Fact-checked by Grok 2 weeks ago

Absolute scale

The absolute scale, also known as the absolute temperature scale, is a system of measuring temperature in which the zero point is defined as absolute zero—the theoretical lowest possible temperature at which all molecular motion ceases and thermal energy is absent. This scale avoids arbitrary reference points, such as the freezing or boiling of water, providing a universal foundation independent of specific substances or materials. The primary examples are the Kelvin scale, the standard in the International System of Units (SI), where degree sizes match those of Celsius (with 0°C equaling 273.15 K), and the Rankine scale, aligned with Fahrenheit increments (where 0°F equals 459.67°R). The development of absolute scales stemmed from 19th-century advances in thermodynamics and the study of gases. William Thomson (later Lord Kelvin) proposed the Kelvin scale in 1848, extrapolating Charles's law observations that the volume of an ideal gas approaches zero at approximately -273.15°C, establishing this as the absolute zero point. This innovation addressed limitations in earlier scales like Celsius, which could yield negative values incompatible with thermodynamic equations. The Rankine scale emerged around the same period, introduced by Scottish engineer William John Macquorn Rankine (1820–1872) as an absolute counterpart to the Fahrenheit scale, primarily for engineering applications in English-speaking countries. Absolute scales are indispensable in physics and because they enable accurate application of fundamental laws, such as the (PV = nRT), where temperature T must be positive and referenced from to reflect proportional relationships between , , and . They also underpin concepts like , which quantifies disorder and requires an absolute reference for integration in the second law of , and efficiency calculations in heat engines via the . Since the 2019 redefinition of the units, the is fixed by the (k = 1.380649 × 10⁻²³ J/K), ensuring precision without reliance on physical artifacts like the of . In practice, the scale dominates scientific and international use, while Rankine persists in some U.S. contexts, such as and power cycles.

Definition and Fundamentals

Core Definition

An absolute temperature scale is a thermometric system in which the zero point is defined as , the lowest conceivable temperature where thermal motion of particles theoretically ceases, thereby eliminating the possibility of negative temperatures and providing a fundamental reference for measurements. This approach anchors the scale to a physically meaningful origin rather than an arbitrary reference, ensuring consistency across thermodynamic contexts. The term "absolute" reflects the scale's direct proportionality to the average kinetic energy of particles, where zero corresponds to the complete cessation of molecular motion under classical thermodynamic assumptions, linking temperature intrinsically to the energetic state of matter. Unlike relative scales, such as Celsius, which employ a zero point based on conventional benchmarks like water's freezing point without thermodynamic significance, absolute scales establish their zero through principles of entropy minimization and energy conservation. On these scales, temperatures are denoted in units like the (K), where the interval represents an equal change in , facilitating precise comparisons of and work in physical processes without needing additive constants for conversions.

Absolute Zero Concept

represents the theoretical minimum achievable in a , defined as the state where the of a perfect attains its minimum value of zero, as stipulated by the third law of . This corresponds to 0 K on the scale, equivalent to approximately -273.15 °C on the scale. At this point, all classical thermal motion ceases, establishing the foundational anchor for absolute temperature scales by providing a universal reference independent of material or environmental conditions. The concept of absolute zero originated from early experimental observations of gas behavior, particularly through the work of Joseph-Louis Gay-Lussac and , who demonstrated that the volume of an is directly proportional to its at constant pressure. When these volume-temperature relationships are plotted and extrapolated linearly, they intersect the temperature axis at -273.15 °C, indicating the point where gas volume would theoretically reduce to zero—a condition unattainable in practice but revealing the finite lower bound of . This extrapolation highlighted the inadequacy of arbitrary scales like Celsius for fundamental physics, paving the way for absolute systems. Physically, absolute zero implies that particles possess only their minimal vibrational energy, with no additional thermal agitation. However, introduces a crucial nuance: due to the , particles confined in a cannot have precisely zero , resulting in a persistent that represents the of quantum systems. This residual energy ensures that even at , microscopic fluctuations remain, preventing a complete halt of all motion and underscoring the quantum limits of classical thermodynamic ideals. Experimentally, approaching involves advanced cooling techniques, such as , where atoms are slowed using precisely tuned laser beams to reach temperatures on the order of microkelvins—fractions of a millionth of a degree above zero. Despite these achievements, the third law imposes fundamental constraints, rendering exact unattainable through any finite process, as the reduction required becomes asymptotically infinite in effort. A key consequence of absolute zero as the scale's lower bound is the prohibition of negative temperatures in equilibrium thermodynamic systems, where temperature monotonically increases with energy. Nonetheless, in specialized non-equilibrium configurations, such as population inversions in media or systems, negative Kelvin temperatures can emerge; these states, first demonstrated experimentally in 1951, exhibit higher and inversion compared to positive s, effectively behaving as "hotter" than infinite temperature by facilitating to any positive-temperature system.

