Fact-checked by Grok 2 weeks ago

Point of zero charge

The point of zero charge (PZC) is the value at which the net total charge on the surface of a solid material dispersed in an aqueous is zero, corresponding to a surface of zero. This condition arises when the activities of charge-determining ions (such as H⁺ and OH⁻) in the bulk solution result in balanced and of surface sites, eliminating any net electrostatic potential difference across the interface. Although often used interchangeably in practice, the PZC is distinct from the isoelectric point (IEP), which is the pH at which the electrophoretic mobility or of particles in suspension is zero. The PZC reflects the total charge on the entire surface (including inner and diffuse layers), whereas the IEP primarily indicates the charge at the shear plane, and discrepancies between the two can occur due to selective adsorption of ions like or that penetrate the electrical double layer. For many metal (hydr)oxides and simple systems without specific ion effects, the PZC and IEP coincide, but in complex materials like activated carbons or soils, the IEP may be lower than the PZC. The PZC is a critical in surface and chemistry, governing the electrostatic interactions that influence , dispersion stability, and the adsorption of ions or molecules onto surfaces. In and environmental applications, it determines the affinity of minerals (e.g., TiO₂ with PZC around 6–7.5) for pollutants, facilitating processes like and soil remediation. Similarly, in and , the PZC affects the binding of reactants to surfaces, enabling selective degradation of contaminants. Recent advancements highlight its role in , where tuning the PZC of materials like activated carbons optimizes charge storage mechanisms in supercapacitors and batteries operating in aqueous electrolytes. Common methods to determine the PZC include , which tracks changes upon acid or base addition to suspensions, and the salt addition method, which identifies the where added does not alter the suspension's equilibrium charge. These techniques, refined over decades since early reviews in the , provide values that vary widely by material—typically 7–10 for many metal oxides, 2–5 for carbons, and 2–5 for silicates—underscoring the PZC's sensitivity to surface composition and preparation.

Definitions and Terminology

Definition of Point of Zero Charge

The point of zero charge (PZC), denoted as pHpzc, is defined as the pH value at which the net surface charge of a solid material immersed in an aqueous solution is zero, arising from the balance between positively and negatively charged surface sites. This condition occurs when the concentrations of H+ and OH- ions adsorbed or dissociated from the surface sites equalize the overall charge, rendering the surface electrically neutral in the absence of specific ion adsorption. The concept of the point of zero charge originated in the early 20th century within colloid and surface science, building on foundational work in the electrical double layer theory. Key contributions came from Louis Georges Gouy in 1910 and David Leonard Chapman in 1913, who described the diffuse distribution of charges at interfaces, laying the groundwork for understanding surface neutrality conditions. The term itself gained prominence in studies of oxide surfaces during the mid-20th century, particularly through experimental investigations of charge balance in aqueous dispersions. Mathematically, at the PZC, the surface charge density \sigma = 0, where \sigma represents the net charge per unit area on the solid-liquid interface. At the PZC, the minimization of electrostatic repulsion between particles promotes aggregation and influences key surface interactions, such as the adsorption of ions or molecules and the stability of colloidal suspensions. This neutrality point is crucial for predicting how surfaces respond to variations, affecting phenomena like in or catalyst performance. For common metal oxides, representative PZC values include silica at approximately pH 2–3 and alumina at pH 8–9, reflecting differences in surface site acidity. The isoelectric point (IEP) is defined as the pH value at which a molecule, such as a protein or dispersed particle, exhibits zero net charge, commonly applied in biochemistry to characterize the charge state of macromolecules like proteins where positive and negative charges balance. In contrast, the point of zero charge (PZC) specifically refers to the pH at which the net surface charge density on a solid material, typically inorganic oxides or minerals, is zero, independent of specific adsorbates and focusing on the intrinsic surface properties without the influence of particle dispersion effects. A key distinction arises in their sensitivity to environmental factors: while the PZC remains relatively fixed for a given surface unless structurally altered, the IEP can shift due to adsorption of ions or other species onto particles, reflecting changes in the effective charge of the entire colloidal system rather than the bare surface alone. Another related concept is the point of zero salt effect (PZSE), which denotes the at which the influence of concentration on the surface charge or pH-dependent behavior becomes negligible, often coinciding with the PZC but differing in cases where ion-specific effects are prominent. Unlike the , which describes the suppression of by added ions of the same type in equilibria, the PZSE and PZC emphasize surface neutrality in the context of variations, without altering the fundamental constants of surface groups. These concepts find distinct applications across scientific fields: the PZC is predominantly used in and to predict adsorption on surfaces, such as in or interfaces, whereas the IEP is central in and science for assessing protein solubility or in suspensions. For instance, in the (α-FeOOH), the PZC is approximately 7.5–8.0, determined from surface-specific methods like , highlighting its role in fixed solid interfaces; in comparison, the IEP for particles in suspension ranges from 7 to 9, varying with measurement techniques like due to potential adsorption influences.

