Potential Difference
1. Definition and Units of Potential Difference
Definition and Units of Potential Difference
Potential difference, often referred to as voltage, is a fundamental concept in electromagnetism and circuit theory. It quantifies the work done per unit charge to move a test charge between two points in an electric field. Mathematically, the potential difference V between points A and B is defined as:
where WAB is the work done to move a charge q from A to B. This scalar quantity is path-independent in conservative electric fields, meaning it depends solely on the endpoints.
Units and Dimensional Analysis
The SI unit of potential difference is the volt (V), equivalent to one joule per coulomb:
In dimensional terms, voltage is expressed as:
where M is mass, L is length, T is time, and I is electric current.
Electrostatic Potential vs. Electromotive Force
Potential difference is distinct from electromotive force (EMF), though both are measured in volts. While EMF describes energy conversion from non-electrical forms (e.g., in batteries), potential difference refers to energy dissipation across circuit elements. For a resistor, Ohm’s Law relates voltage to current:
Practical Measurement
Voltmeters measure potential difference by comparing the energy states of charges at two points. High-precision instruments, such as electrostatic voltmeters or digital multimeters, exploit this principle with minimal current draw to avoid perturbing the system.
Historical Context
The term volt honors Alessandro Volta, inventor of the voltaic pile (1800), the first electrochemical battery. This development underscored the quantifiable nature of electric potential and laid the groundwork for modern circuit analysis.
Applications in Modern Systems
Potential difference drives current in circuits, enabling technologies from microprocessors to power grids. In semiconductor physics, built-in potential (e.g., in PN junctions) governs device behavior. Superconducting systems exploit zero-resistance voltage drops for lossless energy transfer.
Definition and Units of Potential Difference
Potential difference, often referred to as voltage, is a fundamental concept in electromagnetism and circuit theory. It quantifies the work done per unit charge to move a test charge between two points in an electric field. Mathematically, the potential difference V between points A and B is defined as:
where WAB is the work done to move a charge q from A to B. This scalar quantity is path-independent in conservative electric fields, meaning it depends solely on the endpoints.
Units and Dimensional Analysis
The SI unit of potential difference is the volt (V), equivalent to one joule per coulomb:
In dimensional terms, voltage is expressed as:
where M is mass, L is length, T is time, and I is electric current.
Electrostatic Potential vs. Electromotive Force
Potential difference is distinct from electromotive force (EMF), though both are measured in volts. While EMF describes energy conversion from non-electrical forms (e.g., in batteries), potential difference refers to energy dissipation across circuit elements. For a resistor, Ohm’s Law relates voltage to current:
Practical Measurement
Voltmeters measure potential difference by comparing the energy states of charges at two points. High-precision instruments, such as electrostatic voltmeters or digital multimeters, exploit this principle with minimal current draw to avoid perturbing the system.
Historical Context
The term volt honors Alessandro Volta, inventor of the voltaic pile (1800), the first electrochemical battery. This development underscored the quantifiable nature of electric potential and laid the groundwork for modern circuit analysis.
Applications in Modern Systems
Potential difference drives current in circuits, enabling technologies from microprocessors to power grids. In semiconductor physics, built-in potential (e.g., in PN junctions) governs device behavior. Superconducting systems exploit zero-resistance voltage drops for lossless energy transfer.
Relationship Between Electric Field and Potential Difference
The fundamental connection between electric field E and electric potential V arises from the work-energy principle in electrostatics. For a conservative electric field, the potential difference between two points equals the negative line integral of the electric field along the path connecting them:
This integral relationship holds for any path from point a to point b in the field. The negative sign indicates that positive work against the field increases potential energy.
Differential Form
In differential form, the electric field relates to the potential gradient:
This implies:
- The electric field points in the direction of steepest potential decrease
- Field magnitude equals the spatial rate of potential change
- Equipotential surfaces are perpendicular to field lines
Uniform Field Case
For a constant electric field E between parallel plates separated by distance d, the relationship simplifies to:
This linear approximation is widely used in capacitor design, semiconductor devices, and electrostatic applications where field uniformity can be assumed.
Practical Implications
The field-potential relationship enables:
- Calculation of ionization potentials in plasma physics
- Determination of depletion region widths in pn-junctions
- Design of electron optics systems where particle trajectories depend on potential landscapes
In non-uniform fields, numerical methods like finite element analysis are typically employed to solve the inverse problem of determining potential distributions from measured field data.
Relationship Between Electric Field and Potential Difference
The fundamental connection between electric field E and electric potential V arises from the work-energy principle in electrostatics. For a conservative electric field, the potential difference between two points equals the negative line integral of the electric field along the path connecting them:
This integral relationship holds for any path from point a to point b in the field. The negative sign indicates that positive work against the field increases potential energy.
Differential Form
In differential form, the electric field relates to the potential gradient:
This implies:
- The electric field points in the direction of steepest potential decrease
- Field magnitude equals the spatial rate of potential change
- Equipotential surfaces are perpendicular to field lines
Uniform Field Case
For a constant electric field E between parallel plates separated by distance d, the relationship simplifies to:
This linear approximation is widely used in capacitor design, semiconductor devices, and electrostatic applications where field uniformity can be assumed.
Practical Implications
The field-potential relationship enables:
- Calculation of ionization potentials in plasma physics
- Determination of depletion region widths in pn-junctions
- Design of electron optics systems where particle trajectories depend on potential landscapes
In non-uniform fields, numerical methods like finite element analysis are typically employed to solve the inverse problem of determining potential distributions from measured field data.
1.3 Work Done in Moving a Charge
The work done in moving a charge through an electric field is a fundamental concept in electrostatics and circuit theory. When a charge q is displaced by an external force against an electric field E, work must be performed to overcome the Coulomb force. The infinitesimal work dW done in moving the charge by a displacement dl is given by:
where Fext is the applied force equal in magnitude but opposite in direction to the electric force qE. The negative sign indicates that work is done against the field. For a finite displacement from point A to point B, the total work done is the line integral:
Relationship to Potential Difference
This work directly relates to the electric potential difference VAB between the two points. By definition, potential difference is the work done per unit charge:
For a uniform electric field, this simplifies to:
where d is the displacement parallel to the field lines. The work done can then be expressed as:
Energy Considerations
This work represents energy transfer. When W > 0 (positive work), energy is added to the charge-field system, increasing its electric potential energy U:
Conversely, when the field does work on the charge (W < 0), the system loses potential energy. This energy conservation principle is crucial in analyzing circuits and electromagnetic systems.
