Fact-checked by Grok 2 weeks ago

Liquid junction potential

Liquid junction potential, also known as diffusion potential, is the electrical potential difference that arises at the between two solutions of different compositions when they are in direct contact. This potential develops due to the unequal rates of cations and anions across the , leading to a temporary charge separation as faster-moving ions outpace slower ones. For example, in a junction between dilute and concentrated solutions of the same , such as 0.01 M NaOH and 0.001 M NaOH, the higher mobility of ions (approximately five times faster than sodium ions) contributes to the potential buildup. The magnitude of the liquid junction potential typically ranges from 1 to 40 millivolts, depending on factors such as ion mobilities, concentrations, and the geometry of the junction. In electrochemical cells, it manifests as an additive error in measured (), which can significantly impact the accuracy of potentiometric determinations, including measurements and analyses. For instance, in operational cells, the junction between the test solution and the reference electrode's filling solution (often saturated KCl) introduces a residual liquid junction potential that must be accounted for in standard conventions like the Bates-Guggenheim approximation. Liquid junction potentials are particularly pronounced in non-aqueous or mixed-solvent systems, where differences in solvation and properties across the can generate potentials as high as 90 mV, such as between and electrolytes. This phenomenon is critical in fields like electroanalytical chemistry, bioelectrochemistry (e.g., patch-clamp recordings), and conversion devices, where uncompensated potentials can lead to misinterpretation of behaviors or membrane potentials.90031-0) To mitigate these effects, strategies include employing salt bridges filled with equitransferent electrolytes like KCl to equalize fluxes, using computational tools for potential corrections (e.g., Henderson or ), or designing cells with minimal concentration gradients.

Fundamentals

Definition

The liquid junction potential is the potential difference that arises at the between two solutions of different compositions or concentrations, resulting from the unequal mobilities of ions diffusing across the boundary. This charge separation occurs because faster-moving ions outpace slower ones, creating a transient that balances the unequal diffusive fluxes, establishing a steady-state potential difference. Typically, the magnitude of this potential ranges from a few millivolts to about 40 mV, depending on the specific electrolytes and their concentration gradients. It commonly appears in galvanic cells and potentiometric measurements where direct contact between dissimilar solutions is unavoidable, thereby contributing to the measured of the cell. The phenomenon was first described in the late during studies of concentration cells by , who identified it as a key factor in differences in 1888.

Physical Origin

The liquid junction potential arises from the unequal rates of cations and anions across the between two solutions, leading to a transient charge separation that generates an . When two solutions of differing ionic compositions come into contact, ions begin to diffuse down their concentration gradients. However, because cations and anions typically possess different mobilities—often due to variations in size, hydration, or charge— one species migrates faster than the other, resulting in a localized excess of charge on one side of the junction. This charge imbalance creates an opposing that eventually balances the diffusive flux, establishing a steady-state potential difference known as the diffusion potential. The magnitude of this potential is influenced primarily by the mobilities of the involved, as well as their concentrations, valences, and the overall compositions of the solutions. Faster-moving , such as hydrogen in acidic media, exacerbate the charge separation and thus amplify the . Two main types of junctions contribute to this phenomenon: concentration junctions, where the same is present but at different concentrations (e.g., 0.01 M HCl | 0.1 M HCl), and composition junctions, where different are involved, either sharing a common (e.g., 0.1 M HCl | 0.1 M KCl) or not (e.g., 0.05 M KCl | 0.05 M NaNO₃). In both cases, the potential develops rapidly upon contact and persists as long as the diffusion gradient exists. Thermodynamically, the liquid junction potential reflects the driving forces associated with ion mixing and diffusion, stemming from gradients in chemical potential that drive the non-equilibrium transport of ions across the interface. Although the junction does not represent a true thermodynamic equilibrium—lacking zero net flux—it can be linked to changes in arising from the entropy of dilution and the electrochemical potentials of the species involved. This basis underscores how the potential opposes further net ion transfer until a dynamic balance is achieved, with the overall process governed by the principles of irreversible thermodynamics.