Historical Development

Early Theoretical Foundations

In the mid-18th century, Scottish chemist laid foundational groundwork for understanding as a quantifiable property through his investigations into heat capacities and phase changes. Black's experiments, beginning around 1757, demonstrated the existence of —the energy absorbed or released during phase transitions without a change in —which highlighted the distinction between (altering ) and . He utilized fixed points, such as the melting of ice at 0°C and the boiling of water at 100°C, to establish reproducible benchmarks for thermometric scales, thereby advancing the measurement of beyond arbitrary fluid expansions. This work underscored 's role as an intensive property independent of substance quantity, setting the stage for more precise thermodynamic concepts. Building on these ideas, early 19th-century observations of gas behavior provided empirical evidence for an absolute lower limit to temperature. In 1787, French physicist Jacques Charles noted that the volume of a gas at constant pressure expands linearly with temperature, suggesting that extrapolation to zero volume would occur at approximately -273°C, a hypothetical point of no thermal motion. This relationship was independently confirmed and quantified by French chemist Joseph-Louis Gay-Lussac in 1802, who reported that gases expand by about 1/273 of their volume at 0°C for each degree Celsius increase, reinforcing the inference of an absolute zero around -273°C where gas volume would vanish. These findings, though initially empirical, implied a natural zero point for temperature scales, free from the limitations of material-specific thermometers. The prevailing caloric theory, which posited heat as an indestructible fluid (caloric) flowing between bodies, dominated early interpretations but faced challenges from experiments linking heat to mechanical motion. In 1798, American-born physicist (Count Rumford) conducted seminal demonstrations by boring cannon barrels submerged in water, observing that frictional work continuously generated heat sufficient to boil the water, without any finite caloric source—thus suggesting heat as a form of motion rather than a conserved substance. This view gained support in the 1840s through the work of British physicist , whose paddle-wheel experiments quantified the mechanical equivalent of heat, showing that heat arises from molecular and validating the emerging . These developments reframed as the state of minimal , where molecular motion ceases. Synthesizing these advances, British physicist William Thomson (later ) proposed a thermodynamic absolute scale in 1848, grounded in Sadi Carnot's 1824 theory of rather than gas properties. Thomson argued that should be measured proportionally to the motive of heat, with zero defined as the point where Carnot becomes undefined (no heat flow), aligning closely with the -273°C extrapolated from and independent of materials. This scale provided a universal, absolute framework, marking a pivotal shift toward modern .

Modern Standardization

In the early 20th century, significant advancements solidified the concept of through Walther Nernst's heat theorem, formulated between 1906 and 1912. This theorem, later evolving into the Third Law of Thermodynamics, states that as temperature approaches , the of a perfect approaches a minimum value, typically zero, providing a theoretical foundation for unattainability of in finite steps and enabling precise calculations at low temperatures. Nernst's work built on experimental observations of chemical equilibria and specific heats, marking a pivotal shift toward a standardized absolute temperature framework. International standardization of absolute temperature scales advanced in 1954 when the 10th Conférence Générale des Poids et Mesures (CGPM) defined the scale by establishing the of water as the fundamental fixed point. Specifically, the was defined as exactly 1/273.16 of the at water's , setting a reproducible reference that decoupled the scale from arbitrary ice-point measurements while maintaining continuity with prior definitions. This resolution formalized the absolute scale within emerging metrological practices, ensuring global consistency in thermodynamic measurements. A major redefinition occurred in 2019, when the 26th CGPM revised the to be based on a fixed numerical value of the , k = 1.380649 \times 10^{-23} \, \mathrm{J/K}, rather than the water triple point. This change, effective from May 20, 2019, enhanced by linking directly to a fundamental , eliminating dependencies on material properties like water's isotopic composition and improving universality across diverse experimental conditions. The redefinition decoupled the scale from artifact-based references, allowing for more accurate realizations through methods like noise thermometry. Absolute scales, particularly the , are integral to the (SI), where the serves as the base unit for . Adopted as a base unit in the SI framework since 1967 and refined through subsequent revisions, it underpins measurements in physics, , and by providing a scale aligned with natural constants. This integration ensures in scientific data and standards worldwide. Ongoing refinements to the absolute scale's precision are managed through periodic adjustments by the Committee on Data for Science and Technology (CODATA), which evaluates experimental to update values of constants like the . For instance, the 2018 CODATA adjustment incorporated advanced measurements, achieving a relative standard uncertainty of 1.7 × 10^{-7} for k, which informed the exact value fixed in the 2019 redefinition, enhancing metrological consistency and supporting applications in quantum technologies and . These updates, released every four years, maintain the scale's accuracy without altering its definition. The most recent adjustment, the 2022 CODATA, was published in 2024.