Practical Implications
In circuit applications, this work manifests as:
- Energy storage in capacitors (W = ½CV2)
- Power dissipation in resistors (P = IV = I2R)
- Electromotive force in batteries
The concept extends to semiconductor physics, where potential differences govern charge carrier movement in p-n junctions and transistors. In particle accelerators, megavolt potential differences perform work on charged particles, converting electrical energy to kinetic energy.
Generalization for Time-Varying Fields
For non-conservative fields (e.g., induced electric fields from changing magnetic flux), the work integral must include the complete electromagnetic force:
This more general form is essential for analyzing electromagnetic induction and alternating current systems.
1.3 Work Done in Moving a Charge
The work done in moving a charge through an electric field is a fundamental concept in electrostatics and circuit theory. When a charge q is displaced by an external force against an electric field E, work must be performed to overcome the Coulomb force. The infinitesimal work dW done in moving the charge by a displacement dl is given by:
where Fext is the applied force equal in magnitude but opposite in direction to the electric force qE. The negative sign indicates that work is done against the field. For a finite displacement from point A to point B, the total work done is the line integral:
Relationship to Potential Difference
This work directly relates to the electric potential difference VAB between the two points. By definition, potential difference is the work done per unit charge:
For a uniform electric field, this simplifies to:
where d is the displacement parallel to the field lines. The work done can then be expressed as:
Energy Considerations
This work represents energy transfer. When W > 0 (positive work), energy is added to the charge-field system, increasing its electric potential energy U:
Conversely, when the field does work on the charge (W < 0), the system loses potential energy. This energy conservation principle is crucial in analyzing circuits and electromagnetic systems.
Practical Implications
In circuit applications, this work manifests as:
- Energy storage in capacitors (W = ½CV2)
- Power dissipation in resistors (P = IV = I2R)
- Electromotive force in batteries
The concept extends to semiconductor physics, where potential differences govern charge carrier movement in p-n junctions and transistors. In particle accelerators, megavolt potential differences perform work on charged particles, converting electrical energy to kinetic energy.
Generalization for Time-Varying Fields
For non-conservative fields (e.g., induced electric fields from changing magnetic flux), the work integral must include the complete electromagnetic force:
This more general form is essential for analyzing electromagnetic induction and alternating current systems.
2. Voltmeters and Their Operation
2.1 Voltmeters and Their Operation
Basic Principle of Voltmeters
A voltmeter is an instrument designed to measure the potential difference between two points in an electrical circuit. Unlike an ammeter, which must be placed in series with the circuit, a voltmeter is connected in parallel to the component or section across which the voltage is to be measured. This ensures that the voltmeter draws minimal current, thereby minimizing its impact on the circuit's operation.
The fundamental operation of a voltmeter relies on converting the measured voltage into a measurable quantity, typically a current or a deflection in a mechanical indicator. The relationship between voltage and the resulting measurement can be expressed as:
where V is the voltage being measured, Im is the current through the voltmeter, and Rm is the internal resistance of the voltmeter.
Types of Voltmeters
Voltmeters can be broadly classified into two categories based on their working principle:
- Analog Voltmeters – These use a moving coil or moving iron mechanism to indicate voltage. The deflection of a needle is proportional to the voltage applied.
- Digital Voltmeters (DVMs) – These convert the analog voltage into a digital signal using an analog-to-digital converter (ADC) and display the result numerically.
Analog Voltmeter: Moving Coil Mechanism
The most common type of analog voltmeter is the permanent magnet moving coil (PMMC) instrument. Its operation is based on the torque produced by a current-carrying coil in a magnetic field. The torque τ is given by:
where n is the number of turns in the coil, B is the magnetic flux density, A is the area of the coil, and I is the current through the coil. The deflection angle θ is proportional to the current and thus to the voltage being measured.
Digital Voltmeter: Operation and Advantages
A digital voltmeter (DVM) employs an ADC to convert the input voltage into a digital value. The most common types of DVMs use:
- Successive Approximation ADC – Balances the input voltage against a reference voltage using a binary search algorithm.
- Dual-Slope ADC – Integrates the input voltage over time and compares it to a reference, providing high accuracy and noise immunity.
The primary advantages of DVMs include higher precision, automatic range selection, and the ability to interface with digital systems for data logging and processing.
Loading Effect and Input Impedance
An important consideration in voltmeter design is the loading effect, where the voltmeter's internal resistance affects the circuit being measured. For accurate measurements, the voltmeter's input impedance Rin must be significantly higher than the equivalent resistance of the circuit. The error introduced by finite input impedance can be quantified as:
where Req is the Thevenin equivalent resistance of the circuit at the measurement points.
Practical Applications and Modern Developments
Modern voltmeters, particularly DVMs, are integral to laboratory and industrial applications. High-impedance voltmeters (e.g., FET-input or electrometer-grade) are used in sensitive measurements where circuit loading must be minimized. Additionally, specialized voltmeters such as vector voltmeters measure both magnitude and phase of AC voltages, crucial in RF and communication systems.
Recent advancements include the integration of voltmeters into oscilloscopes and multifunction test instruments, enabling simultaneous voltage, current, and frequency measurements with high precision.
2.1 Voltmeters and Their Operation
Basic Principle of Voltmeters
A voltmeter is an instrument designed to measure the potential difference between two points in an electrical circuit. Unlike an ammeter, which must be placed in series with the circuit, a voltmeter is connected in parallel to the component or section across which the voltage is to be measured. This ensures that the voltmeter draws minimal current, thereby minimizing its impact on the circuit's operation.
The fundamental operation of a voltmeter relies on converting the measured voltage into a measurable quantity, typically a current or a deflection in a mechanical indicator. The relationship between voltage and the resulting measurement can be expressed as:
where V is the voltage being measured, Im is the current through the voltmeter, and Rm is the internal resistance of the voltmeter.
Types of Voltmeters
Voltmeters can be broadly classified into two categories based on their working principle:
- Analog Voltmeters – These use a moving coil or moving iron mechanism to indicate voltage. The deflection of a needle is proportional to the voltage applied.
- Digital Voltmeters (DVMs) – These convert the analog voltage into a digital signal using an analog-to-digital converter (ADC) and display the result numerically.
