Fact-checked by Grok 2 weeks ago

Drift current

Drift current is the component of in arising from the directed motion of charge carriers, such as electrons and holes, under the influence of an applied , distinct from which stems from concentration gradients. This phenomenon is fundamental to charge transport in semiconductor devices, where the drift velocity of carriers is proportional to the strength and their , typically yielding current densities expressed as J_{drift} = q (n \mu_n + p \mu_p) E, with q as the , n and p as electron and hole concentrations, \mu_n and \mu_p as their respective mobilities, and E as the . In semiconductors like and , drift current dominates in regions with strong electric fields, such as depletion layers in p-n junctions or Schottky barriers, enabling the operation of diodes, transistors, and photodetectors. Carrier mobility, a key parameter, varies with material and temperature— for instance, electrons in exhibit \mu_n \approx 1400 cm²/V·s at , while holes have \mu_p \approx 450 cm²/V·s— and is limited by scattering mechanisms including and impurities. At high fields exceeding approximately 10⁴ V/cm, drift velocity saturates around 10⁷ cm/s in due to optical , impacting device speed and efficiency in applications like MOSFETs and microwave diodes. The interplay between drift and diffusion currents governs overall carrier transport, with equilibrium occurring when they balance, as in the built-in field of a p-n junction. In optoelectronic devices, such as photodetectors, drift current facilitates rapid carrier collection under reverse , contributing to photocurrents proportional to rates and , while in high-speed transistors, velocity saturation effects must be accounted for to model accurately. Understanding drift current is essential for designing efficient technologies, from integrated circuits to solar cells.

Fundamentals

Definition and Basic Principles

Drift current refers to the arising from the directed movement of charge carriers—such as electrons or holes—within a under the influence of an applied , distinguishing it from the random, thermally driven motion of these carriers. In solid-state s like conductors and semiconductors, free charge carriers respond to the electric field by acquiring a net directional velocity, resulting in a flow of charge that constitutes the current. This phenomenon forms a foundational aspect of charge transport in solids, where carriers exist as mobile electrons in the conduction band of semiconductors or as conduction electrons in metals. The basic principles of drift current stem from the interaction between the and charged particles, imparting a force that biases their otherwise isotropic thermal motion toward a preferred direction. In semiconductors, this leads to a net charge displacement, enabling the material to conduct when doped to increase carrier concentration. The concept gained prominence in the mid-20th century, particularly after the 1947 invention of the by and Walter Brattain at Bell Laboratories, which relied on understanding carrier drift for device operation, followed by William Shockley's theoretical advancements in 1945 on field-effect amplification. The 1949 Haynes-Shockley experiment further validated these ideas by directly measuring carrier drift mobility in , marking a key milestone in physics. Charge carriers in solids include electrons, which are negatively charged and drift opposite to the electric field direction, and holes—effective positive charges arising from missing electrons in the valence band—which drift in the same direction as the field. Mobility, a material-specific parameter, quantifies the ease with which these carriers move under the field, influencing the magnitude of the drift current; higher mobility indicates less resistance to directed motion due to scattering by lattice vibrations or impurities. In intrinsic semiconductors, both electrons and holes contribute equally to drift, while in extrinsic types, the majority carrier dominates. This carrier transport contrasts with diffusion current, which occurs due to concentration gradients rather than fields.

