A word for unit of heat

We are all familiar with the term heat, and we experience the effect of heat almost every day in our lives. As we are familiar with the topic, if we look at the science behind it, heat is the form of energy that changes the temperature of any substance. Heat involves the transfer of energy from an object or an energy source to another medium or an object. Heat is represented by the symbol Q and the heat formula is given as;

(begin{array}{l}Q = mC Delta Tend{array} )

Where Q= heat, m = mass of the body, C = specific heat and Δ T = temperature difference.

While we have learned about what heat is in quick succession, it is also important to learn about how it is measured and the unit of heat. Let’s discuss them below.

Generally, all forms of energy are measured in terms of joules in the SI system. Notably, heat is a form of energy, and therefore the SI unit of heat is also joules (J) which are defined as the amount of energy needed to raise the temperature of a given mass by one degree. Usually, 4.184 joules of heat energy is necessary to increase the temperature of a unit weight (say 1 g) of water from 0 degrees to 1-degree celsius.

Other Heat Units

In the CGS system, heat is expressed in the unit of calories which is further said to be the heat energy needed to increase the temperature of 1 gm of clean water by one degree Celsius. Sometimes kilocalorie (kcal) is referred to as a unit of heat where 1 kcal = 1000 cal.

The British thermal unit (BTU), part of the imperial system, is also used to measure or calculate heat.

Units of Heat
Calorie 1 cal 4.184 J
Joules 1 J 0.000239006 kcal / 0.000947817Btu
BTU 1 Btu 1055.06 J

Related articles:

What is heat? Why do we experience it? How does it travel?

Stay tuned to BYJU’S and Fall in Love with Learning!

‘A UNIT OF HEAT’ is a 11 letter
Phrase
starting with A and ending with T

Crossword answers for A UNIT OF HEAT

Clue Answer

A UNIT OF HEAT
(5)

THERM

Synonyms for THERM

3 letter words

7 letter words

Thanks for visiting The Crossword Solver «A unit of heat».

We’ve listed any clues from our database that match your search for «A unit of heat». There will also be a
list of synonyms for your answer.
The answers have been arranged depending on the number of characters so that they’re easy to
find.

If a particular answer is generating a lot of interest on the site today, it may be highlighted in
orange.

If your word «A unit of heat» has any anagrams, you can find them with our anagram solver or at this
site.

We hope that you find the site useful.

Regards, The Crossword Solver Team

More clues you might be interested in

  1. duration
  2. percolation
  3. loll
  4. sane
  5. edible root
  6. flighty
  7. sound from a nest
  8. jocularity
  9. dangerous bear
  10. mayhem
  11. supreme being
  12. intolerable
  13. ___ xtra (dr pepper alternative)
  14. routine
  15. russian port
  16. russian wolfhound
  17. blood of the gods
  18. improvise
  19. hold on tightly
  20. catch sight of
  21. sketch show
  22. utopia
  23. commissioned soldier
  24. soothes
  25. maritime
  26. moderately warm
  27. protons, neutrons, and electrons combined
  28. in the middle
  29. grand total
  30. time from sunset to sunrise

Generally, heat is anything that provides warmth but scientifically, heat is the flow of energy from a warmer object to a cooler object in comparison till both the objects attain equilibrium. Every matter on earth has some amount of heat energy stored in it. Heat energy flows due to the difference in temperature of the two bodies. In this article, students will learn about the units and conversions of heat energy but first let’s look at a few terms, definitions and concepts.

Heat: Scientifically, heat is defined as the energy that is spontaneously transferred from one object to another due to differences in temperatures. Heat transfer occurs until the bodies attain equilibrium.

Temperature: Temperature is defined as the kinetic energy of molecules of a body.

Internal Energy: The total energy of all the molecules of a body is the internal energy within the object.

Specific Heat: Specific heat, also known as heat capacity, is the amount of energy required to produce a unit change in its temperature.

Difference between Temperature, Heat, and Internal Energy

  • Temperature is the kinetic energy of the molecules of a body. The average kinetic energy of individual molecules is termed temperature.

  • The total energy of all the molecules is the internal energy within the object. Internal energy is an extensive property.

  • Heat is defined as the energy that is spontaneously transferred from one body to another due to its temperature difference.

For example, if a 5 kg of steel, at 100°C, is placed in contact with a 500 kg of steel at 20°C, heat flows from the cube at 300°C to the cube at 20°C, even though, the internal energy of the 20°C cubes is much greater because there is so much more of it. Mathematically heat can be expressed as:

[C=frac{Q}{mtimes Delta T}]

Where m = mass of the body,

C = specific heat,

Δ T = temperature difference.

Q = heat

SI Unit of Heat:

As all the energy is represented in Joules (J), therefore, heat is also represented in Joules. Hence, the SI unit of heat is Joules. Joules can be defined as the amount of energy required to raise the temperature of a given mass by one degree. To increase the temperature of one unit weight of water by one degree, we require 4.184 joules of heat.

Other Heat Units:

Other heat units are:

  • BTU

  • Calorie

  • Joules

BTU:

BTU is a British thermal unit. It is the amount of energy required to raise the temperature of one pound of water by 10 F at sea level.

Conversion:

1 BTU = 1055.06 J = 2.931 x 10-4 kWh = 0.252 kcal = 778.16 ft lbf = 1.055 x 1010 ergs = 252 cal = 0.293 watt-hours

Calorie:

The amount of energy required to raise the temperature of one gram of water by 10 C.

Conversion: 

1 kcal = 4186.8 J = 426.9 kp m = 1.163 x 10-3 kWh = 3.088 ft lbf = 3.9683 BTU = 1000 cal

Joule:

Joule is the SI unit of heat. 

Conversion:

1 J = 0.1020 kpm = 2.778 x 10-7 kcal = 0.7376 ft lb = 1 kg m2 / s2 = 1 watt second = 1 Nm = 9.478 x 10-4 BTU

Conversion Table:

Units of Heat

Calorie

1 cal

4184 J

Joules

1 J

0.000239006 kcal / 0.000947817 Btu

BTU

1 Btu

1055.06 J

Temperature Conversion :

Celsius to Kelvin

K = C+273 

For Example:

1000C = 100+273 = 373 K

Kelvin to Celsius

C = K – 273

For Example:

273 K = 273 – 273 = 00C

Celsius To Fahrenheit

0F = 9/5 (0C ) + 32

Kelvin to Fahrenheit

0F = 9/5 (K-273) +32

Fahrenheit to Celsius

0C = 5/9 (0F-32)

Fahrenheit to Kelvin

K = 5/9 (0F-32) + 273

Example 1:  An electric kettle contains 1.5 kg of water. The specific heat capacity of water is 4180 J kg-1 K-1. Calculate the amount of energy required to raise the temperature of the water from 15 0C to 100 0C.

Solution: 

Given:

Specific heat (C) = 4180 J kg-1 K-1

T1 = 15 0C = 15+273 = 288 K

T2 = 100 0C = 100+273 = 373 K

m = 1.5 kg

Q = m x Δ T x C

Q = 1.5 x 4180 x (373-288)

    = 533 kJ 

Example 2:  Calculate the energy needed to raise the temperature of the water from 20 0C to 90 0C. 

Solution: Q = mcΔθ

= (0.7) (4200) (90-20) = 205.8 kJ

Methods of Transfer of Heat Energy

Convection:  Transfer of heat energy via fluids. When fluids get heated, they form vapours and rise higher up in the environment.

Conduction: Transfer of heat energy through direct contact between two bodies. Such a method of transfer of heat is generally observed in solids.

Radiation: Radiation from hot objects (such as the sun) warms up the air in all directions which are absorbed by molecules all around.

Try Yourself:

  1. Calculate heat required to evaporate 1kg of water at the atmospheric pressure (p = 1.0133 bar) also at the temperature of 100°C.

  2. Calculate heat required to evaporate 1 kg of feed water at the pressure of 6 MPa (p = 60 bar) and the temperature of 275.6°C.

  3. Calculate the specific heat of a 100 kg mass of water if the temperature changes from 150 C to 1000 C. Heat required is 130 BTU.

  4. Calculate the heat required to raise the temperature of 60-milligram mass from 22 K to 273 K. Specific heat given 223 J/K.