Analog Voltmeter: Moving Coil Mechanism
The most common type of analog voltmeter is the permanent magnet moving coil (PMMC) instrument. Its operation is based on the torque produced by a current-carrying coil in a magnetic field. The torque τ is given by:
where n is the number of turns in the coil, B is the magnetic flux density, A is the area of the coil, and I is the current through the coil. The deflection angle θ is proportional to the current and thus to the voltage being measured.
Digital Voltmeter: Operation and Advantages
A digital voltmeter (DVM) employs an ADC to convert the input voltage into a digital value. The most common types of DVMs use:
- Successive Approximation ADC – Balances the input voltage against a reference voltage using a binary search algorithm.
- Dual-Slope ADC – Integrates the input voltage over time and compares it to a reference, providing high accuracy and noise immunity.
The primary advantages of DVMs include higher precision, automatic range selection, and the ability to interface with digital systems for data logging and processing.
Loading Effect and Input Impedance
An important consideration in voltmeter design is the loading effect, where the voltmeter's internal resistance affects the circuit being measured. For accurate measurements, the voltmeter's input impedance Rin must be significantly higher than the equivalent resistance of the circuit. The error introduced by finite input impedance can be quantified as:
where Req is the Thevenin equivalent resistance of the circuit at the measurement points.
Practical Applications and Modern Developments
Modern voltmeters, particularly DVMs, are integral to laboratory and industrial applications. High-impedance voltmeters (e.g., FET-input or electrometer-grade) are used in sensitive measurements where circuit loading must be minimized. Additionally, specialized voltmeters such as vector voltmeters measure both magnitude and phase of AC voltages, crucial in RF and communication systems.
Recent advancements include the integration of voltmeters into oscilloscopes and multifunction test instruments, enabling simultaneous voltage, current, and frequency measurements with high precision.
2.2 Practical Considerations in Measurement
Instrumentation and Loading Effects
Accurate measurement of potential difference requires careful selection of instrumentation to minimize loading effects. Voltmeters, whether analog or digital, exhibit finite input impedance Zin, which forms a parallel circuit with the test points. For a source impedance Zs and measured voltage Vtrue, the observed voltage Vmeas is:
High-impedance sources (e.g., piezoelectric sensors) demand voltmeters with Zin > 10 GΩ to maintain <1% error. Electrometer-grade instruments achieve this through guarded inputs and FET-based amplification.
Ground Loops and Common-Mode Interference
When measuring across non-isolated systems, ground loops introduce spurious potentials due to circulating currents. The resulting error voltage Verror depends on the loop area A, magnetic flux density B, and its rate of change:
Differential measurement techniques with instrumentation amplifiers (CMRR > 100 dB) suppress common-mode noise. For high-frequency interference, twisted-pair cabling and RF shielding become essential.
Thermal EMFs and Contact Potentials
Junctions of dissimilar metals generate Seebeck-effect voltages that compound measurement uncertainty. For copper-constantan connections at ΔT = 1°C:
Gold-plated contacts and isothermal probe designs reduce this effect. Null measurement techniques (e.g., potentiometers) eliminate current flow through junctions, effectively canceling thermal EMFs.
High-Voltage Measurement Challenges
Beyond 1 kV, field distortion and dielectric absorption introduce non-linear errors. Capacitive voltage dividers must account for stray capacitance Cstray:
Precision resistive dividers require temperature-stable materials like bulk metal foil (αR < 2 ppm/°C) to maintain ratio accuracy under thermal gradients.
Dynamic Signal Considerations
For time-varying potentials, instrument bandwidth must exceed the signal's highest significant harmonic. The risetime tr of a measurement system with bandwidth BW follows:
Sampling systems must adhere to Nyquist criteria while accounting for aperture uncertainty. For pulsed measurements, integrating oscilloscopes provide better accuracy than peak-detecting voltmeters.
2.2 Practical Considerations in Measurement
Instrumentation and Loading Effects
Accurate measurement of potential difference requires careful selection of instrumentation to minimize loading effects. Voltmeters, whether analog or digital, exhibit finite input impedance Zin, which forms a parallel circuit with the test points. For a source impedance Zs and measured voltage Vtrue, the observed voltage Vmeas is:
High-impedance sources (e.g., piezoelectric sensors) demand voltmeters with Zin > 10 GΩ to maintain <1% error. Electrometer-grade instruments achieve this through guarded inputs and FET-based amplification.
Ground Loops and Common-Mode Interference
When measuring across non-isolated systems, ground loops introduce spurious potentials due to circulating currents. The resulting error voltage Verror depends on the loop area A, magnetic flux density B, and its rate of change:
Differential measurement techniques with instrumentation amplifiers (CMRR > 100 dB) suppress common-mode noise. For high-frequency interference, twisted-pair cabling and RF shielding become essential.
Thermal EMFs and Contact Potentials
Junctions of dissimilar metals generate Seebeck-effect voltages that compound measurement uncertainty. For copper-constantan connections at ΔT = 1°C:
Gold-plated contacts and isothermal probe designs reduce this effect. Null measurement techniques (e.g., potentiometers) eliminate current flow through junctions, effectively canceling thermal EMFs.
High-Voltage Measurement Challenges
Beyond 1 kV, field distortion and dielectric absorption introduce non-linear errors. Capacitive voltage dividers must account for stray capacitance Cstray:
Precision resistive dividers require temperature-stable materials like bulk metal foil (αR < 2 ppm/°C) to maintain ratio accuracy under thermal gradients.
Dynamic Signal Considerations
For time-varying potentials, instrument bandwidth must exceed the signal's highest significant harmonic. The risetime tr of a measurement system with bandwidth BW follows:
Sampling systems must adhere to Nyquist criteria while accounting for aperture uncertainty. For pulsed measurements, integrating oscilloscopes provide better accuracy than peak-detecting voltmeters.
2.3 Common Sources of Error
Instrumentation Limitations
Voltmeters introduce finite input impedance Zin, causing loading effects when measuring high-impedance circuits. The measured potential difference Vm relates to the true voltage Vt as:
For a source impedance Zsource = 10 kΩ and a voltmeter with Zin = 1 MΩ, this creates a 1% systematic error. Electrometer-grade instruments (>10 GΩ input impedance) are required for nanocurrent measurements.