Physical Mechanism

In semiconductors, charge carriers such as electrons and holes experience a force due to an applied , given by \mathbf{F} = [q](/page/Q) \mathbf{E}, where [q](/page/Q) is the carrier charge and \mathbf{E} is the field; this force accelerates the carriers in the direction opposite to the field for electrons and along the field for holes, superimposing a directed motion on their random velocities. However, this acceleration is frequently interrupted by events, where carriers collide with lattice vibrations (phonons), impurity atoms, or other defects, randomizing their and preventing unbounded speed increase. These collisions balance the field's accelerating effect, leading to a steady-state average motion known as drift, where the net displacement occurs despite the absence of a continuous straight-line path. The paths of carriers under an electric field are characteristically zigzag: between collisions, a carrier accelerates linearly along the field direction, gaining momentum proportional to the field strength and the time elapsed since the last scattering; upon collision, its velocity is altered randomly, often losing the directed component while retaining thermal energy. The mean free time \tau, defined as the average interval between such collisions, governs this steady-state motion by determining how long carriers can accelerate before scattering resets their trajectory; longer \tau allows greater average directed velocity, while frequent scattering diminishes the net drift. Scattering mechanisms vary: phonon interactions dominate in pure crystals at room temperature, while impurities become significant in doped materials, collectively setting the relaxation time \tau that underlies the observed drift. Temperature influences this mechanism primarily through its effect on scattering rates, as higher intensifies vibrations, increasing and thus shortening \tau, which reduces the efficiency of drift for a given field in both semiconductors and metals. In semiconductors, this temperature-induced rise in competes with the exponential increase in intrinsic concentration, but the core drift process—field-driven acceleration moderated by collisions—remains more sensitive to effects than in metals, where is largely temperature-independent and leads to similar but often stronger . In intrinsic semiconductors, where and concentrations are equal, both types undergo this same drift mechanism under , contributing to the total with their respective responses to the field. This average directed motion manifests as the drift velocity, a key outcome detailed mathematically elsewhere.

Mathematical Formulation

Drift Velocity

The v_d represents the acquired by charge carriers in a when subjected to an , superimposed on their random thermal motion. This net directed motion arises because the imparts a consistent on the carriers, leading to a small but measurable despite frequent collisions that randomize their . The derivation of the drift velocity begins with Newton's second law applied to a . For a carrier with charge [q](/page/Q), effective mass [m](/page/M), and velocity v, the net under an electric \mathbf{E} includes the driving q\mathbf{E} and a frictional drag -\frac{m\mathbf{v}}{\tau}, where [\tau](/page/Tau) is the relaxation time (average time between collisions). This yields the equation of motion: m \frac{d\mathbf{v}}{dt} = q\mathbf{E} - \frac{m\mathbf{v}}{\tau}. In , the \frac{d\mathbf{v}}{dt} = 0, so the equation simplifies to q\mathbf{E} = \frac{m\mathbf{v}_d}{\tau}, where \mathbf{v}_d is the steady-state . Solving for \mathbf{v}_d gives: \mathbf{v}_d = \frac{q\tau}{m} \mathbf{E}. This linear relationship holds at low electric fields, where the drift velocity is much smaller than the . For electrons, with charge q = -e (where e > 0 is the ), the drift velocity is \mathbf{v}_{d,n} = -\frac{e\tau_n}{m_n^*} \mathbf{E}, directed opposite to the . For holes, with effective positive charge q = +e, it is \mathbf{v}_{d,p} = +\frac{e\tau_p}{m_p^*} \mathbf{E}, aligned with the field. The relaxation times \tau_n and \tau_p differ due to distinct mechanisms for electrons and holes. The proportionality constant in the drift velocity expression is the carrier mobility \mu, defined as \mu = \frac{|q|\tau}{m^*}, so \mathbf{v}_d = \mu \mathbf{E} (with direction depending on carrier type). Mobility quantifies how easily carriers move under a and is influenced by factors such as scattering (via phonons), scattering, and temperature; higher \tau or lower m^* increases \mu. In at (300 K), electron \mu_n is approximately 1400 cm²/V·s, significantly higher than hole mobility \mu_p at about 470 cm²/V·s, reflecting electrons' lower effective mass and reduced .

