantages, suffers from limitations including electrode fouling, where reaction products or adsorbates accumulate on the surface, degrading response over time. Additionally, dissolved oxygen can interfere by undergoing reduction at similar potentials to the , particularly in oxidase-based biosensors, necessitating or selective mediators.

Voltammetry

Voltammetry encompasses a suite of where the potential applied to a is varied, typically in a linear or pulsed manner, to measure the resulting as a function of potential, yielding characteristic current-potential (i-E) curves that provide insights into processes. These methods differ from constant-potential by actively scanning the potential to probe electrochemical kinetics and , often using inert electrodes such as or glassy carbon. The resulting voltammograms exhibit sigmoidal or peaked shapes depending on the scan type and system reversibility, enabling qualitative identification of analytes and through peak heights or areas. Linear sweep voltammetry (LSV) involves applying a linearly increasing or decreasing potential to the at a constant scan rate, typically from 1 mV/s to 100 V/s, generating a peak-shaped i-E curve for reversible systems where the current rises to a maximum near the before decreasing due to the growing diffusion layer under diffusion-limited mass transport. For reversible transfers, the peak current i_p in LSV is described by the Randles-Ševčík equation: i_p = (2.69 \times 10^5) \, n^{3/2} A D^{1/2} v^{1/2} C where n is the number of electrons transferred, A is the electrode area (cm²), D is the diffusion coefficient (cm²/s), v is the scan rate (V/s), and C is the bulk concentration (mol/cm³); this relation, derived independently by Randles and Ševčík, highlights the square-root dependence on scan rate, confirming diffusion control. The peak potential E_p relates to the formal redox potential, aiding in thermodynamic characterization. Cyclic voltammetry (CV) extends LSV by reversing the potential scan direction after reaching a vertex, producing a characteristic "duck-shaped" voltammogram with anodic and cathodic peaks that reveal both and oxidation processes in a single experiment. For reversible systems, the peak separation \Delta E_p equals $59/n mV at 25°C, indicating Nernstian behavior and rapid , while irreversible systems show larger \Delta E_p (>70 mV) and peak broadening due to slow or coupled chemical reactions. This diagnostic criterion, established through theoretical analysis of single-scan and cyclic methods, distinguishes reversible, quasi-reversible, and irreversible s, with the formal potential E^0 estimated as the midpoint between peaks. Pulse voltammetric techniques enhance by superimposing potential on a linear ramp or waveform, minimizing capacitive currents and sharpening peaks for trace analysis. (DPV) applies small (typically 5-100 mV) to a ramp, measuring the difference in current before and after each , resulting in well-defined peaks whose heights are proportional to concentration, with detection limits down to 10^{-8} M for many analytes. (SWV) uses symmetrical square-wave forward and backward on a , yielding net peak currents that reflect the difference between forward and reverse , offering even higher (up to 10^{-9} M) and faster rates (up to 100 V/s) due to effective rejection of background currents. These variants produce asymmetric or symmetric peak shapes depending on the pulse and frequency, improving signal-to-noise ratios over LSV or for complex matrices. Voltammetric methods are widely applied to determine formal redox potentials by identifying E_{1/2} or midpoints, providing thermodynamic data for like metal ions or organic couples, as seen in the characterization of where E^0 is established at +0.40 V vs. . They also elucidate reaction mechanisms through diagnostic ratios, scan rate dependencies, and follow-up scans; for instance, in , a decreasing anodic-to-cathodic current ratio with increasing scan rate signals an irreversible chemical step following , such as in the oxidation of where coupled alters the mechanism. These applications extend to studying electrode , adsorption effects, and multi-electron transfers in fields like and .

