Fact-checked by Grok 2 weeks ago

Open-circuit voltage

Open-circuit voltage (OCV), also referred to as no-load voltage, is the difference measured across the two terminals of an electrical device, , or power source when no external is flowing through it, equivalent to an infinite or disconnected load condition. This parameter represents the maximum voltage output achievable under zero-current conditions and serves as a fundamental characteristic in , , and power systems. In circuit theory, open-circuit voltage plays a central role in network analysis techniques, such as the Thévenin equivalent theorem, where it defines the ideal value in a simplified series resistor-voltage model that replicates the original circuit's external behavior. For instance, to find the Thévenin voltage, the circuit terminals are treated as open, allowing calculation of the across them without load , which is essential for predicting performance when connecting to external components. This concept extends to practical measurements using high-impedance voltmeters, which approximate open-circuit conditions to avoid loading the source. In photovoltaic devices, open-circuit voltage is the peak voltage generated by a or under standard illumination when the output current is zero, directly tied to the material's bandgap and recombination losses, and critical for determining fill factor and overall power conversion efficiency. Typical values for silicon-based cells range from 0.6 to 0.7 volts per cell at standard test conditions, influencing design to avoid thermal voltage mismatches in arrays. For batteries and electrochemical cells, open-circuit voltage indicates the equilibrium thermodynamic potential between electrodes in the absence of current, providing a reliable proxy for state-of-charge () estimation, as it correlates with the and varies with composition, temperature, and SOC. In lithium-ion batteries, for example, OCV typically spans 3.0 to 4.2 volts per cell depending on SOC, enabling non-invasive monitoring but requiring stabilization time after load disconnection to account for effects. This measurement is standardized in protocols like those from for aerospace applications, where OCV assesses charge status without discharge testing.

Basic Concepts

Definition

Open-circuit voltage (OCV), denoted as V_{oc}, is the potential difference measured across the terminals of a or device when no external load is connected, resulting in zero through the (I = 0). This represents an no-load where the full electrochemical or electromotive potential of the source is observable without any current-induced drops. In , OCV serves as a key parameter for characterizing the maximum voltage output capability of sources such as batteries, generators, or transducers before any load is applied. The concept of open-circuit voltage has roots in 19th-century circuit theory, particularly Kirchhoff's voltage law established in 1845, which describes voltage drops in closed loops but extends naturally to unloaded conditions where no current circulates. (1893) defines the equivalent voltage of a as the open-circuit potential at its terminals. The specific "open-circuit voltage" emerged in early 20th-century literature, notably used by Hans Ferdinand Mayer in 1926, and became standardized in textbooks and standards thereafter as electrical systems grew more complex. Conceptually, open-circuit voltage corresponds to the (EMF) of the source under no-load conditions, representing the energy per unit charge available from the source without current flow. However, in practical devices, the measured OCV may be slightly lower than the ideal theoretical EMF due to internal losses such as , polarization effects, or minor leakage currents inherent to the source material. This distinction highlights OCV as an empirical measurement rather than a purely theoretical value. A basic illustration of open-circuit voltage depicts a symbolic , such as a represented by a long line and short line for the positive and negative terminals, with the path intentionally broken between them to prevent flow; the voltage V_{oc} is indicated across the open terminals using a symbol connected in . This setup visually emphasizes the absence of a path while capturing the potential difference. OCV is complemented by short-circuit as the opposing in source , where maximum flows under zero voltage.

Mathematical Representation

The open-circuit voltage V_{oc} of an electrical source is fundamentally derived from the relationship between the (EMF), denoted as \mathcal{E}, and the R_{int}. According to the basic circuit model, the voltage V_{terminal} under load is given by V_{terminal} = \mathcal{E} - I R_{int}, where I is the current flowing through the . In the open-circuit condition, no external load is connected, resulting in zero current (I = 0). This simplifies the equation to V_{oc} = \mathcal{E}, meaning the open-circuit voltage equals the electromotive force for an ideal source with no residual current. In non-ideal scenarios with a small residual current I_r (such as leakage), the expression becomes V_{oc} = \mathcal{E} - I_r R_{int}, though I_r approaches zero under true open-circuit conditions. This open-circuit voltage also corresponds to the Thevenin equivalent voltage V_{th} in analysis, representing the voltage across the terminals of a linear when disconnected from any load. The Thevenin model simplifies complex to an ideal V_{th} in series with the Thevenin R_{th}, where V_{oc} = V_{th}. For non-ideal devices exhibiting diode-like behavior, such as those involving p-n junctions, the open-circuit voltage follows a logarithmic dependence derived from the Shockley diode equation. The general form is V_{oc} = \frac{n k T}{q} \ln \left( \frac{I_L}{I_0} + 1 \right), where n is the ideality factor, k is Boltzmann's constant, T is the temperature in kelvin, q is the elementary charge, I_L is the photocurrent or light-generated current, and I_0 is the saturation current. This equation highlights how V_{oc} increases logarithmically with the ratio of generated to saturation currents, providing a quantitative link to device physics without assuming specific generation mechanisms. The open-circuit voltage is expressed in volts (V), the SI unit for difference, ensuring consistency in notations and measurements.

