Fact-checked by Grok 2 weeks ago

Seebeck coefficient

The Seebeck coefficient, denoted as S, is a fundamental material property in thermoelectrics that quantifies the magnitude and direction of the thermoelectric voltage generated across a material in response to an applied , typically measured in microvolts per (\mu \mathrm{V/K}). It arises from the diffusion of charge carriers from hotter to cooler regions, driven by the greater average energy of particles at higher temperatures, and is central to the Seebeck effect discovered by physicist in 1821, who observed that joining dissimilar metals at different temperatures produces an capable of deflecting a needle. Mathematically, the Seebeck coefficient is defined as S = -\frac{\Delta V}{\Delta T}, where \Delta V is the difference and \Delta T is the difference across the material, with the negative sign reflecting the conventional for flow in n-type materials. Its value depends on factors such as carrier concentration, effective mass, and ; for instance, metals exhibit small coefficients (typically 1–70 \mu \mathrm{V/K}), while heavily doped semiconductors can reach 100–1000 \mu \mathrm{V/K} or higher, with exceptional cases like certain composites exceeding 3000 \mu \mathrm{V/K}. The of S indicates the dominant : positive for p-type materials where holes prevail, and negative for n-type where s dominate, influencing the efficiency of thermoelectric devices. In practice, the Seebeck coefficient is pivotal for applications in and ; thermocouples, formed by pairing materials with differing S values (e.g., chromel-alumel yielding ~40 \mu \mathrm{V/^\circ C}), enable precise sensing in industrial and scientific settings. Additionally, high-S materials are essential for thermoelectric generators that convert to via the Seebeck effect, offering solid-state alternatives to systems with potential efficiencies tied to the ZT = S^2 \sigma / \kappa, where \sigma is electrical and \kappa is thermal . Ongoing research focuses on enhancing S through nanostructuring and doping to improve ZT values, advancing technologies.

Definition and Fundamentals

Formal Definition

The Seebeck effect is a thermoelectric phenomenon in which a difference applied across a generates an (EMF), producing a voltage without the flow of external current. This arises from the of charge carriers from hotter to cooler regions due to their thermal excitation, creating a net charge separation and an associated . In the framework of linear irreversible , thermoelectric transport is described by coupled equations for density \mathbf{J} and \mathbf{Q}. The is given by \mathbf{J} = \sigma \mathbf{E} - \sigma S \nabla T, where \sigma is the electrical conductivity, \mathbf{E} is the , S is the Seebeck coefficient, and \nabla T is the ; the follows \mathbf{Q} = \Pi \mathbf{J} - \kappa \nabla T, with \Pi the Peltier coefficient and \kappa the thermal conductivity. In the open-circuit condition where \mathbf{J} = 0, the equation simplifies to \mathbf{E} = S \nabla T, indicating that the Seebeck coefficient relates the induced to the . The formal definition of the Seebeck coefficient S is thus S = -\frac{\Delta V}{\Delta T} for a finite difference \Delta T and corresponding open-circuit voltage difference \Delta V, or more precisely in the limit, S = -\left( \frac{\partial V}{\partial T} \right)_{J=0}. This definition emerges directly from the relations, where the negative ensures positive values for p-type materials (where holes dominate) and negative values for n-type materials (where electrons dominate). The units of the Seebeck coefficient are volts per (/), though values are typically on the order of microvolts per kelvin (μV/) for most materials.

Sign Convention

The standard sign convention for the Seebeck coefficient in semiconductors assigns a positive value to p-type materials, where the hot end develops a positive voltage relative to the cold end due to the predominance of positive charge carriers (holes) diffusing from hot to cold. Conversely, n-type materials exhibit a negative Seebeck coefficient, as electrons, the dominant negative charge carriers, diffuse from the hot end to the cold end, making the hot end negatively charged. This convention aligns the sign of the Seebeck coefficient with the sign of the majority charge carriers, facilitating consistent interpretation in thermoelectric device design and analysis. In circuit analysis for thermocouples, the sign of the Seebeck coefficient determines the of the induced flow under a . For a thermocouple formed by two materials A and B, if the relative Seebeck coefficient S_{AB} = S_A - S_B > 0, the flows from material A to B at the cold junction when the other junction is heated, reflecting the established by the . This reversal with sign change is critical for predicting and ensuring proper wiring in circuits, where misinterpreting the sign could lead to inverted readings. A representative example is the iron-copper junction, where iron exhibits a positive Seebeck coefficient relative to (approximately +12.5 μV/K near , derived from absolute values of +19 μV/K for iron and +6.5 μV/K for ). In this setup, heating the junction results in the iron side becoming positively charged relative to , driving from iron to at the cold end. A common pitfall arises from confusing Seebeck coefficients (measured relative to a like a superconductor at zero) with relative ones (between two materials), which can lead to sign errors in multi-material systems. For instance, both iron and have positive absolute coefficients, but their relative sign depends on the difference, potentially inverting expected polarities if not accounted for in experimental setups.

Relation to Other Thermoelectric Coefficients

The Seebeck coefficient S is interconnected with the Peltier coefficient \Pi and the Thomson coefficient \tau through the Kelvin relations, which provide thermodynamic linkages between these thermoelectric transport parameters. The first Kelvin relation states that the Peltier coefficient is the product of the Seebeck coefficient and the absolute temperature: \Pi = S T. The second Kelvin relation connects the temperature derivative of the Seebeck coefficient to the Thomson coefficient: \tau = -T \frac{dS}{dT}. These relations ensure consistency across the thermoelectric effects by linking heat and charge transport phenomena. Physically, the Seebeck, Peltier, and Thomson coefficients all originate from the transported by charge carriers in a under a or current flow. In this framework, the Seebeck coefficient represents the per unit charge carried by the charge carriers, while the Peltier coefficient describes the absorbed or released at a junction due to this flow, and the Thomson coefficient accounts for the reversible production within a . This unified interpretation underscores that the coefficients are manifestations of the same underlying carrier transport mechanism. The interconnections are further grounded in Onsager reciprocity, a principle from that imposes symmetry on the transport coefficient matrix. For thermoelectric effects, this reciprocity yields S_{ij} = -S_{ji}, where S_{ij} is the Seebeck coefficient relating the in direction i to the in direction j. This antisymmetry ensures the thermodynamic consistency of the relations between Seebeck, Peltier, and Thomson coefficients in isotropic and anisotropic materials. These Kelvin relations were derived by William Thomson (later Lord Kelvin) in 1854, building on the earlier observations of Seebeck and Peltier to establish a thermodynamic foundation for thermoelectricity.