Polarography

Polarography is a classical electroanalytical technique that employs a dropping mercury (DME) to measure diffusion-limited currents as a function of applied potential, enabling the qualitative and of electroactive at levels. Developed by Jaroslav Heyrovský in 1922, the method involves electrolyzing a with a linearly increasing potential applied to the DME, which renews the electrode surface periodically through mercury drops, minimizing issues like adsorption and poisoning. In collaboration with Masuzo Shikata, Heyrovský constructed the first automated polarograph in 1924, a device that recorded current-potential curves photographically, revolutionizing electrochemical measurements. This technique draws on voltammetric principles where the observed current arises primarily from the reduction or oxidation of analytes diffusing to the electrode surface./25%3A_Voltammetry/25.05%3A_Polarography) The characteristic output of polarography is the polarographic wave, a sigmoidal current-potential curve where the current rises from a to a diffusion-limited plateau. The half-wave potential, E_{1/2}, marks the point where the current is half the limiting value and provides insight into the analyte's standard E^\circ, often approximated as E_{1/2} \approx E^\circ for reversible systems. For reversible waves, the relationship is described by the logarithmic analysis: \log \left( \frac{i_d - i}{i} \right) = \frac{n}{0.059} (E - E_{1/2}) at 25°C, where i_d is the , i is the measured current, n is the number of electrons transferred, and the slope confirms reversibility. The i_d is quantified by the Ilkovič , derived by Dionýz Ilkovič in 1934: i_d = 708 \, n \, D^{1/2} \, m^{2/3} \, t^{1/6} \, C where D is the coefficient (cm²/s), m is the mass flow rate of mercury (mg/s), t is the drop lifetime (s), and C is the concentration (mmol/L); this establishes the between i_d and C, forming the basis for . Polarography excels in the detection of such as , lead, and in environmental and biological samples, offering detection limits in the parts-per-million range due to the DME's high for hydrogen evolution and renewable surface. For instance, it has been applied to quantify in dental materials at concentrations below 1 , demonstrating its utility in trace-level assessments. To address limitations like the non-faradaic capacitive current from the expanding mercury drop, modern variants such as pulse apply potential pulses and sample the current near the end of each drop's life, where charging currents have decayed, thereby improving signal-to-noise ratios and lowering detection limits to sub-ppm levels.

Coulometric Methods

Controlled-Potential Coulometry

Controlled-potential coulometry is an electroanalytical technique that involves the exhaustive of an at a fixed , where the total charge passed through the is measured to determine the analyte's concentration. A potentiostat maintains the at a constant potential selected to selectively reduce or oxidize the target without interfering with other components. As the electrolysis proceeds, the decreases exponentially and approaches zero once the analyte is fully converted, allowing integration of the current-time curve to yield the total charge. This method ensures high selectivity due to the controlled potential, typically applied in unstirred or stirred solutions using electrodes such as gauze or mercury pools. The fundamental principle relies on , which relate the quantity of substance transformed to the amount of passed. The total charge Q is given by the equation Q = n F N where n is the number of electrons transferred per mole of , F is Faraday's constant ($96485 C/mol), and N is the moles of . In practice, Q is obtained by integrating the i(t) over the electrolysis time t: Q = \int_0^t i(t) \, dt This integration can be performed electronically or computationally, providing a direct measure of the quantity assuming 100% , where all charge contributes to the desired faradaic . For complete , efficiencies exceeding 99% are achievable under optimized conditions, such as sufficient stirring and appropriate potential selection. A calibration involves constructing a working curve of Q versus analyte concentration, which exhibits a linear relationship due to the stoichiometric proportionality in Faraday's law. This allows quantification without external standards in many cases, as the method is absolute. Applications include the determination of redox-active species, such as the reduction of Cu²⁺ to Cu⁰ using a mercury pool , where a suitable reducing potential around -0.25 V vs. ensures selective deposition. Other examples encompass the oxidation of As(III) to As(V) on electrodes and analysis of trace metals like or in nuclear materials, achieving precisions typically on the order of 0.1-2% depending on the system. The primary advantages of controlled-potential coulometry are its status as an absolute method requiring no curves for known stoichiometries and its high accuracy and selectivity for electroactive analytes. It is particularly valuable for purity assessments and mechanistic studies of electrochemical reactions. However, limitations include extended analysis times—often 30-60 minutes—for dilute solutions (e.g., <1 mM), where diffusion-limited currents decay slowly, and the need for corrections for background currents or incomplete efficiencies in complex matrices. Skilled operation is essential to maintain potential control and minimize interferences from secondary reactions.