Abbreviations and Notation

In the literature on surface chemistry and colloid science, several standard abbreviations are employed to denote key concepts related to surface charge neutrality. The point of zero charge is commonly abbreviated as PZC, referring to the condition where the net surface charge density is zero. The isoelectric point, which marks the pH at which a particle's net charge is zero in a dispersion, is abbreviated as IEP. Additional terms include PZSE for the point of zero salt effect, the pH where surface charge remains independent of ionic strength, and PZNPC for the point of zero net proton charge, indicating zero proton-related surface charge. Notation conventions in PZC studies emphasize clarity in distinguishing pH values from other parameters. The lowercase pzc typically denotes the value at the point of zero charge, while the surface at this point is often symbolized as \sigma_0, representing zero net charge per unit area in contexts. Field-specific variations arise due to differing emphases across disciplines. In , particularly for mineral surfaces, PZC is standardly applied to describe charge neutrality in aqueous environments involving oxides and silicates. In , the notation E_pzc is prevalent for the (in volts) at zero charge, highlighting the interfacial potential rather than pH. The terminology has evolved from early 20th-century electrochemical concepts, such as "zero charge potential" introduced by Frumkin around 1930, to the modern standardization of PZC in the mid-20th century. This shift gained prominence post-1950s through influential reviews that unified terms across , promoting PZC for pH-based charge neutrality in colloidal and systems. To maintain consistency, usage guidelines recommend distinguishing pzc (as a metric) from E_pzc (as an ) to prevent misinterpretation in interdisciplinary work, especially when modeling adsorption or interfacial phenomena.

Surface Charge Behavior

pH-Dependent Charging Mechanisms

The pH-dependent charging of solid surfaces, particularly metal oxides and silicates, primarily arises from the and equilibria of amphoteric surface functional groups. These groups, such as groups (>SiOH) on silica surfaces or sites (>MeOH) on oxides like alumina or , respond to changes in solution by gaining or losing protons, thereby developing a net surface charge. For instance, at low , dominates, forming positively charged species like >SiOH₂⁺ or >MeOH₂⁺, while at high , prevails, yielding negatively charged >SiO⁻ or >MeO⁻. This behavior is governed by acid-base equilibria, such as >MeOH ⇌ >MeO⁻ + H⁺ (with dissociation constant K_a) and >MeOH + H₂O ⇌ >MeOH₂⁺ + OH⁻ (related to K_b), where the surface sites act as weak acids or bases. The acid-base character of these surfaces influences their charging profiles and the position of the point of zero charge (PZC), defined as the where net surface charge is zero. Acidic surfaces, exemplified by silica with a low PZC around 2–3, tend to deprotonate readily due to weaker Me–O bonds, resulting in negative charge above the PZC. In contrast, basic surfaces like those of metal oxides (e.g., alumina with PZC ~9 or ZnO with PZC ~9.2) protonate more easily, exhibiting positive charge below the PZC owing to stronger coordination of protons to oxygen ligands. These differences stem from the intrinsic constants: for amphoteric sites, the first constant (pK_{a1}) governs the loss of a proton from the neutral site, and the second (pK_{a2}) from the protonated site, with PZC ≈ (pK_{a1} + pK_{a2})/2. For alumina, typical values are pK_{a1} ≈ 7.1 and pK_{a2} ≈ -9.1 (adjusted for intrinsic conditions). The net surface (σ) can be quantitatively expressed as σ = F (Γ_{+} - Γ_{-}), where F is the , and Γ_{+} and Γ_{-} represent the surface densities of positively and negatively charged sites, respectively. These densities depend on through the fractional occupancies Θ of protonated and deprotonated forms, such that σ_{0,H} = F (Θ_{H^+} - Θ_{OH^-}) \times Γ_{total}, with Θ derived from the equilibria and Boltzmann factors accounting for electrostatic effects. plays a crucial role in this process by forming layers on the surface, where its autoionization (H₂O ⇌ H⁺ + OH⁻) supplies protons and ions that participate in the exchange with surface sites, stabilizing the hydroxylated layer essential for charge development. Factors influencing shifts in the PZC distinguish intrinsic charging, driven solely by pH-dependent proton transfer to/from surface groups, from extrinsic charging caused by specific adsorption of ions (e.g., anions or cations) that alter the effective constants. In inert electrolytes like NaCl, intrinsic mechanisms dominate, but specific ion binding—such as on iron oxides—can shift the PZC downward by enhancing negative charge. This separation is critical for understanding charge behavior independent of effects. Seminal work by Parks established foundational distinctions in surface acid-base properties across oxides, while models by James and Healy formalized the equilibria underlying charge development.

