Fact-checked by Grok 2 weeks ago

Threshold voltage

Threshold voltage, commonly denoted as V_{th} or V_T, is the minimum gate-to-source voltage (V_{GS}) required in a metal-oxide-semiconductor field-effect transistor (MOSFET) to induce strong inversion at the semiconductor surface, forming a conductive channel between the source and drain terminals and allowing significant drain current to flow. This parameter marks the boundary between the subthreshold region, where the transistor conducts weakly via diffusion, and the strong inversion region, where channel conduction dominates via drift, making it essential for MOSFET operation in switching and amplification applications. In n-channel MOSFETs, V_{th} is typically positive (around 0.2–1 V for modern devices), while for p-channel devices it is negative, reflecting the polarity needed to attract opposite charge carriers to the channel. The threshold voltage arises from the physics of the MOS capacitor structure underlying the , where the gate voltage must overcome the flat-band voltage (V_{FB}), twice the Fermi potential ($2\phi_B), and the voltage drop across the depletion layer to achieve inversion. The classical expression for V_{th} in an n-channel is given by
V_{th} = V_{FB} + 2\phi_B + \frac{\sqrt{4\epsilon_s q N_A \phi_B}}{C_{ox}},
where V_{FB} accounts for differences and charges, \phi_B is the potential (approximately (kT/q) \ln(N_A / n_i)), \epsilon_s is the permittivity, q is the charge, N_A is the substrate doping concentration, and C_{ox} = \epsilon_{ox} / t_{ox} is the capacitance per unit area with t_{ox} as thickness. Key factors influencing V_{th} include substrate doping (higher N_A increases V_{th}), thickness (thinner oxides reduce V_{th}), ( V_{th} decreases with rising due to \phi_B variation), and short-channel effects in scaled devices, which can lower V_{th} through charge sharing and drain-induced barrier lowering (DIBL).
Measurement of V_{th} is critical for device and process control, with common methods including constant-current extraction (where V_{GS} is defined at a specific , e.g., 100 μA × W/L), linear from the I_D-V_{GS} in the linear , and second-derivative techniques for precision in subthreshold analysis. Variations in V_{th} due to body bias, known as the body effect, further modulate the parameter as V_{th} = V_{th0} + \gamma (\sqrt{2\phi_B + V_{SB}} - \sqrt{2\phi_B}), where \gamma is the body-effect coefficient and V_{SB} is the source-to-body voltage, impacting circuit performance in applications like logic and .

Fundamentals

Definition and significance

The threshold voltage, denoted as V_{th}, in a () is defined as the minimum gate-to-source voltage (V_{GS}) required to induce a conductive between the source and drain terminals, thereby transitioning the device from the (subthreshold) region to strong inversion where significant drain current flows. This parameter marks the onset of formation at the semiconductor-oxide , enabling the to act as a switch or . The significance of threshold voltage lies in its direct influence on key performance metrics of MOSFET-based integrated circuits, including switching speed, power consumption, and thresholds. It determines the overdrive voltage (V_{GS} - V_{th}), which governs the drive current and thus the transistor's on-state conductance; lower V_{th} enhances speed and efficiency but increases leakage current in the off-state, necessitating careful optimization for low-power designs. In digital circuits, V_{th} affects margins and , with higher values improving immunity to at the cost of slower operation. Historically, the threshold voltage concept emerged with the invention of the by and at Bell Laboratories, who demonstrated the first working silicon-based device in 1959 using a thermally grown gate insulator to overcome surface state issues, with the seminal report published in 1960. This breakthrough enabled scalable fabrication, laying the foundation for modern . MOSFETs exhibit varying V_{th} polarities: positive for enhancement-mode n-channel (NMOS) devices, which require V_{GS} > V_{th} to form a channel, and negative for enhancement-mode p-channel (PMOS) devices; depletion-mode variants, conversely, possess an inherent channel at zero gate bias, with V_{th} negative for NMOS and positive for PMOS to deplete it. In contemporary semiconductor scaling, threshold voltage remains pivotal for enabling low-power operation in advanced nodes, yet it presents challenges below 10 , where maintaining precise control becomes difficult due to quantum effects and variability, potentially increasing off-state leakage and complicating gate stack materials for tunable work functions.