  5. Calculate the specific heat of a 20 dkg mass of water if the temperature changes from 150 C to 260 C. Heat required is 137 BTU.

  6. Calculate the heat required to raise the temperature of 200 kg mass from 2320 C to 300 K. Specific heat given 203 J/K.

  7. Calculate the specific heat of a 1000 kg mass of water if the temperature changes from 15 K to 100 K. Assume the rest data.

  8. Calculate the heat required to raise the temperature of 29 kg mass from 220 C to 273 K. Assume the rest data.

  9. Calculate the specific heat of a 20 kg mass of water if the temperature changes from 1500 C to 1000 C. Heat required is 130 cal.

  10. Calculate the heat required to raise the temperature of 505 kg mass from 320 C to 273 K. Specific heat given 320 J/K.

  11. Explain how heat is transferred in the body?

  12.  Name the other methods for transferring heat.

  13.  What is the SI unit of heat?

Definitions of heat unit

  1. noun

    a unit of measurement for work

    synonyms:

    energy unit, work unit

    see moresee less

    types:

    show 14 types…
    hide 14 types…
    erg

    a cgs unit of work or energy; the work done by a force of one dyne acting over a distance of one centimeter

    eV, electron volt

    a unit of energy equal to the work done by an electron accelerated through a potential difference of 1 volt

    J, joule, watt second

    a unit of electrical energy equal to the work done when a current of one ampere passes through a resistance of one ohm for one second

    calorie, gram calorie, small calorie

    unit of heat defined as the quantity of heat required to raise the temperature of 1 gram of water by 1 degree centigrade at atmospheric pressure

    Calorie, kilocalorie, kilogram calorie, large calorie, nutritionist’s calorie

    a unit of heat equal to the amount of heat required to raise the temperature of one kilogram of water by one degree at one atmosphere pressure; used by nutritionists to characterize the energy-producing potential in food

    B.Th.U., BTU, British thermal unit

    a unit of heat equal to the amount of heat required to raise one pound of water one degree Fahrenheit at one atmosphere pressure; equivalent to 251.997 calories

    therm

    a unit of heat equal to 100,000 British thermal units

    watt-hour

    a unit of energy equal to the power of one watt operating for one hour

    B.T.U., Board of Trade unit, kW-hr, kilowatt hour

    a unit of energy equal to the work done by a power of 1000 watts operating for one hour

    foot-pound

    a unit of work equal to a force of one pound moving through a distance of one foot

    foot-ton

    2240 foot-pounds

    foot-poundal

    a unit of work equal to a force of one poundal moving through a distance of one foot

    horsepower-hour

    a unit of work equal to the work done by one horsepower in one hour

    kilogram-meter

    a unit of work equal to the work done by a one kilogram force operating through a distance of one meter

    type of:

    unit, unit of measurement

    any division of quantity accepted as a standard of measurement or exchange

DISCLAIMER: These example sentences appear in various news sources and books to reflect the usage of the word ‘heat unit’.
Views expressed in the examples do not represent the opinion of Vocabulary.com or its editors.
Send us feedback

EDITOR’S CHOICE

Look up heat unit for the last time

Close your vocabulary gaps with personalized learning that focuses on teaching the
words you need to know.

VocabTrainer - Vocabulary.com's Vocabulary Trainer

Sign up now (it’s free!)

Whether you’re a teacher or a learner, Vocabulary.com can put you or your class on the path to systematic vocabulary improvement.

Get started

Смотреть что такое «unit of heat» в других словарях:

  • Unit of heat — Unit U nit, n. [Abbrev. from unity.] 1. A single thing or person. [1913 Webster] 2. (Arith.) The least whole number; one. [1913 Webster] Units are the integral parts of any large number. I. Watts. [1913 Webster] 3. A gold coin of the reign of… …   The Collaborative International Dictionary of English

  • Unit of heat — Heat Heat (h[=e]t), n. [OE. hete, h[ae]te, AS. h[=ae]tu, h[=ae]to, fr. h[=a]t hot; akin to OHG. heizi heat, Dan. hede, Sw. hetta. See {Hot}.] 1. A force in nature which is recognized in various effects, but especially in the phenomena of fusion… …   The Collaborative International Dictionary of English

  • unit of heat conductance — šiluminio laidžio vienetas statusas T sritis Standartizacija ir metrologija apibrėžtis Šiluminio laidžio matavimo vienetas. atitikmenys: angl. unit of heat conductance vok. Einheit der Wärmeleitwert, f rus. единица теплопроводности, f pranc.… …   Penkiakalbis aiškinamasis metrologijos terminų žodynas

  • unit of heat conduction — šiluminio laidžio vienetas statusas T sritis fizika atitikmenys: angl. unit of heat conduction vok. Einheit der Wärmeleitung, f rus. единица теплопроводности, f pranc. unité de conductance thermique, f …   Fizikos terminų žodynas

  • Unit — U nit, n. [Abbrev. from unity.] 1. A single thing or person. [1913 Webster] 2. (Arith.) The least whole number; one. [1913 Webster] Units are the integral parts of any large number. I. Watts. [1913 Webster] 3. A gold coin of the reign of James I …   The Collaborative International Dictionary of English

  • Unit deme — Unit U nit, n. [Abbrev. from unity.] 1. A single thing or person. [1913 Webster] 2. (Arith.) The least whole number; one. [1913 Webster] Units are the integral parts of any large number. I. Watts. [1913 Webster] 3. A gold coin of the reign of… …   The Collaborative International Dictionary of English

  • Unit jar — Unit U nit, n. [Abbrev. from unity.] 1. A single thing or person. [1913 Webster] 2. (Arith.) The least whole number; one. [1913 Webster] Units are the integral parts of any large number. I. Watts. [1913 Webster] 3. A gold coin of the reign of… …   The Collaborative International Dictionary of English

  • Unit of illumination — Unit U nit, n. [Abbrev. from unity.] 1. A single thing or person. [1913 Webster] 2. (Arith.) The least whole number; one. [1913 Webster] Units are the integral parts of any large number. I. Watts. [1913 Webster] 3. A gold coin of the reign of… …   The Collaborative International Dictionary of English

  • Unit of measure — Unit U nit, n. [Abbrev. from unity.] 1. A single thing or person. [1913 Webster] 2. (Arith.) The least whole number; one. [1913 Webster] Units are the integral parts of any large number. I. Watts. [1913 Webster] 3. A gold coin of the reign of… …   The Collaborative International Dictionary of English

  • Unit of power — Unit U nit, n. [Abbrev. from unity.] 1. A single thing or person. [1913 Webster] 2. (Arith.) The least whole number; one. [1913 Webster] Units are the integral parts of any large number. I. Watts. [1913 Webster] 3. A gold coin of the reign of… …   The Collaborative International Dictionary of English

  • Unit of resistance — Unit U nit, n. [Abbrev. from unity.] 1. A single thing or person. [1913 Webster] 2. (Arith.) The least whole number; one. [1913 Webster] Units are the integral parts of any large number. I. Watts. [1913 Webster] 3. A gold coin of the reign of… …   The Collaborative International Dictionary of English

Crossword Clue Last Updated: 18/02/2023

Below are possible answers for the crossword clue Unit of heat.

7 letter answer(s) to unit of heat

CALORIE

  1. a unit of heat equal to the amount of heat required to raise the temperature of one kilogram of water by one degree at one atmosphere pressure; used by nutritionists to characterize the energy-producing potential in food
  2. a unit of heat, equal to 4.1868 joules (International Table calorie): formerly defined as the quantity of heat required to raise the temperature of 1 gram of water by 1°C under standard conditions.
  3. unit of heat defined as the quantity of heat required to raise the temperature of 1 gram of water by 1 degree centigrade at atmospheric pressure

5 letter answer(s) to unit of heat

THERM

  1. a unit of heat equal to 100,000 British thermal units

Other crossword clues with similar answers to ‘Unit of heat’


Still struggling to solve the crossword clue ‘Unit of heat’?

If you’re still haven’t solved the crossword clue Unit of heat then why not search our database by the letters you have already!