Constant-Current Coulometry

Constant-current coulometry, also known as amperostatic coulometry, involves applying a fixed electrical through an to generate a titrant quantitatively at an , with the electrolysis time measured to determine the charge passed. The fundamental principle relies on Faraday's law of , which relates the charge Q to the produced or consumed: Q = n F N, where n is the number of electrons transferred per of , F is Faraday's constant (approximately 96,485 C/mol), and N is the number of moles of analyte. Since the I is constant, the charge is simply Q = I t, allowing precise calculation of the titrant equivalents from the product of current and electrolysis time until the endpoint is reached. In coulometric titrations, this technique enables the in situ electrogeneration of reagents, eliminating the need for pre-prepared titrant solutions and ensuring high accuracy due to the direct proportionality between charge and titrant amount. A common example is the generation of bromine from bromide ions at a platinum anode for the determination of arsenite, where As(III) is oxidized to As(V) by the electroproduced Br₂ in an acidic medium, with reactions proceeding stoichiometrically: $2 \text{Br}^- \rightarrow \text{Br}_2 + 2e^- and \text{AsO}_3^{3-} + \text{Br}_2 + \text{H}_2\text{O} \rightarrow \text{AsO}_4^{3-} + 2 \text{Br}^- + 2 \text{H}^+. This method achieves errors of less than 0.3% for microgram quantities of arsenic. Other titrants, such as iodine from iodide or cerium(IV) from cerium(III), can be generated similarly for redox titrations of analytes like antimony or uranium. Endpoint detection in constant-current coulometry typically employs a secondary indicator system, often amperometric or potentiometric, to monitor the sudden change in composition at . In amperometric detection, a twin- setup applies a fixed potential to sense the excess titrant's after the , producing a characteristic V-shaped curve. Potentiometric methods, using a reference and indicator pair, track potential shifts due to titrant accumulation, suitable for systems with well-defined potentials. These detection modes ensure sharp without direct measurement of the generated during . The development of constant-current coulometry is closely associated with the work of James J. Lingane in the 1950s, who pioneered its application in analytical chemistry through numerous studies on electrogenerated titrants for precise determinations. A notable application is the coulometric Karl Fischer titration for trace moisture analysis, where iodine is generated at constant current to react with water in a methanol-pyridine medium via \text{H}_2\text{O} + \text{I}_2 + \text{SO}_2 + 3 \text{RN} \rightarrow 2 \text{RNHI} + \text{RNSO}_3, enabling detection limits down to parts per million; commercial instruments based on this became available in the 1970s. This technique has found widespread use in water quality assessment and pharmaceutical analysis due to its automation potential and reagent stability. Recent advancements as of the 2020s include integration with automated flow systems for high-throughput environmental and industrial monitoring.

Conductometric Methods

Direct Conductometry

Direct conductometry is an electroanalytical technique that measures the electrical conductance of a solution to directly assess the total ionic concentration or strength, providing insights into the overall electrolyte content without relying on redox reactions or potential differences. The conductance G of the solution is defined as G = 1/R, where R is the measured electrical resistance, and it is related to the geometry of the measuring cell by G = \kappa A / l, with \kappa denoting the specific conductance, A the effective electrode area, and l the distance between electrodes. The specific conductance \kappa quantifies the solution's ability to conduct electricity and is given by \kappa = \sum \lambda_i c_i, where \lambda_i is the molar ionic conductivity of each ion species i and c_i its concentration. For strong electrolytes, the \Lambda (defined as \Lambda = \kappa / c, with c as the total ) follows Kohlrausch's law, which states that \Lambda = \Lambda_0 - K \sqrt{c}, where \Lambda_0 is the limiting molar conductivity at infinite dilution and K is an empirical constant accounting for interionic interactions. This relationship arises from the independent migration of s at low concentrations and allows direct conductometry to estimate parameters such as total or in aqueous solutions, as higher concentrations linearly correlate with increased conductance up to moderate levels. For instance, in , conductance measurements calibrated against known standards enable rapid assessment of seawater , where \kappa values around 50 mS/cm correspond to typical ionic strengths. To perform accurate measurements, (AC) is preferred over (DC) because AC at frequencies typically between 1 and 100 kHz reduces polarization, where ions accumulate at the surface and distort the . The cell constant l/A is determined by with standard (KCl) solutions of known concentrations (e.g., 0.01, 0.1, or 1 mol kg⁻¹), whose conductance values are standardized by organizations like IUPAC. Electrochemical cells for these measurements generally feature two electrodes coated with to increase the effective surface area and minimize polarization effects. Despite its simplicity and broad applicability, conductometry is limited by its to , with conductance typically increasing by approximately 2% per degree Celsius due to enhanced , necessitating precise (e.g., within 0.01 °C) during measurements. Additionally, the method is inherently non-specific, as it responds to the collective contribution of all present and cannot differentiate between individual species or their charge types, making it unsuitable for selective . These constraints are particularly evident in complex matrices with varying or at high concentrations, where deviations from Kohlrausch's law occur due to pairing and electrostatic interactions.