Influence of Electrolytes and Ions

The presence of background electrolytes in solution modulates the surface charge at the solid-liquid interface by screening the electrostatic potential through the formation of a diffuse electrical double layer. According to the Gouy-Chapman model, this diffuse layer consists of counterions and co-ions distributed according to Boltzmann statistics, which effectively compresses the electric field extending from the charged surface into the electrolyte. The thickness of this diffuse layer is characterized by the Debye length, \kappa^{-1}, given by \kappa^{-1} = \sqrt{\frac{\varepsilon R T}{2 F^2 I}}, where \varepsilon is the permittivity of the medium, R is the gas constant, T is the temperature, F is the Faraday constant, and I is the ionic strength of the electrolyte; higher ionic strength reduces the Debye length, enhancing screening and diminishing the effective range of surface charge interactions. This screening effect influences the point of zero charge (PZC) indirectly by altering the measured zeta potential without changing the intrinsic surface charge density at the PZC itself. Electrolytes can be classified based on their interaction with the surface: indifferent ions, which primarily contribute to electrostatic screening without significant specific adsorption, and potential-determining ions, which directly influence the surface potential and thus the PZC position. Indifferent ions, such as Na^+ and Cl^- at moderate concentrations on oxide surfaces, accumulate in the diffuse layer to neutralize the surface charge via Coulombic forces alone, compressing the double layer without shifting the PZC. In contrast, potential-determining ions like H^+ and OH^- adsorb specifically onto surface sites, establishing the PZC as the pH where their activities balance to yield zero net charge; other ions may act similarly if they bind strongly to the surface. Specific adsorption of ions beyond simple electrostatics can alter the PZC by introducing additional charge at the interface, often following the Hofmeister series, which ranks ions by their lyotropic effects on solubility, stability, and adsorption strength. In this series, chaotropic anions like Cl^- exhibit weaker specific adsorption compared to kosmotropic ones like SO_4^{2-}, leading to differential shifts in the PZC; for instance, on \alpha-alumina surfaces, increasing SO_4^{2-} concentration shifts the PZC to lower pH values more than equivalent Cl^-, due to stronger binding of the divalent anion. These ion-specific effects arise from hydration shell disruptions and direct surface coordination, modulating the effective surface charge density. The point of zero salt effect (PZSE) provides a practical measure of these influences, defined as the at which the net surface charge is independent of concentration, indicating the absence of specific adsorption effects from the background . At the PZSE, curves for different ionic strengths intersect, allowing differentiation between indifferent screening and specific adsorption; for variable-charge minerals like oxides, the PZSE often coincides with the PZC in the absence of adsorbing s. For (\alpha-Fe_2O_3), the PZC is typically around 8.5 in NaCl solutions, and it remains independent of since NaCl is generally considered an indifferent .

Determination Methods

Experimental Techniques

One of the primary experimental techniques for determining the point of zero charge (PZC) involves , which assesses the net surface charge through acid-base additions to a of the material. In this method, a known mass of the solid is dispersed in an inert solution (e.g., 0.01 M NaCl) to maintain constant , and the is titrated with standardized (e.g., HCl) or (e.g., NaOH) while monitoring changes with a . The PZC is identified as the where the slope of the surface (σ) versus plot is zero (dσ/dpH = 0), often calculated from the difference between the titration curve of the and a blank solution. To correct for activity and account for non-ideal behavior, Gran plots are employed; these linearize the titration data by plotting functions such as 10^{pH - pKw}V (for addition) or 10^{-pH}V (for addition), where V is the volume added, allowing extrapolation to the corresponding to free proton binding. A variant, potentiometric mass titration (), enhances efficiency by titrating multiple with varying solid masses simultaneously; the curves intersect at a common point whose is the PZC, applicable across ranges and reducing equilibration time to hours. The addition method involves preparing a of the material in an (e.g., 0.01 M NaNO_3) at various initial values spanning the expected range. A concentrated strong (e.g., NaCl) is then added to perturb the system; the PZC is identified as the initial where the final after addition remains unchanged, indicating zero net surface charge and thus no preferential adsorption of H^+ or OH^- to cause drift. This technique is particularly useful for detecting specific effects and is simple, requiring minimal equipment, though it assumes fast equilibration. Electrophoretic mobility measurements provide an alternative approach by evaluating the isoelectric point (IEP), which often approximates the PZC for oxide surfaces, through (ζ) assessment via microelectrophoresis. The principle relies on applying an to a dilute (typically 0.01–0.1 wt%) in an , where particle velocity is proportional to electrophoretic mobility (μ_e = v/E, with v as velocity and E as field strength); ζ is then derived using the Henry equation, ζ = (η μ_e)/ε f(κa), with η as , ε as , and f(κa) as a . By adjusting across a range (e.g., 2–12) with / additions and measuring μ_e at each point, the IEP/PZC is the where ζ = 0 and mobility vanishes, indicating no net charge driving particle movement. Microelectrophoresis cells, such as flat or cylindrical chambers, facilitate observation of particle trajectories under a , with modern instruments automating tracking for precision. This technique is particularly suited for colloidal particles (0.1–10 μm) and probes the shear plane potential, though it may differ slightly from the true PZC due to diffuse layer effects. Immersion methods detect the PZC by monitoring interfacial changes upon contact with , often indicating zero charge through maximum wettability or charge neutrality signals. In the classic immersion technique for powders or electrodes, the material is immersed in a solution of varying , and the of immersion or transient currents are measured; at PZC, adsorption is minimized, leading to a characteristic minimum in immersion or zero net charge transfer. For wettability-based variants, measurements on compressed pellets or films assess spreading: at PZC, the surface charge is neutral, resulting in minimum hydrophilicity (maximum , often >60°) due to reduced electrostatic attraction to dipoles. sensor approaches, such as those using capillary rise or imbibition in porous , detect maximum penetration at PZC, where electrostatic repulsion is absent, allowing optimal spreading; a records changes as liquid wicks into the material. These methods are advantageous for insoluble solids but require careful surface preparation to avoid artifacts from air entrapment. The batch equilibration, or pH-drift, method offers a simpler, non-titrating alternative by observing pH stabilization in closed systems. A fixed of solid (e.g., 0.1–1 g) is added to series of solutions (e.g., 50 mL of 0.01 M NaCl) adjusted to initial values spanning the expected range (e.g., 2–12 with HCl/NaOH); after equilibration (typically 24–48 h under inert atmosphere), final pH is measured, and the PZC is the of the final versus initial pH plot with the 1:1 line, where no net proton transfer occurs. This approach indirectly reflects adsorption edges by capturing the pH where surface buffering neutralizes added /. It is widely used for adsorbents like oxides due to minimal equipment needs but demands consistent solid-liquid ratios for reproducibility. Despite their utility, these techniques share limitations that can affect accuracy. Results are sensitive to , as finer particles may shift the apparent PZC/IEP due to increased edge site exposure or aggregation effects, with discrepancies up to 1–2 units observed between nano- and micro-scale samples. Impurities, such as surface-active organics or multivalent ions, adsorb preferentially and alter charge balance, leading to erroneous values; purification steps like acid washing are essential. Common errors include CO₂ interference in open systems, which acidifies solutions and depresses measured PZC by 0.5–1 unit during or drift experiments, necessitating N₂ purging or sealed setups. Additionally, method-specific issues arise: potentiometric approaches require high surface area (>1 m²/g) for detectable signals, while electrophoretic methods assume monodisperse particles and can overestimate IEP for irregular shapes. For instance, potentiometric of TiO₂ () suspensions in 0.01 M NaCl yields a PZC of approximately 6.0, consistent with equilibria of surface Ti-OH groups, though values vary slightly (5.8–6.4) with crystal phase or impurities.