Specific Absolute Temperature Scales

Kelvin Scale

The Kelvin scale is the (SI) base unit for , providing an absolute scale that begins at , the theoretical lowest temperature where thermal motion ceases. The unit, the (symbol: ), is defined by fixing the k at exactly $1.380\,649 \times 10^{-23} J ^{-1}, linking temperature directly to thermal energy per degree of freedom in a system. This redefinition, effective since 2019, ensures the scale's precision and universality, with the of —where , , and gas phases coexist in —corresponding exactly to 273.16 . The size of one interval is identical to one , but the Kelvin scale shifts the zero point to , such that 0 °C equals precisely 273.15 . This alignment allows seamless integration with Celsius measurements while eliminating arbitrary reference points, making it ideal for fundamental physical laws. Unlike relative scales, Kelvin temperatures cannot be negative, as all values are measured from the absolute baseline. In usage, absolute temperatures are denoted simply as K without a degree symbol (°), a convention established to differentiate from Celsius (°C) and emphasize the kelvin as a true SI unit rather than a degree interval. Early in the 20th century, the scale was sometimes called the "absolute Celsius" scale, reflecting its Celsius-sized intervals starting from absolute zero, but this term was phased out to honor William Thomson (Lord Kelvin) and standardize nomenclature. Common examples include room temperature at approximately 293 K (corresponding to 20 °C) and the boiling point of water at standard atmospheric pressure at 373.15 K. The Kelvin scale's primary advantages stem from its absolute nature, enabling consistent scientific calculations in fields like and s, where relative scales could introduce errors from negative values or shifting zeros—for instance, avoiding the need to add offsets in equations like the [PV](/page/PV) = nRT. This universality supports precise modeling of phenomena from to stellar interiors, positioning it as the standard for international and .

Rankine Scale

The Rankine scale is an absolute temperature scale that employs degree intervals identical to those of the scale, with its zero point defined at , the theoretical lowest temperature where molecular motion ceases. On this scale, 0°R corresponds to , equivalent to -459.67°F, and the freezing point of at 0°C registers as 491.67°R. The of at standard is 671.67°R. This scale ensures all temperatures are positive values, facilitating thermodynamic calculations in systems using . The was proposed in 1859 by Scottish engineer and physicist William John Macquorn Rankine in his seminal work, A Manual of the Steam Engine and Other Prime Movers, as an absolute counterpart to the scale, paralleling the scale's role for but adapted for English engineering units. Rankine's formulation arose from his foundational contributions to , emphasizing absolute temperatures for accurate analysis of heat engines and energy conversion processes. In practice, the Rankine scale finds primary application in Anglo-American engineering contexts, particularly within U.S.-based thermodynamics education, textbooks, and industries such as and (HVAC) systems, where remains prevalent for temperature measurements. It enables straightforward absolute temperature computations in imperial-unit frameworks, avoiding negative values that complicate efficiency ratios in analyses. Despite its utility in these niches, the is less common globally owing to the widespread adoption of the and the (SI), which favor the scale for international ; this often necessitates conversions in collaborative or cross-border projects.

Thermodynamic Properties

Relation to Entropy and Heat Engines

Absolute temperature scales are essential for analyzing the performance of , as they provide a reference point from that ensures physically meaningful calculations. The maximum theoretical of a heat engine operating between a hot reservoir at T_h and a cold reservoir at T_c is given by the Carnot : \eta = 1 - \frac{T_c}{T_h} where both T_h and T_c must be expressed in absolute units, such as Kelvin, to yield values between 0 and 1. Using absolute scales prevents negative efficiencies or values exceeding 100%, which would violate the laws of thermodynamics; for instance, substituting a negative temperature (as in relative scales below zero) into the formula can produce efficiencies greater than 1, implying impossible perpetual motion. In the context of the second law of thermodynamics, absolute scales underpin calculations, which quantify the directionality of energy processes in heat engines. The change in for a reversible process is defined as \Delta S = \frac{Q_\text{rev}}{T}, where Q_\text{rev} is the reversible and T is the temperature, ensuring that remains a positive, additive . This formulation aligns with the second law, which states that the total of an never decreases, guiding the analysis of irreversible losses in engines where flows from hot to cold reservoirs. Furthermore, the third law of thermodynamics establishes as the state of minimum for a perfect , providing a universal lower bound that reinforces the absolute nature of the scale in assessments. The Clausius inequality formalizes these principles for cyclic processes in heat engines, stating that for any thermodynamic cycle, \oint \frac{\delta Q}{T} \leq 0, with equality holding only for reversible cycles; this relies on absolute temperature to correctly evaluate the and confirm that no engine can achieve perfect without . Relative scales fail here as well, since temperatures below the arbitrary would invert the sign of T in the denominator, leading to nonsensical predictions about flow and cycle viability that contradict experimental observations. Thus, absolute scales ensure that thermodynamic ratios, such as those in and , remain positive and interpretable, enabling reliable design and analysis of practical engines.