Generation Mechanisms

Electrochemical Sources

In electrochemical sources such as batteries and fuel cells, the open-circuit voltage (OCV) originates from the change (ΔG) of the underlying reactions that drive the cell under equilibrium conditions with no net . The relationship is given by the equation \Delta G = -nFE where n is the number of moles of electrons transferred in the reaction, F is the Faraday constant ($96485 C/mol), and E is the cell potential, which equals the OCV at open circuit. This equation quantifies the maximum reversible work extractable from the electrochemical reaction, with a negative ΔG corresponding to a spontaneous process and positive OCV. The OCV in these systems is more precisely described by the Nernst equation, which accounts for non-standard conditions: E = E^\circ - \frac{RT}{nF} \ln Q Here, E^\circ is the standard cell potential (under 1 bar pressure and 1 M concentrations), R is the gas constant ($8.314 J/mol·K), T is the absolute temperature, and Q is the reaction quotient representing the ratio of product to reactant activities. Under standard conditions, where the reaction quotient Q = 1, ln Q = 0 and the cell potential equals the standard potential E°. In electrochemical cells like batteries, the OCV varies with Q: it reaches a maximum when Q is minimized (e.g., fully charged state) and decreases toward 0 as the cell approaches overall chemical equilibrium (Q = K) upon discharge. Deviations from standard conditions, such as changes in reactant concentrations or partial pressures, shift the OCV according to this equation. In real electrochemical cells, the observed OCV often manifests as a mixed potential, arising from the simultaneous occurrence of anodic (oxidation) and cathodic (reduction) reactions at the interfaces, balanced such that the net current is zero. This mixed potential reflects the where exchange currents from forward and reverse processes cancel out, influenced by electrode kinetics and surface phenomena. For example, parasitic reactions like oxygen reduction can lower the effective OCV from its thermodynamic ideal. Representative examples illustrate these principles across battery chemistries. In lead-acid batteries, the OCV is approximately 2 V per cell for a fully charged state, derived from the Pb/PbSO₄ and PbO₂/PbSO₄ half-reactions. Lithium-ion batteries exhibit a nominal OCV of about 3.7 V, though it varies from 3.2 V (discharged) to 4.2 V (charged) depending on the material like LiCoO₂. Nickel-metal (NiMH) batteries show an OCV around 1.2 V per cell, affected by in the NiOOH/Ni(OH)₂ couple. In fuel cells, the OCV follows similar electrochemical . For fuel cells (PEMFC) operating on the hydrogen-oxygen reaction (2H₂ + O₂ → 2H₂O), the theoretical OCV is 1.23 V under standard conditions (25°C, 1 atm), calculated from the standard change of -237.2 kJ/mol for formation. Practical OCVs are lower due to mixed potentials from fuel crossover and side reactions.

Photovoltaic Sources

In photovoltaic sources, such as cells and photodiodes, open-circuit voltage (OCV) arises from the in a p-n junction. Under illumination, photons with energy exceeding the material's bandgap are absorbed, generating electron-hole pairs that diffuse to the junction. The built-in electric field at the p-n interface separates these photo-generated carriers, with electrons accumulating on the n-side and holes on the p-side, establishing a voltage across the device. At open circuit, no external current flows, and the OCV corresponds to the splitting of the quasi-Fermi levels for electrons and holes, which equals the difference divided by the q. This voltage reflects the built-in potential enhanced by the photo-generated carrier density, limited by the absence of net current extraction. The OCV in photovoltaic cells is quantitatively described by the Shockley diode equation adapted for solar cells: V_{oc} = \frac{n k T}{q} \ln \left( \frac{J_{sc}}{J_0} + 1 \right) Here, n is the ideality factor (ideally 1 for diffusion-dominated transport, but often 1-2 due to recombination), k is Boltzmann's constant, T is temperature, J_{sc} is the short-circuit current density (proportional to absorbed photon flux), and J_0 is the dark saturation current density (dependent on material properties like bandgap and defects). Since J_{sc} \gg J_0 under illumination, the equation approximates to V_{oc} \approx \frac{n k T}{q} \ln \left( \frac{J_{sc}}{J_0} \right), highlighting how OCV increases logarithmically with light-generated current while being sensitive to thermal and recombination losses in J_0. This formulation originates from detailed balance principles in p-n junctions. A key factor unique to photovoltaic sources is the bandgap energy E_g, which sets the theoretical maximum OCV near E_g / q minus losses, as it governs photon absorption and the minimum energy for carrier generation. For example, silicon with E_g \approx 1.1 eV typically yields an OCV of 0.6-0.7 V in crystalline cells due to partial bandgap utilization. Perovskite solar cells, with tunable bandgaps around 1.5-1.6 eV, achieve higher OCV values up to 1.2 V, benefiting from low non-radiative recombination. In multi-junction cells, stacking semiconductors with progressively narrower bandgaps (e.g., GaInP/GaAs/Ge) connects junctions in series, cumulatively boosting total OCV to over 2.5 V while optimizing spectrum absorption. OCV also influences the fill factor (FF), the ratio of maximum power to V_{oc} \times J_{sc}, where higher OCV generally supports better FF by reducing relative series resistance impacts. Loss mechanisms primarily stem from recombination, which reduces OCV below the ideal E_g / q. Radiative recombination (band-to-band) sets a fundamental limit per the Shockley-Queisser model, but non-radiative processes—such as Shockley-Read-Hall via defects or in high-carrier densities—dominate practical losses, causing voltage deficits of 0.3-0.4 V in and less in perovskites due to their defect tolerance. Surface and interface recombination further exacerbates this by providing paths for carrier annihilation before extraction, underscoring the need for passivating layers to approach theoretical limits.