Historical Development

Discovery by Seebeck

In 1821, Thomas Johann Seebeck, a Baltic German physicist, conducted pioneering experiments on the interaction between heat and electrical conduction in metals. He formed a closed circuit by joining wires of dissimilar metals, specifically bismuth and copper, creating two junctions. Upon heating one junction while keeping the other at a lower temperature, Seebeck observed a significant deflection of a nearby compass needle, which he attributed to the generation of magnetism induced by the temperature difference. This observation marked the initial discovery of what would later be known as the thermoelectric effect, though Seebeck misinterpreted it as a form of thermomagnetism akin to the recently discovered electromagnetism by Hans Christian Ørsted. Seebeck detailed his findings in a series of papers published between 1822 and 1823, titled Magnetische Polarisation der Metalle und Erze durch Temperatur-Differenz, presented to the and appearing in their Abhandlungen der physikalischen Klasse. In these works, he described extensive tests with various metal combinations, noting the consistent deflections and proposing that gradients could polarize metals magnetically. These publications laid the groundwork for understanding the phenomenon, emphasizing its reproducibility across different materials and ranges, though still framed within a magnetic . The practical implications of Seebeck's observations quickly emerged in the development of the during the 1820s. Inspired by Seebeck's observations, Italian physicist Leopoldo Nobili pioneered this device in the 1820s, often in collaboration with Macedonio Melloni, by connecting multiple thermocouples in series to amplify the voltage generated by temperature differences. The enabled sensitive detection of and heat, serving as an early tool for and pyrometry before the advent of more advanced sensors. Seebeck's magnetic interpretation was soon corrected through further experimentation. Shortly thereafter, Danish physicist recognized that the compass deflection resulted from an induced by the difference, rather than direct thermomagnetism, and termed the phenomenon "thermoelectricity". This clarification shifted the focus from thermomagnetism to the direct conversion of heat into , paving the way for subsequent theoretical and applied advancements.

Evolution of the Concept

Following Thomas Johann Seebeck's discovery of the in 1821, the conceptual understanding of the Seebeck coefficient evolved rapidly in the mid-19th century through thermodynamic linkages to related phenomena. In the 1850s, William Thomson, later , established reciprocal relations that connected the Seebeck effect to the Peltier and Thomson effects, providing a unified thermodynamic framework for thermoelectricity. These relations demonstrated that the Peltier coefficient is proportional to the product of the Seebeck coefficient and absolute temperature, while the Thomson coefficient relates to the temperature derivative of the Seebeck coefficient, enabling predictions of heat absorption or release under current flow in temperature gradients. This work shifted the Seebeck coefficient from an empirical observation to a thermodynamically grounded quantity essential for analyzing irreversible processes in conductors. By the early 20th century, the Seebeck coefficient gained prominence in the emerging field of , where classical models were refined through quantum mechanical interpretations. Arnold Sommerfeld's application of quantum statistics to electron transport in metals during the late provided an initial quantum framework, explaining the Seebeck coefficient as arising from the asymmetric scattering of s near the . This was further advanced in 1931 by Lars Onsager's reciprocal relations for non-equilibrium transport, which generalized the Seebeck coefficient's role in anisotropic media and laid the groundwork for understanding coupled heat and charge flows beyond simple metals. These developments integrated the Seebeck coefficient into band theory, highlighting its dependence on electronic and carrier diffusion, thus bridging classical with quantum solid-state phenomena. A major milestone occurred in the mid-20th century with the recognition of the Seebeck coefficient's potential in semiconductors for thermoelectric energy conversion. In the , and collaborators emphasized semiconductors' large Seebeck coefficients—often orders of magnitude higher than in metals—due to their tunable concentrations and band structures, spurring research into materials like telluride (Bi₂Te₃). This era saw experimental verification that doping could optimize the Seebeck coefficient alongside electrical conductivity, boosting interest in practical thermoelectrics and establishing it as a key parameter in the ZT. 's studies further clarified how band-edge positions and scattering mechanisms influence the coefficient in and , solidifying semiconductors as superior to metals for thermoelectric applications. From the late to 2025, the concept has evolved through integration with , where low-dimensional systems exhibit enhanced Seebeck coefficients due to quantum confinement effects. Seminal theoretical work by L. D. Hicks and M. S. Dresselhaus in 1993 predicted that quantum wells and superlattices could increase the Seebeck coefficient by sharpening the , reducing thermal conductivity while preserving electrical transport. This spurred experimental advances, such as silicon nanowires demonstrating up to 100% enhancement in Seebeck coefficient compared to bulk , attributed to boundary scattering and one-dimensional confinement. Similarly, quantum dots in Ge/Si structures have shown Seebeck enhancements of 40% or more through energy filtering of low-energy carriers, with ongoing research up to 2025 exploring hybrid nanowire-dot arrays for further optimization in nanoscale thermoelectrics. These developments have redefined the Seebeck coefficient as a tunable in confined geometries, extending its theoretical scope beyond bulk materials.

Measurement Methods

Relative Seebeck Coefficient

The relative Seebeck coefficient, denoted as S_{AB}, quantifies the thermoelectric voltage generated by the difference in Seebeck coefficients between two dissimilar materials, A and B, in a configuration. This setup involves forming a closed loop with the two materials joined at two junctions maintained at different temperatures, creating a along their lengths. When a temperature difference \Delta T is applied, an \Delta V is induced, such that S_{AB} = \frac{\Delta V}{\Delta T} = S_A - S_B, where S_A and S_B are the absolute Seebeck coefficients of the respective materials. This differential measurement offers key advantages over absolute methods, as it eliminates the requirement for a universal reference material or precise absolute scale, relying instead on the relative response between the paired conductors. It is the foundational principle for thermometry, enabling reliable sensing in and scientific applications without complex against primary standards. The sign of the generated \Delta V follows the convention where a positive voltage indicates the hot junction drives current from material A to B in the circuit. In practice, the measurement procedure establishes a controlled linear across the sample using heaters and heat sinks, with voltage probes attached at points along the materials where local are monitored via auxiliary thermocouples or sensors. The relative Seebeck coefficient at a target is then determined by plotting the measured voltages against the corresponding temperature differences and extrapolating to the isotherm (where \Delta T = 0) to isolate the intrinsic response at uniform . Potential error sources include at the probe-material interfaces, which can cause voltage offsets, as well as deviations from in the temperature profile due to losses or radiative . Historically, relative Seebeck coefficient measurements have been central to the standardization of types, such as Type K (chromel-alumel) and Type J (iron-constantan), which were developed and formalized in the mid-20th century through efforts by the Instrument Society of America (now part of ) and adopted into ASTM and ANSI standards for consistent performance across applications.