Drift Current Density

The drift current density represents the macroscopic flow of charge carriers in a material driven by an applied electric field, extending the microscopic concept of drift velocity to bulk transport. The drift velocity describes the average directed motion of carriers, with electrons moving opposite to the field and holes in the same direction, leading to a net current when considering carrier concentrations. For electrons, the drift current density is derived as \vec{J_n} = (-q) n \vec{v_{dn}}, where q is the elementary charge, n is the electron concentration, and \vec{v_{dn}} = -\mu_n \vec{E} is the electron drift velocity, with \mu_n the electron mobility and \vec{E} the electric field; the negative sign in the charge reflects the electron charge, yielding \vec{J_n} = q n \mu_n \vec{E}. Similarly, for holes, \vec{J_p} = q p \vec{v_{dp}} = q p \mu_p \vec{E}, where p is the hole concentration and \mu_p the hole mobility. The total drift current density is the vector sum \vec{J_{\text{drift}}} = \vec{J_n} + \vec{J_p} = q (n \mu_n + p \mu_p) \vec{E}. This formulation links directly to electrical , defined as \sigma = q (n \mu_n + p \mu_p), such that \vec{J_{\text{drift}}} = \sigma \vec{E}, establishing the connection to in materials where drift dominates transport. In metals, high electron concentrations (n \approx 10^{22} \, \mathrm{cm^{-3}}) compensate for lower mobilities (\mu_n \approx 10{-}50 \, \mathrm{cm^2/V \cdot s}), resulting in conductivities on the order of $10^5{-}10^7 \, \mathrm{S/cm}; for example, in , \sigma \approx 5.96 \times 10^5 \, \mathrm{S/cm} arises primarily from electron drift despite modest \mu_n. In contrast, semiconductors exhibit tunable conductivity through doping, with carrier concentrations adjusted to $10^{15}{-}10^{18} \, \mathrm{cm^{-3}} and higher mobilities (\mu_n \approx 1000{-}1500 \, \mathrm{cm^2/V \cdot s}, \mu_p \approx 400{-}500 \, \mathrm{cm^2/V \cdot s}) in materials like , enabling \sigma values from $10^{-6} \, \mathrm{S/cm} (intrinsic) to over $10^2 \, \mathrm{S/cm} (heavily doped). This drift current density equation was pivotal in William Shockley's 1950s semiconductor models, particularly in his book Electrons and Holes in Semiconductors, which integrated it into analyses of carrier transport and early transistor designs.

Comparison with Other Currents

Drift versus Diffusion Current

In semiconductors, charge carrier transport is governed by two distinct mechanisms: drift current, which arises from the motion of charged particles under an applied electric field, and diffusion current, which results from the random thermal motion of carriers tending to equalize concentration gradients. Drift current is directly proportional to the electric field strength and carrier density, leading to a linear current-voltage characteristic in uniformly doped materials, consistent with Ohm's law. In contrast, diffusion current depends on the spatial gradient of carrier concentration, driving net flow from regions of higher to lower density, and exhibits non-linear behavior due to its equilibrium-driven nature. Qualitatively, drift current aligns carrier motion with the electric field direction—electrons opposite to the field and holes along it—while diffusion current flows independently of the field, solely dictated by concentration differences, and can oppose or reinforce drift depending on the setup. Drift typically dominates in biased, uniform doping scenarios where the electric field is significant, such as in bulk conduction regions, whereas diffusion prevails in non-uniform doping or high-injection conditions, like near contacts or interfaces. Material dependencies further differentiate them: drift is influenced by mobility, which varies with mechanisms and , while diffusion relates to the diffusion coefficient via the Einstein , tying it closely to properties. The following table summarizes key distinctions:
AspectDrift CurrentDiffusion Current
Driving Force (E)Concentration (∇n or ∇p)
DirectionAlong E for holes, opposite for electronsFrom high to low concentration
Material Dependency (μ), carrier density (n or p)Diffusion coefficient (D),
Typical DominanceHigh , uniform dopingInjection, non-uniform doping
These mechanisms were first systematically separated and modeled in the through the van Roosbroeck framework, which integrated drift and into a unified description for electrons and holes in semiconductors, enabling predictive analysis of device behavior. Neglecting diffusion in drift-dominated analyses, such as low-gradient ohmic regions, can overestimate , while ignoring drift in diffusion-heavy low-field cases fails to account for the electrostatic balancing that maintains charge neutrality at .