Other Electrical Sources

In electromechanical generators, open-circuit voltage arises from Faraday's law of , which states that a changing through a induces an (emf) given by \varepsilon = -\frac{d\Phi_B}{dt}, where \Phi_B is the , even in the absence of load current. This induced voltage represents the open-circuit voltage (OCV), as no current flows to dissipate energy or cause voltage drops across internal resistances. For (AC) generators, or alternators, the OCV corresponds to the peak voltage generated in the armature windings due to the sinusoidal variation of . The root-mean-square () value of this OCV is expressed as E = 4.44 f N \Phi, where f is the , N is the number of turns, and \Phi is the maximum per ; this assumes a sinusoidal and accounts for the of 1.11 for the RMS conversion from peak . In (DC) generators, the OCV approximates the no-load armature voltage, obtained from the magnetization curve under open-circuit conditions, where the generated is proportional to the field and rotational speed without load-induced armature or ohmic drops. Thermoelectric sources generate OCV through the Seebeck effect, where across a material junction produces a voltage V_{oc} = \alpha \Delta T, with \alpha as the and \Delta T as the ; this voltage emerges from diffusion without any current flow in the open circuit. Materials like bismuth telluride (Bi_2Te_3) exhibit a high Seebeck coefficient of approximately 200 \muV/K for p-type variants at room temperature, enabling efficient OCV generation in thermocouples and thermoelectric generators. In piezoelectric sources, OCV results from stress-induced charge separation in certain crystals or ceramics, where mechanical stress \sigma applied to the material generates a polarization D = d \sigma, with d as the piezoelectric coefficient; this charge accumulation produces an open-circuit voltage across the electrodes proportional to the stress magnitude. Such devices, like piezoelectric sensors, convert mechanical deformation directly into voltage without intermediate current flow. Transformers exhibit OCV on the secondary winding equal to the turns ratio times the primary voltage under open-circuit conditions, as the mutual transfers the primary without secondary load affecting the core .

Measurement Methods

Direct Measurement Techniques

Direct measurement of open-circuit voltage (OCV) involves connecting a high-impedance across the terminals of the device under test while ensuring no external load is applied, thereby approximating a true no-load condition. This technique is fundamental for devices such as batteries and , where the 's must exceed 10 MΩ to minimize current draw and loading errors that could otherwise reduce the measured voltage below the actual OCV. The primary equipment includes digital multimeters (DMMs) for steady-state OCV readings and oscilloscopes for capturing transient behaviors during relaxation to equilibrium. For battery testing in systems, standards such as IEC 61427-1 specify procedures that incorporate direct OCV measurements as part of performance evaluation, emphasizing accurate instrumentation under controlled conditions. To perform the measurement, first disconnect any load from the device's terminals to eliminate current flow. Next, attach the voltmeter probes securely across the positive and negative terminals, ensuring proper polarity for DC sources. Allow sufficient stabilization time for the voltage to reach equilibrium: typically seconds for capacitors due to rapid charge redistribution, but 1-4 hours for batteries to account for electrochemical relaxation, with studies showing 3 hours sufficient for over 98% convergence to steady-state OCV in lithium-ion cells like NMC and LFP. Record the voltage once it stabilizes, verifying consistency over multiple readings. Common error sources in direct OCV measurement include voltmeter loading, which introduces a small discharge current proportional to the inverse of input impedance, and contact resistance at probe connections that can cause voltage drops. These are mitigated by selecting DMMs with input impedances greater than 10 MΩ and employing 4-wire Kelvin connections, where separate sense leads measure the true terminal voltage independently of lead and contact resistances. Temperature variations during measurement can also influence readings, though detailed effects are addressed in environmental factors.

Indirect and Advanced Methods

In cases where direct access to the open-circuit terminals is impractical, such as in integrated or operational devices, extrapolation from current-voltage (I-V) load curves provides an indirect estimate of open-circuit voltage (OCV). This method involves measuring voltage across varying load currents, plotting V versus I, and applying a linear fit in the low-current regime to extrapolate the intercept at I=0, which corresponds to V_oc under the Thevenin equivalent circuit model treating the source as a voltage generator in series with internal resistance. This approach is particularly useful for batteries, where nonlinear OCV behavior at low currents can be fitted to predict runtime and state-of-charge without full open-circuit relaxation. Electrochemical impedance spectroscopy (EIS) offers another indirect technique, typically performed around the open-circuit potential to characterize , with the Nyquist plot's high-frequency intercept representing the uncompensated solution resistance rather than directly yielding OCV; however, the voltage at which EIS is conducted serves as the reference OCV approximation in equilibrium conditions. In battery analysis, the full impedance spectrum under near-OCV conditions allows modeling of effects to infer the true OCV by deconvoluting ohmic and kinetic contributions from the measured potential. Simulation methods enable predictive determination of OCV from device parameters without physical measurements, using tools like circuit models or finite element analysis (FEA). In simulations for solar cells, distributed models account for nonhomogeneous current distribution to compute OCV from illuminated I-V characteristics derived from material properties such as bandgap and . FEA, applied to photovoltaic modules, simulates spatial variations in and recombination to forecast OCV under operational conditions, aiding before fabrication. Advanced techniques include pulse-based methods like galvanostatic intermittent technique (GITT) for transient OCV in batteries, where short current pulses are applied followed by relaxation periods to track voltage decay toward OCV, enabling estimation alongside OCV profiling across state-of-charge. In GITT, the OCV is obtained from the steady-state voltage after each relaxation, providing a quasi-equilibrium curve without prolonged open-circuit holds. For photovoltaic cells, optical methods such as spectroscopic indirectly predict OCV by characterizing thin-film thicknesses, refractive indices, and bandgaps, which feed into optical-electrical models to estimate voltage losses and ultimate OCV potential. Recent developments as of 2025 include entropy-driven online identification methods for OCV in lithium-ion batteries, which enable precise state estimation during operation without requiring full open-circuit relaxation by leveraging entropy-based algorithms to track voltage dynamics. For systems, enhanced prediction models using extreme meteorological year data improve OCV forecasting under varying environmental conditions, supporting and performance assessment. In space applications, OCV for cells is often inferred from I-V curves measured under simulated AM0 conditions as per ASTM E948, where the voltage at zero is extrapolated from the illuminated characteristic curve to account for irradiance without actual open-circuit exposure in . This ensures reproducible performance assessment, with reported OCV values like 803 mV for InP-based cells under 1367 W/m² simulation.