Ideal Gas Law Integration

The , expressed as [PV](/page/PV) = n[R](/page/R)T, fundamentally requires the use of T in (or Rankine) for the equation to hold accurately, as the universal R is defined relative to this scale; employing relative scales like would introduce negative values or offsets, leading to erroneous predictions of or . This necessity arises because the law derives from empirical observations and theoretical principles that align with as the baseline where molecular motion theoretically ceases. In the , which underpins the , the average translational per molecule is given by \frac{3}{2} kT, where k is Boltzmann's constant and T is the absolute temperature; this energy is directly proportional to T, implying that at (T = 0), the approaches zero, reflecting the absence of thermal motion. This proportionality ensures that temperature serves as a measure of molecular independent of the gas's composition, validating the law's universality across different gases at the same absolute temperature. Charles's law, stating that the volume V of a fixed amount of gas at pressure is directly proportional to its absolute temperature (V \propto T), emerges from the by holding P and n constant, yielding V = \frac{nR}{P} T. The use of absolute temperature in this derivation allows of the linear V-versus-T relationship to , where the volume would theoretically reach zero, providing early empirical support for the concept of an absolute scale without negative volumes. Extensions to real gases, such as the \left( P + \frac{an^2}{V^2} \right) (V - nb) = nRT, incorporate corrections for intermolecular attractions (a) and molecular volume (b), yet still rely on temperature T to maintain consistency with the framework, particularly in calculating critical points and factors. This preservation of T ensures that the equation accurately models deviations from ideality while anchoring predictions to thermodynamic principles tied to . A practical illustration of absolute temperature's role occurs in defining standard temperature and pressure (STP) conditions for gas volume calculations, where STP is set at 273.15 and 1 (101.325 kPa), yielding a molar volume of approximately 22.414 L for an under the equation V_m = \frac{RT}{P}. This standardization facilitates precise comparisons in chemical and physical measurements, emphasizing the indispensability of the in quantitative gas behavior analyses.

Comparisons and Conversions

Differences from Relative Scales

Relative temperature scales, such as and , define their zero points arbitrarily based on observable phenomena, with the freezing point of set at 0°C or 32°F, respectively, allowing for negative values to represent temperatures below this reference. These scales establish intervals derived from human experiences, like the of at 100°C or 212°F under standard , prioritizing convenience for everyday applications such as reporting. In contrast, absolute scales like and Rankine anchor their zero at —the theoretical point where molecular motion ceases and reaches a minimum—ensuring a physically meaningful baseline without negative values and enabling universal applicability across scientific contexts. This structural difference makes relative scales practical for intuitive comparisons in daily life but unsuitable for thermodynamic calculations, where they can lead to invalid results due to the lack of an absolute reference. Absolute scales offer in scientific endeavors by eliminating shift errors that arise from arbitrary zeros, making them indispensable for laws like the , whereas relative scales excel in providing relatable benchmarks, such as describing 20°C as mild for human comfort. However, relative scales introduce disadvantages in -dependent fields, as their offsets complicate direct energy or interpretations. A common pitfall occurs when relative scales are mistakenly applied in thermodynamic equations, such as substituting Celsius values directly into the , which requires absolute temperature to accurately reflect proportional relationships between , , and ; this error can yield incorrect predictions of gas behavior. Absolute scales mitigate such issues by inherently aligning with physical principles. In practice, hybrid approaches are common, with used for routine measurements due to its familiarity while converting to for computations, as both share identical interval sizes to avoid scaling complications.

Conversion Formulas

The conversion between the and scales is defined by the (SI), where the is the base unit of and the degree is equal in magnitude but offset by the fixed value derived from the historical of . The to convert from to is: T(\mathrm{K}) = T(°\mathrm{C}) + 273.15 Conversely, to convert from Kelvin to Celsius: T(°\mathrm{C}) = T(\mathrm{K}) - 273.15 For the Fahrenheit and Rankine scales, used primarily in the customary FPS system, the Rankine scale is the absolute counterpart to Fahrenheit, with an offset ensuring that 0 °F corresponds to 459.67 °R, based on the definition of absolute zero at -459.67 °F. The conversion from Fahrenheit to Rankine is: T(°\mathrm{R}) = T(°\mathrm{F}) + 459.67 And from Rankine to Fahrenheit: T(°\mathrm{F}) = T(°\mathrm{R}) - 459.67 Cross-system conversions between Fahrenheit/Rankine and Celsius/Kelvin rely on both the interval equality—where the size of 1 K equals 1 °C, and 1 °F equals 5/9 K (or equivalently, 1 K = 9/5 °F = 9/5 °R)—and the zero-point offsets. This interval relationship derives from the historical calibration of the Fahrenheit scale, where the degree is defined as 1/180 of the interval between the freezing and boiling points of water, compared to 1/100 for Celsius, yielding the 9/5 factor. The direct formula to convert Fahrenheit to Kelvin is: T(\mathrm{K}) = \frac{T(°\mathrm{F}) - 32}{1.8} + 273.15 This can be derived step-by-step: first convert Fahrenheit to Celsius using T(°\mathrm{C}) = \frac{T(°\mathrm{F}) - 32}{1.8}, then add 273.15 to reach , incorporating the 1.8 factor from the (9/5 = 1.8). For precision in scientific and applications, the exact value 273.15 must be used in Celsius- conversions, as it is a defining constant in the system post-2019 redefinition; approximations like 273 are suitable only for rough estimates where sub-degree accuracy is unnecessary. Similarly, 459.67 in Fahrenheit-Rankine conversions should be exact, though 460 is sometimes approximated. A common error is neglecting these offsets entirely, treating the scales as purely interval-based without the additive shift, which leads to incorrect absolute temperatures (e.g., assuming 0 °C = 0 K). Representative examples illustrate these formulas: 100 °C converts to $100 + 273.15 = 373.15 K; 212 °F (boiling point of ) converts to $212 + 459.67 = 671.67 °R; and cross-conversion of 98.6 °F (normal body temperature) yields \frac{98.6 - 32}{1.8} + 273.15 = 37 + 273.15 = 310.15 K.