Absolute Seebeck Coefficient

The absolute Seebeck coefficient establishes a universal scale for the thermoelectric response of a material by referencing it to a standard where the Seebeck coefficient is zero, such as a superconductor in its superconducting state at temperatures approaching 0 K, where perfect conductors exhibit S = 0 due to the absence of by charge carriers. This definition allows the direct determination of the , S_absolute, from the measured thermoelectric voltage gradient across the sample when connected via superconducting leads that contribute no voltage. Key techniques for measuring the absolute Seebeck coefficient rely on cryogenic conditions to utilize superconducting references below 1 , where the sample and leads are immersed in a controlled low-temperature to create a precise . Alternative approaches include immersion in variable-temperature baths, which adjust the reference temperature incrementally, or adiabatic demagnetization of paramagnetic materials to reach millikelvin ranges without continuous cooling. These methods contrast with relative measurements, which serve as a prerequisite for extending absolute values to higher temperatures via integration. Challenges in these measurements stem from the need for specialized cryogenic setups, such as dilution refrigerators or cryostats, to maintain stable s and minimize leaks, ensuring reliable voltage detection with nanovolt sensitivity. Achieving high accuracy, typically on the order of 0.1 μV/ or better, requires careful control of resistances, lead contributions, and gradients, with analyses often revealing contributions from finite resolutions and material inhomogeneities. Advancements in the 1970s and 1980s included standardization efforts by the National Institute of Standards and Technology (NIST), which developed reference tables and calibration protocols for thermocouple materials, laying the groundwork for absolute scales through low-temperature benchmarks. More recent progress, extending into the 2020s, encompasses non-cryogenic optical techniques, such as laser-heating methods that enable absolute Seebeck determinations at elevated temperatures up to several hundred by modulating thermal gradients without physical contacts. These innovations, including graphene-based references for room-temperature absolutes, reduce reliance on while preserving precision.

Material Properties

Values for Common Metals

The Seebeck coefficient for metals is generally small in magnitude, typically on the order of a few per (μV/), reflecting the dominance of diffusion with minimal contributions from other mechanisms. At (around 300 ), pure metals exhibit values that are often positive for transition metals and negative for noble metals, with a nearly linear dependence on due to the model of transport. These values are compiled from standardized measurements and are crucial for applications in thermometry and design.
Metal/AlloySeebeck Coefficient (μV/K at 300 K)Notes
(Cu)+1.8Pure annealed; increases slightly with .
Iron (Fe)+15.0Alpha phase; shows stronger dependence due to magnetic effects.
(Ni)-15.0Negative sign indicates electron-like carriers; sensitive to impurities.
(Al)-1.5Low value typical of free-electron metals; linear up to 500 K.
(Pt)-5.0Standard reference; stable and linear over wide range.
(Au)-1.8Similar to ; used in high-purity thermocouples.
(Cu-Ni )-35.0Engineered for low ; nearly -independent.
Type S (Pt-10%Rh vs. Pt)+6.4 to +11.0Differential value; increases with Rh content for higher sensitivity.
These values, derived from steady-state comparative measurements against reference materials like , highlight the low thermoelectric power of metals compared to other material classes. In metals, the absolute value of the Seebeck coefficient tends to increase with impurity scattering, as defects disrupt the symmetric electron distribution near the , enhancing the diffusive voltage gradient; for instance, adding to raises |S| from near zero to tens of μV/K in alloys like . Phonon drag effects, which can amplify S in semiconductors, are negligible in metals due to strong electron-phonon coupling that dissipates momentum quickly. Recent compilations confirm these trends, incorporating cryogenic and high-temperature for industrial alloys. A practical example is the Type S thermocouple, composed of platinum-rhodium versus pure , which leverages the small differential Seebeck coefficient of about 10 μV/K to achieve accurate temperature measurements up to 1600 K in oxidizing environments, as standardized by the International Temperature Scale.

Values for Semiconductors and Other Materials

Semiconductors exhibit significantly higher Seebeck coefficients than metals due to the presence of a band gap, which breaks the symmetry between electrons and holes, leading to an asymmetric near the that enhances the thermoelectric response. This asymmetry results in Seebeck coefficients typically two orders of magnitude larger in semiconductors compared to metals, making them preferable for thermoelectric applications where a large voltage response to temperature gradients is required. In semiconductors, the sign of the Seebeck coefficient indicates the dominant type: positive values for p-type materials ( conduction) and negative values for n-type materials ( conduction). The magnitude depends strongly on doping level, with lower concentrations yielding higher absolute values due to increased per , while higher doping reduces it by bringing the closer to the band edge. Temperature dependence often shows a peak in the Seebeck coefficient near temperatures related to the band gap energy, where thermal balances diffusive and drag contributions, followed by a decrease at higher temperatures due to increased across the gap. Representative examples illustrate these trends across various semiconductor classes. telluride (Bi₂Te₃), a benchmark thermoelectric material, achieves a Seebeck coefficient of approximately 200 μV/K for p-type variants at , benefiting from its narrow and high carrier mobility. at low doping levels (around 10¹⁸ cm⁻³) can reach up to 660 μV/K, highlighting the potential for enhanced values in lightly doped intrinsic-like regimes, though practical applications often involve higher doping for balanced conductivity. like poly(3,4-ethylenedioxythiophene) (PEDOT:PSS) typically exhibit lower values around 50 μV/K, tunable via post-treatments or additives to improve power factors in flexible devices. Emerging materials further expand the range of high Seebeck coefficients. Halide perovskites, such as methylammonium lead iodide (MAPbI₃), demonstrate exceptional values up to 920 μV/K, attributed to their soft and defect-tolerant structures that minimize conductivity while maintaining decent electrical . Two-dimensional materials like show Seebeck coefficients up to 100 μV/K in doped or nanostructured forms, leveraging structures for tunable , though pristine remains limited by its semimetallic nature. Half-Heusler alloys, such as ZrNiSn-based compounds, achieve around 150-165 μV/K for n-type compositions at intermediate temperatures, owing to their robust half-filled gaps and resistance to degradation. Exceptional cases include certain composites, which can exceed 3000 μV/K due to enhanced phonon drag and low carrier concentrations.
MaterialTypeSeebeck Coefficient (μV/K at ~300 K)Key Notes
Bi₂Te₃p-type~200Narrow band gap enhances near-room-temperature performance.
SiLow n- or p-doping~660-1000Peaks at low carrier concentrations; decreases with doping.
PEDOT:PSSp-type~50Organic; tunable via solvents or additives for flexibility.
MAPbI₃ (perovskite)n- or p-type~920High due to defect states; solution-processable.
Graphene (doped)AmbipolarUp to ~1002D Dirac bands; substrate effects influence value.
Half-Heusler (e.g., ZrNiSn)n-type~150-165Stable at high temperatures; band engineering optimizes.