Conductometric Titrations

Conductometric titrations determine the of a by monitoring changes in the electrical conductivity of the solution as titrant is added. The conductivity reflects the total ionic concentration and mobility, which vary as ions are replaced during the . An () of , typically in the kHz range, is employed to measure conductance without inducing electrode polarization or Faradaic reactions at the s. This method was introduced in the late , with early work by Friedrich Kohlrausch and , and significant contributions from researchers like P. Dutoit, who applied it to determinations around 1910. In titrations involving a strong acid and strong base, such as HCl with NaOH, the initial high decreases as H⁺ ions (high mobility) are replaced by Na⁺, forming and NaCl, reaching a minimum at the . Beyond this point, excess OH⁻ ions cause to increase sharply, producing a characteristic V-shaped curve. For weak acid-strong base titrations, like acetic acid with NaOH, the curve shows an initial decrease in due to partial of the weak acid, followed by a minimum near where acetate ions predominate, and then a rise from excess OH⁻; the minimum is less pronounced due to effects. These curves allow endpoint identification at the inflection or minimum point. Precipitation titrations, such as Ag⁺ with Cl⁻ to form AgCl, exhibit stepwise conductance changes: initial high from Ag⁺ drops as low-mobility Cl⁻ replaces it, with a sharp break at equivalence when excess Cl⁻ further decreases due to the sparingly soluble precipitate removing ions from . Complexometric titrations, exemplified by EDTA with Ca²⁺, follow similar patterns, where the stable metal-EDTA reduces free concentration, leading to a conductance minimum at equivalence; the method is effective for divalent cations in ammoniacal buffers. For enhanced , especially in curves with shallow inflections, first-derivative plots of conductance change versus titrant are used, where the corresponds to the maximum or minimum in the derivative curve. This approach improves accuracy over direct curve inspection. Advantages over visual indicator methods include applicability to very dilute solutions (down to 10⁻⁴ M), colored or turbid samples, and reactions lacking suitable indicators, achieving better than 1% without needing initial conductance values from direct conductometry.

Advanced Electroanalytical Techniques

Electrochemical Impedance Spectroscopy

Electrochemical impedance spectroscopy (EIS) is a powerful electroanalytical technique that characterizes electrochemical interfaces by applying a small-amplitude sinusoidal potential over a wide range, typically from millihertz to megahertz, and measuring the resulting current response. This frequency-domain method provides insights into reaction kinetics, mass transport, and capacitive processes that are not easily accessible through steady-state techniques, as it separates contributions based on their characteristic time constants. Unlike time-domain methods such as , EIS yields non-steady-state data that reveal dynamic interfacial behaviors. The impedance Z in EIS is a complex quantity defined as Z(\omega) = Z' + j Z'', where Z' is the real part representing resistive components, Z'' is the imaginary part representing reactive components, j is the , and \omega = 2\pi f is the with f as the in hertz. Data are commonly visualized in Nyquist plots, which graph -Z'' versus Z' to display semicircles indicative of charge transfer and capacitive elements, or Bode plots, which plot the logarithm of the impedance magnitude \log |Z| and phase angle \theta versus \log f to explicitly show dependence and resolve overlapping processes. These representations facilitate the identification of time constants associated with interfacial phenomena. A foundational model for interpreting EIS data is the Randles equivalent circuit, consisting of the solution resistance R_s in series with a parallel combination of the charge-transfer resistance R_{ct} and a constant phase element (CPE), followed by the Warburg impedance W to account for diffusion. This circuit captures the essential elements of faradaic processes: R_s for ohmic drop, R_{ct} for electron transfer kinetics, the CPE for non-ideal capacitance due to surface heterogeneity, and W for semi-infinite linear diffusion, expressed as Z_W = \sigma (1 - j) / \sqrt{\omega} where \sigma is the Warburg coefficient. In Nyquist plots, the high-frequency intercept gives R_s, the semicircle diameter reflects R_{ct}, and the low-frequency 45° line arises from W. The CPE, with impedance Z_{CPE} = 1 / [Y_0 (j \omega)^n] where n (0 < n < 1) quantifies deviation from ideality (e.g., due to rough electrodes), replaces ideal capacitors to better fit real data. Equivalent circuit fitting involves nonlinear least-squares to match experimental data to proposed models, often using software like ZView, with validation via Kramers-Kronig transforms to causality and ; goodness-of-fit is assessed by the chi-squared \chi^2. The models diffusion-limited processes, appearing as a diffusive tail, while the CPE accounts for distributed relaxation times on inhomogeneous surfaces, improving accuracy over ideal components. EIS rose prominently in the with the advent of analyzers, such as the Solartron 1172, enabling measurements down to 0.1 mHz and computational fitting that addressed limitations of earlier methods restricted to higher frequencies, thus filling gaps in probing slow interfacial left by steady-state techniques. Applications of EIS span diverse fields, including studies where it quantifies pitting and passivation on metals like iron in acidic media by analyzing inductive loops from adsorbed species. In , EIS dissects solid-electrolyte and in lithium-ion cells, revealing performance degradation mechanisms through frequency-resolved semicircles. For interface analysis, EIS monitors biorecognition events, such as antigen-antibody binding, by tracking changes in R_{ct} with probes in Randles models, enabling label-free detection in systems like impedimetric immunosensors for pathogens. These uses highlight EIS's role in providing quantitative, mechanistic insights into interfacial dynamics.