Theoretical Framework

Henderson Equation

The Henderson equation provides the foundational theoretical framework for estimating the liquid junction potential arising at the between two solutions of differing . Originally derived by Henderson in 1907, it stems from the integration of the Nernst-Planck equations describing ionic fluxes under steady-state conditions with zero net current. The derivation assumes a sharp boundary that diffuses over time, leading to linear concentration profiles across the junction; the ionic fluxes are then balanced such that the total current vanishes, allowing the to be expressed in terms of concentration gradients and transport properties. This results in an expression for the potential difference obtained by integrating the across the junction region. The general form of the Henderson equation for a multi-ionic system is given by E_j = -\frac{RT}{F} \int_1^2 \sum_i \frac{t_i}{z_i} \, d \ln a_i where R is the , T is the absolute temperature, F is the , t_i is the transport number of i, z_i is its charge, and a_i is its activity, with the taken from solution 1 to solution 2. For a simple 1:1 binary (e.g., MX where M⁺ and X⁻ have constant mobilities u_+ and u_-), the equation simplifies under the assumption of linear gradients to the approximate form E_j \approx \frac{RT}{F} \frac{u_- - u_+}{u_- + u_+} \ln \left( \frac{c_2}{c_1} \right) where c_1 and c_2 are the concentrations on either side of the junction (activities may replace concentrations for greater accuracy). Transport numbers t_\pm relate to mobilities via t_+ = u_+ / (u_+ + u_-) and t_- = u_- / (u_+ + u_-) for univalent ions. Key assumptions underlying the include dilute solutions where activity coefficients are unity or constant, no convective flows, a planar junction with linear concentration gradients, and constant ionic mobilities independent of concentration and unaffected by interactions with other ions. These hold reasonably for small concentration differences, typically yielding errors below 1 mV in simple systems like KCl dilutions. However, the equation has notable limitations: it becomes inaccurate for high concentrations where vary or s deviate significantly, and for multivalent ions where the assumptions of constant transport numbers fail. Initially, it neglects variations, though later refinements incorporate mean ionic activities; errors can reach several millivolts in complex mixtures like those involving divalent cations.

Advanced Models

Planck's exact equation provides a rigorous theoretical foundation for calculating liquid junction potentials by fully integrating the Nernst-Planck equations without the simplifying assumptions of dilute solutions or linear concentration gradients inherent in earlier models like Henderson's. This approach accounts for the diffusion of all ionic species across the junction, yielding the potential difference as E_j = \frac{RT}{F} \int \sum_i \frac{t_i}{z_i} \frac{d\mu_i}{RT}, where R is the , T is , F is Faraday's constant, t_i is the transference number of ion i, z_i is its charge, and \mu_i is its , integrated over the junction from one solution to the other. The equation captures the contributions from all ions, making it suitable for systems where ionic interactions and varying mobilities play significant roles, though it requires numerical evaluation for complex cases due to the form. For concentrated electrolyte solutions, where non-ideal behaviors dominate, the Pitzer model has been integrated into liquid junction potential calculations to accurately incorporate activity coefficients through ion-specific interaction parameters. This semi-empirical approach uses coefficients (\beta^{(0)}, \beta^{(1)}, and C^\phi) to model short- and long-range electrostatic interactions, enabling precise prediction of transference numbers and potentials in high-ionic-strength mixtures without assuming ideality. By replacing mean ionic activities in the Planck with Pitzer-derived values, the model extends applicability to brines and electrolytes, improving accuracy over dilute approximations in concentrated systems. Numerical simulations have advanced the modeling of liquid junction potentials in non-ideal conditions, particularly for junctions involving , irregular geometries, or time-dependent dynamics, using methods like finite element and solutions to the coupled Nernst-Planck-Poisson equations. Finite element methods discretize the junction to resolve local and ion fluxes, providing detailed profiles of concentration gradients and potentials that analytical models overlook, such as in ion-selective membranes or microchannels. simulations complement these by offering atomistic insights into ion and short-range interactions at the junction, though they are computationally intensive and typically used for validation rather than routine prediction. Refinements for multivalent ions and mixed electrolytes, developed primarily in the 1980s and 2000s, build on variants of the Scatchard equation to handle complex transference in systems, incorporating higher-order interaction terms for accurate activity and corrections. These extensions address limitations in binary approximations by parameterizing mixing effects for ions with differing valences, such as in CaCl₂-KCl junctions, where multivalent contributions can amplify potentials by factors of 2–3 compared to monovalent cases. Such models prioritize ion-pairing and specific short-range forces, enhancing predictive power for geophysical and biochemical applications involving diverse compositions.