Mathematical formulation

The threshold voltage V_{th} of a long-channel metal-oxide- (MOSFET) represents the gate-source voltage at which strong inversion occurs at the surface, enabling significant channel conduction. This baseline model is derived from the MOS capacitor structure by solving in the under the depletion approximation, assuming no mobile carriers in the and a uniform across the oxide. The derivation considers the of the and the p-type for an n-channel MOSFET (NMOS), where the gate voltage must counteract the difference, achieve the required surface potential for inversion, and support the depletion charge. The flat-band voltage V_{FB} is the first component, defined as the gate voltage needed to produce zero band bending at the semiconductor surface in the absence of charges. It is expressed as V_{FB} = \phi_{ms} - \frac{Q_{ox}}{C_{ox}}, where \phi_{ms} is the work function difference between the gate material and the semiconductor, Q_{ox} is the fixed oxide charge per unit area (typically positive and located near the silicon interface), and C_{ox} = \frac{\epsilon_{ox}}{t_{ox}} is the oxide capacitance per unit area with \epsilon_{ox} as the oxide permittivity and t_{ox} as the oxide thickness. This term accounts for inherent material asymmetries and any trapped charges in the oxide. The second component is the surface potential at the onset of strong inversion, which is $2\phi_f, where \phi_f is the Fermi potential given by \phi_f = \frac{kT}{q} \ln \left( \frac{N_A}{n_i} \right) for a p-type substrate. Here, k is Boltzmann's constant, T is the absolute temperature, q is the elementary charge, N_A is the uniform acceptor doping concentration, and n_i is the intrinsic carrier concentration of . The factor of 2 arises because strong inversion occurs when the minority carrier () concentration at the surface equals the majority carrier () concentration in the bulk, requiring the surface potential to be twice the Fermi potential relative to the bulk. The third component compensates for the voltage drop across the due to the . The depletion charge density is Q_d = -\sqrt{2 [q](/page/Q) \epsilon_s N_A (2 \phi_f)}, obtained by integrating \frac{d^2 \psi}{dx^2} = \frac{[q](/page/Q) N_A}{\epsilon_s} (with \psi as the electrostatic potential and \epsilon_s as the ) under the from the surface to the depletion edge. The corresponding voltage contribution is \frac{|Q_d|}{C_{ox}} = \frac{\sqrt{2 [q](/page/Q) \epsilon_s N_A (2 \phi_f)}}{C_{ox}}, which can also be written using the body-effect coefficient \gamma = \frac{\sqrt{2 [q](/page/Q) \epsilon_s N_A}}{C_{ox}} as \gamma \sqrt{2 \phi_f}. This term increases with higher substrate doping or thinner , reflecting greater depletion charge that the gate must induce. Combining these contributions yields the full threshold voltage equation for a long-channel NMOSFET under zero : V_{th} = V_{FB} + 2 \phi_f + \frac{\sqrt{2 q \epsilon_s N_A (2 \phi_f)}}{C_{ox}} This expression assumes an ideal long-channel device with uniform doping, no short-channel effects, negligible traps, and operation at where quantum-mechanical effects on the threshold are minor. The model relies on the gradual-channel approximation, where the vertical field dominates over the lateral field along the . For a p-channel MOSFET (PMOS) with an n-type (donor concentration N_D), the counterpart has opposite due to the inversion of carrier types and substrate doping: V_{th} is negative, \phi_f = -\frac{kT}{q} \ln \left( \frac{N_D}{n_i} \right) (or taken as positive with sign adjustments), and the depletion term changes sign to -\frac{\sqrt{2 q \epsilon_s N_D (-2 \phi_f)}}{C_{ox}}, resulting in V_{th} = V_{FB} + 2 \phi_f - \gamma \sqrt{-2 \phi_f}. This ensures the gate voltage must be sufficiently negative to attract holes to the surface.

Physical dependencies

Oxide thickness dependence

The threshold voltage V_{th} in MOSFETs depends on the gate oxide capacitance per unit area C_{ox}, defined as C_{ox} = \epsilon_{ox}/t_{ox}, where \epsilon_{ox} is the oxide permittivity and t_{ox} is the physical oxide thickness. The depletion charge contribution to V_{th} is inversely proportional to C_{ox}, so decreasing t_{ox} enhances the gate-to-channel capacitive coupling, thereby reducing V_{th}. This relationship arises because a thinner oxide allows stronger electrostatic control over the channel potential, shifting the onset of strong inversion to lower gate biases. Historical scaling of t_{ox} has directly lowered V_{th} to enable higher in integrated circuits, but it has also introduced trade-offs. In the , t_{ox} typically ranged from 50–100 nm for gate lengths around 7.5 µm, but aggressive reduced it to 1.1–1.6 nm by the 45 nm node in the mid-2000s, with modern sub-5 nm nodes achieving equivalent thicknesses below 1 nm. While this trend decreases V_{th} and boosts drive current, it exacerbates direct tunneling leakage through the , limiting further reductions without alternative materials. For t_{ox} < 2 nm, quantum mechanical effects counteract the classical reduction in V_{th}. Carrier quantization in the inversion layer causes the electron charge centroid to shift away from the Si/SiO₂ interface and induces effective bandgap widening, increasing the effective V_{th} by 0.1–0.2 V depending on the surface electric field and doping. These effects, prominent in ultrathin oxides, necessitate quantum-corrected models for accurate device simulation. To sustain thin effective thicknesses without excessive leakage, high-κ dielectrics like were adopted at the 45 nm node in 2007, replacing . , with a dielectric constant of ~25, enables physical thicknesses of several nanometers while maintaining an equivalent oxide thickness (EOT) comparable to 1 nm , thereby stabilizing V_{th} and minimizing tunneling current. This transition preserved the capacitive benefits of scaling for V_{th} control. Contemporary V_{th} predictions incorporate EOT in the C_{ox} term, where EOT = t_{ox} \cdot (\epsilon_{SiO_2}/\epsilon_{high-κ}) + t_{IL} (with t_{IL} as any interfacial layer thickness), ensuring the formulation accounts for hybrid stacks and accurately reflects the effective capacitive coupling.