Anodic Stripping Voltammetry

Anodic stripping voltammetry (ASV) is an ultrasensitive electroanalytical technique primarily used for the detection of ions, relying on a two-step process that enhances sensitivity through electrochemical preconcentration. In the first step, known as cathodic deposition or preconcentration, metal ions in the analyte solution are reduced and accumulated onto the surface of a at a constant negative potential (E_{dep}) for a fixed duration (t_{dep}), typically ranging from 1 to 30 minutes, often with stirring to promote mass transport. This deposition forms an amalgam or deposit proportional to the analyte concentration, governed by principles derived from . The second step involves anodic stripping, where the potential is scanned positively (e.g., using linear sweep or pulse voltammetry), oxidizing the deposited metal and producing a characteristic current peak (i_p) whose height is directly proportional to the deposition time and the original metal ion concentration (i_p \propto t_{dep} \times [M^{n+}]). This preconcentration step can amplify detection sensitivity by up to 10,000-fold compared to non-accumulation methods, enabling quantification at sub-ppb levels. Commonly employed electrodes include the hanging mercury drop electrode (HMDE), which provides a renewable, smooth surface ideal for amalgam formation with metals like copper, lead, cadmium, and zinc, though its use has declined due to mercury toxicity concerns. Alternatives such as thin-film electrodes, including mercury film electrodes (MFEs) or solid-state options like bismuth, tin, or gold films deposited on glassy carbon substrates, offer similar performance without mercury, extending applicability to non-amalgam-forming elements like arsenic or silver. A widely adopted variant is differential pulse anodic stripping voltammetry (DPASV), which applies a series of potential pulses during the stripping phase to minimize capacitive currents and improve signal resolution, achieving detection limits in the ppb range (e.g., 1-10 ppb for lead and cadmium) with high selectivity in complex matrices. These electrode configurations, combined with controlled deposition parameters, allow for simultaneous multi-element analysis while maintaining low limits of detection, such as 8 × 10^{-11} mol L^{-1} for thallium with a deposition time of 180 s. ASV finds extensive applications in , particularly for detecting toxic such as lead () and () in natural waters, sediments, and at concentrations relevant to regulatory limits (e.g., <10 ppb for in ). It addresses modern trace analysis needs by enabling field-portable instrumentation for real-time assessment of pollution sources, such as industrial effluents or atmospheric deposition, and has been integrated into hyphenated systems like ASV-ICP-MS for enhanced . from co-existing ions or , which can adsorb on the or alter deposition efficiency, is mitigated through techniques like medium exchange—replacing the sample solution with a clean (e.g., acetate buffer) post-deposition to remove effects—or by adding chelating agents to selectively complex interfering species. ASV's development accelerated in the with contributions from researchers like J. F. van der Pol and others, building on earlier polarographic foundations to establish it as a standard for ultrasensitive detection.