Quantification

Calculation Methods

The calculation of liquid junction potential begins with identifying the relevant ions in the two solutions forming the junction and obtaining their transport numbers, which are derived from ionic mobilities or limiting conductivities reported in literature databases such as those compiled by Robinson and Stokes. Transport numbers t_i for each ion i are calculated as the fraction of total current carried by that ion, t_i = \frac{c_i u_i |z_i|}{\sum_j c_j u_j |z_j|}, where c_i is the concentration, u_i is the mobility, and z_i is the charge. Next, select an appropriate model, such as the Henderson equation for dilute solutions or more advanced formulations for concentrated systems, referencing the theoretical framework established in seminal works. Input the concentrations or activities of ions in each solution, adjusting for ionic strength using Debye-Hückel theory to compute mean activity coefficients if necessary. For non-linear concentration gradients, numerically integrate the potential expression over the junction path using methods like Simpson's rule or finite difference approximations implemented in computational tools. A representative example is the junction between 0.1 M HCl and 0.1 M KCl at 25°C. Using the Henderson approximation with transport numbers t_{\ce{H+}} \approx 0.83, t_{\ce{Cl-}} \approx 0.17 in HCl, and t_{\ce{K+}} \approx 0.49, t_{\ce{Cl-}} \approx 0.51 in KCl, the calculated potential is approximately +27 mV, with the positive sign indicating the HCl side is at higher potential due to the high mobility of H⁺ ions. Dedicated software facilitates these computations, particularly for multi-ion systems. Programs like JPCalc implement the generalized Henderson equation, allowing users to input ion concentrations, mobilities, and temperature to automatically compute the potential and apply corrections for electrophysiological recordings. For complex geochemical or environmental applications, PHREEQC can be used to first calculate activities and speciation under varying ionic strengths, which are then fed into custom scripts or Excel-based calculators for the junction potential integration. These tools often include built-in mobility databases to streamline the process. Common errors in these calculations arise from neglecting temperature dependence, as vary significantly with (e.g., increasing by about 2% per °C), leading to inaccuracies up to 5-10 mV in non-isothermal junctions. Similarly, ignoring effects can overestimate potentials by failing to account for activity corrections, especially in solutions above 0.1 M where Debye-Hückel assumptions break down.

Experimental Measurement

The experimental measurement of liquid junction potential typically employs an configuration with two identical reversible electrodes flanking the junction, such as silver-silver (Ag/AgCl) electrodes immersed in solutions of varying concentrations. In this setup, known as a concentration , the total electromotive force () measured across the includes contributions from both the differences due to the concentration gradients and the liquid junction potential. The junction potential is isolated by subtracting the theoretical concentration (calculated assuming no junction, based on Nernstian response to activity differences) from the measured under zero current conditions. For instance, a of the form Ag/AgCl | KCl (c₁) || KCl (c₂) | Ag/AgCl allows determination of the junction potential in this manner. To ensure accuracy, calibration methods involve comparisons with reference cells featuring negligible or zero junction potentials, such as those using high-concentration bridges that minimize gradients, or techniques from measurements across a series of concentration ratios approaching unity. Historical approaches from the , including Niels Bjerrum's method, relied on plotting EMF values against concentration differences to linearly extrapolate to the zero-junction condition, providing validation against theoretical predictions. These methods have been refined to achieve precisions of 0.1 mV or better, often cross-verified by subtracting calculated contributions from total in asymmetric cells. Instrumentation for these measurements requires high-impedance voltmeters or potentiometers with input resistances exceeding 10^{12} Ω to prevent current flow that could distort the potential, ensuring negligible at the electrodes. Challenges in such experiments include preventing contamination from impurities that alter mobilities, which can be mitigated by using freshly prepared solutions and porous barriers like frits; additionally, gradients across the must be controlled to below 0.1°C to avoid thermo-electric effects contributing spurious voltages. Early 20th-century methods, evolving from the , emphasized flowing junctions to reduce layers and , laying the foundation for modern precision techniques.