Influencing Factors

Environmental Influences

Temperature significantly influences the open-circuit voltage (OCV) across various sources, primarily through its effect on reaction kinetics and carrier concentrations. In lead-acid batteries, the OCV exhibits a negative temperature coefficient of approximately -2 mV/°C per cell, meaning higher temperatures reduce the equilibrium voltage due to enhanced self-discharge and altered electrolyte properties. For photovoltaic (PV) devices, the OCV decreases with increasing temperature because the dark saturation current density J_0 follows an Arrhenius dependence, J_0 \propto \exp(-E_a / kT), where E_a is the activation energy, leading to higher recombination rates and lower V_{oc}. Humidity and atmospheric pressure also impact OCV, particularly in electrochemical sources with liquid electrolytes. Excessive moisture can lead to side reactions or corrosion, which degrade performance and affect OCV stability over time. In contrast, these effects are minimal for solid-state cells, where sealed structures limit environmental ingress. Pressure variations, such as those from altitude changes, have negligible direct influence on most OCV values, though mechanical compression in pouch-type batteries can slightly enhance contact and reduce , indirectly supporting OCV. For sources specifically, illumination intensity plays a key role, with OCV increasing logarithmically with up to saturation levels under standard conditions. This behavior stems from the short-circuit J_{sc} scaling linearly with photon flux, while recombination limits further gains at high intensities. Environmental factors accelerate aging and of OCV over time. In lead-acid batteries, exposure to elevated temperatures like 40°C promotes faster sulfation and grid corrosion, resulting in accelerated OCV decline due to irreversible . Similarly, combined humidity and temperature stress exacerbates in electrochemical systems, reducing long-term OCV stability. To ensure consistency in OCV measurements and comparisons, industry standards specify testing at 25°C, allowing isolation of environmental influences from intrinsic device performance.

Material and Device-Specific Factors

The open-circuit voltage (OCV) in electrochemical and photovoltaic devices is profoundly influenced by the intrinsic properties of the constituent materials, particularly the of s and semiconductors. In lithium-ion batteries, the choice of electrode materials directly determines the potential difference that defines the OCV; for instance, cells employing a cathode paired with a exhibit a nominal OCV ranging from 3.7 V to 4.2 V, reflecting the high of LiCoO₂ (approximately 4.0 V vs. /Li⁺) contrasted with the lower potential of lithiated graphite (around 0.1 V vs. /Li⁺). This combination yields a higher OCV compared to alternatives like LiFePO₄ cathodes, which operate at about 3.4 V vs. /Li⁺ in similar configurations. In photovoltaic semiconductors, dopants modulate the and carrier concentrations, thereby tuning the built-in potential; for example, doping in n-type solar cells enhances the junction field, potentially increasing OCV by optimizing minority carrier lifetimes, though excessive doping can induce narrowing and reduce it. Device architecture further shapes the effective OCV by dictating how voltage potentials are combined or dissipated. In series-connected photovoltaic modules, the OCV scales linearly with the number of cells, as each contributes its individual potential additively under open-circuit conditions; a typical module with 60 cells, each at ~0.6 V, achieves a total OCV of approximately 36 V. Conversely, low shunt —arising from defects or pinholes in thin films—provides parallel leakage paths that shunt photogenerated current, thereby lowering the OCV; reductions of up to 10% have been observed in cells with shunt resistances below 100 Ω·cm². Material purity and structural defects play a critical role in OCV degradation through enhanced non-radiative recombination. In solar cells, impurities such as transition metals act as recombination centers, trapping charge carriers and diminishing the quasi-Fermi level splitting; for instance, iron impurities in can reduce OCV by 10-20 mV per decade of concentration increase, translating to overall losses of 5-15% in high-purity baselines. Similarly, defects in perovskites introduce states that accelerate Shockley-Read-Hall recombination, suppressing OCV by increasing the dark . Scale effects become prominent in miniaturized or thin-film devices, where layers dominate performance. In thin-film batteries, such as those with oxynitride electrolytes, interfacial impedance from solid-solid contacts can cause variations in OCV compared to bulk counterparts, due to incomplete or space-charge layers that hinder ion transport at the electrode-electrolyte boundary. A key interplay in rechargeable systems is the correlation between OCV and state-of-charge (), where material-specific OCV-SOC curves enable precise estimation. For LiCoO₂/ batteries, the OCV rises monotonically from ~3.0 V at 0% to ~4.2 V at 100% , following a sigmoidal profile governed by intercalation ; this curve is foundational for algorithms in battery management systems.