Theoretical and Computational Approaches

The site-binding model provides a foundational theoretical framework for predicting the point of zero charge (PZC) by treating surface charging as a result of and reactions at specific binding sites on surfaces. In this model, surface hydroxyl groups (SOH) act as amphoteric sites that can protonate to form SOH₂⁺ or deprotonate to SO⁻, with equilibrium constants pKₐ and pK_b governing these processes. The PZC occurs when the concentrations of positively and negatively charged sites balance, typically at = (pKₐ + pK_b)/2 for symmetric sites. This approach assumes discrete binding sites and neglects diffuse layer effects initially, enabling predictions of pH-dependent surface charge without direct measurement. An extension, the triple-layer model, incorporates by dividing the electrical layer into three planes: the surface plane (o-plane) for binding sites, the inner Helmholtz plane (β-plane) for specifically adsorbed ions, and the outer Helmholtz plane (d-plane) marking the start of the diffuse layer. Surface site densities and capacitances between planes are key parameters, allowing the model to account for ion-specific effects on PZC shifts. The surface σ₀ is given by \sigma_0 = F \left( \Gamma_{\ce{SOH_2^+}} - \Gamma_{\ce{SO^-}} \right), where F is the Faraday constant, and \Gamma_{\ce{SOH_2^+}} and \Gamma_{\ce{SO^-}} represent the surface site densities for protonated and deprotonated species, derived from mass-action laws tied to pKₐ and pK_b. This formulation predicts PZC by solving charge neutrality at σ₀ = 0. Integration of the site-binding or triple-layer models with DLVO theory extends predictions to colloidal stability near the PZC, where net surface charge is zero but residual electrostatic interactions persist due to ion distributions. DLVO combines van der Waals attractions with electrostatic repulsions, using the PZC-derived potential to compute the interaction energy minimum; at PZC, stability is often lowest as repulsion vanishes, promoting aggregation unless steric effects intervene. This hybrid approach has been applied to oxide suspensions, revealing how electrolyte concentration modulates the energy barrier for particle coagulation. Density functional theory (DFT) computations refine PZC predictions by calculating site-specific protonation energies and pKₐ values for oxide surfaces from first principles. For instance, DFT simulations of MgO, TiO₂, and γ-Al₂O₃ interfaces yield pH-dependent speciation, showing how undercoordinated sites dominate charging and shift PZC based on metal-oxygen bond strengths. These methods bypass empirical fitting by optimizing surface geometries in vacuum or implicit solvent, though explicit inclusion via hybrid DFT-molecular dynamics enhances accuracy for solvation-influenced sites. Molecular dynamics (MD) simulations complement DFT by modeling dynamic ion adsorption and water structuring at oxide-water interfaces, directly probing PZC through charge equilibration. Classical MD, often using software like , simulates cation adsorption on charged or surfaces, revealing outer-sphere complexes near PZC and charge reversal at high ion concentrations. These trajectories quantify adsorption isotherms and diffuse layer , validating site-binding assumptions under realistic conditions. Theoretical models like site-binding and triple-layer rely on assumptions of ideal, defect-free surfaces and uniform site densities, which limit applicability to real, heterogeneous materials where roughness alters charging. Solvation effects pose further challenges, as implicit models underestimate hydrogen bonding and screening at interfaces, leading to overestimated pKₐ shifts and PZC errors up to 2-3 units compared to experiments. Hybrid explicit-implicit approaches mitigate this but increase computational cost. Recent advances employ to predict PZC directly from material composition, bypassing detailed simulations. Post-2020 studies train models on datasets of stoichiometries and potentials, achieving predictions within 0.5 units by correlating elemental electronegativities and coordination numbers to surface acidity; for metal-organic frameworks, regressors tuned charge distributions to optimize PZC for electrocatalytic applications. These methods accelerate screening of hypothetical materials while incorporating effects from augmented features.