Also found in: Medical.

the quantity of heat required to raise, by one degree, the temperature of a unit mass of water, initially at a certain standard temperature. The temperature usually employed is that of 0° Centigrade, or 32° Fahrenheit.
(Physics) a determinate quantity of heat adopted as a unit of measure; a thermal unit (see under Thermal). Water is the substance generally employed, the unit being one gram or one pound, and the temperature interval one degree of the Centigrade or Fahrenheit scale. When referred to the gram, it is called the gram degree. The British unit of heat, or thermal unit, used by engineers in England and in the United States, is the quantity of heat necessary to raise one pound of pure water at and near its temperature of greatest density (39.1° Fahr.) through one degree of the Fahrenheit scale.

See also: Heat, Unit

Webster’s Revised Unabridged Dictionary, published 1913 by G. & C. Merriam Co.

Mentioned in
?

  • B.Th.U.
  • British thermal unit
  • BTU
  • calorie
  • calorific
  • energy unit
  • gram calorie
  • Gram degree
  • heat unit
  • kilocalorie
  • kilogram calorie
  • large calorie
  • nutritionist’s calorie
  • small calorie
  • Specific heat of a substance
  • therm
  • -therm
  • therm-
  • Thermometic conductivity

References in periodicals archive
?

New standards for food safety are being pursued by the Australian unit of Heat and Control, Mt.

Dictionary browser
?

  • unissued
  • unit
  • unit aircraft
  • unit cell
  • unit character
  • unit combat readiness
  • unit commitment status
  • unit cost
  • Unit deme
  • unit designation list
  • unit emplaning officer
  • unit equipment
  • unit factor
  • unit holder
  • unit identification code
  • unit investment trust
  • Unit jar
  • unit line number
  • unit loading
  • unit magnetic pole
  • unit matrix
  • unit movement control center
  • unit movement data
  • unit of account
  • unit of ammunition
  • Unit of heat
  • Unit of illumination
  • unit of issue
  • Unit of measure
  • unit of measurement
  • Unit of power
  • Unit of resistance
  • unit of time
  • unit of viscosity
  • Unit of work
  • unit personnel and tonnage table
  • unit price
  • unit pricing
  • unit process
  • unit readiness
  • unit reserves
  • unit rule
  • unit strength
  • Unit stress
  • unit training assembly
  • unit trust
  • unit type code
  • unit vector
  • Unit.
  • UNITA
  • Unitable

Full browser
?

  • Unit of Assessment
  • unit of blood
  • unit of blood
  • unit of blood
  • unit of blood
  • unit of blood
  • unit of capacity
  • unit of coal
  • Unit of Comparative Plant Ecology
  • Unit of Competency
  • Unit of Constant Purchasing Power
  • unit of convergence
  • Unit of Credit
  • Unit Of Delivery
  • Unit of Dental Activity
  • Unit of Destruction
  • Unit of distance
  • Unit of Dynamic Pressure
  • Unit of Employment
  • Unit of energy
  • Unit of energy
  • Unit of Entitlement
  • Unit of Execution
  • Unit of Expertise
  • Unit of Fire
  • Unit of Fire
  • unit of force
  • Unit of Functionality
  • Unit of General Local Government
  • Unit of Health-Care Epidemiology
  • Unit of heat
  • Unit of illumination
  • Unit of Inquiry
  • Unit of Interchange
  • unit of intermedin
  • unit of issue
  • unit of issue
  • unit of issue
  • Unit of Issue Conversion Factor
  • Unit of Learning
  • Unit of Least Precision
  • Unit of length
  • unit of light
  • unit of light
  • unit of light
  • unit of luminous flux
  • unit of luminous flux
  • unit of luminous flux
  • unit of luminous intensity
  • unit of luminous intensity
  • unit of luminous intensity
  • unit of luteinizing activity
  • unit of magnetic field intensity
  • unit of magnetic field intensity
  • unit of magnetic field intensity
  • unit of magnetic flux intensity
  • unit of magnetic flux intensity
  • unit of magnetic flux intensity
  • unit of mass
  • Unit of measure
  • Unit of measure

see caption

A glowing-hot metal bar showing incandescence, the emission of light due to its temperature, is often recognized as a source of heat.

In thermodynamics, heat is defined as the form of energy crossing the boundary of a thermodynamic system by virtue of a temperature difference across the boundary.[1] A thermodynamic system does not contain heat. Nevertheless, the term is also often used to refer to the thermal energy contained in a system as a component of its internal energy and that is reflected in the temperature of the system. For both uses of the term, heat is a form of energy.

An example of formal vs. informal usage may be obtained from the right-hand photo, in which the metal bar is «conducting heat» from its hot end to its cold end, but if the metal bar is considered a thermodynamic system, then the energy flowing within the metal bar is called internal energy, not heat. The hot metal bar is also transferring heat to its surroundings, a correct statement for both the strict and loose meanings of heat. Another example of informal usage is the term heat content, used despite the fact that physics defines heat as energy transfer. More accurately, it is thermal energy that is contained in the system or body, as it is stored in the microscopic degrees of freedom of the modes of vibration.[2]

Heat is energy in transfer to or from a thermodynamic system, by a mechanism that involves the microscopic atomic modes of motion or the corresponding macroscopic properties.[3] This descriptive characterization excludes the transfers of energy by thermodynamic work or mass transfer. Defined quantitatively, the heat involved in a process is the difference in internal energy between the final and initial states of a system, and subtracting the work done in the process.[4] This is the formulation of the first law of thermodynamics.

The measurement of energy transferred as heat is called calorimetry, performed by measuring its effect on the states of interacting bodies. For example, heat can be measured by the amount of ice melted, or by change in temperature of a body in the surroundings of the system.[5]

In the International System of Units (SI) the unit of measurement for heat, as a form of energy, is the joule (J).

Notation and units

As a form of energy, heat has the unit joule (J) in the International System of Units (SI). In addition, many applied branches of engineering use other, traditional units, such as the British thermal unit (BTU) and the calorie. The standard unit for the rate of heating is the watt (W), defined as one joule per second.

The symbol Q for heat was introduced by Rudolf Clausius and Macquorn Rankine in c. 1859.[6]

Heat released by a system into its surroundings is by convention a negative quantity (Q < 0); when a system absorbs heat from its surroundings, it is positive (Q > 0). Heat transfer rate, or heat flow per unit time, is denoted by {dot {Q}}, but it is not a time derivative of a function of state (which can also be written with the dot notation) since heat is not a function of state.[7] Heat flux is defined as rate of heat transfer per unit cross-sectional area (watts per square metre).

Classical thermodynamics

Heat and entropy

In 1856, Rudolf Clausius, referring to closed systems, in which transfers of matter do not occur, defined the second fundamental theorem (the second law of thermodynamics) in the mechanical theory of heat (thermodynamics): «if two transformations which, without necessitating any other permanent change, can mutually replace one another, be called equivalent, then the generations of the quantity of heat Q from work at the temperature T, has the equivalence-value[8][9]

{displaystyle {frac {Q}{T}}.}

In 1865, he came to define the entropy symbolized by S, such that, due to the supply of the amount of heat Q at temperature T the entropy of the system is increased by

{displaystyle Delta S={frac {Q}{T}}}

(1)

In a transfer of energy as heat without work being done, there are changes of entropy in both the surroundings which lose heat and the system which gains it. The increase, ΔS, of entropy in the system may be considered to consist of two parts, an increment, ΔS that matches, or ‘compensates’, the change, −ΔS, of entropy in the surroundings, and a further increment, ΔS′′ that may be considered to be ‘generated’ or ‘produced’ in the system, and is said therefore to be ‘uncompensated’. Thus

{displaystyle Delta S=Delta S'+Delta S''.}

This may also be written

Delta S_{mathrm {system} }=Delta S_{mathrm {compensated} }+Delta S_{mathrm {uncompensated} },,,,{text{with}},,,,Delta S_{mathrm {compensated} }=-Delta S_{mathrm {surroundings} }.

The total change of entropy in the system and surroundings is thus

Delta S_{mathrm {overall} }=Delta S^{prime }+Delta S^{prime prime }-Delta S^{prime }=Delta S^{prime prime }.