Theoretical Explanations

Charge Carrier Diffusion

The diffusion mechanism is the primary contributor to the Seebeck coefficient in metals and degenerate semiconductors, arising from the response of s to a . In a subjected to a temperature difference, the hotter region excites a greater number of s (s or holes) above the , leading to a higher local carrier density compared to the colder region. This density imbalance drives a net of carriers from the hot to the cold end, resulting in charge accumulation that generates an opposing . The Seebeck coefficient S quantifies this effect as the induced per unit , S = -E / \nabla T, with the negative for transport. The sign of S is negative for n-type s ( diffusion makes the cold end negative) and positive for p-type s (hole diffusion makes the cold end positive). Theoretically, this diffusion process is described within the framework of the Boltzmann transport equation (BTE) under the relaxation time approximation. The BTE yields the electrical current density \mathbf{j} = \sigma \mathbf{E} - \sigma S \nabla T, where \sigma is the electrical conductivity, and in open circuit conditions (\mathbf{j} = 0), the Seebeck coefficient emerges as S = \frac{1}{eT} \frac{\int (\epsilon - \mu) \sigma(\epsilon) \left( -\frac{\partial f}{\partial \epsilon} \right) d\epsilon}{\int \sigma(\epsilon) \left( -\frac{\partial f}{\partial \epsilon} \right) d\epsilon}, with e the elementary charge, T temperature, \mu chemical potential, \epsilon energy, f the Fermi-Dirac distribution, and \sigma(\epsilon) the energy-dependent conductivity. At low temperatures (k_B T \ll E_F, where E_F is the Fermi energy), a Sommerfeld expansion approximates the integrals, leading to the Mott formula: S = -\frac{\pi^2 k_B^2 T}{3e} \left. \frac{d \ln \sigma(E)}{dE} \right|_{E=E_F}, where k_B is Boltzmann's constant. This expression highlights how S depends on the energy derivative of conductivity near the Fermi level, reflecting the asymmetric scattering or density of states that biases carrier transport. For a simple free-electron metal with constant relaxation time, \sigma(E) \propto E^{3/2}, yielding S \approx -\frac{\pi^2 k_B^2 T}{2 e E_F}, a small value scaling linearly with T and inversely with E_F. The derivation assumes elastic scattering and degenerate statistics, making it applicable to metals and heavily doped semiconductors. Key factors influencing the contribution include carrier concentration and , which enter through \sigma(E) = e^2 n(E) \tau(E) / m^*, with n(E) carrier , \tau(E) relaxation time, and m^* effective mass. Higher carrier concentration raises E_F, reducing |S| in degenerate systems, while energy-dependent (e.g., via mechanisms) amplifies the in the Mott , enhancing |S|. In non-degenerate semiconductors, where k_B T \gg E_g (bandgap), the extends to S \approx \frac{k_B}{e} \left( \frac{E_c - \mu}{k_B T} + A \right) for electrons, with E_c conduction edge and A a , emphasizing the role of carrier relative to the edge over Fermi-level . The Mott holds well at low T but breaks down in non-degenerate regimes or high T, where full BTE solutions are needed. Recent extensions in the 2020s address disordered systems, such as Mott conduction, where scaling arguments modify the temperature dependence to S \propto T^{d/(d+1)} in d-dimensions, accounting for localized states and inelastic processes beyond the elastic Mott assumption. This relates briefly to the per carrier transported, as S approximates -s/e ( s) in diffusive regimes.

Phonon Drag

In the phonon drag mechanism, a across a material creates a nonequilibrium distribution of phonons, with hotter phonons near the hot end possessing greater momentum. These phonons interact with charge carriers through electron-phonon scattering, transferring momentum and "dragging" the carriers toward the cold end, which generates an additional and enhances the overall Seebeck voltage beyond the pure diffusive contribution from charge carriers. This effect, first theoretically analyzed by Clarence Herring, adds a positive contribution to the Seebeck coefficient for both n-type and p-type materials, as the drag direction aligns with the diffusive flow. A simple estimate for the phonon drag contribution S_\mathrm{ph} to the Seebeck coefficient, derived from kinetic theory considerations in Herring's framework, is S_\mathrm{ph} \approx \left( \frac{k_\mathrm{B}}{e} \right) \left( \frac{C_\mathrm{ph}}{C_\mathrm{el}} \right) \left( \frac{v_\mathrm{s}}{v_\mathrm{F}} \right) \times f, where k_\mathrm{B} is Boltzmann's constant, e is the elementary charge, C_\mathrm{ph} and C_\mathrm{el} are the volumetric heat capacities of phonons and electrons, v_\mathrm{s} is the speed of sound, v_\mathrm{F} is the Fermi velocity, and f is a numerical factor of order unity accounting for scattering details. The phonon drag effect is most prominent in pure or lightly doped semiconductors at low temperatures (10–100 K), where long phonon mean free paths enable efficient momentum transfer, but it becomes negligible in metals due to frequent Umklapp scattering that dissipates phonon momentum. Experimental observations confirm this through characteristic peaks or rapid increases in the Seebeck coefficient as a function of , S(T), in materials like and at cryogenic s. For instance, early measurements on silicon samples showed the phonon drag term dominating the total S below approximately 100 , with S rising sharply before decreasing at higher temperatures due to increased . Similar behavior was reported in , where the effect leads to enhanced thermopower in high-purity crystals. The ratio of the phonon drag contribution to the diffusive term typically ranges from 0.1 to 0.3 in these semiconductors under moderate doping conditions. Studies from the and have extended investigations to nanostructures, demonstrating that phonon drag can be tuned for thermoelectric improvements. In silicon nanowires, comparisons between bulk and nanostructured samples have quantified the drag effect's magnitude in bulk material at , revealing contributions up to one-third of the total Seebeck coefficient. simulations of -based alloys and nanostructures indicate that engineering —such as through designs—can enhance the Seebeck coefficient by optimizing drag while suppressing thermal conductivity, potentially boosting the thermoelectric by over an at low temperatures. Furthermore, experiments on doped crystals have measured phonon drag accounting for 50–60% of the total Seebeck coefficient at optimal carrier densities.