Mitigation Strategies

Salt Bridges

Salt bridges serve as a primary method to minimize liquid junction potential in electrochemical cells by incorporating a high-concentration solution of an equitransferent , such as (KCl), where the transport numbers of the cation (K⁺) and anion (Cl⁻) are approximately equal at 0.5. This near-equality in ionic mobilities ensures that the rates of the ions are balanced, thereby reducing the potential difference that arises at the interfaces between the salt bridge and the solutions in the half-cells. The use of saturated or highly concentrated KCl further diminishes the junction potential by swamping the ionic gradients with a uniform, high-conductivity medium. Various designs of salt bridges are employed to facilitate ionic conduction while restricting convective mixing of solutions. Agar gel bridges, formed by gelling the electrolyte with agar, provide a semi-solid pathway that prevents bulk flow. Fritted glass bridges utilize porous glass disks as barriers, allowing slow ion diffusion through fine pores. Fiber junctions, often made from ceramic or asbestos fibers saturated with the electrolyte, offer a flexible alternative for compact setups. Saturated KCl remains the standard electrolyte in these designs, typically yielding residual liquid junction potentials below a few millivolts. KCl is the preferred material for salt bridges due to the close matching of K⁺ and Cl⁻ mobilities, which outperforms alternatives like NaCl where Na⁺ mobility is about 70% that of K⁺, leading to greater potential imbalances. In cases where chloride ions may interfere with sensitive analytes, such as silver-containing solutions, potassium nitrate (KNO₃) is selected as an alternative for its similar equitransferent properties without introducing Cl⁻. Overall, salt bridges effectively reduce liquid junction potentials by 90–99% in most applications, a technique historically adopted in the early 1900s to enable precise potentiometric measurements in electrochemistry.

Alternative Techniques

Junction-free designs represent innovative approaches to eliminate liquid junction potentials by avoiding direct interfaces altogether. One prominent method employs ionic liquid-based electrodes, where membranes incorporating ionic liquids such as bis(trifluoromethanesulfonyl)imide with imidazolium or cations define the interfacial potential through limited ion partitioning, thereby bypassing traditional liquid junctions. These electrodes, available in liquid-contact or solid-contact configurations, demonstrate signal comparable to Ag/AgCl references and perform effectively in potentiometric titrations for ions like Pb²⁺ and measurements. Another technique utilizes hydrophobic ion-doped polymeric membranes in current pulse-based electrodes, where brief electrical pulses release controlled amounts of low-mobility cations (e.g., ions) into a thin Nernst layer adjacent to the sample, preventing significant mixing while maintaining a stable potential independent of sample composition. This design offers advantages over salt bridges, including reduced clogging, minimal leakage, and high (e.g., <0.5 mV deviation over days in biological fluids like ). Flowing junctions provide dynamic control over layers to suppress junction potentials without static bridges. In flow-through sensors, spatiotemporal manipulation of —such as inducing via low-frequency vibration (e.g., 2 Hz) in narrow channels—mixes adjacent solutions briefly to disrupt charge separation, stabilizing measurements with times reduced to 15 seconds and errors below 0.06 units. Microfluidic implementations further refine this by integrating free- junctions with continuous replenishment of internal via convective in nanochannels, achieving potential drifts as low as 1 mV over 100 hours and insensitivity to external or variations. These designs leverage precise to maintain junction stability, particularly in miniaturized electrochemical systems. Software compensation enables post-measurement correction of liquid junction potentials by calculating their contributions from known compositions and mobilities. Tools like JPCalcWin apply the generalized Henderson equation to estimate potentials (often 3–10 mV in electrophysiological setups) across configurations such as patch-clamp or intracellular recordings, allowing users to input concentrations, temperatures, and types for accurate adjustments (e.g., V_corrected = V_measured - E_j). This method is essential for in complex solutions, supporting up to multiple ions and validating corrections against experimental benchmarks, thereby improving precision without hardware modifications. Specialized electrolytes minimize liquid junction potentials through compositional matching that reduces ion mobility differences. Equimolal solutions, where cation and anion concentrations are equal across phases, limit diffusion gradients and thus potential errors, particularly in pH determinations when paired with stable bridges like ionic liquids. Matched-mobility salts, such as quaternary ammonium chlorides (e.g., tetrabutylammonium chloride), serve as low-interference alternatives in reference systems, preloading channels to counteract ionic strength variations and enable ionic strength-independent potentiometric sensing with deviations under 1 mV. These electrolytes are especially useful in low-conductivity or biological samples to avoid ion-specific artifacts. Emerging methods post-2000 harness to control diffusion layers at microscales. Microfluidic channels with integrated salt bridges maintain reproducible potentials (e.g., 19.3 ± 6 mV vs. Ag/AgCl) by confining flow and minimizing junction variability in compact devices. electrolytes, such as cross-linked poly() diacrylate networks infused with liquid salts, immobilize anions to suppress uneven transport, achieving conductivities over 1 mS/cm at 40°C and stable with reduced interfacial potentials in batteries. These approaches enhance portability and reliability in applications like sensors and .