Substrate bias (body effect)

In MOSFETs, the substrate bias effect, commonly known as the body effect, arises when a reverse bias voltage (V_{BS} < 0 for NMOS devices) is applied between the source and the substrate, modifying the threshold voltage V_{TH}. This bias widens the depletion region beneath the channel, increasing the concentration of fixed charges that the gate must overcome to induce strong inversion at the silicon surface. As a result, a higher gate-to-source voltage V_{GS} is required to form the inversion layer, thereby elevating V_{TH}. The modified threshold voltage under substrate bias is given by: V_{TH} = V_{TH0} + \gamma \left( \sqrt{|2\phi_F - V_{BS}|} - \sqrt{|2\phi_F|} \right) where V_{TH0} is the zero-bias threshold voltage, \gamma is the body effect coefficient, and \phi_F is the Fermi potential. This equation accounts for the bias-induced shift, with the square root terms reflecting the change in surface potential and depletion charge. The derivation stems from the increased depletion charge density under reverse bias. The bulk charge per unit area Q_B in the depletion region is Q_B = -\sqrt{2 \epsilon_{Si} q N_A |2\phi_F - V_{BS}|}, where \epsilon_{Si} is the permittivity of silicon, q is the elementary charge, and N_A is the substrate acceptor doping concentration. This charge contributes to the gate voltage as -Q_B / C_{ox}, where C_{ox} is the oxide capacitance per unit area, leading to the \gamma term defined as \gamma = \sqrt{2 q \epsilon_{Si} N_A} / C_{ox}. The body effect thus directly ties to the enhanced electrostatic control needed to counter the widened depletion width. In bulk CMOS technology, the body effect coefficient \gamma typically ranges from 0.3 to 0.5 V^{1/2}, depending on oxide thickness and doping levels, allowing predictable V_{TH} adjustments for circuit design. This effect is leveraged in stacked transistor configurations, where reverse biasing the body of off-transistors in a series stack raises their V_{TH}, significantly suppressing subthreshold leakage currents—often by factors of 10 or more per stacked device—while maintaining performance in active paths. In silicon-on-insulator (SOI) devices, the body effect is altered due to the isolated body region, which reduces the magnitude of \gamma compared to bulk MOSFETs because of the thinner silicon film and limited depletion charge modulation; however, floating body dynamics can introduce additional variability in effective bias response.

Temperature dependence

The temperature dependence of the threshold voltage in MOSFETs arises primarily from the variation of the Fermi potential, \phi_f, with temperature. The Fermi potential is expressed as \phi_f = \frac{kT}{q} \ln\left(\frac{N_a}{n_i}\right), where N_a is the acceptor doping concentration, n_i is the intrinsic carrier concentration, k is Boltzmann's constant, T is the absolute temperature, and q is the elementary charge. The intrinsic carrier concentration n_i increases strongly with temperature, following n_i \propto T^{3/2} \exp\left(-\frac{E_g}{2kT}\right), where E_g is the bandgap energy. This exponential growth in n_i dominates, causing \phi_f to decrease with increasing T, which in turn reduces the threshold voltage V_{th} by approximately 1-2 mV/K in typical silicon devices. A comprehensive model for the temperature coefficient incorporates both the Fermi potential shift and the depletion charge term: \frac{dV_{th}}{dT} \approx -\frac{d(2\phi_f)}{dT} - \frac{1}{2} \frac{\sqrt{2\epsilon_s q N_a (2\phi_f)}}{2\phi_f} \cdot \frac{d(2\phi_f)}{dT}, where \epsilon_s is the semiconductor permittivity. The first term reflects the direct decrease in surface potential, while the second accounts for the temperature sensitivity of the space-charge contribution. For silicon MOSFETs, this yields a negative coefficient dV_{th}/dT typically ranging from -1.5 to -4 mV/°C, with the exact value depending on doping and oxide thickness. This inherent negative dependence means V_{th} decreases as temperature rises, shifting device characteristics across operating conditions. Operationally, the reduced V_{th} at elevated temperatures increases subthreshold leakage current exponentially, posing challenges for power efficiency in integrated circuits, particularly in standby modes. Conversely, the on-state drive current is diminished because carrier mobility \mu degrades with temperature, following \mu \propto T^{-1.5} due to intensified phonon scattering, which partially offsets the leakage increase but limits performance at high T. To mitigate these effects and maintain stability over automotive and industrial ranges (-40°C to 125°C), compensation techniques include dual V_{th} processes, which assign low-V_{th} transistors to speed-critical paths and high-V_{th} ones to leakage-sensitive areas, and adaptive body biasing, which dynamically adjusts substrate voltage to tune V_{th} against thermal drifts. In advanced technology nodes, FinFET architectures exhibit weaker temperature dependence of V_{th} compared to planar MOSFETs, attributed to enhanced gate electrostatics that better confine carriers and reduce thermal perturbations in the channel potential. This results in more stable V_{th} variation, typically under 2 mV/°C, aiding reliability in scaled devices for high-performance computing.