Connection to Entropy

The Seebeck coefficient provides a thermodynamic perspective on the transport of by charge carriers in a . Thermodynamically, the Seebeck coefficient can be expressed as S = -\frac{1}{eT} \left( \frac{\partial \mu}{\partial \ln T} \right)_N, with the derivative taken at constant particle number N. This formulation highlights how temperature variations in the drive the thermoelectric response through changes. This relation interprets the Seebeck coefficient as the entropy flow per unit charge , quantifying the "dragged" entropy accompanying charge motion under a . In optimized thermoelectrics, particularly semiconductors, the maximum |S| scales approximately as \frac{k_B}{e} \cdot \frac{\Delta E_g}{k_B T}, where k_B is Boltzmann's constant and \Delta E_g represents the relevant energy scale, such as half the band gap, emphasizing the role of material band structure in enhancing . The diffusive motion of charge carriers and drag contribute to this total entropy flow, unifying microscopic mechanisms in a macroscopic thermodynamic framework. The Seebeck coefficient's connection to directly influences the thermoelectric ZT = \frac{S^2 \sigma T}{\kappa}, where \sigma is electrical conductivity and \kappa is thermal conductivity; higher per carrier boosts S^2, thereby improving device efficiency limits approaching the Carnot efficiency. In recent developments during the 2020s, theories incorporating non-equilibrium have addressed transient thermoelectric phenomena, such as time-dependent responses in nanoscale systems where assumptions fail, providing insights into dynamic beyond steady-state conditions.

Applications and Significance

In Thermoelectric Devices

Thermoelectric generators exploit the Seebeck effect to convert into electrical power by establishing a across a module composed of p-type and n-type legs connected electrically in series and thermally in parallel. The generated voltage is proportional to the temperature difference ΔT and the difference in Seebeck coefficients between the materials, enabling direct energy harvesting from sources such as or vehicle exhausts. The efficiency of these generators is fundamentally limited by the material properties encapsulated in the dimensionless ZT = (S² σ / κ) T, where S is the Seebeck coefficient, σ is electrical conductivity, κ is thermal conductivity, and T is absolute ; higher S contributes quadratically to ZT, enhancing overall performance. The maximum conversion η is given by \eta = \frac{\Delta T}{T_h} \cdot \frac{\sqrt{1 + ZT_m} - 1}{\sqrt{1 + ZT_m} + T_c / T_h}, where ΔT = T_h - T_c is the difference, T_h and T_c are the hot- and cold-side s, and ZT_m is evaluated at the mean (T_h + T_c)/2. In thermoelectric coolers, the Seebeck effect underpins the Peltier process, where an applied through the junctions of dissimilar materials induces heat absorption at one side and rejection at the other, with the heat pumping rate Q = S T I (where I is current) directly depending on the Seebeck coefficient S to achieve effective cooling without . A high Seebeck coefficient is crucial for optimizing ZT > 1, as it amplifies output and while minimizing thermal losses; for instance, commercial telluride (Bi₂Te₃)-based modules typically achieve ZT ≈ 1 at 300 K, enabling practical efficiencies around 5-7% for generation and coefficients of performance up to 0.7 for cooling near . Recent advances in the have focused on flexible thermoelectric devices using organic materials, which offer mechanical compliance for wearable applications; notable examples include achieving Seebeck coefficients exceeding 300 μV/K, for instance, certain undoped thiophene-based polymers with values up to 1215 μV/K, facilitating lightweight generators for body heat harvesting. In 2024, further enhanced Seebeck coefficients in molecular junctions for improved flexible thermoelectrics.

In Temperature Sensing

The Seebeck coefficient enables precise through , devices consisting of two dissimilar conductors joined at junctions maintained at different , generating a voltage proportional to the . The () produced is described by the equation \Delta V = \int_{T_1}^{T_2} [S_A(T) - S_B(T)] \, dT, where S_A(T) and S_B(T) are the temperature-dependent Seebeck coefficients of the two materials, T_1 is the reference (cold) junction , and T_2 is the measurement (hot) junction . This relative measurement relies on the in Seebeck coefficients between the paired materials. curves, derived from standardized tables, convert the measured voltage to , accounting for the nonlinear variation of S(T). A widely used example is the Type K , made from (90% , 10% ) and (95% , 2% manganese, 2% aluminum, 1% ), with a Seebeck coefficient of approximately 40 μV/K near . These curves ensure reliable conversion across operating ranges, typically from -200°C to 1250°C for Type K. With proper cold junction compensation—often using a precision reference sensor or ice-point equivalent—accuracies as fine as 0.1°C can be achieved over wide ranges, minimizing errors from ambient variations at the reference junction. Thermocouples find extensive use in industrial pyrometry for monitoring high-temperature processes like operations and , where ruggedness and wide ranges are essential. In medical thermometry, they measure tissue or fluid temperatures during procedures such as or , providing real-time feedback with biocompatible configurations. For extreme conditions, Type B thermocouples (platinum-30% vs. platinum-6% ) support measurements up to 1800°C intermittently, ideal for applications like processing or monitoring, though they exhibit low output at lower temperatures. Recent enhancements in the 2020s incorporate micro-thermocouples fabricated via techniques, employing thin-film materials with enhanced Seebeck coefficients (e.g., up to 48 μV/°C for NiCr/NiSi films) for compact, fast-response sensors. These enable seamless integration into platforms for wireless, remote temperature monitoring in smart systems, such as wearable health devices or environmental sensors, surpassing traditional bulk designs in and .