Significance

Electrochemical Applications

In galvanic cells, the liquid junction potential (LJP) arises at the between solutions of differing compositions, contributing to deviations from the ideal (EMF) and necessitating corrections for accurate determination of standard potentials. This non-ideal contribution stems from differential ion mobilities across the junction, which generate a diffusion potential that alters the overall cell voltage, particularly in cells with transference where ions migrate under an . For instance, in electrochemical cells interfacing solutions via porous media, the LJP can distort measured EMFs, requiring theoretical models like the Nernst-Planck equations to quantify and subtract its effect for thermodynamic consistency. The historical significance of LJP traces back to Walther Nernst's foundational work in the on concentration cells, where he first derived quantitative expressions for potential differences driven by diffusion across junctions, laying the groundwork for modern . Nernst's 1888 analysis of potentials in such cells provided atomistic explanations that integrated LJP into the framework, influencing subsequent developments in understanding non-equilibrium electrochemical processes. In potentiometry, LJP significantly impacts ion concentration measurements by introducing an uncontrolled voltage offset at the reference-sample solution interface, often leading to errors in (ISE) readings. Within batteries and fuel cells, LJPs occur at junctions between multi-phase electrolytes or compartments with varying ion concentrations, resulting in internal voltage drops that reduce overall cell efficiency and output. In lithium-ion batteries, these potentials manifest as additional ohmic losses across electrolyte interfaces, contributing to diminished performance during charge-discharge cycles.

Practical Implications

In pH measurements using glass electrodes, the liquid junction potential arising between the sample solution and the filling solution can introduce significant errors, typically ranging from 5 to 20 mV, which equates to shifts of 0.1 to 0.3 units depending on the ionic composition and concentration differences. These errors are particularly pronounced in samples with low or extreme values, where diffusion rates exacerbate the potential mismatch. To mitigate this, standard protocols employ reference electrodes filled with saturated (KCl), leveraging the near-equal mobilities of K⁺ and Cl⁻ ions to minimize the junction potential and stabilize readings. In environmental monitoring, liquid junction potentials pose challenges for sensors assessing seawater salinity and pH, where salinity variations can induce uncorrected errors of approximately 0.028 pH units over a 10-unit salinity change, leading to inaccuracies in oceanographic data critical for and studies. Similarly, in biological sensors for blood analysis, residual liquid junction potentials contribute to systematic errors, such as negative biases in (K⁺) measurements when chloride ions are substituted by , potentially affecting clinical diagnoses of imbalances. Industrial applications, particularly in chemical within , are highly sensitive to these potentials, as uncorrected liquid junction effects in high-ionic-strength streams can produce errors exceeding 0.5 units, resulting in delayed equilibration times of hours and compromised quality monitoring that disrupts production efficiency and product consistency.