Applications and Significance

In Device Characterization

Open-circuit voltage (OCV) serves as a fundamental parameter in the datasheets of electrochemical power sources, providing a direct indicator of the device's maximum potential output under no-load conditions. For lithium-ion batteries, the fully charged OCV is typically specified as 4.2 V per cell, reflecting the completion of the charging process before any relaxation occurs. This value settles to 3.7–3.9 V after , serving as the nominal voltage for system design and performance benchmarking. In photovoltaic devices, OCV contributes to calculations, where the power conversion efficiency η is determined by η = (V_oc × J_sc × FF) / P_in, with J_sc as the short-circuit , FF as the fill factor, and P_in as the incident light power (standardized at 100 mW/cm²). This metric allows researchers and manufacturers to quantify how closely a device approaches its theoretical performance limits. In during , OCV testing is a non-destructive method to identify defects in power sources. For modules, OCV measurements under test conditions help detect issues like cell cracks or poor interconnections, with individual cells rejected if V_oc falls below approximately 0.6 V, ensuring module-level performance meets specifications. Similarly, in production, deviations in OCV from expected values signal imbalances or inconsistencies, prompting rejection to maintain batch reliability. These tests are often integrated into automated production lines for . OCV loss analysis is a key tool in research and development for evaluating device performance against theoretical ideals. By comparing measured OCV to the bandgap-derived limit, engineers quantify losses from mechanisms like recombination; in early photovoltaic cells, such losses reduced V_oc by around 20% relative to predictions due to non-radiative processes. This approach has driven iterative improvements, as seen in the historical evolution of battery OCV from 1.5 V in zinc-carbon cells commercialized in the 1890s to 3.7 V nominal in lithium-ion batteries developed in the 1990s, enabling higher energy densities and broader applications. A practical application of OCV in characterization involves its role in (MPPT) algorithms for , where the maximum power point voltage is estimated as 70–80% of the measured OCV to approximate optimal operating conditions without full I-V sweeps. This estimation technique simplifies real-time performance assessment in field testing and lab validation.

In System Design and Diagnostics

In system design, open-circuit voltage (OCV) plays a pivotal role in equivalent circuit models (ECMs) used for simulating battery behavior in battery management systems (BMS). These models represent the battery as an ideal voltage source corresponding to OCV in series with internal resistances and capacitances, enabling accurate prediction of voltage responses under various loads and currents. For instance, in lithium-ion BMS for electric vehicles, OCV-SOC (state of charge) relationships—often modeled via polynomial, exponential, or tabular functions—are integrated into ECMs to estimate remaining runtime by correlating OCV with SOC, allowing simulations to forecast discharge duration with errors below 1% when using high-resolution tabular data. This approach facilitates design optimization for applications like renewable energy storage, where precise runtime predictions ensure system reliability without overprovisioning capacity. For diagnostics, OCV measurements serve as a key indicator of faults in battery systems, particularly in detecting degradation mechanisms and imbalances. In lead-acid batteries, sulfation—caused by prolonged undercharging—manifests as reduced capacity and performance under load due to the formation of insoluble lead crystals, signaling the need for equalization charging or . In series-connected strings, OCV variations across cells reveal SOC imbalances, which can lead to uneven aging or if unaddressed; diagnostic algorithms in BMS use these OCV disparities to trigger balancing circuits, preventing premature pack failure. Such OCV-based diagnostics enhance fault isolation in large-scale systems like packs or storage, reducing downtime by identifying issues non-invasively during rest periods. OCV is instrumental in optimization strategies, particularly through the , which guides load matching in power sources such as photovoltaic panels or batteries. The theorem states that maximum power delivery to the load occurs when the load resistance equals the source's , resulting in the load voltage being half the OCV (V_load = V_oc / 2) and power P_max = (V_oc)^2 / (4 R_internal). In practical , engineers use OCV measurements to approximate internal resistance via Thevenin equivalents, setting R_load ≈ R_internal to achieve this condition; for example, in solar charge controllers, this matching boosts energy harvest efficiency to around 50% at peak power, though it trades off for higher overall system efficiency by adjusting for varying environmental conditions. This principle is widely applied in renewable and portable power systems to maximize utilizable output without complex in simpler setups. In safety applications, monitoring within BMS frameworks is essential for preventing overcharge and deep in batteries, where real-time estimation relies on OCV thresholds to enforce protective cutoffs. For lithium-ion s, BMS algorithms track OCV to halt charging if it exceeds 4.2 V per , averting electrolyte decomposition and thermal instability, while an OCV below 2.5 V signals deep , triggering discharge termination to avoid copper in the . These thresholds, derived from OCV-SOC curves, enable proactive measures like balancing during rest states, significantly reducing risks in high-voltage packs. In renewable-integrated systems, such as setups, advanced OCV models further support safe operation by improving accuracy. A notable benefit of OCV-based approaches in is enhanced state estimation precision; for instance, OCV models in packs for services can improve estimation accuracy by up to 10% compared to methods relying solely on voltage under load, leading to better and grid stability.