References

  1. [1]
    None
    ### Definition and Basic Principles of Electroanalytical Chemistry
  2. [2]
    Electroanalytical Method - an overview | ScienceDirect Topics
    Electroanalytical methods are defined as techniques that evaluate electrical quantities, such as electric current and applied potential, in relation to ...
  3. [3]
    A synthetic chemist's guide to electroanalytical tools for studying ...
    A range of electroanalytical tools can be applied to studying redox reactions, probing key mechanistic questions in synthetic chemistry.
  4. [4]
    A review on analytical methods with special reference to ...
    Jul 1, 2021 · This review aims to provide basic principles of chromatography, spectroscopy, and electroanalytical methods for the analysis of anticancer drugs published in ...
  5. [5]
    Three-Electrode Setups - Pine Research Instrumentation
    Sep 26, 2024 · In a traditional three-electrode cell, three different electrodes (working, counter, and reference) are placed in the same electrolyte solution.
  6. [6]
    [PDF] The Potentiostat and the Voltage Clamp - The Electrochemical Society
    In the early 1940s, Archie Hickling at the. University of Leicester, England, who was working in the field of electrochemistry, invented the potentiostat and ...
  7. [7]
    Construction and Working of Dropping Mercury Electrode & Rotating ...
    The dropping mercury electrode (DME) is a silver electrode used in polarography, with its capillary tubes used to continuously feed mercury into the ...
  8. [8]
    Construction of Ag/AgCl Reference Electrode from Used Felt-Tipped ...
    Here we report a simple, easy-to-fabricate, inexpensive, reliable, unbreakable, and reproducible Ag/AgCl reference electrode.
  9. [9]
    [PDF] Chapter 22 An Introduction to Electroanalytical Chemistry
    ✓ Except potentiometric methods all electroanalytical methods involve electrical currents and current measurements. We need to consider the behavior of cells ...
  10. [10]
    [PDF] Electrochemistry: The Nernst Equation
    Electrochemical Oxidation-Reduction Reactions. Reduction on the Right. Page 4. Nernst Equation Derivation. ΔG = − nFE cell. ΔG = ΔG0 + RT lnQ. Oxidation.
  11. [11]
    standard hydrogen electrode (S05917) - IUPAC Gold Book
    For solutions in protic solvents, the universal reference electrode for which, under standard conditions, the standard electrode potential ( H A + / H A 2 ) ...
  12. [12]
    [PDF] Standard Electrode Potentials and Temperature Coefficients in ...
    Oct 15, 2009 · Table 1 contains standard electrode potentials and tem- perature coefficients in water at 298.15 K for nearly 1700 half-reactions at pH ...
  13. [13]
    [PDF] ELECTROCHEMICAL METHODS
    Apr 5, 2022 · Electrochemical methods : fundamentals and applications / Allen J. Bard, Larry R. Faulkner.— 2nd ed. p. cm. Includes index. ISBN 0-471 ...Missing: excerpt | Show results with:excerpt
  14. [14]
    A Practical Beginner's Guide to Cyclic Voltammetry - ACS Publications
    Nov 3, 2017 · A short introduction to cyclic voltammetry is provided to help the reader with data acquisition and interpretation. Tips and common pitfalls are provided.
  15. [15]
    Potentiometry - an overview | ScienceDirect Topics
    Potentiometry is defined as a method used in electroanalytical chemistry to determine the concentration of a solute in solution by measuring the potential ...
  16. [16]
    11.2: Potentiometric Methods - Chemistry LibreTexts
    Sep 11, 2021 · In potentiometry we measure the potential of an electrochemical cell under static conditions. Because no current—or only a negligible current— ...
  17. [17]
    [PDF] All-Solid-State Ion-Selective Electrodes: A Tutorial for Correct Practice
    Oct 15, 2021 · At ambient temperature, an ideal ISE exhibits a Nerns- tian slope of 59.2 mV/decade for monovalent ions, and of. 29.5 mV/decade for divalent ...
  18. [18]
    [PDF] POTENTIOMETRY
    For example, at room temperature a 59.1-mV change in the electrode potential should result from a 10-fold change in the activity of a monovalent ion (z = 1).
  19. [19]
    [PDF] The Glass pH Electrode - The Electrochemical Society
    Eglass electrode = E' – RT/2.303F pH. (6). The Eglass electrode is measured with a pH meter, a voltmeter with circuitry allowing direct readout of pH. The ...
  20. [20]
    Electrode for Sensing Fluoride Ion Activity in Solution - Science
    Electrodes constructed from single-crystal sections of rare earth fluorides respond to fluoride ion activity over more than five orders of magnitude.
  21. [21]
    [PDF] 8.3.2.3 Constants and symbols Equation for emf Response of Ion ...
    The Nikolsky-Eisenman equation: E is the experimentally determined galvanic potential difference of ISE cell (in V) when the only variables are activities in ...
  22. [22]
    Determination of the equivalence point in potentiometric titrations ...
    Determination of the equivalence point in potentiometric titrations. Part II. G. Gran, Analyst, 1952, 77, 661. DOI: 10.1039/AN9527700661.Missing: original | Show results with:original
  23. [23]