References

  1. [1]
    liquid junction (L03584) - IUPAC
    Any junction between two electrolyte solutions of different composition. Across such a junction there arises a potential difference, called the liquid junction ...Missing: electrochemistry | Show results with:electrochemistry<|control11|><|separator|>
  2. [2]
  3. [3]
  4. [4]
  5. [5]
    Junction Potentials - Chemistry LibreTexts
    Aug 29, 2023 · A potential develops at any interface, or junction, where there is a separation of charge. For example, a potential can develop when a metal electrode comes in ...
  6. [6]
    Electrochemistry Encyclopedia -- Nernst
    Another of his important contributions was the discovery of the liquid junction potential , also published in the above journal in 1888 (available on the WWW).
  7. [7]
    [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 ...
  8. [8]
    [PDF] Liquid-junction potentials, and relative activity coefficients of chloride ...
    2 For work within 0.01 pH, therefore, it is desirable to know how to reduce the liquid-junction potential to less than 1 mv toward average and specific types ...<|control11|><|separator|>
  9. [9]
    None
    ### Summary of Henderson Equation Derivation for Liquid Junction Potential
  10. [10]
    [PDF] A new open source software for the calculation of the liquid junction ...
    Mar 24, 2014 · The calculation is made by solving the Nernst–Planck equations for the stationary state in conditions of local electrical quasi-neutrality at ...<|separator|>
  11. [11]
    Calculation of liquid junction potentials - ScienceDirect.com
    Equations and methods for the calculation of liquid junction potentials are reported. They are applicable to homoionic junctions over large molality ranges.
  12. [12]
    [PDF] A COMPUTER PROGRAM INCORPORATING PITZER'S ...
    PHRQPITZ calculates geochemical reactions in brines, including aqueous speciation, mineral solubility, mixing, irreversible reactions, and reaction path.
  13. [13]
    Computer Simulation of Ion-Selective Membrane Electrodes and ...
    Evidently, the liquid-junction potentials derived from the finite-element method are in excellent agreement with the values predicted either from Planck's or ...
  14. [14]
    Dynamic Theory of Type 3 Liquid Junction Potentials
    Mar 15, 2010 · A Nernst−Planck−Poisson finite difference simulation is used to model the dynamic evolution of a series of liquid junctions of the type A + X − |B + Y −.
  15. [15]
    Conversion of Parameters Among Variants of Scatchard's Neutral ...
    Aug 6, 2025 · In this report we present an analytical method for transforming the mixing parameters of neutral-electrolyte-type models with larger numbers of ...Missing: multivalent junction
  16. [16]
    JPCalc, a software package for calculating liquid junction potential ...
    The JPCalc program has been designed to graphically illustrate how junction potential contributions arise in various electrophysiological situations.Missing: tools PHREEQC
  17. [17]
    PHREEQC Version 3 | U.S. Geological Survey - USGS.gov
    PHREEQC Version 3 is a computer program written in the C++ programming language that is designed to perform a wide variety of aqueous geochemical calculations.PHREEQC Version 3 Downloads · Description of input and · Phreeqc3 htmlMissing: liquid junction potential
  18. [18]
    How Do Liquid-Junction Potentials and Medium Polarity at Electrode ...
    Feb 3, 2023 · In electrochemical cells, such liquid-junction potentials (ELJ) formed along the electrical paths between the working and the reference ...
  19. [19]
    A Simple Method for Decreasing the Liquid Junction Potential in a ...
    Apr 1, 2015 · The liquid junction potential (LJP), the phenomenon that occurs when two electrolyte solutions of different composition come into contact, ...
  20. [20]
    [PDF] Effects of corrections for liquid-junction potentials of saturated ...
    Errors in pH, dissociation constants, ionic strength, signs of "salt- ing-out" coefficients, and hydrogen-ion activity, due to neglect of or partial corrections ...
  21. [21]
    [PDF] Electrodes in Potentiometry - Monograph
    The potentiometric measurement is carried out practically currentless with a voltmeter with a high impedance measuring input (pH. Meter). U1: Galvani potential ...<|separator|>
  22. [22]
  23. [23]
    Low-temperature water electrolysis: fundamentals, progress, and ...