References

  1. [1]
    Seebeck Coefficient - an overview | ScienceDirect Topics
    In subject area: Engineering. The Seebeck coefficient is defined as the ratio of the generated Seebeck voltage to the temperature difference across a ...
  2. [2]
    Seebeck Effect - an overview | ScienceDirect Topics
    The Seebeck effect is a thermoelectric phenomenon discovered in 1821 by Thomas Johann Seebeck [KAS 97]. Locally heating a junction between two different ...
  3. [3]
  4. [4]
    Thermocouple Principles—the Seebeck Effect ... - All About Circuits
    Jul 10, 2022 · Seebeck Voltage—the Seebeck Coefficient. The open-circuit voltage produced by a temperature gradient along a wire is called Seebeck voltage and ...
  5. [5]
    Thermoelectric materials and applications for energy harvesting ...
    Thermoelectrics utilize the Seebeck effect for solid state conversion of heat to electricity. A gauge representing the performance of thermoelectric materials ...
  6. [6]
    Evolving Strategies Toward Seebeck Coefficient Enhancement
    Thermoelectric (TE) materials, interconverting electrical energy with heat energy, could improve energy utilization efficiency, optimize the energy structure, ...Introduction · Conclusion and Perspectives · Biographies · References
  7. [7]
    Transport Function Analysis of Thermoelectric Properties
    We like to think of the Seebeck coefficient as primarily a measure of the Fermi Level, relative to the band edge. This can be seen in the common, simplified ...
  8. [8]
    3.5.12 Seebeck Coefficient - IuE
    The Seebeck coefficient is defined as the ratio of the resulting voltage and the temperature difference.
  9. [9]
    [PDF] Apparatus for the high temperature measurement of the Seebeck ...
    Jun 1, 2012 · This convention ensures agreement between the Seebeck coefficient and the sign of the carriers.
  10. [10]
    [PDF] Simple Demonstration of the Seebeck Effect - ERIC
    Metals have different thermoelectric sensitivities, or Seebeck coefficients. For example, iron has a Seebeck coefficient of 19 μV/°C at 0°C, which means that ...<|control11|><|separator|>
  11. [11]
    Thermocouple Basics—Using the Seebeck Effect for Temperature ...
    Aug 14, 2022 · For example, with copper that has a Seebeck coefficient of +1.5 μV/°C at 0 °C, we can use a constantan wire with an absolute Seebeck ...
  12. [12]
    The Seebeck Coefficient | Electronics Cooling
    Nov 1, 2006 · The Seebeck coefficient is defined as the Seebeck voltage per unit temperature and is a material property. However, this is not the whole story ...
  13. [13]
    [1602.07036] On the fundamental aspect of the first Kelvin's relation ...
    Feb 21, 2016 · Kelvin's relations may be considered as cornerstones in the theory of thermoelectricity. Indeed, they gather together the three thermoelectric ...
  14. [14]
    Entropy flow in thermoelectric/thermochemical transport | MRS Bulletin
    Jul 10, 2024 · Based on linear kinetics, the Seebeck coefficient is essentially the entropy per unit charge of the transporting electronic species ...
  15. [15]
    Thermoelectricity, metallic liquidity, and magnetohydrodynamics
    Jun 24, 2024 · In 1948, Callen showed that Onsager reciprocity provided a firm basis for this Kelvin relation between Seebeck and Peltier coefficients (3).<|control11|><|separator|>
  16. [16]
    On the fundamental aspect of the first Kelvin's relation in ...
    On the other hand, the first Kelvin's relation is traditionally introduced only as a convenient mathematical relation between Seebeck and Thomson coefficients.
  17. [17]
    History of Thermoelectrics
    The Peltier coefficient is simply the Seebeck coefficient times absolute temperature. This thermodynamic derivation led Thomson to predict a third ...
  18. [18]
    Seebeck's couples | Opinion - Chemistry World
    Jul 31, 2020 · In 1821 he soldered copper and bismuth wires into a circular loop. He placed a compass needle inside the loop and then heated one of the ...<|control11|><|separator|>
  19. [19]
    Thermo-Electric Generators - Douglas Self
    Jan 23, 2023 · Thermo-electric generators convert heat directly into electricity, using the voltage generated at the junction of two different metals.
  20. [20]
    Thermoelectric couple | fisicaalmuseo.it
    Thermopiles made up of a number of couples were used in around 1830 by Italian physicists Leopoldo Nobili (1784–1835) and Macedonio Melloni (1798-1854) to ...
  21. [21]
    History of Thermoelectrics - The Morelli Research Group
    Thomas Johann Seebeck discovered in 1821 that a circuit made of two different metals could deflect the needle of a compass when the junctions of the two ...
  22. [22]
    Effect of quantum-well structures on the thermoelectric figure of merit
    Effect of quantum-well structures on the thermoelectric figure of merit. L. D. Hicks · M. S. Dresselhaus. Department of Physics, Massachusetts ...
  23. [23]
    Enhancing the Seebeck effect in Ge/Si through the combination of ...
    Nov 8, 2019 · Our results show that the Seebeck coefficient is enhanced up to ≈40% in a 3-fold CQD material with respect to 2-fold Ge/Si CQDs.
  24. [24]
    [PDF] Thermocouple: Facts and Theories
    i) Seebeck Effect: Few months later the Oersted's discovery, in 1820, of magnetic effect of electric current. (i.e., magnetism is an electrical phenomenon) ...<|control11|><|separator|>
  25. [25]
    [PDF] A Basic Guide to Thermocouple Measurements - Texas Instruments
    By joining dissimilar metals that have different Seebeck voltages at a temperature sensing junction, a thermocouple voltage (VTC) is generated.
  26. [26]
    A setup to measure the Seebeck coefficient and electrical ...
    Oct 20, 2020 · This work documents an all-in-one custom setup that allows us to measure the in-plane Seebeck coefficients and electrical conductivities of anisotropic thin ...II. SETUP · Method · Isotropic reference samples · Anisotropic films
  27. [27]
    Data analysis for Seebeck coefficient measurements - AIP Publishing
    Jun 3, 2013 · The sample Seebeck coefficient is determined in a measurement sequence as shown in Figure 3. First gradient heater 1 is powered and a ...
  28. [28]
    Errors Associated in Seebeck Coefficient Measurement for ...