Applications

Scientific Research

In physics, absolute temperature scales are crucial for describing quantum phenomena near . Bose-Einstein condensates, a predicted by quantum statistics, form when a dilute gas of bosons is cooled to temperatures on the order of nanokelvins, enabling macroscopic quantum coherence and applications in precision sensing. In cosmology, the radiation, a relic of the , maintains a uniform blackbody of approximately 2.726 K, providing a fundamental reference for the universe's thermal history and expansion. In chemistry, absolute scales underpin the temperature dependence of reaction kinetics and equilibria. The Arrhenius equation models reaction rates as k = A e^{-E_a / RT}, where T is the absolute in , R is the , A is the , and E_a is the , allowing prediction of rate constants across temperature ranges without negative values that would arise from relative scales. Similarly, equilibrium constants vary with absolute according to the van't Hoff equation, \ln(K_2 / K_1) = -\Delta H^\circ / R (1/T_2 - 1/T_1), which integrates thermodynamic principles to quantify shifts in chemical balance under changing conditions. Astronomy relies on absolute temperatures to characterize stellar atmospheres and . The Sun's effective surface is 5772 , determined from its fitting solar measurements, which informs models of solar output and helioseismology. follows , \lambda_{\max} T = b, where b \approx 2.898 \times 10^{-3} m· is Wien's constant and T is absolute , enabling estimation of stellar temperatures from peak emission wavelengths observed in . Cryogenics highlights the precision required at low absolute temperatures for material properties. Superconductivity in mercury emerges abruptly at 4.2 K, where electrical resistance drops to zero, a discovery that revealed quantum pairing mechanisms essential for understanding low-temperature phase transitions. This threshold underscores why absolute scales are indispensable, as relative scales would misrepresent the proximity to zero and associated quantum effects. Recent advances in ultracold atom research have pushed temperatures to 38 picokelvins (3.8 × 10^{-11} K) in Bose-Einstein condensates, achieved through evaporative cooling in optical traps, facilitating studies of quantum many-body dynamics and simulations of condensed matter systems with unprecedented control.

Engineering and Industry

In engineering applications, absolute temperature scales such as the Kelvin are essential for calculating the efficiency of thermodynamic cycles in power plants, where steam turbines operate at inlet temperatures around 800 K to maximize energy extraction. For instance, the available energy in steam is directly proportional to the absolute inlet temperature for given pressures, enabling precise determination of thermal efficiencies that can reach up to 45% in supercritical plants operating at approximately 873 K. In HVAC and refrigeration systems, absolute scales underpin the design of vapor-compression cycles, where the coefficient of performance (COP) for a Carnot refrigerator is given by COP = T_c / (T_h - T_c), with T_c and T_h as the absolute cold and hot reservoir temperatures in , respectively. This formulation ensures accurate compressor ratio selections and system sizing, as relative scales would yield negative or invalid values, leading to errors in predicting cooling capacities and energy consumption. For example, units maintaining food storage at around 273 rely on these calculations to achieve COP values of 2-4 under conditions. Materials science employs absolute temperatures in heat treatment processes to precisely control phase transformations and prevent errors such as unintended formation in steels. The Larson-Miller parameter, P = T (C + log_{10} t_r), where T is in , C is a material constant (often 20 for steels), and t_r is rupture time in hours, correlates time and to predict life and optimize annealing or tempering schedules at temperatures like 873-1273 K. This approach is critical for components in high-stress environments, ensuring microstructural stability without relying on arbitrary relative scales. Aerospace engineering utilizes absolute scales for rocket propulsion, where combustion chamber temperatures exceed 3000 , dictating nozzle design and specific impulse calculations via the ideal rocket equation involving exhaust velocity proportional to sqrt(T_c / M), with T_c in . Re-entry heat shields must withstand peak surface temperatures up to 2000-3000 , where absolute temperature governs ablation rates and thermal protection modeling to safeguard payloads during atmospheric friction. In the energy sector, absolute scales are integral to (LNG) storage and handling, maintained at approximately 111 K to keep in liquid form at , minimizing boil-off rates calculated using thermodynamic properties at reference. Standards for safety and efficiency, such as those from the National Institute of Standards and Technology, mandate absolute for predictions and cryogenic vessel integrity assessments, preventing risks like over-pressurization during fluctuations.