Thermal EMF Effects
Dissimilar metal junctions in test leads generate Seebeck-effect voltages (0.1 μV/K to 10 μV/K per junction). For copper-constantan thermocouples at ΔT = 5°C:
This becomes significant in sub-millivolt measurements. Strategies include:
- Using identical copper leads for all connections
- Implementing offset-nulling techniques
- Maintaining isothermal conditions
Ground Loops
Multiple ground references create circulating currents through finite conductor impedance. The error voltage Vloop follows:
At 60 Hz with 10 cm of 22 AWG wire (R = 52 mΩ, L = 100 nH), a 1 mA ground current produces 52 μV of 60 Hz ripple plus harmonic content.
Dielectric Absorption
Charge trapping in cable insulation creates hysteresis in fast voltage measurements. The relaxation current follows a stretched exponential:
Where 0 < β < 1 characterizes material disorder. PTFE-insulated cables exhibit τ ≈ 100 ms and β ≈ 0.7, while PVC shows τ ≈ 10 s with β ≈ 0.5.
Quantum Limitations
At cryogenic temperatures, Johnson-Nyquist noise sets fundamental limits. The spectral density for a 1 kΩ resistor at 4 K is:
This corresponds to 470 fV/√Hz noise density, requiring SQUID-based amplification for sub-nanovolt measurements.
2.3 Common Sources of Error
Instrumentation Limitations
Voltmeters introduce finite input impedance Zin, causing loading effects when measuring high-impedance circuits. The measured potential difference Vm relates to the true voltage Vt as:
For a source impedance Zsource = 10 kΩ and a voltmeter with Zin = 1 MΩ, this creates a 1% systematic error. Electrometer-grade instruments (>10 GΩ input impedance) are required for nanocurrent measurements.
Thermal EMF Effects
Dissimilar metal junctions in test leads generate Seebeck-effect voltages (0.1 μV/K to 10 μV/K per junction). For copper-constantan thermocouples at ΔT = 5°C:
This becomes significant in sub-millivolt measurements. Strategies include:
- Using identical copper leads for all connections
- Implementing offset-nulling techniques
- Maintaining isothermal conditions
Ground Loops
Multiple ground references create circulating currents through finite conductor impedance. The error voltage Vloop follows:
At 60 Hz with 10 cm of 22 AWG wire (R = 52 mΩ, L = 100 nH), a 1 mA ground current produces 52 μV of 60 Hz ripple plus harmonic content.
Dielectric Absorption
Charge trapping in cable insulation creates hysteresis in fast voltage measurements. The relaxation current follows a stretched exponential:
Where 0 < β < 1 characterizes material disorder. PTFE-insulated cables exhibit τ ≈ 100 ms and β ≈ 0.7, while PVC shows τ ≈ 10 s with β ≈ 0.5.
Quantum Limitations
At cryogenic temperatures, Johnson-Nyquist noise sets fundamental limits. The spectral density for a 1 kΩ resistor at 4 K is:
This corresponds to 470 fV/√Hz noise density, requiring SQUID-based amplification for sub-nanovolt measurements.
3. Potential Difference in Series Circuits
3.1 Potential Difference in Series Circuits
In a series circuit, components are connected end-to-end, forming a single path for current flow. The potential difference (voltage) across each component depends on its resistance and the total current flowing through the circuit. Kirchhoff's Voltage Law (KVL) governs the distribution of voltage, stating that the sum of potential differences around any closed loop must equal the applied voltage.
Mathematical Derivation
Consider a series circuit with N resistors R₁, R₂, ..., Rₙ connected to a voltage source V. The total resistance Rtotal is the sum of individual resistances:
Using Ohm's Law, the current I through the circuit is:
The potential difference Vi across each resistor Ri is then:
This shows that the voltage divides proportionally to the resistance in a series configuration.
Practical Implications
Series circuits are widely used in voltage divider networks, sensor biasing, and precision measurement systems. For example, in a resistive voltage divider, two resistors in series split the input voltage into a fraction determined by their ratio. This principle is foundational in analog signal conditioning and reference voltage generation.
Case Study: Precision Voltage Reference
High-precision voltage references often employ series-connected resistors to generate stable, low-noise outputs. If R₁ = 9 kΩ and R₂ = 1 kΩ are connected in series to a 10 V source, the output voltage across R₂ is:
Tolerance and temperature stability of resistors directly affect accuracy, making material selection critical.
Non-Ideal Considerations
Real-world components introduce parasitic effects. Resistors exhibit:
- Temperature dependence: Resistance varies with ambient conditions.
- Parasitic capacitance: High-frequency performance degrades due to unintended capacitive coupling.
- Noise: Thermal (Johnson-Nyquist) and flicker noise contribute to signal instability.
For high-frequency or high-precision applications, these factors necessitate careful circuit design, including the use of low-temperature-coefficient (LTC) resistors and shielding.
Experimental Validation
To empirically verify voltage division, measure the potential difference across each resistor in a series network using a high-impedance voltmeter. Discrepancies from theoretical values may indicate:
- Non-ideal meter loading (finite input impedance).
- Resistor tolerance deviations.
- Contact resistance at junctions.
For rigorous validation, use a four-wire Kelvin measurement to eliminate lead resistance errors.
3.1 Potential Difference in Series Circuits
In a series circuit, components are connected end-to-end, forming a single path for current flow. The potential difference (voltage) across each component depends on its resistance and the total current flowing through the circuit. Kirchhoff's Voltage Law (KVL) governs the distribution of voltage, stating that the sum of potential differences around any closed loop must equal the applied voltage.
Mathematical Derivation
Consider a series circuit with N resistors R₁, R₂, ..., Rₙ connected to a voltage source V. The total resistance Rtotal is the sum of individual resistances:
Using Ohm's Law, the current I through the circuit is:
The potential difference Vi across each resistor Ri is then:
This shows that the voltage divides proportionally to the resistance in a series configuration.
Practical Implications
Series circuits are widely used in voltage divider networks, sensor biasing, and precision measurement systems. For example, in a resistive voltage divider, two resistors in series split the input voltage into a fraction determined by their ratio. This principle is foundational in analog signal conditioning and reference voltage generation.
Case Study: Precision Voltage Reference
High-precision voltage references often employ series-connected resistors to generate stable, low-noise outputs. If R₁ = 9 kΩ and R₂ = 1 kΩ are connected in series to a 10 V source, the output voltage across R₂ is:
Tolerance and temperature stability of resistors directly affect accuracy, making material selection critical.
Non-Ideal Considerations
Real-world components introduce parasitic effects. Resistors exhibit:
- Temperature dependence: Resistance varies with ambient conditions.
- Parasitic capacitance: High-frequency performance degrades due to unintended capacitive coupling.