Doping concentration effects

The threshold voltage V_{th} of a MOSFET is significantly influenced by the substrate doping concentration N_a, which primarily affects the depletion charge term in the device. In the standard long-channel model, the contribution from the bulk charge is given by \frac{\sqrt{2 \epsilon_s q N_a (2 \phi_f)}}{C_{ox}}, where \epsilon_s is the permittivity of silicon, q is the elementary charge, \phi_f is the Fermi potential, and C_{ox} is the oxide capacitance per unit area. Higher N_a increases this term, thereby raising V_{th} because a greater gate voltage is required to overcome the increased positive space charge in the p-type substrate for n-channel devices. Typical substrate doping levels in modern CMOS processes range from $10^{16} to $10^{18} cm^{-3} , balancing short-channel control with mobility degradation. Channel doping, often implemented through halo or pocket implants, further modulates V_{th} to counteract short-channel effects while introducing specific trade-offs. Halo implants, typically using boron for nMOS, create non-uniform doping profiles with higher concentrations near the source and drain junctions, effectively increasing the average channel doping and elevating V_{th} by enhancing the depletion barrier. However, these implants can lead to increased junction leakage currents due to band-to-band tunneling at the higher doped regions. In advanced nodes, such as 90 nm and below, halo doping doses are optimized around $10^{13} cm^{-2} to maintain V_{th} stability without excessive parasitic effects. The gate work function difference \phi_{ms} between the gate material and the semiconductor also plays a key role in setting V_{th}, as it directly shifts the flat-band voltage V_{fb}. Traditional n^+ polysilicon gates have a work function of approximately 4.05 eV, while metal gates in high-k/metal-gate stacks can be tuned from 4.1 to 5.2 eV, resulting in V_{th} shifts of 0.5 to 1 V compared to polysilicon. For CMOS integration, dual work function metals—such as for nMOS (low work function) and for pMOS (high work function)—enable balanced V_{th} values around ±0.3 to 0.4 V without heavy channel doping. This approach reduces reliance on implants and improves performance in sub-45 nm technologies. Threshold voltage adjust implants provide post-fabrication fine-tuning of V_{th}. For nMOS, boron ions are implanted into the channel to increase p-type doping and raise V_{th}, with doses typically in the $10^{12} to $10^{13} cm^{-2} range to achieve adjustments of 50-200 mV. In pMOS, BF_2 implants are preferred over pure boron to minimize penetration through the gate oxide while similarly shifting V_{th} negatively, avoiding excessive fluorine-related degradation. These implants are performed after gate definition but before spacer formation, ensuring precise control in dual-threshold CMOS processes. Doping concentration choices involve inherent trade-offs in device performance and variability. Higher doping levels reduce threshold voltage variation by minimizing depletion width fluctuations but exacerbate short-channel effects like drain-induced barrier lowering due to increased lateral electric fields. Conversely, lower doping supports low-power operation by enabling near-ideal subthreshold slopes and higher carrier mobility but amplifies random dopant fluctuation impacts, leading to greater device-to-device V_{th} spread in scaled technologies. Optimal profiles, often combining uniform substrate and localized channel doping, are determined through process simulation to balance these factors for specific applications like low-standby-power logic.

Variations and fluctuations

Random dopant fluctuation

Random dopant fluctuation (RDF) arises from the statistical variation in the number and position of dopant atoms within the channel depletion region of MOSFETs, leading to device-to-device mismatches in threshold voltage (Vth). This phenomenon is governed by Poisson statistics, where the variance in the number of dopants scales with the mean number, resulting in fluctuations that become increasingly significant as device dimensions shrink. The standard deviation of Vth due to RDF can be approximated as \sigma_{V_{th}} \approx \frac{q t_{ox}}{\epsilon_{ox}} \sqrt{\frac{N_A W_{dep}}{3 W L}}, where q is the elementary charge, t_{ox} is the oxide thickness, \epsilon_{ox} is the oxide permittivity, N_A is the channel doping concentration, W_{dep} is the depletion width, W is the channel width, and L is the channel length. This formula highlights how higher doping and smaller channel areas amplify the variability, as fewer dopant atoms lead to larger relative fluctuations in the electrostatic potential barrier. Pelgrom's law empirically describes the scaling of these mismatches, stating that \sigma_{V_{th}} \propto 1/\sqrt{W L}, indicating that threshold voltage matching improves with larger transistor areas due to averaging over more dopant atoms. This relationship, derived from measurements across various CMOS technologies, underscores RDF as a key contributor to mismatch in analog circuits and static random-access memory (SRAM) cells, where precise Vth control is essential for functionality and yield. RDF becomes the dominant source of Vth variability below 65 nm technology nodes, where the number of dopant atoms in the channel drops to tens or fewer, exacerbating statistical noise. In 28 nm nodes, typical \sigma_{V_{th}} values range from 20-50 mV for minimum-sized transistors, significantly limiting manufacturing yield and necessitating design margins in high-density circuits like SRAM arrays. To mitigate RDF, strategies focus on reducing channel doping concentration, such as employing lightly doped or undoped channels in conjunction with high-k/metal gate stacks to maintain Vth control via work function tuning. Advanced structures like undoped further suppress RDF by minimizing the depletion region's dopant dependence, achieving \sigma_{V_{th}} below 10 mV in sub-20 nm nodes and improving overall variability. Monte Carlo simulations are widely used to model RDF by randomly placing dopant atoms according to Poisson distributions and solving the 3D Poisson equation to compute Vth distributions, enabling statistical analysis of variability without exhaustive fabrication testing. These methods reveal non-Gaussian tails in Vth histograms and correlations with other parameters, guiding process optimization.