References

  1. [1]
    Lord Kelvin
    The Kelvin scale, however, is an absolute scale. Zero is not arbitrarily defined; it is the lowest possible temperature that can be achieved. Thus ...
  2. [2]
    Absolute Temperature Scale - an overview | ScienceDirect Topics
    The absolute temperature scale is defined as a temperature measurement system that is independent of the physical properties of any specific substance, ...
  3. [3]
    Rankine scale – Knowledge and References - Taylor & Francis
    The Rankine scale (denoted by °R) is an absolute temperature scale whose degree is 1.8 times smaller than the Kelvin degree. It is thus defined by Eqs. 2.4 and ...
  4. [4]
    What Is Absolute Temperature? Definition and Scales - Science Notes
    May 4, 2021 · Except for absolute zero, every temperature on an absolute scale is a positive value. This is the main reason many equations require absolute ...
  5. [5]
    Absolute temperature scale - (Physical Chemistry I) - Fiveable
    The use of the absolute temperature scale is crucial in defining various thermodynamic concepts such as entropy and enthalpy.
  6. [6]
  7. [7]
    Kelvin: Introduction | NIST
    May 14, 2018 · The absolute, thermodynamic temperature of an object provides information on how much average energy of motion (kinetic energy) its atoms and ...
  8. [8]
    Entropy and the 2nd & 3rd Laws of Thermodynamics
    Third law: The entropy of a perfect crystal is zero when the temperature of the crystal is equal to absolute zero (0 K). The crystal must be perfect, or else ...
  9. [9]
    How Low Can Temperature Go? Lord Kelvin and the Science of ...
    Jun 26, 2024 · The third law of thermodynamics states that while one can never reach absolute zero, there is no fundamental limit as to how close you can get.
  10. [10]
    10.3: The Gas Laws - Chemistry LibreTexts
    Jul 20, 2016 · Plots of the volume of gases versus temperature extrapolate to zero volume at −273.15°C, which is absolute zero (0 K), the lowest temperature ...<|separator|>
  11. [11]
    How do scientists cool objects to absolute zero?
    May 31, 2013 · Every quantum system has a non-zero ground state energy which is the lowest possible state. This is nature's way of enforcing the Heisenberg ...
  12. [12]
    [PDF] Lecture 13: March 10, 2021 Heisenberg's Uncertainty Principle
    (I) Zero Point Energy: According to Heisenberg uncertainty principle, even at absolute zero temperature, particles cannot be at rest when confined in a finite ...
  13. [13]
    [PDF] Laser-Cooling Atoms to Within micro-Kelvins of Absolute Zero
    Using the interaction of the atoms of a gas with light beams one can control their degrees of freedom and reduce dramatically their thermal motion in order to ...Missing: unattainability third
  14. [14]
    [PDF] thermodynamics and superfluidity of a strongly interacting fermi gas
    The limit is known as the Doppler cooling limit and is roughly 140 µK for 6Li. While 140 millionths of a degree above absolute zero seems like an impressively.
  15. [15]
    [PDF] Consistent thermostatistics forbids negative absolute temperatures
    Dec 8, 2013 · negative absolute temperature were first reported in 1951 by Purcell and Pound5 in seminal work on the population inversion in nuclear spin ...
  16. [16]
    Joseph Black and Latent Heat - American Physical Society
    The latent heat that Black discovered greatly slows the melting of snow and ice. He gave the first account of this work on April 23, 1762 at the University of ...Missing: fixed | Show results with:fixed
  17. [17]
    Joseph Black, carbon dioxide, latent heat, and the beginnings of the ...
    Black found that the water warmed to 40° in half an hour. However, the ice took ten and a half hours to obtain the same temperature, or in other words 21 times ...
  18. [18]
    Sketching The Life Of Joseph Black - C&EN
    Oct 22, 2012 · His theory of latent heat, a change in a system's heat content at constant temperature, marks the beginning of the study of thermodynamics.
  19. [19]
    Ideal Gas Law
    According to Charles's law, which is another experimental result that was obtained by Jacques Charles in 1787, the volume of an ideal gas is proportional to ...
  20. [20]
    Joseph Louis Gay-Lussac | Science History Institute
    From his first major program of research in 1801–1802, he concluded that equal volumes of all gases expand equally with the same increase in temperature: this ...
  21. [21]
    What Gay-Lussac didn't tell us | American Journal of Physics
    He determined that when any of several gases is heated from 0 to 100 ° C ⁠, its volume increases by 37.5%. This result is 2.5% higher than modern values, and ...
  22. [22]
    Count Rumford on Heat - chemteam.info
    A very considerable quantity of heat may be excited in the friction of two metallic surfaces and given off in a constant stream or flux, in all directions.
  23. [23]
    Heat, work and subtle fluids: a commentary on Joule (1850) 'On the ...
    Apr 13, 2015 · Joule's experiments in the 1840s. James Joule had no doubt that the caloric theory was fundamentally flawed. He arrived at this conclusion ...