Applications

In Electrochemistry

In , the point of zero charge is referred to as the potential of zero charge (E_pzc), defined as the at which the net charge in the electrical double layer at the metal-electrolyte is zero. This condition marks a state where the surface neither attracts nor repels excessively, resulting in a minimum in the differential double-layer , expressed as C = \frac{d\sigma}{dE}, with \sigma denoting the surface and E the applied potential. The E_pzc serves as a critical reference point for understanding interfacial structure and charging behavior in electrochemical systems, influencing phenomena from ion adsorption to electrocatalytic activity. The Lippmann equation, a cornerstone of interfacial electrochemistry, quantifies the relationship between potential variation and surface charge as \Delta V = -\frac{\sigma}{C}, where the E_pzc establishes the baseline for the uncharged interface (\sigma = 0). This equation arises from the thermodynamic dependence of interfacial tension on potential and is particularly applicable to ideally polarizable electrodes, enabling predictions of charge accumulation and double-layer expansion away from E_pzc. By defining the neutral interface, E_pzc facilitates modeling of capacitive charging and potential-dependent responses in energy devices. In supercapacitors, the E_pzc guides the selection of operating voltage windows to maximize while avoiding reactions, as deviations from E_pzc modulate and packing efficiency in the double layer. For instance, aligning the cell voltage symmetrically around the average E_pzc of the electrodes optimizes charge storage without exceeding stability limits. In corrosion protection, E_pzc dictates the adsorption dynamics of organic inhibitors; anions adsorb preferentially on positively charged surfaces above E_pzc, forming protective layers that inhibit metal . A classic example is on electrodes, where for the Au(111) facet in dilute perchloric acid, E_pzc ≈ 0.25 V vs. SHE, providing a for such interactions. The E_pzc also impacts electrocatalytic processes like the (HER) and (ORR), as the double-layer configuration near E_pzc alters reactant adsorption and proton transfer kinetics; for HER, potentials below E_pzc enhance H+ attraction, while for ORR on metals like Au(111), E_pzc proximity influences binding stability. Recent studies have highlighted pH-dependent shifts in E_pzc for electrodes in alkaline media, such as on Cu(111), where increasing from 13 to 14 lowers the potential of zero free charge by approximately 0.09 V, affecting interfacial OH adsorption and overall device performance in pH-variable environments. These findings underscore the need for tailored electrolytes to align E_pzc with operational potentials in advanced energy storage systems.

In Environmental Geochemistry

In environmental geochemistry, the point of zero charge (PZC) plays a critical role in governing adsorption processes on mineral surfaces, which dictate the fate of contaminants in natural systems. At pH values above the PZC, mineral surfaces acquire a negative charge due to deprotonation of surface hydroxyl groups, leading to electrostatic repulsion of anions and enhanced attraction for cations such as heavy metals. For instance, lead (Pb²⁺) sorption onto clay minerals like kaolinite is maximized at pH > PZC (typically 4-6 for clays), where positive metal ions bind via electrostatic and specific inner-sphere complexation, reducing their mobility in soils and sediments. In natural environments, PZC values for soils generally range from 4 to 7, influenced by mineral composition and , directly affecting mobility and . Below the PZC, positively charged surfaces promote anion adsorption (e.g., ), limiting leaching, while above it, negative charges facilitate cation release, enhancing mobility in variable-charge soils like . In aquifers, dissolved humic acids can shift the effective PZC downward by adsorbing onto mineral surfaces and altering charge distribution, thereby increasing the transport potential of sorbed metals through complexation and reduced electrostatic retention. The environmental impact of PZC extends to controlling colloid-facilitated transport of pollutants, where charge compatibility between colloids and grains determines attachment or . At < PZC, positively charged colloids (e.g., iron oxide nanoparticles carrying heavy metals) exhibit reduced deposition on negatively charged sand grains, enhancing subsurface migration and risking groundwater contamination. Geochemical modeling tools like PHREEQC incorporate PZC data within surface complexation models to simulate these interactions, predicting contaminant speciation and transport under varying ionic strengths and conditions in natural waters. A prominent case study involves arsenic adsorption on iron oxides such as goethite and ferrihydrite, which have PZC values of approximately 8-9, making them effective sorbents in neutral to alkaline groundwaters. At pH < PZC, arsenate (As(V)) forms strong inner-sphere complexes on positively charged surfaces, facilitating natural attenuation and informing remediation strategies like permeable reactive barriers in arsenic-contaminated aquifers. This mechanism has been pivotal in addressing groundwater arsenic issues in regions like Bangladesh, where iron oxide amendments leverage PZC-driven adsorption to immobilize the toxin. Climate-driven soil acidification, exacerbated by increased CO₂ levels and altered precipitation patterns, lowers soil pH below typical values, thereby enhancing trace element bioavailability through increased positive surface charge and metal desorption. Recent reviews highlight how this shift amplifies the mobilization of elements like cadmium and zinc in agricultural soils, posing risks to ecosystems and food chains, with pH decreases of 0.5-1 unit potentially doubling metal solubility in acidic environments.