Short-channel effects

In short-channel MOSFETs, where the channel length L becomes comparable to the depletion region widths of the source and drain junctions (typically L < 100 nm), electrostatic control by the gate diminishes, leading to deterministic degradations in threshold voltage V_{th}. These short-channel effects (SCEs) primarily manifest as V_{th} roll-off and drain-induced barrier lowering (DIBL), which reduce the gate's ability to modulate the channel potential uniformly and increase off-state leakage. Charge sharing between the source/drain depletion regions and the channel depletion region under the gate reduces the effective depletion charge, thereby lowering V_{th}. This roll-off becomes prominent as L scales down, with an approximate shift given by \Delta V_{th} \approx -\left( \frac{\epsilon_{si} t_{ox}}{\epsilon_{ox} L} \right) V_{ds}, where \epsilon_{si} and \epsilon_{ox} are the permittivities of silicon and oxide, t_{ox} is the oxide thickness, and V_{ds} is the drain-to-source voltage; this expression captures the two-dimensional field penetration from the drain. DIBL further exacerbates V_{th} reduction by allowing the high drain potential to lower the source-channel potential barrier, facilitating easier carrier injection into the channel even at low gate voltages. The DIBL coefficient \eta, defined as \eta = \Delta V_{th} / \Delta V_{ds}, quantifies this effect and typically ranges from 50-100 mV/V in bulk silicon MOSFETs, indicating significant barrier lowering with increasing V_{ds}. Multi-gate architectures, such as FinFETs, suppress DIBL to below 50 mV/V by enhancing gate control through multiple sides of the channel. These SCEs became evident in commercial processes around the 0.25 \mum technology node in the late 1990s, where V_{th} roll-off exceeded 100 mV and DIBL contributed to subthreshold leakage currents rising by over an order of magnitude compared to longer-channel devices. The introduction of in the early 2000s provided superior three-dimensional electrostatic control, reducing SCEs by factors of 2-5 relative to planar bulk at equivalent nodes, enabling scaling to 22 nm. structures, emerging in the 2010s, further mitigate SCEs by fully encircling the channel, achieving \eta < 30 mV/V in sub-10 nm devices. By 2025, are in production at 3 nm and 2 nm nodes, achieving \eta < 20 mV/V through enhanced gate control and undoped channels. The onset of SCEs is governed by the natural length \lambda = \sqrt{\frac{\epsilon_{si} t_{ox} W_{dep}}{\epsilon_{ox}}}, where W_{dep} is the channel depletion width; significant degradations occur when L \lesssim 5\lambda, as this allows substantial sub-surface leakage paths. To counteract SCEs, halo (pocket) doping introduces higher dopant concentrations near the source/drain junctions to steepen the potential barrier and restore gate dominance, while raised source/drain structures shallow the junctions to minimize charge sharing—though both techniques trade off with increased random dopant fluctuation variability.

Measurement and modeling

Extraction methods

The extraction of threshold voltage (Vth) in MOSFETs is typically performed from experimental current-voltage (I-V) characteristics, as the precise definition of Vth can vary depending on the operational regime and device scaling. Common methods focus on identifying the gate-source voltage (Vgs) at which the device transitions from weak to strong inversion, often using drain current (Id) versus Vgs curves measured in either the linear or saturation regions. These techniques address ambiguities arising from non-ideal behaviors, such as mobility degradation and subthreshold leakage, but require careful selection to ensure consistency across devices and conditions. One widely adopted approach is the constant current method, which defines Vth as the Vgs value where Id reaches a predefined criterion, typically Id = 0.1 μA × (W/L), with W and L being the channel width and length, respectively. This method is straightforward and applicable to both linear and saturation regions, making it suitable for automated extraction in circuit simulations and production testing. However, it is sensitive to carrier mobility variations, as changes in mobility can shift the Id level without altering the underlying inversion threshold, leading to inaccuracies in devices with non-uniform doping or interface scattering. An adjusted variant normalizes the current criterion to account for substrate bias, improving reliability for short-channel devices. Another established technique is the linear extrapolation method, which relies on the square-law model of MOSFET operation in saturation. Here, the square root of Id (√Id) is plotted against Vgs, and Vth is determined by extrapolating the linear portion of this curve to the Vgs-axis intercept. This approach assumes ideal strong inversion behavior where Id ∝ (Vgs - Vth)^2, providing a physically grounded estimate that mitigates some subthreshold contributions. It is particularly effective for long-channel devices but can be distorted in nanoscale MOSFETs due to velocity saturation and series resistance, necessitating corrections for accurate results. The transconductance change method offers a derivative-based alternative, identifying Vth at the gate voltage where the transconductance (gm = dId/dVgs) reaches its maximum or where the second derivative d²Id/dVgs² equals zero. This technique captures the onset of strong inversion by highlighting the inflection point in the Id-Vgs curve, often using low drain voltage to emphasize channel conductance changes. It is less dependent on absolute current levels than the constant current method and provides insights into interface quality, though it requires high-resolution measurements to resolve the peak accurately, especially in noisy data. Variants based on the transconductance-to-current ratio (gm/Id) further refine extraction by minimizing drain voltage dependencies, making it robust for analog design applications. Challenges in Vth extraction arise particularly in the subthreshold regime, where the current follows an exponential form Id = I0 exp((Vgs - Vth)/(n kT/q)), with n as the subthreshold swing factor, k as , T as temperature, and q as the elementary charge; Vth is thus defined at a specific low Id level, but the choice of this level introduces ambiguity. In nanoscale devices, short-channel effects (SCE) exacerbate this by blurring the weak-to-strong inversion transition through drain-induced barrier lowering, leading to extracted Vth values that vary by up to 100-200 mV across methods and requiring bias-dependent adjustments. For advanced characterization, capacitance-based methods like provide an alternative by measuring gate-to-channel and channel-to-substrate capacitances separately as functions of Vgs, from which Vth is inferred at the point of minimum capacitance or flat-band shift. This technique is valuable for isolating oxide and interface contributions without relying on current transport, offering high precision in ultra-thin body devices, though it demands specialized probing to decouple parasitic effects. Additionally, cryogenic measurements enable temperature-dependent studies, revealing Vth shifts due to freeze-out and trap effects down to 4.2 K, with extraction often combining I-V and C-V data to validate models for quantum computing applications.