Role in Total Carrier Transport

In semiconductors, the total current density arises from the superposition of drift and diffusion contributions for both electrons and holes. For electrons, it is expressed as \mathbf{J}_n = q \mu_n n \mathbf{E} + q D_n \nabla n, where q is the elementary charge, \mu_n is the electron mobility, n is the electron concentration, \mathbf{E} is the electric field, and D_n is the electron diffusion coefficient; a similar form holds for holes as \mathbf{J}_p = q \mu_p p \mathbf{E} - q D_p \nabla p, with p the hole concentration and \mu_p, D_p the corresponding mobilities and diffusivities. The overall total current density is then \mathbf{J} = \mathbf{J}_n + \mathbf{J}_p. The coefficients are linked to the via the , D_{n,p} = \frac{[kT](/page/KT)}{q} \mu_{n,p}, where k is Boltzmann's and T is the , ensuring thermodynamic consistency in non-degenerate semiconductors. This originates from the balance between drift and under conditions and is fundamental to the drift-diffusion model. The drift-diffusion model describes carrier transport through these current equations coupled with the continuity equations for electrons and holes, \frac{\partial n}{\partial t} = -\frac{1}{q} \nabla \cdot \mathbf{J}_n + G - R and \frac{\partial p}{\partial t} = -\frac{1}{q} \nabla \cdot \mathbf{J}_p + G - R, where G and R are generation and recombination rates, respectively. In steady state, the time derivatives vanish, and assuming quasi-neutrality or solving Poisson's equation for the electric field, the model enforces current continuity via \nabla \cdot \mathbf{J} = 0. This framework governs the flow of both majority and minority carriers, with drift typically dominating majority carrier transport in high-field regions and diffusion driving minority carrier injection across concentration gradients. In extrinsic semiconductors, where doping creates an abundance of one carrier type, drift often dominates under reverse bias conditions, as the sweeps minority carriers generated by thermal processes, while currents remain negligible due to low minority concentrations. Solving the coupled nonlinear drift-diffusion equations typically requires numerical methods, such as finite element approximations, which discretize the spatial domain and iteratively converge to steady-state solutions while preserving physical bounds like positivity of carrier densities.

Applications in Semiconductor Devices

Drift Current in p-n Junctions

In a p-n junction, the abrupt change in doping concentration between the p-type region (high acceptor density N_A) and n-type region (high donor density N_D) leads to carrier diffusion across the interface, uncovering ionized dopants that form a space-charge . This charge separation generates a built-in directed from the n-side to the p-side, with magnitude determined by \frac{dE}{dx} = \frac{\rho}{\epsilon_s}, where \rho is the net charge density from dopants. The field strength peaks at the metallurgical junction and creates a built-in potential barrier \phi_{bi} = \frac{kT}{q} \ln\left(\frac{N_A N_D}{n_i^2}\right), typically around 0.7 V for at , which halts net carrier flow at equilibrium. At zero , the built-in drives majority carriers (holes in the p-region and electrons in the n-region) in opposite directions to the flux, resulting in equal and opposite drift and currents that sum to zero net current across the junction. This balance maintains , with the drift component arising from the field-induced v_d = \mu E, where \mu is the carrier mobility. In reverse bias, the applied voltage augments the built-in field, expanding the depletion width to W \approx \sqrt{\frac{2\epsilon_s (V_{bi} + V_R)}{q} \left(\frac{1}{N_A} + \frac{1}{N_D}\right)} and strongly enhancing minority carrier drift toward the contacts once they enter the via thermal generation or . The resulting reverse I_s \approx q A \left( \frac{D_p p_{n}}{L_p} + \frac{D_n n_{p}}{L_n} \right), where p_n = n_i^2 / N_D and n_p = n_i^2 / N_A are equilibrium minority concentrations, D denotes coefficients, and L denotes lengths, is dominated by across the high-field depletion zone, though limited by the supply of minorities diffusing to its edge. For p-n diodes, this drift-contributed leakage is typically on the order of $10^{-12} A/cm² at , independent of further increases in reverse voltage until breakdown. Under forward bias, the external voltage reduces the effective field in the depletion region, narrowing its width and lowering the barrier to minority carrier injection, allowing diffusion currents to surge exponentially as I \propto e^{qV / kT}. Drift remains essential for rapidly sweeping these injected carriers across the thinned depletion zone to maintain current continuity, but its relative contribution diminishes compared to diffusion. The overall I-V curve thus exhibits exponential forward conduction primarily from diffusion, transitioning to linear saturation in reverse from the field-driven drift of minorities.