In Colloid and Materials Science

In colloid and materials science, the point of zero charge (PZC) plays a pivotal role in governing colloidal stability by modulating electrostatic interactions between particles. At the PZC, the net surface charge is zero, eliminating electrostatic repulsion and allowing van der Waals attractive forces to dominate, which promotes particle aggregation and flocculation. This behavior is quantitatively described by the Derjaguin-Landau-Verwey-Overbeek (DLVO) theory, where the total interaction energy between particles is the sum of attractive van der Waals forces and repulsive electrostatic double-layer forces; near the PZC, the absence of the latter leads to reduced energy barriers for coagulation, as observed in hematite nanoparticles with a PZC range of pH 5.5–9.5. Tuning the PZC is essential in materials design, particularly for nanoparticles used in targeted applications like drug delivery. Mesoporous silica nanoparticles (MSNs), with a PZC of approximately pH 2–3, exhibit a negatively charged surface at physiological pH (around 7.4), facilitating electrostatic adsorption and loading of positively charged (cationic) therapeutic agents such as doxorubicin or antimicrobial peptides. This charge-based loading enhances controlled release and bioavailability while minimizing premature drug dissociation in biological fluids. In adsorption technologies for water purification, the PZC determines the selectivity and efficiency of materials toward organic pollutants. Activated carbon, typically with a PZC in the range of pH 7–9, becomes positively charged below this value, promoting adsorption of anionic organic compounds like dyes or phenolic pollutants through electrostatic attraction and π-π interactions. For instance, rice husk-derived biochar with a PZC around pH 8–9 effectively removes cationic textile dyes under neutral to alkaline conditions by leveraging surface charge to enhance binding capacity. Specific examples illustrate PZC's practical utility in processing and . In ceramic processing, slip casting relies on adjusting slurry away from the PZC (often 1–2 for titania) to maximize electrostatic repulsion and achieve high solids loading with low viscosity; for Ti₃SiC₂ powders, deflocculation at > pzc (around 6–7) enables dense green bodies with minimal defects. In , the PZC of zeolites (typically 2–5 for acidic types) influences reactant adsorption by altering surface charge; at above the PZC, negative charges enhance binding of cationic precursors, boosting activity in reactions like CO oxidation on ceria-zeolite composites. Emerging applications leverage PZC engineering in for development, particularly through doping to fine-tune surface charge for improved selectivity. Recent advancements (post-2022) in doped silica or metal oxide , such as nitrogen-doped oxide composites with adjustable PZC via incorporation, enable pH-responsive detection of analytes like or biomolecules by optimizing electrostatic interactions at the interface.