- Noise: Thermal (Johnson-Nyquist) and flicker noise contribute to signal instability.
For high-frequency or high-precision applications, these factors necessitate careful circuit design, including the use of low-temperature-coefficient (LTC) resistors and shielding.
Experimental Validation
To empirically verify voltage division, measure the potential difference across each resistor in a series network using a high-impedance voltmeter. Discrepancies from theoretical values may indicate:
- Non-ideal meter loading (finite input impedance).
- Resistor tolerance deviations.
- Contact resistance at junctions.
For rigorous validation, use a four-wire Kelvin measurement to eliminate lead resistance errors.
3.2 Potential Difference in Parallel Circuits
In parallel circuits, the potential difference (voltage) across each branch is identical and equal to the total voltage supplied by the source. This fundamental property arises from Kirchhoff’s Voltage Law (KVL), which states that the sum of potential differences around any closed loop in a circuit must be zero. For a parallel configuration, all branches share the same two nodes, enforcing equal voltage across them.
Mathematical Derivation
Consider a parallel circuit with N resistors connected across a voltage source V. By definition, the potential difference across each resistor Ri is:
This equality holds regardless of the individual resistances or currents flowing through each branch. The current through each resistor is determined by Ohm’s Law:
The total current Itotal drawn from the source is the sum of the branch currents:
This leads to the equivalent resistance Req of the parallel network:
Practical Implications
The uniform potential difference in parallel circuits has critical applications:
- Power Distribution: Household wiring uses parallel connections to ensure all appliances receive the same voltage (e.g., 120V or 230V).
- Circuit Redundancy: If one branch fails (e.g., an open circuit), the remaining branches continue operating at full voltage.
- Current Sharing: High-power systems, such as LED arrays or server power supplies, distribute current across parallel paths to avoid overloading individual components.
Non-Ideal Considerations
In real-world systems, parasitic resistances (e.g., wire resistance, contact resistance) introduce minor voltage drops across branches. For precision applications, these effects are modeled as:
where Rparasitic includes trace resistances and connector losses. In high-current designs, Rparasitic must be minimized to maintain voltage uniformity.
Case Study: Superconducting Parallel Links
In superconducting quantum computers, parallel Josephson junctions exhibit zero voltage drop when current-sharing. This ideal behavior is exploited to create ultra-low-loss interconnects, where:
Deviations occur only if a junction transitions to a resistive state, highlighting the importance of uniform critical current densities in parallel junction fabrication.
3.2 Potential Difference in Parallel Circuits
In parallel circuits, the potential difference (voltage) across each branch is identical and equal to the total voltage supplied by the source. This fundamental property arises from Kirchhoff’s Voltage Law (KVL), which states that the sum of potential differences around any closed loop in a circuit must be zero. For a parallel configuration, all branches share the same two nodes, enforcing equal voltage across them.
Mathematical Derivation
Consider a parallel circuit with N resistors connected across a voltage source V. By definition, the potential difference across each resistor Ri is:
This equality holds regardless of the individual resistances or currents flowing through each branch. The current through each resistor is determined by Ohm’s Law:
The total current Itotal drawn from the source is the sum of the branch currents:
This leads to the equivalent resistance Req of the parallel network:
Practical Implications
The uniform potential difference in parallel circuits has critical applications:
- Power Distribution: Household wiring uses parallel connections to ensure all appliances receive the same voltage (e.g., 120V or 230V).
- Circuit Redundancy: If one branch fails (e.g., an open circuit), the remaining branches continue operating at full voltage.
- Current Sharing: High-power systems, such as LED arrays or server power supplies, distribute current across parallel paths to avoid overloading individual components.
Non-Ideal Considerations
In real-world systems, parasitic resistances (e.g., wire resistance, contact resistance) introduce minor voltage drops across branches. For precision applications, these effects are modeled as:
where Rparasitic includes trace resistances and connector losses. In high-current designs, Rparasitic must be minimized to maintain voltage uniformity.
Case Study: Superconducting Parallel Links
In superconducting quantum computers, parallel Josephson junctions exhibit zero voltage drop when current-sharing. This ideal behavior is exploited to create ultra-low-loss interconnects, where:
Deviations occur only if a junction transitions to a resistive state, highlighting the importance of uniform critical current densities in parallel junction fabrication.
3.3 Kirchhoff’s Voltage Law
Kirchhoff’s Voltage Law (KVL) states that the algebraic sum of all voltages around any closed loop in a circuit is zero. This principle arises from the conservation of energy in an electric field and is foundational for analyzing complex circuits. Mathematically, for a loop with n voltage drops:
Derivation from Energy Conservation
Consider a charge q traversing a closed loop. The net work done by the electric field must be zero, as the charge returns to its initial potential. The work done per unit charge is the voltage, leading to:
Practical Application: Multi-Loop Circuit Analysis
To apply KVL in a multi-loop circuit:
- Define a loop direction (clockwise/counterclockwise).
- Assign polarities to voltage drops across components (e.g., resistors: drop in the direction of current).
- Sum voltages algebraically, treating rises as positive and drops as negative.
Example: Series RC Circuit
For a loop with a voltage source V_s, resistor R, and capacitor C, KVL yields:
Differentiating with respect to time gives the differential equation for the circuit’s transient response.
Limitations and Assumptions
KVL assumes:
- Lumped-element model: No significant varying magnetic fields linking the loop (violations require inclusion of induced EMF via Faraday’s Law).
- Quasi-static conditions: Circuit dimensions are small compared to the wavelength of operating frequencies.
Advanced Context: Mesh Analysis
KVL underpins mesh analysis, where currents in independent loops (mesh currents) are solved simultaneously. For a circuit with m meshes, the system of equations is:
where R_ij represents resistances common to meshes i and j, and V_i is the net voltage source in mesh i.
3.3 Kirchhoff’s Voltage Law
Kirchhoff’s Voltage Law (KVL) states that the algebraic sum of all voltages around any closed loop in a circuit is zero. This principle arises from the conservation of energy in an electric field and is foundational for analyzing complex circuits. Mathematically, for a loop with n voltage drops:
Derivation from Energy Conservation
Consider a charge q traversing a closed loop. The net work done by the electric field must be zero, as the charge returns to its initial potential. The work done per unit charge is the voltage, leading to:
Practical Application: Multi-Loop Circuit Analysis
To apply KVL in a multi-loop circuit:
- Define a loop direction (clockwise/counterclockwise).