This may also be written

Delta S_{mathrm {overall} }=Delta S_{mathrm {compensated} }+Delta S_{mathrm {uncompensated} }+Delta S_{mathrm {surroundings} }=Delta S_{mathrm {uncompensated} }.

It is then said that an amount of entropy ΔS has been transferred from the surroundings to the system. Because entropy is not a conserved quantity, this is an exception to the general way of speaking, in which an amount transferred is of a conserved quantity.

From the second law of thermodynamics it follows that in a spontaneous transfer of heat, in which the temperature of the system is different from that of the surroundings:

{displaystyle Delta S_{mathrm {overall} }>0.}

For purposes of mathematical analysis of transfers, one thinks of fictive processes that are called reversible, with the temperature T of the system being hardly less than that of the surroundings, and the transfer taking place at an imperceptibly slow rate.

Following the definition above in formula (1), for such a fictive reversible process, a quantity of transferred heat δQ (an inexact differential) is analyzed as a quantity T dS, with dS (an exact differential):

T,mathrm {d} S=delta Q.

This equality is only valid for a fictive transfer in which there is no production of entropy, that is to say, in which there is no uncompensated entropy.

If, in contrast, the process is natural, and can really occur, with irreversibility, then there is entropy production, with dSuncompensated > 0. The quantity T dSuncompensated was termed by Clausius the «uncompensated heat», though that does not accord with present-day terminology. Then one has

{displaystyle T_{surr},mathrm {d} S=delta Q+T,mathrm {d} S_{mathrm {uncompensated} }>delta Q.}

This leads to the statement

{displaystyle T_{surr},mathrm {d} Sgeq delta Qquad {text{(second law)}},.}

which is the second law of thermodynamics for closed systems.

In non-equilibrium thermodynamics that makes the approximation of assuming the hypothesis of local thermodynamic equilibrium, there is a special notation for this. The transfer of energy as heat is assumed to take place across an infinitesimal temperature difference, so that the system element and its surroundings have near enough the same temperature T. Then one writes

{displaystyle mathrm {d} S=mathrm {d} S_{mathrm {e} }+mathrm {d} S_{mathrm {i} },,}

where by definition

delta Q=T,mathrm {d} S_{mathrm {e} },,,,,{text{and}},,,,,mathrm {d} S_{mathrm {i} }equiv mathrm {d} S_{mathrm {uncompensated} }.

The second law for a natural process asserts that[10][11][12][13]

{displaystyle mathrm {d} S_{mathrm {i} }>0.}

Heat and enthalpy

For a closed system (a system from which no matter can enter or exit), one version of the first law of thermodynamics states that the change in internal energy ΔU of the system is equal to the amount of heat Q supplied to the system minus the amount of thermodynamic work W done by system on its surroundings. The foregoing sign convention for work is used in the present article, but an alternate sign convention, followed by IUPAC, for work, is to consider the work performed on the system by its surroundings as positive. This is the convention adopted by many modern textbooks of physical chemistry, such as those by Peter Atkins and Ira Levine, but many textbooks on physics define work as work done by the system.

Delta U=Q-W,.

This formula can be re-written so as to express a definition of quantity of energy transferred as heat, based purely on the concept of adiabatic work, if it is supposed that ΔU is defined and measured solely by processes of adiabatic work:

Q=Delta U+W.

The thermodynamic work done by the system is through mechanisms defined by its thermodynamic state variables, for example, its volume V, not through variables that necessarily involve mechanisms in the surroundings. The latter are such as shaft work, and include isochoric work.

The internal energy, U, is a state function. In cyclical processes, such as the operation of a heat engine, state functions of the working substance return to their initial values upon completion of a cycle.

The differential, or infinitesimal increment, for the internal energy in an infinitesimal process is an exact differential dU. The symbol for exact differentials is the lowercase letter d.

In contrast, neither of the infinitesimal increments δQ nor δW in an infinitesimal process represents the change in a state function of the system. Thus, infinitesimal increments of heat and work are inexact differentials. The lowercase Greek letter delta, δ, is the symbol for inexact differentials. The integral of any inexact differential in a process where the system leaves and then returns to the same thermodynamic state does not necessarily equal zero.

As recounted above, in the section headed heat and entropy, the second law of thermodynamics observes that if heat is supplied to a system in a reversible process, the increment of heat δQ and the temperature T form the exact differential

mathrm {d} S={frac {delta Q}{T}},

and that S, the entropy of the working body, is a state function. Likewise, with a well-defined pressure, P, behind a slowly moving (quasistatic) boundary, the work differential, δW, and the pressure, P, combine to form the exact differential

mathrm {d} V={frac {delta W}{P}},

with V the volume of the system, which is a state variable. In general, for systems of uniform pressure and temperature without composition change,

mathrm {d} U=Tmathrm {d} S-Pmathrm {d} V.

Associated with this differential equation is the concept that the internal energy may be considered to be a function U (S,V) of its natural variables S and V. The internal energy representation of the fundamental thermodynamic relation is written as[14][15]

U=U(S,V).

If V is constant

Tmathrm {d} S=mathrm {d} U,,,,,,,,,,,,(V,,{text{constant)}}

and if P is constant

Tmathrm {d} S=mathrm {d} H,,,,,,,,,,,,(P,,{text{constant)}}

with the enthalpy H defined by

H=U+PV.

The enthalpy may be considered to be a function H(S, P) of its natural variables S and P. The enthalpy representation of the fundamental thermodynamic relation is written[15][16]

H=H(S,P).

The internal energy representation and the enthalpy representation are partial Legendre transforms of one another. They contain the same physical information, written in different ways. Like the internal energy, the enthalpy stated as a function of its natural variables is a thermodynamic potential and contains all thermodynamic information about a body.[16][17]

If a quantity Q of heat is added to a body while it does only expansion work W on its surroundings, one has

Delta H=Delta U+Delta (PV),.

If this is constrained to happen at constant pressure, i.e. with ΔP = 0, the expansion work W done by the body is given by W = P ΔV; recalling the first law of thermodynamics, one has

Delta U=Q-W=Q-P,Delta V{text{   and   }}Delta (PV)=P,Delta V,.

Consequently, by substitution one has

{displaystyle {begin{aligned}Delta H&=Q-P,Delta V+P,Delta V\&=Qqquad qquad ,,,,,,,,,,,,,,,,,,,,{text{at constant pressure without electrical work.}}end{aligned}}}

In this scenario, the increase in enthalpy is equal to the quantity of heat added to the system. This is the basis of the determination of enthalpy changes in chemical reactions by calorimetry. Since many processes do take place at constant atmospheric pressure, the enthalpy is sometimes given the misleading name of ‘heat content’[18] or heat function,[19] while it actually depends strongly on the energies of covalent bonds and intermolecular forces.

In terms of the natural variables S and P of the state function H, this process of change of state from state 1 to state 2 can be expressed as

{displaystyle {begin{aligned}Delta H&=int _{S_{1}}^{S_{2}}left({frac {partial H}{partial S}}right)_{P}mathrm {d} S+int _{P_{1}}^{P_{2}}left({frac {partial H}{partial P}}right)_{S}mathrm {d} P\&=int _{S_{1}}^{S_{2}}left({frac {partial H}{partial S}}right)_{P}mathrm {d} S,,,,,,,,,,,,,,,{text{at constant pressure without electrical work.}}end{aligned}}}

It is known that the temperature T(S, P) is identically stated by

left({frac {partial H}{partial S}}right)_{P}equiv T(S,P),.

Consequently,

{displaystyle Delta H=int _{S_{1}}^{S_{2}}T(S,P)mathrm {d} S,,,,,,,,,,,,,,,,,,,,{text{at constant pressure without electrical work.}}}

In this case, the integral specifies a quantity of heat transferred at constant pressure.

History

As a common noun, English heat or warmth (just as French chaleur, German Wärme, Latin calor, Greek θάλπος, etc.) refers to (the human perception of) either thermal energy or temperature. Speculation on thermal energy or «heat» as a separate form of matter has a long history, identified as caloric theory, phlogiston theory, and fire.