References

  1. [1]
    point of zero charge (P04704) - IUPAC Gold Book
    A surface charge is at its point of zero charge when the surface charge density is zero. It is a value of the negative logarithm of the activity in the bulk ...
  2. [2]
    The pH dependent surface charging and points of zero charge. IX ...
    This review updates a compilation of the points of zero charge (PZC) and isoelectric points (IEP) of various materials by the same author [1]. In principle PZC ...
  3. [3]
    Isoelectric points and points of zero charge of metal (hydr)oxides
    The point of zero charge (PZC) and isoelectric point (IEP) became standard parameters to characterize metal oxides in aqueous dispersions.
  4. [4]
  5. [5]
    THE ZERO POINT OF CHARGE OF OXIDES 1 - ACS Publications
    Relating Catalytic Activity of d0 Semiconducting Metal Oxides to the Fermi Level Position. The Journal of Physical Chemistry C 2014, 118 (13) , 6873-6881.
  6. [6]
    Electronic Effects in the Electric Double Layer | Chemical Reviews
    The development of microscopic models for the double layer began with the works of Gouy 3 and Chapman, 4 who proposed a space charge theory for the solution ...
  7. [7]
    Protein pI and Intracellular Localization - PMC - NIH
    The isoelectric point (pI) of a protein is defined as the pH at which the net charge of a protein molecule is zero. Accordingly, proteins are positively ...
  8. [8]
    [PDF] Iso-electric Point (IEP) and Point of Zero Charge (PZC)
    IEP is the pH where zeta potential reverses, while PZC is where surface potential reverses. They are often the same, but differ with indifferent ions.
  9. [9]
    [PDF] Point Of Zero Charge Determination in Soils and Minerals via ...
    The point of zero charge (PZC) is the pH where the net total particle charge is zero. It's determined using methods like potentiometric titration and ion ...
  10. [10]
    Point of zero charge determination in soils and minerals via ...
    The pH where the net total particle charge is zero is called the point of zero charge (PZC), which is one of the most important parameters used to describe ...
  11. [11]
    A Comparison of Point of Zero Charge Measurement Methodology
    Apr 1, 2012 · The objective of the present study was to measure and compare the point of zero charge (PZC) using different methods.
  12. [12]
    Change of the point of zero net proton charge (pH PZNPC ) of clay ...
    Aug 20, 2018 · We studied the relationship of the pH PZNPC of three clay minerals (kaolinite, illite, and montmorillonite) at seven ionic strengths (from 0.001 to 0.1 M).Missing: PZSE salt
  13. [13]
    Potential of zero charge and surface charging relation of metal ...
    Mar 23, 2020 · The potential of zero charge (pzc) is where there is no free-electrode charge. When potential deviates from pzc, the interphase becomes ...
  14. [14]
    Potentials of Zero Charge - SpringerLink
    The pzc was proved to be an important electrochemical characteristic of metal and to play a major role in electrocapillary and electrokinetic phenomena.
  15. [15]
    (PDF) pH-dependent surface charging of metal oxides - ResearchGate
    Aug 6, 2025 · In this work, the pH-dependent surface charging of metal oxides due to the specific adsorption of H+/OH in the presence of indifferent and specific ions,
  16. [16]
  17. [17]
    The pH dependent surface charging and points of zero charge. X ...
    The point of zero charge (PZC) is a key surface characteristic, especially for adsorbents, and is related to pH-dependent surface charging. It is often ...<|control11|><|separator|>
  18. [18]
  19. [19]
    Modeling the Gouy–Chapman Diffuse Capacitance with Attractive ...
    The Gouy–Chapman-like contribution to the capacitance depends on the bulk ion concentration c0 through eq 2. The contribution of ion–surface attraction to the ...
  20. [20]
    Mechanisms of Surface Charge Modification of Carbonates ... - MDPI
    In other words, PDIs are the ions that both produce and control the surface charge, while indifferent ions control expansion of the double layer in the solution ...
  21. [21]
    Specific ion effect on the point of zero charge of α-alumina and on ...
    The Hofmeister series of inorganic anions with a common cation, Na+, are originally arranged based on their relative influence on the precipitation and ...
  22. [22]
    Potentiometric Mass Titrations: Experimental and Theoretical ...
    In this paper, we present a novel methodology, called the potentiometric mass titration (PMT) technique, for determining the point of zero charge (pzc) of ...Missing: history | Show results with:history
  23. [23]
    A COMPARISON OF POINT OF ZERO CHARGE MEASUREMENT ...
    Apr 1, 2011 · The IEP measurement of PZC resulted in an appreciably more acidic PZC of 4.4±0.1, a value that is in agreement with previous reported values ...Abstract · INTRODUCTION · EXPERIMENTAL · RESULTS<|control11|><|separator|>
  24. [24]
    Time-resolved determination of the potential of zero charge at ...
    Feb 8, 2018 · In this work, we present a method that combines electrode immersion transient and impedance measurements for the determination of the PZC.Missing: E_pzc notation
  25. [25]
    Switchable imbibition in nanoporous gold | Nature Communications
    Jul 1, 2014 · Here we show for aqueous electrolyte imbibition in nanoporous gold that the fluid flow can be reversibly switched on and off through electric potential control.
  26. [26]
    Understanding electrochemical interfaces through comparing ...
    Apr 23, 2024 · The potential of zero charge (pzc) is a fundamental property of the electrode–electrolyte interface, whose knowledge is a key requirement for a ...Missing: history | Show results with:history
  27. [27]
    A Comparison of Point of Zero Charge Measurement Methodology
    The present work presents the first reported instance where MT and PA have been applied to measure the point of zero charge of either pyrolusite or goethite.
  28. [28]
    TiO2 point of zero charge determined by the acid–base titration ...
    Previous studies have indicated that the PZC of TiO 2 P25 is located at pH = 6.35 ( Figure S2) and confirmed that the surface of the photocatalyst is also ...Missing: example | Show results with:example
  29. [29]
    Site-binding model of the electrical double layer at the oxide/water ...
    This model is used to calculate theoretical surface charge densities of the potential-determining (H+/OH–) ions and the potential at the Outer Helmholtz Plane, ...
  30. [30]
    Evaluation of internally consistent parameters for the triple-layer ...
    The analysis yields an internally consistent set of triple-layer parameters which will be used in developing a predictive model for electrolyte adsorption.Missing: seminal papers
  31. [31]
    Surface Force Analysis and Application of DLVO Theory | ACS Omega
    Dec 2, 2020 · There is a specific pH, where the zeta potential becomes zero, which is called the point of zero charge (PZC) or the isoelectric point (IEP), as ...
  32. [32]
    pH Dependence of MgO, TiO2, and γ-Al2O3 Surface Chemistry from ...
    Jun 14, 2022 · We investigate the pH-dependent surface chemistry of three relevant and widely used oxides, MgO, TiO 2 , and γ-Al 2 O 3 , from ab initio molecular dynamics ...
  33. [33]
    Molecular Dynamics Simulations of Quartz (101)-Water and ...