Drift Current in Field-Effect Transistors

In field-effect transistors (FETs), particularly metal-oxide-semiconductor field-effect transistors (MOSFETs), drift current serves as the dominant mechanism for charge carrier transport within the channel. Upon applying a gate-to-source voltage exceeding the threshold voltage, an inversion layer of minority carriers forms at the semiconductor-oxide interface, enabling conduction. A drain-to-source voltage then induces a longitudinal electric field along the channel, causing carriers to drift and produce the drain current, expressed as the sheet current density J_{\text{drift}} = q \mu_n n_s E_{\text{channel}}, where q is the elementary charge, \mu_n the electron mobility, n_s the inversion layer sheet carrier density, and E_{\text{channel}} the channel electric field. The gate voltage modulates n_s, typically approximated as n_s \approx C_{\text{ox}} (V_{\text{GS}} - V_{\text{T}}) / q (with C_{\text{ox}} the oxide capacitance per unit area and V_{\text{T}} the ), thereby controlling the drift current magnitude for . In the linear regime at low V_{\text{DS}}, the is uniform, yielding I_{\text{D}} \propto V_{\text{DS}} behavior akin to a . At higher V_{\text{DS}}, the device transitions to , where pinch-off occurs near the , limiting further current increase. In short-channel MOSFETs, velocity caps the drift velocity at approximately $10^7 cm/s in due to optical , modifying the saturation current to I_{\text{D}} = W C_{\text{ox}} v_{\text{sat}} (V_{\text{GS}} - V_{\text{T}}) (with W the width). High lateral fields in the lead to hot effects, where accelerated gain sufficient energy to cause and interface state generation, degrading through enhanced scattering and reducing long-term device performance. Unlike bipolar junction transistors, where dominates minority transport across the narrow base, FET conduction relies primarily on field-driven , facilitating voltage control and high . This drift-dominated transport enables efficient high-speed switching in complementary () circuits, supporting GHz-frequency operation in processors and RF applications by minimizing delay through rapid carrier acceleration under controlled fields.