The modern understanding of thermal energy originates with Thompson’s 1798 mechanical theory of heat (An Experimental Enquiry Concerning the Source of the Heat which is Excited by Friction), postulating a mechanical equivalent of heat.
A collaboration between Nicolas Clément and Sadi Carnot (Reflections on the Motive Power of Fire) in the 1820s had some related thinking along similar lines.[20] In 1845, Joule published a paper entitled The Mechanical Equivalent of Heat, in which he specified a numerical value for the amount of mechanical work required to «produce a unit of heat».
The theory of classical thermodynamics matured in the 1850s to 1860s. John Tyndall’s Heat Considered as Mode of Motion (1863) was instrumental in popularizing the idea of heat as motion to the English-speaking public.
The theory was developed in academic publications in French, English and German. From an early time, the French technical term chaleur used by Carnot was taken as equivalent to the English heat and German Wärme (lit. «warmth», while the equivalent of heat would be German Hitze).

The process function Q was introduced by Rudolf Clausius in 1850.
Clausius described it with the German compound Wärmemenge, translated as «amount of heat».[21]

James Clerk Maxwell in his 1871 Theory of Heat outlines four stipulations for the definition of heat:

  • It is something which may be transferred from one body to another, according to the second law of thermodynamics.
  • It is a measurable quantity, and so can be treated mathematically.
  • It cannot be treated as a material substance, because it may be transformed into something that is not a material substance, e.g., mechanical work.
  • Heat is one of the forms of energy.[22]

The process function Q is referred to as Wärmemenge by Clausius, or as «amount of heat» in translation.
Use of «heat» as an abbreviated form of the specific concept of «quantity of energy transferred as heat» led to some terminological confusion by the early 20th century. The generic meaning of «heat», even in classical thermodynamics, is just «thermal energy».[23]
Since the 1920s, it has been recommended practice to use enthalpy to refer to the «heat content at constant volume», and to thermal energy when «heat» in the general sense is intended, while «heat» is reserved for the very specific context of the transfer of thermal energy between two systems.
Leonard Benedict Loeb in his Kinetic Theory of Gases (1927) makes a point of using «quantity of heat»
or «heat–quantity» when referring to Q:[24]

After the perfection of thermometry […] the next great advance made in the field of heat was the definition of a term which is called the quantity of heat. [… after the abandonment of caloric theory,] It still remains to interpret this very definite concept, the quantity of heat, in terms of a theory ascribing all heat to the kinetics of gas molecules.

[25]

Today’s narrow definition of heat in physics contrasts with its use in common language, in some engineering disciplines, and in the historical scientific development of thermodynamics in the caloric theory. The terminology of heat in these instances may be replaced accurately with entropy.[26][27]

Richard Feynman introduced heat with a physical depiction, as associated with the jiggling motion of atoms and molecules, with faster motion corresponding to increased temperature.[28] To explain physics further, he used the term «heat energy,»[29] along with «heat».[30]

Carathéodory (1909)

A frequent definition of heat is based on the work of Carathéodory (1909), referring to processes in a closed system.[31][32][33][34][35][36]

The internal energy UX of a body in an arbitrary state X can be determined by amounts of work adiabatically performed by the body on its surroundings when it starts from a reference state O. Such work is assessed through quantities defined in the surroundings of the body. It is supposed that such work can be assessed accurately, without error due to friction in the surroundings; friction in the body is not excluded by this definition. The adiabatic performance of work is defined in terms of adiabatic walls, which allow transfer of energy as work, but no other transfer, of energy or matter. In particular they do not allow the passage of energy as heat. According to this definition, work performed adiabatically is in general accompanied by friction within the thermodynamic system or body. On the other hand, according to Carathéodory (1909), there also exist non-adiabatic, diathermal walls, which are postulated to be permeable only to heat.

For the definition of quantity of energy transferred as heat, it is customarily envisaged that an arbitrary state of interest Y is reached from state O by a process with two components, one adiabatic and the other not adiabatic. For convenience one may say that the adiabatic component was the sum of work done by the body through volume change through movement of the walls while the non-adiabatic wall was temporarily rendered adiabatic, and of isochoric adiabatic work. Then the non-adiabatic component is a process of energy transfer through the wall that passes only heat, newly made accessible for the purpose of this transfer, from the surroundings to the body. The change in internal energy to reach the state Y from the state O is the difference of the two amounts of energy transferred.

Although Carathéodory himself did not state such a definition, following his work it is customary in theoretical studies to define heat, Q, to the body from its surroundings, in the combined process of change to state Y from the state O, as the change in internal energy, ΔUY, minus the amount of work, W, done by the body on its surrounds by the adiabatic process, so that Q = ΔUYW.

In this definition, for the sake of conceptual rigour, the quantity of energy transferred as heat is not specified directly in terms of the non-adiabatic process. It is defined through knowledge of precisely two variables, the change of internal energy and the amount of adiabatic work done, for the combined process of change from the reference state O to the arbitrary state Y. It is important that this does not explicitly involve the amount of energy transferred in the non-adiabatic component of the combined process. It is assumed here that the amount of energy required to pass from state O to state Y, the change of internal energy, is known, independently of the combined process, by a determination through a purely adiabatic process, like that for the determination of the internal energy of state X above. The rigour that is prized in this definition is that there is one and only one kind of energy transfer admitted as fundamental: energy transferred as work. Energy transfer as heat is considered as a derived quantity. The uniqueness of work in this scheme is considered to guarantee rigor and purity of conception. The conceptual purity of this definition, based on the concept of energy transferred as work as an ideal notion, relies on the idea that some frictionless and otherwise non-dissipative processes of energy transfer can be realized in physical actuality. The second law of thermodynamics, on the other hand, assures us that such processes are not found in nature.

Before the rigorous mathematical definition of heat based on Carathéodory’s 1909 paper,
historically, heat, temperature, and thermal equilibrium were presented in thermodynamics textbooks as jointly primitive notions.[37] Carathéodory introduced his 1909 paper thus: «The proposition that the discipline of thermodynamics can be justified without recourse to any hypothesis that cannot be verified experimentally must be regarded as one of the most noteworthy results of the research in thermodynamics that was accomplished during the last century.» Referring to the «point of view adopted by most authors who were active in the last fifty years», Carathéodory wrote: «There exists a physical quantity called heat that is not identical with the mechanical quantities (mass, force, pressure, etc.) and whose variations can be determined by calorimetric measurements.» James Serrin introduces an account of the theory of thermodynamics thus: «In the following section, we shall use the classical notions of heat, work, and hotness as primitive elements, … That heat is an appropriate and natural primitive for thermodynamics was already accepted by Carnot. Its continued validity as a primitive element of thermodynamical structure is due to the fact that it synthesizes an essential physical concept, as well as to its successful use in recent work to unify different constitutive theories.»[38][39] This traditional kind of presentation of the basis of thermodynamics includes ideas that may be summarized by the statement that heat transfer is purely due to spatial non-uniformity of temperature, and is by conduction and radiation, from hotter to colder bodies. It is sometimes proposed that this traditional kind of presentation necessarily rests on «circular reasoning»; against this proposal, there stands the rigorously logical mathematical development of the theory presented by Truesdell and Bharatha (1977).[40]

This alternative approach to the definition of quantity of energy transferred as heat differs in logical structure from that of Carathéodory, recounted just above.

This alternative approach admits calorimetry as a primary or direct way to measure quantity of energy transferred as heat. It relies on temperature as one of its primitive concepts, and used in calorimetry.[41] It is presupposed that enough processes exist physically to allow measurement of differences in internal energies. Such processes are not restricted to adiabatic transfers of energy as work. They include calorimetry, which is the commonest practical way of finding internal energy differences.[42] The needed temperature can be either empirical or absolute thermodynamic.