    In classical MD simulations, ion adsorption at different pH is usually studied by modifying the model of solid surface, typically changing the surface charge ...
  34. [34]
    Theoretical investigation on potential of zero free charge of (111 ...
    The study used a hybrid solvation method with DFT to calculate the potential of zero free charge (PZFC) for Group 10 and 11 metals, finding that surface ...1. Introduction · 3. Results And Discussion · 3.2. Hybrid Solvation Model<|separator|>
  35. [35]
    Machine learning guided tuning charge distribution by composition ...
    Mar 18, 2024 · This study uses machine learning to accelerate MOF synthesis, train models with zeta potentials, and predict MOFs for oxygen evolution, showing ...Missing: zero | Show results with:zero
  36. [36]
    Zero Charge Potential - an overview | ScienceDirect Topics
    The potential E of metal electrodes at which the interfacial charge σ is zero (hence, g ion = 0) is the potential of zero charge (the zero charge potential), E ...
  37. [37]
    On the electrode charge at the metal/solution interface with specific ...
    Pztc is the potential at which the total charge is zero (the capacitive charge compensates the faradaic charge). Pzfc is the potential at which the free charge ...
  38. [38]
    Cell voltage versus electrode potential range in aqueous ... - Nature
    Apr 21, 2015 · Potential of zero voltage (PZV). When a supercapacitor is fully discharged, its voltage is zero if EN2 ≥ EP1 and the potentials of the two ...
  39. [39]
    Corrosion Inhibition and the Zero Charge Potential of Metals - Nature
    De suggested that inhibitive anions are adsorbed preferentially on metal/solution interfaces that are positive to the zero charge potential (potential of ...
  40. [40]
    First Step of the Oxygen Reduction Reaction on Au(111)
    Aug 29, 2023 · Free energy of O 2 * adsorption at Au(111) as a function of surface charge density (bottom x-axis) or applied potential vs SHE (top x-axis).
  41. [41]
    Origin of Solvent Dependency of the Potential of Zero Charge
    Nov 15, 2023 · In this study, we aim to link the solvent dependency to the microscopic phenomenon of electron spillover occurring at the metal–solution interface.
  42. [42]
    The Potential of Zero Charge and the Electrochemical Interface ...
    Mar 1, 2021 · We present the first two-stranded correlation of the potential of zero free charge (pzfc) of Cu(111) in alkaline electrolyte at different pH values.
  43. [43]
    Adsorption of heavy metals in glacial till soil
    The point of zero charge (pH PZC ) of glacial till was found to be 7.0±2.5. Surface charge experiments revealed the high buffering capacity of the glacial till.
  44. [44]
    A review on adsorption characteristics and influencing mechanism ...
    3.1.​​ It has been found that the adsorption of most heavy metals in soil can be increased from 0 to nearly 100% in a small pH range.<|control11|><|separator|>
  45. [45]
    [PDF] Nutrient mobility in variable- and permanent-charge soils - KBS LTER
    Because soils contain a mixture of. Al, Fe, Si and organic hydroxyls, bulk soil PZC's tend to range between pH. 3.5 and 5 (Parfitt 1980; Uehara & Gillman 1981).
  46. [46]
    [PDF] Humic Acid Facilitates the Transport of ARS-Labeled Hydroxyapatite ...
    oxyhydroxides are amphoteric with relatively high points of zero charge (pHPZC). Often, pH values in groundwater and soils are below the pHPZC of iron (7.5) ...
  47. [47]
    [PDF] Chemical Factors Influencing Colloid Mobilization and Th(IV)
    Aug 4, 2012 · 2.9 Colloid-Facilitated Contaminant Transport ... The pH at which a mineral has a zero charge is called the zero point charge, or pHpzc (Ryan et ...
  48. [48]
    Surface Complexation Modeling of Fluoride Adsorption by Soil and ...
    Jul 5, 2018 · Based on the fact that F− adsorbs onto the net negative surface in the pH range studied (considering that the point of zero charge [PZC] of this ...
  49. [49]
    Arsenic removal from aqueous solutions by adsorption onto iron ...
    Mar 6, 2014 · The point of zero charge (pHpzc) of the prepared iron oxide was found to be 7.9. Over the PZC value, iron oxide is present in the monomeric ...
  50. [50]
    Bioavailable soil Pb minimized by in situ transformation to ... - PNAS
    Dec 21, 2020 · This conversion decreased soil Pb bioavailability greater than 90%, even after pH neutralization to revitalize the soil to agronomic health. PLJ ...Bioavailable Soil Pb... · Results · Materials And Methods<|control11|><|separator|>
  51. [51]
    Climate induced microbiome alterations increase cadmium ... - Nature
    Oct 30, 2024 · Here we show that future climate affects soil cadmium through altered soil microbiome and nutrient cycles, with soil pH as critical factor.Missing: zero | Show results with:zero
  52. [52]
    Changes in soil pH and mobility of heavy metals in contaminated soils
    Dec 6, 2021 · Soil acidification increases heavy metal mobility, especially Cd, Zn, and Pb. Even small amounts of acid cause pH changes, increasing metal ...Missing: climate zero
  53. [53]
    Dispersion and stability of bare hematite nanoparticles - NIH
    Hematite nanoparticles have a point of zero charge (PZC) ranging from pH 5.5 to 9.5 depending on the method of synthesis and experimental conditions ( ...Missing: shift | Show results with:shift
  54. [54]
    An overview of surface forces and the DLVO theory | ChemTexts
    Jul 22, 2023 · It is necessary, however, to keep in mind that the equation assumes that the ions have no volume (i.e., they are treated as point charges).
  55. [55]
    Mesoporous silica nanoparticles for therapeutic/diagnostic ...
    Since the point of zero charge for MSNs is 2–3, their surface is negatively charged in the absence of specific ion adsorption under biological conditions ...
  56. [56]
    [PDF] Role of Point of Zero Charge in the Adsorption of Cationic Textile ...
    May 19, 2022 · The point of zero charge (PZC) is an electrokinetic property of biochars that influences dye adsorption. Rice husk biochar was effective in ...
  57. [57]
    Point of zero charge and intrinsic equilibrium constants of activated ...
    The point of zero charge for ACC in KNO3 solutions, determined by batch equilibrium method, is at pH=7. In the presence of zinc and cadmium ions, pHPZC ...Missing: purification | Show results with:purification
  58. [58]
    Surface Chemistry, Dispersion Behavior, and Slip Casting of Ti 3 ...
    Aug 4, 2009 · The point of zero charge (pzc) is achieved when the relative densities of the surface groups inline image are satisfied, provided that the ...
  59. [59]
    Article Boosting catalytic performance via electron transfer effect for ...
    Feb 16, 2023 · In this work, we achieve this goal by taking advantage of the discrepancy in the point of zero charge between acidic zeolite support and ceria ...
  60. [60]
    A review of the emerging role of engineered nanomaterials as ...
    Sep 20, 2025 · However, its low Point of Zero Charge (PZC) (pH 1.6–3.5) limits its effectiveness for anionic pollutants. Studies have shown a correlation ...
  61. [61]
    Recent advances in nitrogen-doped graphene oxide nanomaterials
    Jun 21, 2023 · This review covers recent advances on production techniques, unique properties and novel applications of nitrogen-doped graphene oxide (NGO).