    May 17, 2022 · (ii) It is challenging to accurately determine the liquid junction potentials. (iii) The real E0Ref (versus SHE) may drift from ...
  24. [24]
  25. [25]
    Current Pulse Based Reference Electrodes Without Liquid Junctions
    Reference electrodes with hydrophobic ion-doped polymeric membranes offer a promising alternative to reference electrodes with conventional salt bridges.Missing: techniques mitigate excluding
  26. [26]
  27. [27]
    None
    ### Step-by-Step Process for Calculating Liquid Junction Potential Using JPCalcWin
  28. [28]
    [PDF] Accurate Determination of pH by Use of Ionic Liquid Salt Bridge ...
    Mar 26, 2012 · ... liquid junction potential (LJP) between the salt bridge and ... equimolal phosphate buffer solution at 5 - 60. ◦. C. The results suggest that.
  29. [29]
    Ionic strength-independent potentiometric cation concentration ...
    Nov 1, 2021 · In this work, a direct concentration output is realized for a potassium-selective electrode by using a TBA + (tetrabutylammonium)-selective membrane as the ...
  30. [30]
    A miniaturized electrochemical system with a novel polyelectrolyte ...
    The novel reference electrode system using a polyelectrolyte salt bridge on a microfluidic chip maintained a reproducible potential of 19.3 ± 6 mV versus a ...Missing: post- | Show results with:post-
  31. [31]
    Cross-Linked Gel Electrolytes with Self-Healing Functionalities for ...
    Nov 10, 2022 · We propose a gel electrolyte based on a uniform and highly cross-linked network, hosting a high amount of liquid electrolyte, with multiple advantages.
  32. [32]
    A closer look at the Nernst-Planck model for liquid junctions in ...
    A closer look at the Nernst-Planck model for liquid junctions in galvanic cells and the concept of EMF. December 2018; Journal of Electroanalytical Chemistry ...<|separator|>
  33. [33]
    Junction potentials in electrochemical cells with transference
    The junction potential is defined as the Galvani potential between two (bulk) phases. When both phases are liquid and miscible it is known as a LJP, generally ...
  34. [34]
    (PDF) Genesis of the Nernst Equation - ResearchGate
    Aug 10, 2025 · Nernst's three seminal contributions of 1888 and 1889 provided for the first time clear, atomistic explanations of, and quantitative expressions for, potential ...
  35. [35]
    Potentiometric Sensing | Analytical Chemistry - ACS Publications
    Oct 27, 2020 · As a consequence, the reference potential depends on the composition of the bridge electrolyte, the pore size of the frit, and the concentration ...Reference Electrodes · New Ion-Selective Readout... · Miniaturized Ion-Selective...
  36. [36]
    Overpotential from Cosolvent Imbalance in Battery Electrolytes - NIH
    May 26, 2023 · One standard method to quantify thermodynamic factors is based on measuring the so-called liquid-junction potential—the voltage drop across a ...Missing: fuel | Show results with:fuel
  37. [37]
    Ultrahigh voltage and energy density aluminum‐air battery based on ...
    Jul 20, 2020 · ... battery reduces to 1.87 V. The reason might be the liquid junction potential caused by the difference of concentrations for ions on both ...Missing: internal | Show results with:internal
  38. [38]
    [PDF] Manual: Theory and Practice of pH Measurement
    Liquid junction potential. A liquid junction potential exists when two dissimilar electrolyte solutions come into contact but are not allowed to mix. The ...
  39. [39]
    Spectrophotometric Calibration of pH Electrodes in Seawater Using ...
    The procedures developed in this study account for salinity-induced variations in liquid junction potentials that, if not taken into account, would create pH ...
  40. [40]
    Effects of residual liquid junction potential in direct potentiometry of ...
    Replacing Cl- with HCO3- leads to an increase in the RLJP, which in turn contributes to a negative error in K+ analysis. A small decrease in RLJP is observed as ...
  41. [41]
    Chemistry: What you don't know can hurt you | Control Global
    Dec 16, 2016 · In high-ionic-strength streams, the liquid junction potential can cause an error of 0.5 pH or more, and require hours to equilibrate (stop ...
  42. [42]
    Liquid Junction Potential (LJP) Theory and Correction
    LJPcalc calculates the liquid junction potential according to the stationary Nernst-Planck equation which is typically regarded as superior to the simpler ...<|separator|>