In contrast, the Carathéodory way recounted just above does not use calorimetry or temperature in its primary definition of quantity of energy transferred as heat. The Carathéodory way regards calorimetry only as a secondary or indirect way of measuring quantity of energy transferred as heat. As recounted in more detail just above, the Carathéodory way regards quantity of energy transferred as heat in a process as primarily or directly defined as a residual quantity. It is calculated from the difference of the internal energies of the initial and final states of the system, and from the actual work done by the system during the process. That internal energy difference is supposed to have been measured in advance through processes of purely adiabatic transfer of energy as work, processes that take the system between the initial and final states. By the Carathéodory way it is presupposed as known from experiment that there actually physically exist enough such adiabatic processes, so that there need be no recourse to calorimetry for measurement of quantity of energy transferred as heat. This presupposition is essential but is explicitly labeled neither as a law of thermodynamics nor as an axiom of the Carathéodory way. In fact, the actual physical existence of such adiabatic processes is indeed mostly supposition, and those supposed processes have in most cases not been actually verified empirically to exist.[43]

Heat transfer

Heat transfer between two bodies

Referring to conduction, Partington writes: «If a hot body is brought in conducting contact with a cold body, the temperature of the hot body falls and that of the cold body rises, and it is said that a quantity of heat has passed from the hot body to the cold body.»[44]

Referring to radiation, Maxwell writes: «In Radiation, the hotter body loses heat, and the colder body receives heat by means of a process occurring in some intervening medium which does not itself thereby become hot.»[45]

Maxwell writes that convection as such «is not a purely thermal phenomenon».[46] In thermodynamics, convection in general is regarded as transport of internal energy. If, however, the convection is enclosed and circulatory, then it may be regarded as an intermediary that transfers energy as heat between source and destination bodies, because it transfers only energy and not matter from the source to the destination body.[47]

In accordance with the first law for closed systems, energy transferred solely as heat leaves one body and enters another, changing the internal energies of each. Transfer, between bodies, of energy as work is a complementary way of changing internal energies. Though it is not logically rigorous from the viewpoint of strict physical concepts, a common form of words that expresses this is to say that heat and work are interconvertible.

Cyclically operating engines that use only heat and work transfers have two thermal reservoirs, a hot and a cold one. They may be classified by the range of operating temperatures of the working body, relative to those reservoirs. In a heat engine, the working body is at all times colder than the hot reservoir and hotter than the cold reservoir. In a sense, it uses heat transfer to produce work. In a heat pump, the working body, at stages of the cycle, goes both hotter than the hot reservoir, and colder than the cold reservoir. In a sense, it uses work to produce heat transfer.

Heat engine

In classical thermodynamics, a commonly considered model is the heat engine. It consists of four bodies: the working body, the hot reservoir, the cold reservoir, and the work reservoir. A cyclic process leaves the working body in an unchanged state, and is envisaged as being repeated indefinitely often. Work transfers between the working body and the work reservoir are envisaged as reversible, and thus only one work reservoir is needed. But two thermal reservoirs are needed, because transfer of energy as heat is irreversible. A single cycle sees energy taken by the working body from the hot reservoir and sent to the two other reservoirs, the work reservoir and the cold reservoir. The hot reservoir always and only supplies energy and the cold reservoir always and only receives energy. The second law of thermodynamics requires that no cycle can occur in which no energy is received by the cold reservoir. Heat engines achieve higher efficiency when the ratio of the initial and final temperature is greater.

Heat pump or refrigerator

Another commonly considered model is the heat pump or refrigerator. Again there are four bodies: the working body, the hot reservoir, the cold reservoir, and the work reservoir. A single cycle starts with the working body colder than the cold reservoir, and then energy is taken in as heat by the working body from the cold reservoir. Then the work reservoir does work on the working body, adding more to its internal energy, making it hotter than the hot reservoir. The hot working body passes heat to the hot reservoir, but still remains hotter than the cold reservoir. Then, by allowing it to expand without passing heat to another body, the working body is made colder than the cold reservoir. It can now accept heat transfer from the cold reservoir to start another cycle.

The device has transported energy from a colder to a hotter reservoir, but this is not regarded as by an inanimate agency; rather, it is regarded as by the harnessing of work . This is because work is supplied from the work reservoir, not just by a simple thermodynamic process, but by a cycle of thermodynamic operations and processes, which may be regarded as directed by an animate or harnessing agency. Accordingly, the cycle is still in accord with the second law of thermodynamics. The ‘efficiency’ of a heat pump (which exceeds unity) is best when the temperature difference between the hot and cold reservoirs is least.

Functionally, such engines are used in two ways, distinguishing a target reservoir and a resource or surrounding reservoir. A heat pump transfers heat to the hot reservoir as the target from the resource or surrounding reservoir. A refrigerator transfers heat, from the cold reservoir as the target, to the resource or surrounding reservoir. The target reservoir may be regarded as leaking: when the target leaks heat to the surroundings, heat pumping is used; when the target leaks coldness to the surroundings, refrigeration is used. The engines harness work to overcome the leaks.

Macroscopic view

According to Planck, there are three main conceptual approaches to heat.[48] One is the microscopic or kinetic theory approach. The other two are macroscopic approaches. One is the approach through the law of conservation of energy taken as prior to thermodynamics, with a mechanical analysis of processes, for example in the work of Helmholtz. This mechanical view is taken in this article as currently customary for thermodynamic theory. The other macroscopic approach is the thermodynamic one, which admits heat as a primitive concept, which contributes, by scientific induction[49] to knowledge of the law of conservation of energy. This view is widely taken as the practical one, quantity of heat being measured by calorimetry.

Bailyn also distinguishes the two macroscopic approaches as the mechanical and the thermodynamic.[50] The thermodynamic view was taken by the founders of thermodynamics in the nineteenth century. It regards quantity of energy transferred as heat as a primitive concept coherent with a primitive concept of temperature, measured primarily by calorimetry. A calorimeter is a body in the surroundings of the system, with its own temperature and internal energy; when it is connected to the system by a path for heat transfer, changes in it measure heat transfer. The mechanical view was pioneered by Helmholtz and developed and used in the twentieth century, largely through the influence of Max Born.[51] It regards quantity of heat transferred as heat as a derived concept, defined for closed systems as quantity of heat transferred by mechanisms other than work transfer, the latter being regarded as primitive for thermodynamics, defined by macroscopic mechanics. According to Born, the transfer of internal energy between open systems that accompanies transfer of matter «cannot be reduced to mechanics».[52] It follows that there is no well-founded definition of quantities of energy transferred as heat or as work associated with transfer of matter.

Nevertheless, for the thermodynamical description of non-equilibrium processes, it is desired to consider the effect of a temperature gradient established by the surroundings across the system of interest when there is no physical barrier or wall between system and surroundings, that is to say, when they are open with respect to one another. The impossibility of a mechanical definition in terms of work for this circumstance does not alter the physical fact that a temperature gradient causes a diffusive flux of internal energy, a process that, in the thermodynamic view, might be proposed as a candidate concept for transfer of energy as heat.

In this circumstance, it may be expected that there may also be active other drivers of diffusive flux of internal energy, such as gradient of chemical potential which drives transfer of matter, and gradient of electric potential which drives electric current and iontophoresis; such effects usually interact with diffusive flux of internal energy driven by temperature gradient, and such interactions are known as cross-effects.[53]

If cross-effects that result in diffusive transfer of internal energy were also labeled as heat transfers, they would sometimes violate the rule that pure heat transfer occurs only down a temperature gradient, never up one. They would also contradict the principle that all heat transfer is of one and the same kind, a principle founded on the idea of heat conduction between closed systems. One might to try to think narrowly of heat flux driven purely by temperature gradient as a conceptual component of diffusive internal energy flux, in the thermodynamic view, the concept resting specifically on careful calculations based on detailed knowledge of the processes and being indirectly assessed. In these circumstances, if perchance it happens that no transfer of matter is actualized, and there are no cross-effects, then the thermodynamic concept and the mechanical concept coincide, as if one were dealing with closed systems. But when there is transfer of matter, the exact laws by which temperature gradient drives diffusive flux of internal energy, rather than being exactly knowable, mostly need to be assumed, and in many cases are practically unverifiable. Consequently, when there is transfer of matter, the calculation of the pure ‘heat flux’ component of the diffusive flux of internal energy rests on practically unverifiable assumptions.[54][quotations 1][55] This is a reason to think of heat as a specialized concept that relates primarily and precisely to closed systems, and applicable only in a very restricted way to open systems.