Modeling in devices

In compact models for circuit simulation, the threshold voltage V_{th} serves as a fundamental parameter that governs the transition between weak and strong inversion regimes in MOSFETs. These models integrate V_{th} dependencies on physical factors such as temperature, substrate bias, and channel dimensions to predict device behavior accurately across operating conditions. Widely adopted industry-standard models like BSIM and EKV exemplify this approach, enabling efficient simulation in tools such as SPICE for analog, digital, and mixed-signal circuits. The BSIM family, developed by the Berkeley Short-channel IGFET Model group, treats V_{th} as a core parameter with explicit dependencies. In BSIM4, a threshold-voltage-based model for bulk MOSFETs, the long-channel V_{th} at nominal conditions, denoted V_{TH0}, is adjusted for temperature via V_{th}(T) = V_{TH0} + KT1 \left( \frac{T}{T_{NOM}} - 1 \right) + KT2 (V_{bseff} + 2 V_{TH0}) \left( \frac{T}{T_{NOM}} - 1 \right), where KT1 (default -0.11 V) captures the primary temperature coefficient, KT2 (default 0.022 V) accounts for body-bias interaction with temperature, T is the operating temperature, T_{NOM} is the nominal temperature, and V_{bseff} is the effective source-to-body voltage. The body effect is incorporated through coefficients K1 (default 0.5 V^{1/2}) and K2 (default 0), yielding terms like K1 (\sqrt{|V_{bs} + 2\phi_F|} - \sqrt{2\phi_F}) - K2 V_{bs}, where \phi_F is the Fermi potential. Short-channel and narrow-width adjustments, such as drain-induced barrier lowering (DIBL) via \Delta V_{th,DIBL} = \theta_{th}(L_{eff}) (V_{ds} - \phi_s - V_{bi}), ensure scalability with effective channel length L_{eff} and width W_{eff}. BSIM6 extends this to a charge-based framework, maintaining similar temperature and body dependencies while emphasizing symmetry for analog/RF applications, with V_{th} derived from pinch-off potential calculations that include non-uniform doping effects. The EKV model, a charge-based formulation continuous across inversion levels, defines V_{th} implicitly through the pinch-off voltage V_P, which marks the boundary where the inversion coefficient i_c = I_D / I_S = 1, transitioning from weak (i_c < 0.1) to strong inversion (i_c > 10). Here, V_P = V_{G'} - n V_T \ln(1 + \exp((V_{G'} - 2\phi_F - V_{ch})/ (n V_T))), with effective gate overdrive V_{G'} = V_{GS} - V_{FB} - \phi_s, slope factor n, and thermal voltage V_T; the long-channel V_{th} (VTO, default 0.5 V) shifts with as V_{TO}(T) = V_{TO} - TCV (T - T_{nom}). This avoids abrupt discontinuities, making EKV suitable for low-voltage designs where subthreshold operation is prevalent. Parameter extraction for these models typically involves fitting to measured current-voltage (I-V) characteristics. Threshold voltage V_{th}, body-effect \gamma (or [K1](/page/K1)), and \mu are optimized using nonlinear least-squares methods on curves (I_D vs. V_{GS}) at various V_{BS} and V_{DS}, often starting with linear or constant-current methods for initial V_{th} estimates, followed by to match saturation and linear regimes. For statistical variations, random fluctuation (RDF) is incorporated via Pelgrom's , where the standard deviation \sigma(\Delta V_{th}) = A_{VT} / \sqrt{W L}, with A_{VT} (typically 1-5 mV·μm) derived from simulations calibrated to measured mismatch data; this extends compact models for variability-aware design in scaled technologies. Modern extensions address non-planar devices, such as FinFETs, where fin width W_{fin} influences V_{th} due to quantum confinement and corner effects. The BSIM-CMG model for multi-gate FinFETs augments the core V_{th} with fin geometry terms, like \Delta V_{th,fin} \propto 1/W_{fin} from volume inversion, calibrated via 3D technology computer-aided design (TCAD) simulations that solve in full 3D structures to capture electrostatics and dopant distributions. TCAD tools, such as Sentaurus, enable predictive modeling by simulating V_{th} and DIBL in fin arrays before compact model tuning. Reliability effects, including (HCI) in nMOSFETs and (NBTI) in pMOSFETs, induce V_{th} shifts that degrade performance over time. HCI generates traps under high V_{DS}, causing a \Delta V_{th} decrease of 50-100 over a 10-year lifetime at accelerated , while NBTI creates traps under negative V_{GS} and elevated temperature, increasing |\Delta V_{th}| by similar magnitudes in pMOSFETs. These are modeled as power-law time dependencies, \Delta V_{th} \propto t^n, with n \approx 1/4 for NBTI (reaction-diffusion ) and n \approx 1/2 for HCI (lucky-electron model), integrated into compact models via bias- and temperature-dependent lookup tables or empirical functions for lifetime prediction in s.