References

  1. [1]
    Open Circuit and Short Circuit | Ultimate Electronics Book
    The open circuit voltage is the voltage difference measured between two terminals when no current is drawn or supplied. The short circuit current is the current ...
  2. [2]
    Open Circuit Voltage - an overview | ScienceDirect Topics
    The open circuit voltage (Voc) represents the maximum voltage obtainable from the cell, which occurs at zero current. This value corresponds directly with the ...
  3. [3]
    Thevenin Equivalent and Norton Equivalent Circuits
    ### Explanation of Open-Circuit Voltage in Thevenin Equivalent Circuit
  4. [4]
  5. [5]
    Open-Circuit Voltage - PVEducation
    The open-circuit voltage, V OC , is the maximum voltage available from a solar cell, and this occurs at zero current.
  6. [6]
    Assessing battery performance: Compared to what?
    May 9, 2019 · The open-circuit voltage is the voltage available from a battery with no current flow. It represents the battery's maximum voltage. The ...
  7. [7]
    [PDF] Practical Aspects of Battery Technology - Idaho National Laboratory
    May 13, 2018 · Open-circuit voltage is the voltage across the terminals of a cell or babery when no external current flows. It is usually close to the.
  8. [8]
    [PDF] N93-13 84
    The OCV of a battery is voltage measured when no load is applied to the battery. Measurement of the OCV is an indicator of the state of charge for many types ...
  9. [9]
    Open Circuit Voltage - an overview | ScienceDirect Topics
    Open-circuit voltage is the voltage across the solar cell when there is no current, meaning that the load is not connected. Any change in the open-circuit ...
  10. [10]
    What is Open Circuit Voltage? Explained - The Electricity Forum
    Open circuit voltage is the potential difference measured across the terminals of a device when no external load is applied.
  11. [11]
    [PDF] Origins of the Equivalent Circuit Concept: The Current-Source ...
    Sep 2, 2002 · Mayer is perhaps the first to point out that the equivalent voltage and current source values equal the open-circuit voltage and short-circuit ...
  12. [12]
    Thevenin Equivalent Voltage & Resistance: What is it ... - Electrical4U
    Apr 11, 2021 · Thevenin equivalent voltage (Veq) is equal to the open-circuit voltage measured across two terminals of load. This value of the ideal voltage ...How to Find The Thevenin... · Thevenin Equivalent Circuit...
  13. [13]
    17.2: The Gibbs Free Energy and Cell Voltage - Chemistry LibreTexts
    Apr 12, 2020 · A spontaneous redox reaction is characterized by a negative value of ΔG°, which corresponds to a positive value of E°cell.
  14. [14]
    Gibbs Free Energy Change - an overview | ScienceDirect Topics
    (2.55) Δ G = − nFE. where ΔG is the Gibbs energy change of the cell reaction ... Hence, by measuring the open-circuit voltage E, one can determine ΔG of the cell ...
  15. [15]
    9.4: Measures of Batteries and Fuel Cells - Engineering LibreTexts
    May 22, 2022 · The Nernst equation is useful to chemists because it can be used to solve for the amount concentration of reaction products and reactants. The ...
  16. [16]
    Open circuit voltage - DoITPoMS
    The EMF calculated using Nernst equations is called a Nernst Voltage and is the reversible cell voltage which would exist at a given temperature and pressure.Missing: batteries | Show results with:batteries
  17. [17]
    Open Circuit (Mixed) Potential Changes Upon Contact Between ...
    We investigate the principle of the open circuit potential (OCP) change upon a particle collision event based on mixed potential theory.
  18. [18]
    Theoretical analysis of mixed open-circuit potential for high ...
    In this paper, analytical equations for calculation of the theoretical OCV are developed, using the mixed potential theory.
  19. [19]
    BU-303: Confusion with Voltages - Battery University
    The nominal voltage of lead acid is 2 volts per cell, however when measuring the open circuit voltage, the OCV of a charged and rested battery should be 2.1V/ ...
  20. [20]
    What Are the Different Types of Battery Voltages? - YIC Electronics
    For example, lithium-ion batteries operate between 3.2 and 4.2 volts per cell, offering stable voltage but a sharp drop near full discharge. Lead-acid batteries ...
  21. [21]
    Evaluation of hysteresis expressions in a lumped voltage prediction ...
    A phenomenon that has a large impact on battery model accuracy for NiMH batteries is open circuit voltage (OCV) hysteresis, which causes the OCV to differ not ...
  22. [22]
    How to Predict Fuel Cell Performance
    Jul 2, 2019 · The potential between the oxygen cathode where the reduction occurs and the hydrogen anode at which the oxidation occurs will be 1.229 V at ...
  23. [23]
    The effect of hydrogen crossover on open-circuit voltage in polymer ...
    We theoretically analyze below the role of hydrogen crossover in PEM fuel cells, in shortchanging the OCV from its promised value of around 1.23 V, and show ...
  24. [24]
  25. [25]
    The high open-circuit voltage of perovskite solar cells: a review
    One of the features of PSCs that make them stand out among all photovoltaics (PVs) is their high open-circuit voltage (VOC) although they are made by solution ...
  26. [26]
    Recombination Losses - PVEducation
    Recombination losses effect both the current collection (and therefore the short-circuit current) as well as the forward bias injection current.