References

  1. [1]
    [PDF] Motion and Recombination of Electrons and Holes
    The total drift current density is the sum of the electron and the hole components: ... In addition to the drift current, there is a second component of current ...
  2. [2]
    [PDF] Physics of Semiconductor Devices
    ... Sze. Department of Electronics Engineering. National Chiao Tung University ... drift current exactly balances the diffusion current, qn,un8 = - qDn--. dn.
  3. [3]
    [PDF] Lecture 3 - MIT
    What did we learn today? • Electrons and holes in semiconductors are mobile and charged. – ⇒ Carriers of electrical current! • Drift current: produced by ...
  4. [4]
    [PDF] ME 432 Fundamentals of Modern Photovoltaics
    current in a semiconductor. 1. Drift Current : flow of charge in response to an applied electric field. 2. Diffusion Current : flow of charge in response to a ...
  5. [5]
    A revision of the semiconductor theory from history to applications
    Jun 8, 2024 · In a semiconductor, the total drift current density corresponds to the sum of Eqs. 34a and 34b. According to this result and knowing Ohm's ...
  6. [6]
    [PDF] William Shockley - Nobel Lecture
    In order to test the model of carrier injection, J. R. Haynes and the author collaborated in the drift-mobility experiment or « Haynes' experiment » on.
  7. [7]
  8. [8]
    Drift | PVEducation
    The net acceleration in the case of steady state current flow is balanced by the decelerations of the collision processes. If N(t) is the number of ...
  9. [9]
    [PDF] Lecture 3 Electron and Hole Transport in Semiconductors Review
    Since the electrons move about randomly in all directions (Brownian motion), as time goes on more electrons will move from regions of higher electron.
  10. [10]
    [PDF] Lecture 7 Drift and Diffusion Currents Reading: Pierret 3.1-3.2
    Thus, the drift velocity increases with increasing applied electric field. Where <τ> is the average time between “particle” collisions in the semiconductor.
  11. [11]
    Drift Current Density - an overview | ScienceDirect Topics
    Drift current density is defined as the sum of the electron and hole drift current densities in a semiconductor, expressed in terms of the electric field and ...
  12. [12]
    Current Density in Metal and Semiconductor | Electrical4U
    May 24, 2024 · Current density, denoted by J, is given by J = I/A, where 'I' is the current and 'A' is the cross-sectional area. If N electrons pass through a ...
  13. [13]
    [PDF] Drift and Diffusion Currents
    Drift and diffusion currents are how charge carriers move in an electric field and concentration gradient. Drift current is related to Ohm's Law.
  14. [14]
    [PDF] 6.012 Microelectronic Devices and Circuits, Lecture 3
    • Electrons and holes in semiconductors are mobile and charged. – ⇒ Carriers of electrical current! • Drift current: produced by electric field drift ∝ E drift.
  15. [15]
    [PDF] Lecture 4 Semiconductor Physics Review - Index of /
    ➢ Drift current is in the same direction as the electric field. ➢ Diffusion ... ➢ Definition of inversion. ➢ Point at which density of electrons on ...
  16. [16]
    Theory of the Flow of Electrons and Holes in Germanium and Other ...
    Jul 29, 2013 · Authors. W. Van Roosbroeck. First published: October 1950 Full ... drift, diffusion, and recombination. This formulation is specialized ...
  17. [17]
    The Transport of Added Current Carriers in a Homogeneous ...
    This equation is derived in a form which exhibits the ambipolar nature of the diffusion, drift, and recombination mechanisms under electrical neutrality.
  18. [18]
    Basic Equations | PVEducation
    I = J × Area and for a 1 cm2 device J and I are equal. The first term in each equation is for drift and the second term is for diffusion. Continuity Equations.
  19. [19]
    [PDF] ECE 440 Lecture 33 : Test 2 Review
    Apr 9, 2010 · In reverse bias, the current is due mostly to the drift of minority carriers. •Carriers arise mainly from thermal generation near the.
  20. [20]
    Bias of PN Junctions - PVEducation
    Reverse Bias​​ As in forward bias, the drift current is limited by the number of minority carriers on either side of the p-n junction and is relatively unchanged ...
  21. [21]
    Analysis of a finite element method for the drift-diffusion ...
    Summary. An explicit finite element method for numerically solving the drift-diffusion semiconductor device equations in two space dimensions is analyzed.
  22. [22]
    [PDF] PN and Metal–Semiconductor Junctions
    If the current is limited to a reasonable value by the external circuit so that heat dissipation in the PN junction is not excessive, the PN junction can be.
  23. [23]
    None
    ### Summary of Drift Current in p-n Junctions
  24. [24]
    [PDF] MOSFET Drain Current Modeling - MIT OpenCourseWare
    current, iD, as the product of q. N. * (y) , the inversion layer sheet ... drift velocity of the inversion layer carriers there (electrons in the n ...Missing: mu n_s E
  25. [25]
    [PDF] A Review of MOSFET Fundamentals - nanoHUB
    velocity cm/s --->. 107. 104 υ = μE υ =υ sat. V. DS. L. = 1.0V. 100nm. ≈1×105. V/cm. 105. Page 13. 13. MOSFET IV: velocity saturation ... MOSFET theory is based ...Missing: regimes | Show results with:regimes
  26. [26]
    The correct account of nonzero differential conductance in the ...
    Jan 29, 2014 · ... saturation velocity in silicon, which is approximately 107 cm/s. Characteristic electric field for velocity saturation effect is. urn:x-wiley ...
  27. [27]
    Hot Carrier Injection - an overview | ScienceDirect Topics
    The typical effect of hot-carrier, or commonly hot-electron, degradation is to reduce the on-state current in an n-MOSFET and increase the off-state current ...
  28. [28]
    physics of semiconductor devices
    we have related the flow of current to drift and diffusion mechanisms. The ... In an "ideal" bipolar transistor the base current should be much smaller.<|separator|>
  29. [29]
    Gigahertz and terahertz transistors for 5G, 6G, and beyond mobile ...
    Aug 19, 2024 · The RF input signal of a GHz–THz transistor causes a change of the charge distribution in the device and, to achieve a gain >1 at high ...<|control11|><|separator|>