  24. [24]
    Lord Kelvin | On an Absolute Thermometric Scale...
    The principle according to which the scale of the air-thermometer is graduated, is simply that equal absolute expansions of the mass of air or gas in the ...<|separator|>
  25. [25]
    Resolution 3 of the 10th CGPM (1954) - BIPM
    The 10th Conférence Générale des Poids et Mesures decides to define the thermodynamic temperature scale by choosing the triple point of water as the fundamental ...
  26. [26]
    Resolution 1 of the 26th CGPM (2018) - BIPM
    The kelvin, symbol K, is the SI unit of thermodynamic temperature. It is defined by taking the fixed numerical value of the Boltzmann constant k to be 1.380 649 ...<|separator|>
  27. [27]
    Kelvin: Present Realization | NIST
    May 15, 2018 · In May of 2019, the kelvin was redefined in terms of a fundamental quantity in nature known as the Bolzmann constant.Today, the kelvin is ...
  28. [28]
    - kelvin - BIPM
    SI base unit: kelvin (K). The kelvin, symbol K, is the SI unit of thermodynamic temperature. It is defined by taking the fixed numerical value of the ...
  29. [29]
    CODATA Recommended Values of the Fundamental Physical ...
    Jan 10, 2022 · We report the 2018 self-consistent values of constants and conversion factors of physics and chemistry recommended by the Committee on Data ...
  30. [30]
    SI Units – Temperature | NIST
    The kelvin (K) is defined by taking the fixed numerical value of the Boltzmann constant k to be 1.380 649 ×10 −23 when expressed in the unit JK −1.
  31. [31]
    temperature | Communication Standards - NREL
    Nov 18, 2024 · Use a degree symbol (°) with temperatures expressed in the Celsius and Fahrenheit scales but not with kelvins (just use K).
  32. [32]
    Heat Engines: the Carnot Cycle - Galileo
    efficiency =WQH=QH−QCQH. This is the answer, but it's not particularly useful: measuring heat flow, especially the waste heat, ...
  33. [33]
    It's a negative on negative absolute temperatures
    Dec 20, 2013 · If one inserts such negative temperatures into an equation for the efficiency of a heat engine, known as Carnot's formula, then one can obtain ...
  34. [34]
    Second Law of Thermodynamics
    The change in entropy (delta S) is equal to the heat transfer (delta Q) divided by the temperature (T). For a given physical process, the entropy of the ...Missing: Q_rev / absolute
  35. [35]
  36. [36]
  37. [37]
    Equation Of State (Ideal Gas) | Glenn Research Center - NASA
    Jul 7, 2025 · Also be aware that the temperature given in the equation of state must be an absolute temperature that begins at absolute zero. In the metric ...
  38. [38]
    Ideal Gas Behavior - StatPearls - NCBI Bookshelf
    May 6, 2024 · The temperature value in the ideal gas law must be in absolute units, either Rankine (°R) or Kelvin (K), to prevent the right-hand side of the ...Introduction · Function · Issues of Concern · Clinical Significance
  39. [39]
    Ideal Gas Law - HyperPhysics Concepts
    An ideal gas can be characterized by three state variables: absolute pressure (P), volume (V), and absolute temperature (T).
  40. [40]
    Temperature and Kinetic Energy - Chemistry 301
    The key to the kinetic theory of gases is the notion that the average kinetic energy ( ) for the particles in a gas sample is proportional to the temperature. ...
  41. [41]
    Kinetic Theory of Gases: A Brief Review - Galileo
    ... gas law that the average molecular kinetic energy is proportional to the absolute temperature, . Boltzmann's constant k = 1.38.10-23 joules/K. Maxwell finds ...
  42. [42]
    Kinetic theory - Physics
    Dec 1, 1999 · The absolute temperature of an ideal gas is proportional to the average kinetic energy per gas molecule.
  43. [43]
    Charles' Law (Jacques-Alexandre-César Charles)
    Absolute zero can therefore be more accurately defined as the temperature at which the pressure and the volume of a gas extrapolate to zero. When data obtained ...
  44. [44]
    Deviations from the Ideal Gas Law
    The van der Waals equation predicts that the pressure will have to reach 1620 atm to achieve the same results. The van der Waals equation gives results that ...
  45. [45]
    The van der Waals Equation | PNG 520
    P = absolute pressure v = molar volume. T = absolute temperature. R = universal gas constant. Equation (7.11b) demonstrates that vdW EOS is explicit in pressure ...
  46. [46]
    Standard Temperature and Pressure (STP) for a gas is defined by ...
    Standard Temperature and Pressure (STP) for a gas is defined by IUPAC as 0° C (273.15 K), 100 kPa. This differs from an earlier definition that used 1 atm (101 ...
  47. [47]
    Methane Gas Volume Expansion Ratios and Ideal Gas Deviation ...
    Jan 6, 2006 · Standard temperature and pressure (STP) are defined as T = 0°C or 273.15 K, P=1 atmosphere or 0.101325 MPa. At STP, Vm = 22,346 cm3/mol = VSTP.
  48. [48]
    Temperature