References

  1. [1]
    Threshold Voltage - an overview | ScienceDirect Topics
    4. MOSFET threshold voltage is defined as the gate voltage at which significant current starts to flow from the source to the drain (Fig. 5).
  2. [2]
    [PDF] Derivation of MOSFET Threshold Voltage from the MOS Capacitor
    Threshold voltage is the gate voltage needed to turn on a MOSFET. It's the voltage where the interface potential equals and opposes the substrate potential.
  3. [3]
    The threshold voltage of MOSFET and its influence on digital circuits
    The MOSFET threshold voltage value will have influence in behaviour of electronic device which contain MOSFET transistors. The results obtained emphasize the ...
  4. [4]
    1960: Metal Oxide Semiconductor (MOS) Transistor Demonstrated
    John Atalla and Dawon Kahng fabricate working transistors and demonstrate the first successful MOS field-effect amplifier.Missing: original | Show results with:original
  5. [5]
    The MOSFET and Metal Oxide Semiconductor Tutorial
    Both the Depletion and Enhancement type MOSFETs use an electrical field produced by a gate voltage to alter the flow of charge carriers, electrons for n-channel ...
  6. [6]
    Challenges for 10 nm MOSFET process integration - ResearchGate
    Aug 6, 2025 · The threshold voltage requirements pose a severe challenge to the process technology for appropriate gate materials with the tunable work ...Missing: modern | Show results with:modern
  7. [7]
    [PDF] Physics of Semiconductor Devices
    In this Third Edition of Physics of Semiconductor Devices, over 50% of the material has been revised or updated, and the material has been totally reorganized.
  8. [8]
    [PDF] MOSFETs in ICs—Scaling, Leakage, and Other Topics
    This understanding gives us a simple equation for Vt roll-off,. (7.3.1) where Vt-long is the threshold voltage of a long-channel transistor, for which Cd = 0.
  9. [9]
    [PDF] MOSFET DEVICE SCALING: A (BIASED) HISTORY OF GATE STACKS
    Kahng and M.M. Atalla to report the first MOSFET on a silicon substrate in. 1960 [6]. Following J. Kilby's invention in 1958 of the IC using germanium [7], R ...
  10. [10]
    [PDF] The New Era of Scaling in an SoC World - UCSD CSE
    Classical MOSFET scaling was first described in 1974. Page 7. 7. 30 Years of MOSFET Scaling. Gate Length: 1.0 µm. 35 nm. Gate Oxide Thickness: 35 nm. 1.2 nm.
  11. [11]
    [PDF] 10-nm CMOS - UC San Diego
    This effect shifts threshold voltage up for another 0.1 to 0.3 V depending on the electric field. The Vt shift can be obtained through theoretical calculation ...
  12. [12]
    [PDF] Hafnium oxide-based dielectrics by atomic layer deposition
    In 2007, when the MOSFET technology node had reached 45 nm, the. CMOS fabrication industry moved to replace the incumbent dielectric layer, SiO2, with a high-κ ...
  13. [13]
    Experimental determination of equivalent oxide thickness of gate ...
    May 15, 2002 · The EOT is defined as the physical thick- ness of thermal oxide having the same capacitance, and is usually derived from the capacitance-voltage ...<|separator|>
  14. [14]
    [PDF] MOS Transistor
    The supply voltage Vdd = 5 V. ○ Body Effect ○. 6.18 P-channel MOSFET with heavily doped P-type poly-Si gate has a threshold voltage of. –1.5 V with Vsb = 0 V.
  15. [15]
    [PDF] Optimal Body Bias Selection for Leakage Improvement and Process ...
    Reverse substrate (or body) bias in the. “off”-state is one leakage reduction technique that has been successfully employed to reduce IOFF [1]. By selectively ...
  16. [16]
    [PDF] Electrical Characteristics Comparison Between Fully-Depleted SOI ...
    The advantage of FD SOI MOSFET include the elimination of the floating-body effect and better short channel behaviour. For Partially Depleted SOI device (Fig 2- ...
  17. [17]
    [PDF] an2386-how-to-achieve-the-threshold-voltage-thermal-coefficient-of ...
    The threshold voltage thermal coefficient (see Equation 41) depends on the Fermi potential divided by T, minus a constant and minus a term function of the ...
  18. [18]
  19. [19]
    [PDF] What's not in the power MOSFET data sheet, part 1: temperature ...
    The threshold voltage has a negative temperature coefficient and the on-resistance has a positive temperature coefficient. Figure 2. The electrical ...
  20. [20]
    [PDF] Dual-Threshold Voltage Design of Sub-Threshold Circuits by Jia Yao
    Aug 2, 2014 · Threshold voltage of MOSFET technology represents the value of the gate-source voltage when the current in a MOS transistor starts to ...<|control11|><|separator|>
  21. [21]
    [PDF] Body Bias Voltage Computations for Process and Temperature ...
    Body biasing, as a means of threshold voltage modulation provides an effective knob to alter the delay and leakage of the circuit.
  22. [22]
  23. [23]
    Comparison of Temperature Dependent Carrier Transport in FinFET ...
    A similar temperature dependence of threshold voltage (VTH) and subthreshold swing (SS) is observed for both devices. However, effective mobility (μeff) shows ...
  24. [24]
    Influence of high substrate doping levels on the threshold voltage ...
    Abstract: The high levels of substrate doping needed in deep-submicrometer MOS devices affect device properties strongly. The authors present a detailed ...
  25. [25]
    Threshold voltage model for deep-submicrometer MOSFETs
    The model considers gate oxide thickness, source/drain junction depth, and channel doping. LDD devices show less V/sub th/ dependence on drain voltage.
  26. [26]
    Degradation of MOSFETs drive current due to halo ion implantation
    We found that the degradation of saturation current of halo device is directly related to the less roll-off of saturation threshold voltage due to the increased ...
  27. [27]
    Analyze of input process parameter variation on threshold voltage in ...
    While the halo implant dose was identified as an adjustment factor to get the nominal values of VTH for NMOS device equal to 0.289V at tox= 1.06nm.
  28. [28]
  29. [29]
    Dual work function metal gate CMOS transistors by Ni-Ti interdiffusion