In many writings in this context, the term «heat flux» is used when what is meant is therefore more accurately called diffusive flux of internal energy; such usage of the term «heat flux» is a residue of older and now obsolete language usage that allowed that a body may have a «heat content».[56]

Microscopic view

In the kinetic theory, heat is explained in terms of the microscopic motions and interactions of constituent particles, such as electrons, atoms, and molecules.[57] The immediate meaning of the kinetic energy of the constituent particles is not as heat. It is as a component of internal energy.
In microscopic terms, heat is a transfer quantity, and is described by a transport theory, not as steadily localized kinetic energy of particles. Heat transfer arises from temperature gradients or differences, through the diffuse exchange of microscopic kinetic and potential particle energy, by particle collisions and other interactions. An early and vague expression of this was made by Francis Bacon.[58][59] Precise and detailed versions of it were developed in the nineteenth century.[60]

In statistical mechanics, for a closed system (no transfer of matter), heat is the energy transfer associated with a disordered, microscopic action on the system, associated with jumps in occupation numbers of the energy levels of the system, without change in the values of the energy levels themselves.[61] It is possible for macroscopic thermodynamic work to alter the occupation numbers without change in the values of the system energy levels themselves, but what distinguishes transfer as heat is that the transfer is entirely due to disordered, microscopic action, including radiative transfer. A mathematical definition can be formulated for small increments of quasi-static adiabatic work in terms of the statistical distribution of an ensemble of microstates.

Calorimetry

Quantity of heat transferred can be measured by calorimetry, or determined through calculations based on other quantities.

Calorimetry is the empirical basis of the idea of quantity of heat transferred in a process. The transferred heat is measured by changes in a body of known properties, for example, temperature rise, change in volume or length, or phase change, such as melting of ice.[62][63]

A calculation of quantity of heat transferred can rely on a hypothetical quantity of energy transferred as adiabatic work and on the first law of thermodynamics. Such calculation is the primary approach of many theoretical studies of quantity of heat transferred.[31][64][65]

Engineering

A red-hot iron rod from which heat transfer to the surrounding environment will be primarily through radiation.

The discipline of heat transfer, typically considered an aspect of mechanical engineering and chemical engineering, deals with specific applied methods by which thermal energy in a system is generated, or converted, or transferred to another system. Although the definition of heat implicitly means the transfer of energy, the term heat transfer encompasses this traditional usage in many engineering disciplines and laymen language.

Heat transfer is generally described as including the mechanisms of heat conduction, heat convection, thermal radiation, but may include mass transfer and heat in processes of phase changes.

Convection may be described as the combined effects of conduction and fluid flow. From the thermodynamic point of view, heat flows into a fluid by diffusion to increase its energy, the fluid then transfers (advects) this increased internal energy (not heat) from one location to another, and this is then followed by a second thermal interaction which transfers heat to a second body or system, again by diffusion. This entire process is often regarded as an additional mechanism of heat transfer, although technically, «heat transfer» and thus heating and cooling occurs only on either end of such a conductive flow, but not as a result of flow. Thus, conduction can be said to «transfer» heat only as a net result of the process, but may not do so at every time within the complicated convective process.

Latent and sensible heat

In an 1847 lecture entitled On Matter, Living Force, and Heat, James Prescott Joule characterized the terms latent heat and sensible heat as components of heat each affecting distinct physical phenomena, namely the potential and kinetic energy of particles, respectively.[66][quotations 2]
He described latent energy as the energy possessed via a distancing of particles where attraction was over a greater distance, i.e. a form of potential energy, and the sensible heat as an energy involving the motion of particles, i.e. kinetic energy.

Latent heat is the heat released or absorbed by a chemical substance or a thermodynamic system during a change of state that occurs without a change in temperature. Such a process may be a phase transition, such as the melting of ice or the boiling of water.[67][68]

Heat capacity

Heat capacity is a measurable physical quantity equal to the ratio of the heat added to an object to the resulting temperature change.[69] The molar heat capacity is the heat capacity per unit amount (SI unit: mole) of a pure substance, and the specific heat capacity, often called simply specific heat, is the heat capacity per unit mass of a material. Heat capacity is a physical property of a substance, which means that it depends on the state and properties of the substance under consideration.

The specific heats of monatomic gases, such as helium, are nearly constant with temperature. Diatomic gases such as hydrogen display some temperature dependence, and triatomic gases (e.g., carbon dioxide) still more.

Before the development of the laws of thermodynamics, heat was measured by changes in the states of the participating bodies.

Some general rules, with important exceptions, can be stated as follows.

In general, most bodies expand on heating. In this circumstance, heating a body at a constant volume increases the pressure it exerts on its constraining walls, while heating at a constant pressure increases its volume.

Beyond this, most substances have three ordinarily recognized states of matter, solid, liquid, and gas. Some can also exist in a plasma. Many have further, more finely differentiated, states of matter, such as glass and liquid crystal. In many cases, at fixed temperature and pressure, a substance can exist in several distinct states of matter in what might be viewed as the same ‘body’. For example, ice may float in a glass of water. Then the ice and the water are said to constitute two phases within the ‘body’. Definite rules are known, telling how distinct phases may coexist in a ‘body’. Mostly, at a fixed pressure, there is a definite temperature at which heating causes a solid to melt or evaporate, and a definite temperature at which heating causes a liquid to evaporate. In such cases, cooling has the reverse effects.

All of these, the commonest cases, fit with a rule that heating can be measured by changes of state of a body. Such cases supply what are called thermometric bodies, that allow the definition of empirical temperatures. Before 1848, all temperatures were defined in this way. There was thus a tight link, apparently logically determined, between heat and temperature, though they were recognized as conceptually thoroughly distinct, especially by Joseph Black in the later eighteenth century.

There are important exceptions. They break the obviously apparent link between heat and temperature. They make it clear that empirical definitions of temperature are contingent on the peculiar properties of particular thermometric substances, and are thus precluded from the title ‘absolute’. For example, water contracts on being heated near 277 K. It cannot be used as a thermometric substance near that temperature. Also, over a certain temperature range, ice contracts on heating. Moreover, many substances can exist in metastable states, such as with negative pressure, that survive only transiently and in very special conditions. Such facts, sometimes called ‘anomalous’, are some of the reasons for the thermodynamic definition of absolute temperature.

In the early days of measurement of high temperatures, another factor was important, and used by Josiah Wedgwood in his pyrometer. The temperature reached in a process was estimated by the shrinkage of a sample of clay. The higher the temperature, the more the shrinkage. This was the only available more or less reliable method of measurement of temperatures above 1000 °C (1,832 °F). But such shrinkage is irreversible. The clay does not expand again on cooling. That is why it could be used for the measurement. But only once. It is not a thermometric material in the usual sense of the word.

Nevertheless, the thermodynamic definition of absolute temperature does make essential use of the concept of heat, with proper circumspection.

«Hotness»

The property of hotness is a concern of thermodynamics that should be defined without reference to the concept of heat. Consideration of hotness leads to the concept of empirical temperature.[70][71] All physical systems are capable of heating or cooling others.[72] With reference to hotness, the comparative terms hotter and colder are defined by the rule that heat flows from the hotter body to the colder.[73][74][75]

If a physical system is inhomogeneous or very rapidly or irregularly changing, for example by turbulence, it may be impossible to characterize it by a temperature, but still there can be transfer of energy as heat between it and another system. If a system has a physical state that is regular enough, and persists long enough to allow it to reach thermal equilibrium with a specified thermometer, then it has a temperature according to that thermometer. An empirical thermometer registers degree of hotness for such a system. Such a temperature is called empirical.[76][77][78] For example, Truesdell writes about classical thermodynamics: «At each time, the body is assigned a real number called the temperature. This number is a measure of how hot the body is.»[79]

Physical systems that are too turbulent to have temperatures may still differ in hotness. A physical system that passes heat to another physical system is said to be the hotter of the two. More is required for the system to have a thermodynamic temperature. Its behavior must be so regular that its empirical temperature is the same for all suitably calibrated and scaled thermometers, and then its hotness is said to lie on the one-dimensional hotness manifold. This is part of the reason why heat is defined following Carathéodory and Born, solely as occurring other than by work or transfer of matter; temperature is advisedly and deliberately not mentioned in this now widely accepted definition.