  27. [27]
    17 The Laws of Induction - Feynman Lectures - Caltech
    The general law for the electric field associated with a changing magnetic field is ∇×E=−∂B∂t. We will call this Faraday's law. It was discovered by Faraday ...
  28. [28]
    13.6 Electric Generators and Back Emf – University Physics Volume 2
    The peak emf of a generator is ϵ 0 = N B A ω . Any rotating coil produces an induced emf. In motors, this is called back emf because it opposes the emf input to ...
  29. [29]
    [PDF] AC Generator
    The emf induced is c = -. dΦB dt. = 2πfNAB sin(2πft). (2). So to get a larger amplitude you could ... 1 point for three significant figures in equation (3)
  30. [30]
    [PDF] DC GENERATORS 10/00
    CALCULATIONS: 1. Plot the No-load saturation (magnetization) curve for VT vs IF. [ open circuit voltage versus increasing and decreasing field current ] ...
  31. [31]
    [PDF] design and fabrication of a hybrid nanoparticle-wick heat sink ...
    This is known as the Seebeck effect, where the Seebeck coefficient of a material is defined as a measure of the magnitude of induced thermoelectric voltage in ...
  32. [32]
    [PDF] Carbon Nanotube-based Thermoelectric Fabric for Wearable and ...
    Aug 27, 2024 · The p-type Sb-doped Bi2Te3 had a. Seebeck of 200 µV/K and in CNT array over 750 µV/K. This shows that CNTs are not only useful in stand-alone ...
  33. [33]
    [PDF] A study of low level vibrations as a power source for wireless sensor ...
    When a piezoelectric material is placed under a mechanical stress, an open circuit voltage (a charge separation) appears across the material. Likewise, if a.Missing: formula | Show results with:formula
  34. [34]
    Methods to Measure Open Circuit Voltage on a Battery Pack
    This application note describes several ways of measuring open circuit voltage on a battery pack including at the full pack level, on individual cells that are ...
  35. [35]
    Characteristics of Open Circuit Voltage Relaxation in Lithium-Ion ...
    Jul 26, 2022 · A relaxation period of 3 h is sufficient to achieve above 98% of the SS-OCV for SOC estimation from the captured data for NMC and LFP batteries.
  36. [36]
    Kelvin Connection - an overview | ScienceDirect Topics
    A four-wire (or Kelvin) connection to the device under test allows the voltage across the device to be measured by avoiding voltage drops along the wiring in ...
  37. [37]
    [PDF] An Accurate Electrical Battery Model Capable of Predicting Runtime ...
    The proposed model predicts all the im- portant properties and is compatible with lead-acid, NiCd,. NiMH, Li-Ion, polymer Li-Ion, and other electrochemical bat-.
  38. [38]
    [PDF] Accurate Electrical Battery Model Capable of Predicting Runtime ...
    This model accounts for all dynamic characteristics of the battery, from nonlinear open-circuit voltage, current-, temperature-, cy- cle number-, and storage ...
  39. [39]
    Electrochemical Impedance Spectroscopy A Tutorial
    The Nyquist and Bode plots for various simple electrical circuits containing a single passive element (R, C, or L) as well as combinations of them in ...
  40. [40]
    [PDF] Electrochemical Impedance Spectroscopy Primer
    The Nyquist plot in Figure 2-3 results from the electrical circuit of Figure 2-4. The semicircle is characteristic of a single "time constant". Electrochemical ...
  41. [41]
    Finite element modeling for analysis of electroluminescence and ...
    We investigate the potential of computationally efficient finite element simulation of solar cells and modules by considering planar electrodes.2. Results And Discussion · 2.1. Simulation Model · 2.3. Thermal Imaging
  42. [42]
    Galvanostatic intermittent titration technique (GITT) for Li-ion batteries
    GITT has three major applications for lithium-ion (Li-ion) batteries: the determination of the diffusion coefficient, open-circuit voltage (OCV) analysis, and ...
  43. [43]
    [PDF] Principles and Applications of Galvanostatic Intermittent Titration ...
    The GITT measures the transient voltage change and open-circuit voltage (OCV) change during the charging and discharging processes using only a constant current ...
  44. [44]
    The novel usage of spectroscopic ellipsometry for the development ...
    We present the novel use of spectroscopic ellipsometry (SE) for the development of a-Si:H solar cell. SE is a very fast and useful tool to measure various ...
  45. [45]
    [PDF] N91-30205 - NASA Technical Reports Server
    b) Open-circuit voltage, c) Shod-circuit current density, d). Fill factor. As ... Photovoitaic Cells Using Reference. Cells," ASTM Standard E948. Coutts ...
  46. [46]
  47. [47]
    How to Charge a Lead Acid Battery | Power Sonic
    The temperature coefficient is -2mV/cell/ºC below 20°C (68°F) in float use and -6mV/cell/ ºC below 20°C in cyclic use. For higher temperatures the charge ...
  48. [48]
    [PDF] Temperature dependence of solar cell performance—an analysis
    Mar 10, 2012 · The temperature dependence of reverse saturation current density, Jo and its effect on the para- meters, Voc, FF and Z is discussed.
  49. [49]
    [PDF] Effect of Humidity on Properties of Lithium-ion Batteries
    Mar 31, 2021 · The results of these analyses show that imperfect solid electrolyte interface formation increases the direct current resistance. This ...
  50. [50]
  51. [51]
    Effect of Light Intensity - PVEducation