    Potentiometric titrations using Gran plots: A textbook omission
    Describes Gran's graphical method of end point determination for the precise analysis of acids and bases.
  24. [24]
    Application of modified Gran functions and derivative methods to ...
    This paper presents an evaluation of the modified Gran functions (MGFs) and derivative methods for assessing the concentrations of inorganic carbon species, ...
  25. [25]
    Critical evaluation of potentiometric redox titrations in enology
    This review will focus on voltammetric techniques that have been developed for the electrochemical characterisation of wines, including modified electrodes.
  26. [26]
    Titration Methods: Manual vs. Potentiometric vs. Thermometric
    Feb 24, 2023 · Potentiometric titration can be used for virtually any type of titration reaction (acid-base, redox, precipitation and complexometric), but ...
  27. [27]
    Potentiometric Titration: Learn Definition, Types & Advantages
    It has a few limitations in range. · Potentiometric titration is highly dependent on the sensitivity of the electrode. · The electrodes used in potentiometric ...
  28. [28]
    Amperometry - an overview | ScienceDirect Topics
    Amperometry is defined as an electrochemical method used to determine the electric current generated by the oxidation or reduction of an analyte at an electrode ...
  29. [29]
    An electrochemical method for measurement of mass transport in ...
    For a membrane covered planar electrode in the steady state method, a final steady state current Is = nFADC0/L (n is electrons transferred, F is Faraday ...
  30. [30]
    Amperometric Method - an overview | ScienceDirect Topics
    The amperometric method measures steady-state current from a constant potential applied to an electrode, where current depends on the concentration of oxidized ...
  31. [31]
    Influence of viscosity on chronoamperometry of reversible redox ...
    The current transient response in single potential step experiment is given by Cottrell equation[66], I C ( E , t ) = n F D O A 0 C s ( E ) π D O t where n is ...
  32. [32]
    Chronoamperometry - an overview | ScienceDirect Topics
    In this context, the Cottrell equation is the most useful equation in chronoamperometry due to the ET process over the period respecting the electrode at a ...
  33. [33]
    Calibration-Free Analysis with Chronoamperometry at Microelectrodes
    Current values calculated using the Cottrell equation (blue dots), and the Mahon and Oldham equation for short times (black dots) and for long times (red dots).
  34. [34]
    Development of Amperometric Glucose Biosensor Based on ...
    Nov 4, 2014 · However, the oxidase-based amperomertic biosensors are sensitive to the variable oxygen concentrations present in the solution. In addition, a ...
  35. [35]
    A Set of Undergraduate Laboratory Titration Experiments
    May 22, 2025 · A set of four amperometric titration experiments for the determination of iron in dietary supplement tablets, permanganate in a pond cleaner liquid, iodine in ...
  36. [36]
    Amperometry - an overview | ScienceDirect Topics
    Amperometry is an electroanalytical technique measuring the current resulting from the electrochemical oxidation or reduction of an electroactive species at a ...
  37. [37]
    Dissolved oxygen: the electroanalytical chemists dilemma
    Dissolved oxygen is electrochemically reduced at potentials less negative than −0.05 V vs. standard calomel electrode (SCE) and, therefore, interferes with ...Missing: limitations | Show results with:limitations<|control11|><|separator|>
  38. [38]
  39. [39]
    [PDF] Jaroslav Heyrovsky - Nobel Lecture
    To accelerate the plotting of the curves we have constructed with Shikata in 19241 an automatic device, the "polaro- graph" (Fig. 3) by rotating the Kohlrausch ...Missing: Masao invention 1920s
  40. [40]
    THE POLAROGRAPH | Jaroslav Heyrovský | NJH (English)
    Jan 31, 2006 · In order to accelerate the measurements with dropping mercury electrode Heyrovský with his Japanese coworker Masuzo Shikata constructed an ...Missing: Masao invention 1920s
  41. [41]
    Polarography Can Successfully Quantify Heavy Metals in Dentistry
    In general, polarography is a type of voltammetry where the working electrode is a dropping mercury electrode (DME) to monitor changes in the current upon the ...
  42. [42]
    Normal Pulse Polarography - an overview | ScienceDirect Topics
    The pulse increases in potential with every drop. The current is sampled just before the end of the drops lifetime, minimizing the effect of the capacitance ...
  43. [43]
    [PDF] Coulometric Methods of Analysis
    This method employs two or three electrodes, and either a constant current or a constant potential is applied to the pre-weighed working electrode. ♤ Coulometry ...
  44. [44]
    [PDF] Pharmaceutical Analytical Chemistry: Open Access
    Constant potential coulometry is more commonly used than constant current coulometry, as it allows for better control of the reaction conditions. Coulometry can ...<|control11|><|separator|>
  45. [45]