- Assign polarities to voltage drops across components (e.g., resistors: drop in the direction of current).
- Sum voltages algebraically, treating rises as positive and drops as negative.
Example: Series RC Circuit
For a loop with a voltage source V_s, resistor R, and capacitor C, KVL yields:
Differentiating with respect to time gives the differential equation for the circuit’s transient response.
Limitations and Assumptions
KVL assumes:
- Lumped-element model: No significant varying magnetic fields linking the loop (violations require inclusion of induced EMF via Faraday’s Law).
- Quasi-static conditions: Circuit dimensions are small compared to the wavelength of operating frequencies.
Advanced Context: Mesh Analysis
KVL underpins mesh analysis, where currents in independent loops (mesh currents) are solved simultaneously. For a circuit with m meshes, the system of equations is:
where R_ij represents resistances common to meshes i and j, and V_i is the net voltage source in mesh i.
4. Potential Difference in Batteries and Power Supplies
4.1 Potential Difference in Batteries and Power Supplies
Electrochemical Basis of Battery Potential
The potential difference in a battery arises from electrochemical reactions at the anode and cathode. The Nernst equation describes the equilibrium potential for a single electrode:
where E is the electrode potential, E0 is the standard electrode potential, R is the gas constant, T is temperature, n is the number of electrons transferred, F is Faraday's constant, and Q is the reaction quotient. For a complete cell, the net potential difference is:
Internal Resistance and Terminal Voltage
Real batteries exhibit internal resistance (rint), causing the terminal voltage (Vterm) to differ from the open-circuit potential (Voc):
where I is the discharge current. This relationship explains voltage sag under load and the importance of low-impedance designs for high-current applications.
Power Supply Regulation
Modern power supplies use feedback control to maintain constant output voltage despite load variations. A basic linear regulator maintains potential difference through:
where Vref is a reference voltage. Switching regulators achieve higher efficiency by rapidly modulating the duty cycle (D) of a transistor:
Practical Considerations
- Battery aging increases internal resistance and reduces available potential difference
- Load matching maximizes power transfer when Rload = rint
- Transient response becomes critical in high-speed digital systems
Measurement Techniques
Four-point probe measurements eliminate lead resistance errors when characterizing battery potential:
Precision measurements require null detection methods or instrumentation amplifiers with high common-mode rejection ratios.
4.1 Potential Difference in Batteries and Power Supplies
Electrochemical Basis of Battery Potential
The potential difference in a battery arises from electrochemical reactions at the anode and cathode. The Nernst equation describes the equilibrium potential for a single electrode:
where E is the electrode potential, E0 is the standard electrode potential, R is the gas constant, T is temperature, n is the number of electrons transferred, F is Faraday's constant, and Q is the reaction quotient. For a complete cell, the net potential difference is:
Internal Resistance and Terminal Voltage
Real batteries exhibit internal resistance (rint), causing the terminal voltage (Vterm) to differ from the open-circuit potential (Voc):
where I is the discharge current. This relationship explains voltage sag under load and the importance of low-impedance designs for high-current applications.
Power Supply Regulation
Modern power supplies use feedback control to maintain constant output voltage despite load variations. A basic linear regulator maintains potential difference through:
where Vref is a reference voltage. Switching regulators achieve higher efficiency by rapidly modulating the duty cycle (D) of a transistor:
Practical Considerations
- Battery aging increases internal resistance and reduces available potential difference
- Load matching maximizes power transfer when Rload = rint
- Transient response becomes critical in high-speed digital systems
Measurement Techniques
Four-point probe measurements eliminate lead resistance errors when characterizing battery potential:
Precision measurements require null detection methods or instrumentation amplifiers with high common-mode rejection ratios.
4.2 Potential Difference in Electronic Components
Fundamentals of Potential Difference in Components
Potential difference (V) across an electronic component arises due to charge separation governed by the component's electrical properties. In a resistor, it is directly proportional to current (I) via Ohm's Law:
where R is resistance. For capacitors, potential difference builds as charge accumulates:
where Q is stored charge and C is capacitance. Inductors generate a potential difference in response to changing current:
where L is inductance.
Semiconductor Junctions and Built-in Potential
In p-n junctions, the contact potential difference (V₀) arises from Fermi-level alignment:
where NA and ND are doping concentrations, ni is intrinsic carrier density, k is Boltzmann's constant, and T is temperature. This potential barrier critically determines diode behavior under bias.
Transistor Operating Voltages
In MOSFETs, the threshold voltage (Vth) defines the gate potential required for inversion layer formation:
where VFB is flat-band voltage, φB is bulk potential, εs is semiconductor permittivity, and Cox is oxide capacitance.
Practical Measurement Considerations
When measuring potential differences in circuits:
- Voltmeter loading effects must be considered—the instrument's finite input resistance (Rin) forms a parallel circuit, altering the measured voltage.
- High-frequency measurements require compensation for parasitic inductances and capacitances that introduce phase shifts.
- Differential measurements using instrumentation amplifiers reject common-mode noise in low-voltage applications.
Thermal Voltage Effects
At small scales, thermal noise generates random potential differences described by the Johnson-Nyquist formula:
where B is bandwidth. This becomes significant in high-impedance circuits and low-noise amplifier design.
Nonlinear Components
In devices like varactors, the potential difference varies nonlinearly with charge:
where C0 is zero-bias capacitance and α, β are nonlinear coefficients. This behavior enables voltage-controlled oscillators in RF systems.
4.2 Potential Difference in Electronic Components
Fundamentals of Potential Difference in Components
Potential difference (V) across an electronic component arises due to charge separation governed by the component's electrical properties. In a resistor, it is directly proportional to current (I) via Ohm's Law:
where R is resistance. For capacitors, potential difference builds as charge accumulates:
where Q is stored charge and C is capacitance. Inductors generate a potential difference in response to changing current:
where L is inductance.
Semiconductor Junctions and Built-in Potential
In p-n junctions, the contact potential difference (V₀) arises from Fermi-level alignment:
where NA and ND are doping concentrations, ni is intrinsic carrier density, k is Boltzmann's constant, and T is temperature. This potential barrier critically determines diode behavior under bias.
Transistor Operating Voltages
In MOSFETs, the threshold voltage (Vth) defines the gate potential required for inversion layer formation:
where VFB is flat-band voltage, φB is bulk potential, εs is semiconductor permittivity, and Cox is oxide capacitance.