    Aug 7, 2025 · ... Seebeck coefficient that includes various sources of errors from contact geometry, sensors, measurement techniques and thermocouple. In ...
  29. [29]
    Chapter 3—Thermocouple Materials - ASTM Digital Library
    The commonly used thermocouple types are identified by letter designations originally assigned by the Instrument Society of America (ISA) and adopted as an ...
  30. [30]
    [PDF] NIST Monograph 175
    NIST, originally founded as the National Bureau of Standards in 1901, works to strengthen U.S. industry's competitiveness; advance science and engineering; and ...
  31. [31]
    Precise absolute Seebeck coefficient measurement and uncertainty ...
    Jan 13, 2020 · The intrinsic properties of superconductors enable the direct determination of the absolute Seebeck coefficient at low temperature due to ...
  32. [32]
    Precise measurement of absolute Seebeck coefficient from ...
    Jun 17, 2019 · The Seebeck coefficient is an essential indicator of the potential performance of thermoelectric materials used for thermoelectric power ...INTRODUCTION · III. EXPERIMENTS · IV. RESULTS AND DISCUSSION
  33. [33]
    [PDF] nbsir-85/3133 - NIST Technical Series Publications
    Seebeck and Peltier Effects. 2. The Laws of Thermocouple Circuits. 3. Standard Thermocouple Thermometers and Their Calibration. 4. Special Problems in ...
  34. [34]
    Laser-based setup for simultaneous measurement of the Seebeck ...
    Nov 5, 2018 · It allows a simultaneous measurement of the electrical conductivity (σ) and the Seebeck coefficient (S), from room temperature up to 250 °C.
  35. [35]
    Direct Measurement of the Absolute Seebeck Coefficient Using ...
    Jul 17, 2024 · In this work, we present a simple method with graphene transistors to measure the absolute Seebeck coefficient at and above room temperature.Missing: extrapolate | Show results with:extrapolate
  36. [36]
    Thermoelectric properties of semimetals | Phys. Rev. Materials
    Sep 3, 2019 · The large Seebeck coefficient is the result of the presence of the band gap which breaks the symmetry between electrons and holes.
  37. [37]
    Semi-metals as potential thermoelectric materials | Scientific Reports
    Jun 29, 2018 · The reason for their low Seebeck coefficient is the symmetry of the density of states around the chemical potential. The number of hot electrons ...Results And Discussion · Electrical Transport · Thermal Transport
  38. [38]
    Seebeck Effect - an overview | ScienceDirect Topics
    The polarity of the Seebeck effect depends on whether the semiconductor is of the p or n type. Intrinsic semiconductors have zero Seebeck coefficients. For ...Missing: trends | Show results with:trends
  39. [39]
    Understanding the Temperature Dependence of the Seebeck ...
    The Seebeck coefficient shows more complicated temperature dependence than conventional systems, with both monotonic increases and nonmonotonic behavior.Missing: convention | Show results with:convention<|control11|><|separator|>
  40. [40]
    Room-Temperature Thermoelectric Performance of n-Type ...
    The Bi2Te2.75S0.25 sample with carrier concentration similar to Bi2Te3 (∼1.0 × 1019) have shown the Seebeck coefficient of ∼150 μV/K, much higher than the ...
  41. [41]
    Effect of Nanographene Coating on the Seebeck Coefficient ... - MDPI
    Apr 1, 2023 · Low doping of silicon of the order of 1018 cm−3 can demonstrate a Seebeck coefficient of 660 µV/K [19] while mitigating the other parameters, ...
  42. [42]
    Great Enhancement in the Seebeck Coefficient and Thermoelectric ...
    Jul 28, 2023 · The incorporation of zwitterions into PEDOT:PSS can greatly enhance the Seebeck coefficient from 21.2 to 61.6 μV K−1.Abstract · Introduction · Results and Discussion · Conclusion
  43. [43]
    Ultra-high Seebeck coefficient and low thermal conductivity of a ...
    The thermoelectric properties of this MAPbI3 single crystal were studied, showing that the Seebeck coefficient of 920 ± 91 μV K−1 almost remained unchanged ...
  44. [44]
    High thermoelectricpower factor in graphene/hBN devices - PNAS
    We further show that the Seebeck coefficient provides a direct measure of substrate-induced random potential fluctuations and that their significant reduction ...
  45. [45]
    Half-Heusler-like compounds with wide continuous compositions ...
    Jan 10, 2022 · The Seebeck coefficient of the sample with x = 0.6 reaches −165 μV/K at 470 K, suggesting the promising potential as an n-type thermoelectric ...
  46. [46]
    doped silicon from first principles via the solution of the Boltzmann ...
    Aug 25, 2016 · Figure 7 shows the Seebeck coefficient as a function of temperature at low doping. ... coefficients of n -doped silicon via the exact ...
  47. [47]
    Large Seebeck effect by charge-mobility engineering - Nature
    Jun 25, 2015 · The Mott expression relates the Seebeck coefficient to the logarithmic energy derivatives of N(ɛ) and τ(ɛ) at the Fermi energy,
  48. [48]
    [PDF] First-Principles Thermodynamic Theory of Seebeck Coefficients Yi ...
    Seebeck coefficients have been based on the Boltzmann kinetic transport theory. ... For instance, in the Cutler-Mott theory 9 the Seebeck coefficient is computed ...
  49. [49]
    Thermoelectric Response in Strongly Disordered Systems
    Nov 7, 2022 · ... Seebeck coefficient for Mott variable-range hopping in a d-dimensional system varies as S ∝ Td/(d+1), which is different from the ...Missing: extensions 2020s
  50. [50]
    Theory of the Thermoelectric Power of Semiconductors
    This paper develops the recently suggested explanation that the thermoelectric power Q is the sum of the usual electronic term Qe, resulting from the ...Missing: Seebeck | Show results with:Seebeck
  51. [51]
    Ab initio optimization of phonon drag effect for lower-temperature ...
    Our simulation explores the coupled electron phonon transport and uncovers the possibility of optimizing the phonon drag for better thermoelectrics.
  52. [52]
    Determination of Seebeck coefficient originating from phonon-drag ...
    Aug 18, 2023 · The phonon-drag effect is useful for improving the thermoelectric performance, especially the Seebeck coefficient.