    [PDF] Electroanalysis and Coulometric Analysis - Allen J. Bard
    Controlled potential coulometry con- tinues to find application to the ... current or titration efficiency from 100% will be the most serious limitation to very ...<|control11|><|separator|>
  46. [46]
    The Coulometric Titration of Arsenic by Means of Electrolytically ...
    Lingane, Allen J. Bard. Coulometric titration of antimony with electrogenerated iodine. Analytica Chimica Acta 1957, 16 , 271-273. https://doi.org/10.1016 ...
  47. [47]
    A sensitive amperometric endpoint detection system for ...
    An indirect amperometric endpoint detection ... Potentiometric determination of nanogram quantities of halides at extreme dilution by constant current coulometry.
  48. [48]
    Constant Current Source for Coulometric Titrations - ACS Publications
    James J Lingane. Coulometric titration of gold with electrogenerated chloro-cuprous ion. Analytica Chimica Acta 1958, 19 , 394-401. https://doi.org/10.1016 ...
  49. [49]
    [PDF] Karl Fischer Titration - School experiments - Mettler Toledo
    1970. First coulometric KF titration instruments are commercially available. 1980. Pyridine-free KF reagents are commercially available. 1984. First ...
  50. [50]
    [PDF] Principles of Electrochemistry Second Edition
    Oct 5, 2021 · Chapter 1 Equilibrium Properties of Electrolytes. 1. 1.1 Electrolytes: Elementary Concepts. 1. 1.1.1 Terminology.<|control11|><|separator|>
  51. [51]
    Electrical Conductivity Measurement of Electrolyte Solution - J-Stage
    The specific electrical conductivity of each material is defined as κ (or σ) (S m−1) and is calculated by the following equation. The specific resistance, ρ, is ...
  52. [52]
    [PDF] Conductivity Theory and Practice
    Conductivity measurements can only be performed after calibration, as the cell constant value is used to calculate the conductivity. Low conductivity ...Missing: electrochemistry | Show results with:electrochemistry
  53. [53]
    [PDF] Conductometric Titrations - CORE
    TITRATION OF A WEAK ACID WITH A WEAK BASE. Here the full value of the conductometric method of titration is brought out. It is very difficult to titrate a ...
  54. [54]
    Systemical investigation on the determination of sulfate in cement ...
    Nov 15, 2022 · Dutoit in 1910 first recommended the conductometric titration to determine sulfate in solution [23]. Anderson determines the sulfate in solution ...
  55. [55]
    [PDF] 6.2.4 Conductometric Titrations
    The principle of conductometric titration is based on the fact that during the titration, one of the ions is replaced by the other and invariably these two ...Missing: history Dutoit von Wattenwyl
  56. [56]
    [PDF] Conductometric Titration for Determination of Trazodone ...
    equivalence point. The first derivative data were fitted to a built-in nonlinear regression model while approximations to Gaussians of the second derivative.
  57. [57]
    Electrochemical Impedance Spectroscopy A Tutorial
    The user is allowed to change the values of the circuits' components and to see how the respective Nyquist and (magnitude and phase) Bode plots are transformed.
  58. [58]
    Basics of Electrochemical Impedance Spectroscopy
    The Bode Plot for the electric circuit of Figure 4 is shown in Figure 5. Unlike the Nyquist Plot, the Bode Plot does show frequency information.
  59. [59]
    [PDF] A Brief History of Electrochemical Impedance Spectroscopy
    Over the past two decades, electrochemical impedance spectroscopy (EIS) has emerged as the most powerful of electrochemical techniques for defining reaction.
  60. [60]
    Anodic Stripping Voltammetry - an overview | ScienceDirect Topics
    Anodic stripping voltammetry is defined as an electrochemical technique used for the determination of ionic metals, involving the electrodeposition of the ...Missing: seminal papers
  61. [61]
    Anodic stripping voltammetry | Journal of Chemical Education
    Considers the theory, techniques, applications, and instrumentation of anodic stripping voltammetry.Missing: principles seminal
  62. [62]
    Thin Film Electrodes for Anodic Stripping Voltammetry: A Mini-Review
    Feb 2, 2022 · Anodic stripping voltammetry (ASV) is an important electrochemical technique with a history stretching back to the earliest days of ...
  63. [63]
    Metals Analysis by Anodic Stripping Voltammetry
    Aug 29, 2023 · You have seen that ASV is an extremely sensitive technique, with detection limits commonly in the ppb range. Part per trillion detection limits ...Learning Outcomes · Purpose · Theory · Sample Preparation and Data...
  64. [64]
    Addressing the practicalities of anodic stripping voltammetry for ...
    Oct 15, 2019 · Anodic Stripping Voltammetry (ASV) has the capability to detect heavy metals at sub ppb-level with portable and cheap instrumentation.Missing: seminal | Show results with:seminal
  65. [65]
    Anodic stripping voltammetry with medium exchange in trace ...
    Medium exchange, where the test solution is replaced after electrodeposition by a simple electrolyte such as acetate buffer, before the oxidation (stripping) ...