Practical Measurement Considerations
When measuring potential differences in circuits:
- Voltmeter loading effects must be considered—the instrument's finite input resistance (Rin) forms a parallel circuit, altering the measured voltage.
- High-frequency measurements require compensation for parasitic inductances and capacitances that introduce phase shifts.
- Differential measurements using instrumentation amplifiers reject common-mode noise in low-voltage applications.
Thermal Voltage Effects
At small scales, thermal noise generates random potential differences described by the Johnson-Nyquist formula:
where B is bandwidth. This becomes significant in high-impedance circuits and low-noise amplifier design.
Nonlinear Components
In devices like varactors, the potential difference varies nonlinearly with charge:
where C0 is zero-bias capacitance and α, β are nonlinear coefficients. This behavior enables voltage-controlled oscillators in RF systems.
4.3 Safety Considerations in High Voltage Applications
Electrical Breakdown and Insulation Requirements
High voltage systems operate at potentials where dielectric breakdown becomes a critical concern. The breakdown voltage Vb of an insulating material is given by:
where Eb is the dielectric strength (in V/m) and d is the thickness of the insulation. For air at standard temperature and pressure (STP), Eb ≈ 3 MV/m. However, this value decreases with humidity, pressure variations, and contamination. Practical systems must incorporate safety margins, typically designing for 2-3 times the expected operating voltage.
Creepage and Clearance Distances
Two key safety parameters govern physical layout:
- Clearance: The shortest air path between conductors
- Creepage: The shortest path along insulating surfaces
International standards (IEC 60664-1) specify minimum distances based on:
- Voltage level (peak and RMS)
- Pollution degree (1-4)
- Material group (I-III)
For example, a 10 kV system in pollution degree 2 requires at least 25 mm clearance and 32 mm creepage distance.
Arc Flash Hazards
High voltage arcs release tremendous energy through:
Arc temperatures exceed 20,000 K, producing:
- Intense UV/IR radiation
- Blast pressures >100 kPa
- Molten metal expulsion
The incident energy E (in cal/cm²) determines required personal protective equipment (PPE) levels per NFPA 70E standards.
Grounding and Shielding
Proper grounding strategies must account for:
Where Ifault is the maximum expected fault current and Rground is the grounding system resistance. For HV systems, multi-point grounding with equipotential bonding is essential to limit step and touch potentials below 50 V.
Electrostatic shielding becomes critical above 5 kV to prevent:
- Corona discharge
- Electromagnetic interference (EMI)
- Stray capacitance coupling
Safety Interlocks and Procedures
Engineered safeguards should include:
- Positive-opening disconnect switches
- Capacitor discharge circuits with verified voltage monitors
- Two-hand control systems for access
- Lockout/tagout (LOTO) protocols
For pulsed power systems, the stored energy E = ½CV² requires special consideration. A 10 μF capacitor charged to 50 kV stores 12.5 kJ - equivalent to 3 kg of TNT.
5. Recommended Textbooks
5.1 Recommended Textbooks
- The Best Online Library of Electrical Engineering Textbooks — Force, Energy, and Potential Difference 5.8; Independence of Path 5.9; Kirchoff's Voltage Law for Electrostatics - Integral Form 5.10; Kirchoff's Voltage Law for Electrostatics - Differential Form 5.11; Electric Potential Field Due to Point Charges 5.12; Electric Potential Field due to a Continuous Distribution of Charge 5.13
- PDF Electrical and Electronic Principles and Technology — 1.6 Electrical potential and e.m.f. 5 1.7 Resistance and conductance 6 1.8 Electrical power and energy 6 1.9 Summary of terms, units and their symbols 7 2 An introduction to electric circuits 9 2.1 Electrical/electronic system block diagrams 10 2.2 Standard symbols for electrical components 11 2.3 Electric current and quantity of electricity 11
- 19.1 Electric Potential Energy: Potential Difference — When such a battery moves charge, it puts the charge through a potential difference of 12.0 V, and the charge is given a change in potential energy equal to ΔPE = q Δ V ΔPE = q Δ V. So to find the energy output, we multiply the charge moved by the potential difference. Solution. For the motorcycle battery, q = 5000 C q = 5000 C and Δ V ...
- Electric Potential Difference - Physics Book - gatech.edu — Claimed and Written by Daniel Kurniawan for PHYS2212 The figure above shows a voltmeter measuring the potential difference in the battery. Electric Potential Difference, also known as voltage, is the difference in electric potential energy between two points per unit of electric charge. The voltage between two points is equal to the work done per unit of charge against an unchanging electric ...
- 5.1: Introduction - Physics LibreTexts — The ratio of the charge stored on the plates to the potential difference \(V\) across them is called the capacitance \(C\) of the capacitor. Thus: \[Q=CV.\label{5.1.1}\] If, when the potential difference is one volt, the charge stored is one coulomb, the capacitance is one farad, F. Thus, a farad is a coulomb per volt.
- IB Physics Notes - 5.1 Electric potential difference, current and ... — IB Physics notes on 5.1 Electric potential difference, current and resistance
- PDF 5.1.2 Potential difference and power - FLIPPED AROUND PHYSICS — Potential difference (!.#.) is the energy per unit charge given up by charges as they pass through a device (e.g. a resistor). The electrical energy given up by the charges is transferred into other forms of energy (e.g. a resistor transfers the energy to heat, a bulb transfers the energy to
- Electric Field as the Gradient of Potential - CircuitBread — In Section 5.12, we defined the scalar electric potential field . as the electric potential difference at . relative to a datum at infinity. In this section, we address the "inverse problem" - namely, how to calculate . given . Specifically, we are interested in a direct "point-wise" mathematical transform from one to the other.
- 3: Electric Potential - Physics LibreTexts — Furthermore, spherical charge distributions (such as charge on a metal sphere) create external electric fields exactly like a point charge. The electric potential due to a point charge is, thus, a case we need to consider. 3.5: Determining Field from Potential In certain systems, we can calculate the potential by integrating over the electric ...