    The Celsius (oC) and Fahrenheit (oF) scales measure relative temperatures. These scales compare the temperature being measured with arbitrary standards such as ...
  49. [49]
    1.2 Thermometers and Temperature Scales - UCF Pressbooks
    Thermometers measure temperature according to well-defined scales of measurement. The three most common temperature scales are Fahrenheit, Celsius, and Kelvin.
  50. [50]
    [PDF] The Ideal Gas Law and Absolute Zero - UMD Physics
    The absolute temperature of an ideal gas is proportional to the kinetic ... If you are measuring temperature on the Celsius scale, then the ideal gas law must be.<|separator|>
  51. [51]
    Kinetic theory of gases - Physics
    Jun 23, 1998 · The absolute temperature of an ideal gas is proportional to the average kinetic energy per gas molecule.
  52. [52]
    [PDF] SI Brochure - 9th ed./version 3.02 - BIPM
    May 20, 2019 · 10th CGPM, 1954: degree Kelvin adopted as a base unit. 163. 13th CGPM, 1967/68: decides formal definition of the kelvin, symbol K. 169. CIPM ...
  53. [53]
    [PDF] The International System of Units (SI) – Conversion Factors for ...
    The formula for converting degree Rankine to thermodynamic temperature is: TK = TR /1.8. A temperature interval may be expressed in SI either in kelvin or in ...
  54. [54]
    NIST Guide to the SI, Appendix B.8: Factors for Units Listed ...
    Feb 1, 2016 · Appendix B.8 lists factors for units listed alphabetically, generally not for use in NIST publications, except for italicized units.
  55. [55]
    2.3.2: Temperature conversions - Chemistry LibreTexts
    Jul 20, 2023 · Learn about the various temperature scales that are commonly used in chemistry. Define density and use it as a conversion factor. There are ...
  56. [56]
    Observation of Bose-Einstein Condensation in a Dilute Atomic Vapor
    The condensate fraction first appeared near a temperature of 170 nanokelvin and a number density of 2.5 × 1012 per cubic centimeter and could be preserved for ...
  57. [57]
  58. [58]
    The origin and status of the Arrhenius equation - ACS Publications
    Contributions of Svante Arrhenius to reaction kinetics. ACS Publications. © American Chemical Society and Division of Chemical Education, Inc ...
  59. [59]
    Hoff Equation - an overview | ScienceDirect Topics
    The van't Hoff equation depicts the relationship between the solubility of a specific solute and the temperature of solution.
  60. [60]
    [PDF] 2. The Sun as a star
    Effective temperature ( e. T ). 5772 K. 3 76. 10 . K. 1 arc sec. 726 km. 7 86. 10 . cm. Page 4. Sun's age. The of the Sun is estimated from meteorites, the ...
  61. [61]
    (PDF) WIEN 1893 FINAL - ResearchGate
    Aug 9, 2023 · PDF | The upper limits of the wavelength, which come from the heat radiation of solid bodies, follow from the second main law of thermodynamics;
  62. [62]
    [PDF] Onnes 1911 - Physics
    KAMERLINGH ONNES. -"Further experiments with Liquid. Helium. G. On the Electrical Resistance of Pure Metals, etc. VI. On the Sudden Change ...
  63. [63]
    New record set for lowest temperature—38 picokelvins - Phys.org
    Oct 13, 2021 · New record set for lowest temperature—38 picokelvins ... A team of researchers affiliated with several institutions in Germany and two in France ...
  64. [64]
    Steam Turbines - an overview | ScienceDirect Topics
    The energy available (Btu/lb) for given inlet and exhaust pressure is proportional to the absolute inlet steam temperature. Frequently economics justify the ...
  65. [65]
    Industrial steam turbines - Siemens Energy
    The inlet steam pressure is 103 to 160 bar(a) and the inlet steam temperature is 400 to 565 °C. Download our information package about our industrial steam ...
  66. [66]
    6.4 Carnot cycles – Introduction to Engineering Thermodynamics
    The thermodynamic temperature scale, or absolute temperature scale is defined so that ( Q L Q H ) r e v = T L T H . In summary, the thermal efficiency of an ...
  67. [67]
    [PDF] Thermodynamic Cycles and the Second Law Absolute Temperature ...
    Refrigeration cycle coefficient of performance. COP1&2 ... • The Kelvin absolute temperature scale is defined as one where the function is,.
  68. [68]
    Propellant Combustion Charts - Rocket and Space Technology
    ... chamber temperature (Tc) in a rocket engine. Below we see a graph of ... temperature of approximately 3,545 Kelvin. Gas Molecular Weight. The exhaust ...
  69. [69]
    Protecting space crew from extreme heat during reentry
    May 9, 2018 · Atomistic simulations of the evolving pyrolysis fragments at 1,200 to 2,000 degrees Kelvin (1,700 to “The material we used for the model was ...
  70. [70]
    [PDF] LNG measurement - NIST Technical Series Publications
    The temperature interval one degree Celsius equals one kelvin exactly. Celsius temperature (t) is related to thermodynamic temperature (T) by the equation ...
  71. [71]
    Measuring the LNG process with confidence - LNG Industry
    Oct 2, 2023 · This is done by cooling the gas to 111 Kelvin. The energy-intensive process requires accurate and reliable flow measurement to ensure that the ...