    Therefore, under concentration, Voc increases logarithmically with light intensity, as shown in the equation below; where X is the concentration of sunlight.
  52. [52]
    (PDF) Failure analysis of lead-acid batteries at extreme operating ...
    Jun 7, 2023 · Environmental aging results in shorter cycle life due to the degradation of electrode and grid materials at higher temperatures (25°C and 40°C), ...
  53. [53]
    Battery open circuit voltage testing and application - TYCORUN
    Dec 26, 2023 · OCV test time: OCV test time refers to the time required to wait for the chemical reaction inside the battery to reach equilibrium after ...
  54. [54]
    Open-Circuit Voltage Models Should Be Thermodynamically ...
    Jan 24, 2024 · We propose a simple thermodynamically consistent OCV model enabled by differentiable thermodynamic modeling, which obeys the second law of thermodynamics.Supporting Information · Author Information · Biographies · References
  55. [55]
    A Parametric Open Circuit Voltage Model for Lithium Ion Batteries
    common electrode materials used in Li-ion cells, such as graphite,24. LiMnO2. 17 and LiCoO2. 25 and was also observed with the NMC cells used in this work (see ...
  56. [56]
    Effect of doping on the short-circuit current and open-circuit voltage ...
    Oct 17, 2014 · The results show that doping density is one of the decisive factors affecting the photocurrent of the PSCs.Missing: dopants | Show results with:dopants
  57. [57]
    Shunt Resistance | PVEducation
    Low shunt resistance causes power losses in solar cells by providing an alternate current path for the light-generated current.
  58. [58]
    Quantifying Losses in Open-Circuit Voltage in Solution-Processable ...
    Jul 28, 2015 · While most inorganic and perovskite solar cells show voltage losses due to nonradiative recombination, in organic solar cells the relatively ...
  59. [59]
    Exploring deep defect state impact on open circuit voltage of ...
    Sep 11, 2018 · The impurities in the donor acceptor materials and incomplete phase separation are likely to contribute in SRH recombination. These ...
  60. [60]
    Unraveling the LiMn2O4/LiPON Interface Degradation in All-Solid ...
    May 5, 2025 · The LMO-TFB exhibits a lower open circuit voltage (<2.8 V) than that of the LMO-LIB (∼3 V), which can be attributed to the infusion of Li+ ions ...
  61. [61]
    Integratable all-solid-state thin-film microbatteries - PNAS
    Apr 22, 2025 · This work advances the large-scale integration of microbatteries with microdevices, facilitating space-efficient and high-performance energy storage solutions.
  62. [62]
    BU-409: Charging Lithium-ion - Battery University
    The charger may kick in when the open circuit voltage drops to 4.05V/cell and turn off again at 4.20V/cell. Chargers made for operational readiness, or standby ...
  63. [63]
    Open circuit voltage characterization of lithium-ion batteries
    Dec 10, 2014 · A novel normalized OCV modeling approach is developed. The normalized OCV modeling is stable over temperature changes and battery aging.
  64. [64]
    Solar Cell Efficiency - PVEducation
    FF is the fill factor and η is the efficiency. The input power for efficiency calculations is 1 kW/m2 or 100 mW/cm2. Thus the input power for a 100 × 100 mm ...
  65. [65]
  66. [66]
    Advances in Lithium-Ion Batteries
    the open circuit voltage, OCV, of these early cells: Advances in Lithium-Ion Batteries. Edited by W. van Schalkwijk and B. Scrosati, Kluwer Academic/Plenum ...
  67. [67]
    Maximum Power Tracking Based Open Circuit Voltage Method for ...
    A new technique for tracking a pseudo-maximum power point for creating power reserve is presented. By using developed fractional open circuit voltage algorithm ...
  68. [68]
    Open-Circuit Voltage Models for Battery Management Systems - MDPI
    This paper presents a review of existing OCV–SOC models and proposes a systematic approach to select a suitable OCV–SOC for implementation based on various ...
  69. [69]
    Determination of SoH of Lead-Acid Batteries by Electrochemical ...
    It can be assumed that the decrease in measured C r e s is provoked not only by sulfation but also by the SoC decrease. In the test cells, the OCV changes only ...<|separator|>
  70. [70]
    Voltage equalization control algorithm for monitoring and balancing ...
    Mar 28, 2016 · This paper deals with the development of a voltage equalization control algorithm for individual cell monitoring and balancing of series connected Li-Ion ...
  71. [71]
    4.6: Maximum Power Transfer Theorem - Engineering LibreTexts
    Jun 5, 2021 · While matching the resistance produces the maximum load power, it does not produce maximum load current or maximum load voltage. In fact ...
  72. [72]
    Lithium-Ion Battery Management System for Electric Vehicles - MDPI
    It manages cell monitoring under discharge and charge conditions, overcharge and undercharge protection, temperature and heat monitoring, fault detection ...
  73. [73]
    Why do most battery protection circuits have 2.5V discharge cap?
    Nov 7, 2017 · Li-ion batteries can be discharged to a minimum of 2.5V, but it is recommended that the lowest voltage that a Li-ion or LiPo battery be ...<|separator|>
  74. [74]
    Optimal open-circuit voltage (OCV) model for improved electric ...
    From simulation results, proposed OCV model improves battery SOC by 10% after V2G operating period (2 p.m.) compared to a case without the model. Also, proposed ...