  53. [53]
    Thermoelectric Properties of Materials
    The Seebeck coefficient of a narrow band system, often used for insulating transition metal and rare earth oxides is given by the Heikes formula. ... where c c is ...
  54. [54]
    First-principles thermodynamic theory of Seebeck coefficients
    Dec 3, 2018 · The Seebeck coefficient is a thermodynamic quantity, related to the change of chemical potential with temperature, and is part of the figure of ...
  55. [55]
    Understanding the Thermoelectric Transport Properties of Organic ...
    Nov 26, 2024 · The units of entropy on the other hand are. This implies that the Seebeck coefficient is a measure of the entropy per unit charge carrier, ...
  56. [56]
    Generic Seebeck effect from spin entropy - ScienceDirect.com
    May 28, 2021 · The Seebeck coefficient (α) is a key parameter determining the efficiency of useful thermoelectric devices. It measures temperature-difference- ...
  57. [57]
    [PDF] What is measured when measuring a thermoelectric coefficient? - HAL
    Mar 14, 2023 · The Onsager picture of reciprocity puts the Kelvin relation , between Peltier and Seebeck coefficients, on firm grounds.
  58. [58]
    Thermoelectric generator (TEG) technologies and applications
    Thermoelectric generators (TEGs) have demonstrated their capacity to transform thermal energy directly into electric power through the Seebeck effect.
  59. [59]
    How Thermoelectric TEG Generators Work - Tecteg
    Dec 16, 2020 · TEG generators use the Seebeck effect, creating power by heating one side and cooling the other, with power output proportional to the ...
  60. [60]
    Formula for energy conversion efficiency of thermoelectric generator ...
    Oct 6, 2020 · Thermoelectric conversion efficiency in a thermoelectric generator has been discussed using a formula3,4 ... (ZT)PR. Figure 2(a) shows a ...II. PHYSICS AND... · III. FORMULA FOR... · IV. COMPARISON WITH...
  61. [61]
    ZT – figure of merit - Linseis
    ZT = Seebeck coefficient squared * absolute temperature * electrical conductivity / specific thermal conductivity the material values ​​must be taken into ...
  62. [62]
    The central role of the Peltier coefficient in thermoelectric cooling
    Mar 27, 2014 · This paper discusses the special role that the Peltier effect plays in thermoelectric cooling. From a particular energy balance for a ...
  63. [63]
    Peltier Effect - an overview | ScienceDirect Topics
    The Peltier effect is the basis for many modern day TE refrigeration devices and the Seebeck effect is the basis for TE power generation devices (see ...
  64. [64]
    Realizing high thermoelectric performance in N-type Bi 2 Te 3 ...
    Oct 4, 2021 · Here, we report a peak ZT over 1.2 at 400 K in an n-type Bi2Te2.7Se0.3 sample doped by 5 wt. % Sb under the spark plasma sintering method. This ...
  65. [65]
  66. [66]
    Organic Thermoelectric Materials as the Waste Heat Remedy - PMC
    The compounds had high Seebeck coefficient values of over –1000 μV∙K−1. A-DCV-DPPTT without the dopant achieved a Seebeck coefficient ( α ) value of –1215 μV∙K ...
  67. [67]
    Recent advances, design guidelines, and prospects of flexible ...
    Jun 24, 2020 · The Seebeck coefficient of the film increased from 9 μV K−1 to 230 μV K−1 after deposition, leading to a higher power factor (320 μW (m K2)−1).
  68. [68]
    [PDF] Conversion of Thermocouple Voltage to Temperature
    The Seebeck coefficient is a property of the wire material. The value of ¯σ ... The ref- erence temperature is Tr = 21.23◦C and the thermocouple EMF (relative to ...
  69. [69]
    [PDF] The Basics of Thermocouples
    Dec 2, 1999 · Thermocouples produce a small. Seebeck voltage. For example, a type. K thermocouple produces about 40 µV per degree Celsius when both junc ...
  70. [70]
    [PDF] Temperature Scales, Thermocouples, and Resistance Thermometers
    1.2 Seebeck coefficients for each type are shown in Fig. 4a-1. 40. SEEBECK COEFFICIENT, µV/K. 20. 1. Type E. Type T. EP vs Au-0.07 Fe. O. 100. 200. TEMPERATURE, ...
  71. [71]
    Modern Thermocouples and a High-Resolution Delta-Sigma ADC ...
    Feb 7, 2023 · Accuracies up to ±0.1°C are now possible over a very wide -270°C to +1750°C range. To utilize all the new thermocouple capabilities, high- ...
  72. [72]
    Thermocouple Cold Junction Compensation - tc-inc.com
    Here temperature errors are typically less than 0.1°C. The use of an ice point reference, or its equivalent (however generated), is still preferable to the ...
  73. [73]
    Thermocouple Surface Pyrometers - FDA
    Aug 26, 2014 · A band or ribbon type surface pyrometer is used by processors utilizing the Steriflamme to monitor container surface temperatures.
  74. [74]
    Medical devices in-depth review: medical-grade thermocouples
    Jan 22, 2021 · Thermocouples are commonly used to treat tachycardia and atrial fibrillation, as well as to measure the tissue temperature during radiofrequency ...
  75. [75]
    Type B Thermocouple - Temperature Sensors / Probes - tc-inc.com
    It can be used continuously up to 1,600°C and intermittently up to around 1,800°C.
  76. [76]
    Temperature measurement performance of thin-film thermocouple ...
    Jan 15, 2023 · Its Seebeck coefficient was 48.13 μV/°C, and the highest temperature difference measured between cold and heat regions was 600 °C. The cutting ...Missing: IoT | Show results with:IoT
  77. [77]
    Research on NiCr/NiSi thin film thermocouple sensor for measuring ...
    Mar 18, 2025 · This study describes a method for preparing thermocouple sensor based on NiCr/NiSi thin films. Using magnetron sputtering technology.Missing: micro- IoT
  78. [78]
    Flexible thin film thermocouples: From structure, material, fabrication ...
    Flexible thin-film thermocouples (TFTCs) have been garnering interest as temperature sensors due to the advantages of being flexible, ultrathin, and ultralight.
  79. [79]
    Thermocouple-integrated resonant microcantilever for on-chip ...
    Mar 26, 2025 · We integrated two pairs of thermocouples, heating resistors, and resonant drive/detection resistors into a single microcantilever, where the ...Missing: IoT | Show results with:IoT