- PDF Chapter 5 — potential •V = V (x,y,z) electric potential in a region of space that do not contain any electric charges Note: this is a 2-D motion •Different from previous cases: here boundary conditions are needed in place of initial conditions. Partial differential equation (PDE) Eq. (5.1)
5.2 Online Resources
- Unit 5.2.2: Current, Potential Difference and Electromotive Force — P5.2.2 Current, potential difference and electromotive force (e.m.f.) Core\u000B1 Demonstrate understanding of current, potential difference, e.m.f. and resistance\u000B2 State that current is related to the flow of charge\u000B4 State that current in metals is due to a flow of electrons\u000B5 State that the
- 19.1 Electric Potential Energy: Potential Difference — To say we have a 12.0 V battery means that its terminals have a 12.0 V potential difference. When such a battery moves charge, it puts the charge through a potential difference of 12.0 V, and the charge is given a change in potential energy equal to . So to find the energy output, we multiply the charge moved by the potential difference. Solution
- 19.1 Electric Potential Energy: Potential Difference — An electron accelerated through a potential difference of 1 V is given an energy of 1 eV. It follows that an electron accelerated through 50 V is given 50 eV. A potential difference of 100,000 V (100 kV) will give an electron an energy of 100,000 eV (100 keV), and so on.
- 19.1: Electric Potential Energy- Potential Difference — An electron accelerated through a potential difference of 1 V is given an energy of 1 eV. It follows that an electron accelerated through 50 V is given 50 eV. A potential difference of 100,000 V (100 kV) will give an electron an energy of 100,000 eV (100 keV), and so on.
- Electromotive Force & Potential Difference — The potential difference is the difference in the electrical potential across each component: 5 volts for the bulb (on the left) and 7 volts for the resistor (on the right) The definition of p.d. can also be expressed using an equation. Where. V = potential difference (p.d.) (V) W = energy transferred to the components from the charge carriers (J)
- Potential Divider Circuits (5.2.9) | DP IB Physics: SL Revision Notes ... — The input voltage V in is applied to the top and bottom of the series resistors; The output voltage V out is measured from the centre to the bottom of resistor R 2; The potential difference V across each resistor depends upon its resistance R:. The resistor with the largest resistance will have a greater potential difference than the other one from V = IR; If the resistance of one of the ...
- Potential Difference (Voltage) | iCalculator™ — b) The negative result shows the work is done by the electric force, not by the external forces. Hence, we obtain the value 2 × 10-3 J for the work done by the electric force.. You have reached the end of Physics lesson 14.5.2 Potential Difference (Voltage).There are 8 lessons in this physics tutorial covering Electric Potential, you can access all the lessons from this tutorial below.
- IGCSE Physics 4.2.3 - EMF & Potential Difference — 4.2.3 Electromotive force and potential difference. electromotive force (e.m.f.) and potential difference (p.d.) Electromotive Force (e.m.f.) Definition: The electrical work done by a source in moving a unit charge around a complete circuit. Measured in: Volts (V). Explanation:
- 7.2 Electric Potential and Potential Difference — Voltage is not the same as energy. Voltage is the energy per unit charge. Thus, a motorcycle battery and a car battery can both have the same voltage (more precisely, the same potential difference between battery terminals), yet one stores much more energy than the other because [latex]\text{Δ}U=q\text{Δ}V.[/latex] The car battery can move more charge than the motorcycle battery, although ...
- Electric Potential Difference Instruction Flashcards - Quizlet — Study with Quizlet and memorize flashcards containing terms like A point charge of 4.0 µC is placed at a distance of 0.10 m from a hard rubber rod with an electric field of 1.0 × 103 . What is the electric potential energy of the point charge? ⇒ -- × 10-4 J What is the electric potential energy of the point charge at 1.3 m? ⇒ -- × 10-3 J, Identify the charge(s) with increasing electric ...
5.3 Research Papers and Articles
- 7.3: Electric Potential and Potential Difference — This page titled 7.3: Electric Potential and Potential Difference is shared under a CC BY 4.0 license and was authored, remixed, and/or curated by OpenStax via source content that was edited to the style and standards of the LibreTexts platform.
- Physics Tutorial: Electric Potential Difference — This difference in electric potential is represented by the symbol ΔV and is formally referred to as the electric potential difference. By definition, the electric potential difference is the difference in electric potential (V) between the final and the initial location when work is done upon a charge to change its potential energy.
- 3.5: Electric Potential - Physics LibreTexts — Learning Objectives By the end of this section, you will be able to: Define electric potential, voltage, and potential difference. Calculate electric potential and potential difference from potential energy. Define the electron-volt. Describe systems in which the electron-volt is a useful unit. Apply conservation of energy to systems with electric charges.
- Pumping electrons from chemical potential difference — Recent advances in development of triboelectric generators have demonstrated a variety of potential applications [15], [16]. Making use of the chemical potential difference and temperature difference between a metal tip and testing sample surface, thermoelectric potential variations were characterized with the atomic resolution [24].
- Betavoltaics using scandium tritide and contact potential difference — Tritium-powered betavoltaic micropower sources using contact potential difference (CPD) are demonstrated. Thermally stable scandium tritide thin films with a surface activity of 15 mCi ∕ cm 2 were used as the beta particle source. The electrical field created by the work function difference between the ScT film and a platinum or copper electrode was used to separate the beta-generated ...
- 5.3: Potential Energy Curves - Chemistry LibreTexts — The energy for the electronic excitation of an iodine atom E (I*) is known quite accurately from atomic spectroscopy, the value being 7603 cm -1. This energy is just the separation in energy between the iodine molecule X and B state potential curves in the limit where R approaches ∞ ∞ (See Figure 5.3.3).
- Electric Potential, Capacitors, and Dielectrics | SpringerLink — This chapter continues our study of electrostatics, introducing the concepts of electric potential and capacitance. We analyze electrical circuits containing capacitors in parallel and in series, and learn how energy, electric potential, and electric charge are...
- Performance of Li-Ion Batteries: Contribution of Electronic Factors to ... — Specifically, we relate the ionization potential obtained by photoelectron spectroscopy to the electron chemical potential at the onset of the charging plateau, and its difference to the ionization potential of lithium to the electronic contribution of the cell voltage.
- Teaching electric circuits with a focus on potential differences — [This paper is part of the Focused Collection on Curriculum Development: Theory into Design.] Developing a solid understanding of simple electric circuits represents a major challenge to most students in middle school. In particular, students tend to reason exclusively with current and resistance when analyzing electric circuits as they view voltage as a property of the electric current and ...
- Brief overview of electrochemical potential in lithium ion batteries — The problems related to electrochemical potential in LIBs are reviewed, including the output voltage of electrodes, the spatial distribution of electrochemical potential in the full cell and its possible effects on the performance. The relevant factors affecting the output voltage are presented.