    The threshold voltage of the n-MOS devices is determined solely by Ti, while the threshold voltage of the p-MOS devices is determined by the Ni-rich alloy ...
  30. [30]
    Modeling and Optimizing of Threshold Voltage of 32nm NMOS ...
    After the gate is built, the next step is to implant boron difluoride (bf2) at the well active area in 544 Page 3 order to adjust VTH value.
  31. [31]
  32. [32]
    Random dopant induced threshold voltage lowering and fluctuations ...
    The detailed analysis shows that the threshold voltage fluctuations are determined not only by the fluctuation in the dopant number, but also in the dopant ...Missing: formula | Show results with:formula
  33. [33]
  34. [34]
    [PDF] Study of Variability in Advanced Transistor Technologies
    May 1, 2015 · This study examines how transistor variability, especially in threshold voltage, increases as dimensions scale down, impacting circuit ...Missing: Pelgrom's | Show results with:Pelgrom's
  35. [35]
    [PDF] Thin-Body Silicon FET Devices and Technology - UC Berkeley EECS
    May 9, 2007 · undoped channel is used to reduce the effect of statistical dopant fluctuations. They also have additional benefits of better short channel ...<|separator|>
  36. [36]
  37. [37]
    A simple theory to predict the threshold voltage of short-channel ...
    A simple theory to predict the threshold voltage of short-channel IGFET's. Author links open overlay panel L.D. Yau. Show more. Add to Mendeley. Share. Cite.
  38. [38]
    Drain induced barrier lowering (DIBL) accurate model for nanoscale ...
    In this paper, an accurate new model for drain induced barrier lowering (DIBL) tunneling in silicon on insulator (SOI) metal oxide semiconductor field ...
  39. [39]
  40. [40]
    Constant-Current Threshold Voltage Extraction in HSPICE for ...
    We present a substantially enhanced HSPICE feature that extracts MOSFET threshold voltage. (VT) based on the constant-current definition universally adopted ...
  41. [41]
  42. [42]
    MOSFET Threshold Voltage Extraction: How to Get Power ... - Keysight
    Aug 8, 2023 · It means the gate voltage where a MOSFET transitions from weak to strong inversion. In simple terms, Vth is the voltage where conduction of ...Threshold Voltage: A... · Linear Extrapolation Method... · Constant Current Method (cc)
  43. [43]
    Comparison of MOSFET-threshold-voltage extraction methods
    Aug 5, 2025 · ... The two devices show similar short-channel performance. The threshold voltage (V T ) of the devices is about 0.4 V. V T is extracted from ...
  44. [44]
    [PDF] STABLE EXTRACTION OF THRESHOLD VOLTAGE USING ...
    In the transconductance change method, the threshold voltage is defined as the gate voltage at which the derivative of the low drain voltage transconductance ...Missing: Id^{ | Show results with:Id^{<|separator|>
  45. [45]
    On the gm/ID-based approaches for threshold voltage extraction in ...
    In this paper, we investigate the transconductance-to-current ratio (gm/ID) methods for the threshold voltage extraction using two popular threshold voltage ...
  46. [46]
  47. [47]
    Subthreshold-Conduction | Digital-CMOS-Design || Electronics Tutorial
    The subthreshold conduction effect can be formulated as : IDS = I0 exp VGSn KTq. Where n > 1 is a non ideality factor. • In digital circuit designs the ...
  48. [48]
    Does the Threshold Voltage Extraction Method Affect Device ...
    Apr 26, 2021 · We have considered five different extraction methods: the second derivative (SD), constant current (CC), linear extrapolation (LE), third ...<|separator|>
  49. [49]
    (PDF) MOS transistors characterization by split C-V method
    This paper shows split CV measurement results of MOS transistors made in different technologies. The idea was to show that this method is widely applicable.
  50. [50]
    A new characterization technique for SOI wafers: Split C(V) in ...
    We demonstrate for the first time the feasibility of split C–V measurements on as-fabricated SOI wafers using pseudo-MOSFET configuration. An adapted ...
  51. [51]
    [PDF] Cryogenic MOSFET Threshold Voltage Model - arXiv
    Apr 22, 2019 · Abstract—This paper presents a physics-based model for the threshold voltage in bulk MOSFETs valid from room down to cryogenic temperature (4.2 ...
  52. [52]
    [PDF] BSIM4.3.0 MOSFET Model - CMOSedu.com
    Threshold Voltage Model 2-1. 2.1 Long-Channel Model With Uniform Doping 2-1 ... BSIM4 considers the influence of stress on mobility, velocity saturation, ...
  53. [53]
    [PDF] The EPFL-EKV MOSFET Model Equations for Simulation - Ngspice
    Mar 29, 1999 · It accounts also for charge-sharing and reverse short-channel effects on the threshold voltage. For P-channel devices, is given a negative sign.
  54. [54]
    [PDF] BSIM6.0 MOSFET Compact Model - WRcad
    May 4, 2013 · This chapter presents the BSIM6 tempera- ture dependence models for threshold voltage, mobility, saturation velocity, source/drain.
  55. [55]
    (PDF) BSIM3v3 to EKV2.6 Model Parameter Extraction and ...
    Oct 21, 2025 · The model parameters are extracted for different bins and optimisation is done using nonlinear optimisation LM algorithm. The optimised EKV2.6 ...
  56. [56]
    Models | BSIM GROUP
    ... model for silicon-on-insulator devices BSIM-BULK A charge-based model for bulk MOSFETs BSIM4 A threshold voltage-based model for bulk MOSFET (Link for BSIM3),
  57. [57]
    BSIM-CMG: Standard FinFET compact model for advanced circuit ...
    BSIM-CMG is an industry standard compact model for FinFETs, capturing advanced physical effects for advanced circuit design.
  58. [58]
    Mechanism of nitrogen-enhanced negative bias temperature ...
    Typically, NBTI could cause a considerable threshold voltage shift (ΔVth) by as much as 50–100 mV in a pMOSFET due to the creation of interface states (Nit or ...
  59. [59]
    A computational model of NBTI and Hot Carrier Injection time ...
    The power law model is similar to threshold voltage shift due to hot-carrier degradation with stress time [18, 25] and NBTI. This power law model can be ...