This is also the reason that the zeroth law of thermodynamics is stated explicitly. If three physical systems, A, B, and C are each not in their own states of internal thermodynamic equilibrium, it is possible that, with suitable physical connections being made between them, A can heat B and B can heat C and C can heat A. In non-equilibrium situations, cycles of flow are possible. It is the special and uniquely distinguishing characteristic of internal thermodynamic equilibrium that this possibility is not open to thermodynamic systems (as distinguished amongst physical systems) which are in their own states of internal thermodynamic equilibrium; this is the reason why the zeroth law of thermodynamics needs explicit statement. That is to say, the relation ‘is not colder than’ between general non-equilibrium physical systems is not transitive, whereas, in contrast, the relation ‘has no lower a temperature than’ between thermodynamic systems in their own states of internal thermodynamic equilibrium is transitive. It follows from this that the relation ‘is in thermal equilibrium with’ is transitive, which is one way of stating the zeroth law.

Just as temperature may be undefined for a sufficiently inhomogeneous system, so also may entropy be undefined for a system not in its own state of internal thermodynamic equilibrium. For example, ‘the temperature of the solar system’ is not a defined quantity. Likewise, ‘the entropy of the solar system’ is not defined in classical thermodynamics. It has not been possible to define non-equilibrium entropy, as a simple number for a whole system, in a clearly satisfactory way.[80]

See also

  • Effect of sun angle on climate
  • Heat death of the Universe
  • Heat diffusion
  • Heat equation
  • Heat exchanger
  • Heat wave
  • Heat flux sensor
  • Heat transfer coefficient
  • History of heat
  • Orders of magnitude (temperature)
  • Sigma heat
  • Shock heating
  • Thermal battery
  • Thermal management of electronic devices and systems
  • Thermometer
  • Relativistic heat conduction
  • Uniform Mechanical Code
  • Uniform Solar Energy and Hydronics Code
  • Waste heat

References

  1. ^ Van Wylen, Gordon; Sonntag, Richard (1978). Fundamentals of Classical Thermodynamics (Second edition, SI Version, Revised Printing ed.). Chapter 4.7, Definition of Heat: John Wiley & Sons. p. 76. ISBN 0-471-04188-2.{{cite book}}: CS1 maint: location (link)
  2. ^ D.V. Schroeder (1999). An Introduction to Thermal Physics. Addison-Wesley. p. 15. ISBN 0-201-38027-7.
  3. ^ Herbert B. Callen (1985). Thermodynamics and an Introduction to Thermostatics (2 ed.). John Wiley & Sons. http://cvika.grimoar.cz/callen/ Archived 17 October 2018 at the Wayback Machine or http://keszei.chem.elte.hu/1alapFizkem/H.B.Callen-Thermodynamics.pdf Archived 30 December 2016 at the Wayback Machine , p. 8: Energy may be transferred via … work. «But it is equally possible to transfer energy via the hidden atomic modes of motion as well as via those that happen to be macroscopically observable. An energy transfer via the hidden atomic modes is called heat
  4. ^ Callen, p.19
  5. ^ Maxwell, J.C. (1871), Chapter III.
  6. ^ Macquorn Rankine
    in the same year used the same symbol. The two physicists were in correspondence at the time, so that it is difficult to say which of the two first introduced the symbol. (Kenneth L. Caneva, Helmholtz and the Conservation of Energy: Contexts of Creation and Reception (2021), p. 562.
  7. ^ Baierlein, R. (1999), p. 21.
  8. ^ Clausius, R. (1854).
  9. ^ Clausius, R. (1865), pp. 125–126.
  10. ^ De Groot, S.R., Mazur, P. (1962), p. 20.
  11. ^ Kondepudi, D, Prigogine, I. (1998), p. 82.
  12. ^ Kondepudi, D. (2008), p. 114.
  13. ^ Lebon, g., Jou, D., Casas-Vásquez, J. (2008), p. 41.
  14. ^ Callen, H.B., (1985), Section 2-3, pp. 40–42.
  15. ^ a b Adkins, C.J. (1983), p. 101.
  16. ^ a b Callen, H.B. (1985), p. 147.
  17. ^ Adkins, C.J. (1983), pp. 100–104.
  18. ^ Adkins, C.J. (1968/1983), p. 46.
  19. ^ Bailyn, M. (1994), p. 208.
  20. ^ Lervig, P. Sadi Carnot and the steam engine:Nicolas Clément’s lectures on industrial chemistry, 1823–28. Br. J Hist. Sci. 18:147, 1985.
  21. ^
    Die Wärmemenge, welche dem Gase mitgetheilt werden muss, während es aus irgend einem früheren Zustande auf einem bestimmten Wege in den Zustand übergeführt wird, in welchem sein Volumen = v und seine Temperatur = t ist, möge Q heissen
    (R. Clausius, Ueber die bewegende Kraft der Wärme und die Gesetze, welche sich daraus für die Wärmelehre selbst ableiten lassen Archived 17 April 2019 at the Wayback Machine, communication to the Academy of Berlin, February 1850, published in Pogendorff’s Annalen vol. 79, March/April 1850, first translated in Philosophical Magazine vol. 2, July 1851, as «First Memoir» in: The Mechanical Theory of Heat, with its Applications to the Steam-Engine and to the Physical Properties of Bodies, trans. John Tyndall, London, 1867, p. 25).
  22. ^ Maxwell, J.C. (1871), p. 7.
  23. ^ «in a gas, heat is nothing else than the kinetic or mechanical energy of motion of the gas molecules». B.L. Loeb, The Kinetic Theory of Gases (1927), p. 14.
  24. ^ From this terminological choice may derive a tradition to the effect that the letter Q represents «quantity», but there is no indication that Clausius had this in mind when he selected the letter in what seemed to be an ad hoc calculation in 1850.
  25. ^ B.L. Loeb, The Kinetic Theory of Gases (1927), p. 426 Archived 24 June 2018 at the Wayback Machine.
  26. ^ Hans U. Fuchs (2010). The Dynamics of Heat–A Unified Approach to Thermodynamics and Heat Transfer (2 ed.). Springer. p. 3. ISBN 978-1-4419-7603-1.
  27. ^ Friedrich Herrmann, Entropy from the Beginning, Plenary Lecture
  28. ^ Feynman, Richard; Leighton, Robert; Sands, Matthew (1963). The Feynman Lectures on Physics, Volume 1 (Library of Congress number 63-20717, fourth printing, 1966 ed.). Chapter 1-2, Matter is made of atoms: Addison-Wesley Publishing Company. p. 1-3.{{cite book}}: CS1 maint: location (link)
  29. ^ Feynman, Richard; Leighton, Robert; Sands, Matthew (1963). The Feynman Lectures on Physics, Volume 1 (Library of Congress number 63-20717, fourth printing, 1966 ed.). Chapter 4-1, What is energy: Addison-Wesley Publishing Company. p. 4-2.{{cite book}}: CS1 maint: location (link)
  30. ^ Feynman, Richard; Leighton, Robert; Sands, Matthew (1963). The Feynman Lectures on Physics, Volume 1 (Library of Congress number 63-20717, fourth printing, 1966 ed.). Chapter 13: Addison-Wesley Publishing Company. p. 13-3.{{cite book}}: CS1 maint: location (link)
  31. ^ a b Carathéodory, C. (1909).
  32. ^ Adkins, C.J. (1968/1983).
  33. ^ Münster, A. (1970).
  34. ^ Pippard, A.B. (1957).
  35. ^ Fowler, R., Guggenheim, E.A. (1939).
  36. ^ Buchdahl, H.A. (1966).
  37. ^ Lieb, E.H., Yngvason, J. (1999), p. 10.
  38. ^ Serrin, J. (1986), p. 5.
  39. ^ Owen, D.R. (1984), pp. 43–45.
  40. ^ Truesdell, C., Bharatha, S. (1977).
  41. ^ Maxwell, J.C. (1871), p.v.
  42. ^ Atkins, P., de Paula, J. (1978/2010), p. 54.
  43. ^ Pippard, A.B. (1957/1966), p. 15.
  44. ^ Partington, J.R. (1949), p. 118.
  45. ^ Maxwell, J.C. (1871), p. 10.
  46. ^ Maxwell, J.C. (1871), p. 11.
  47. ^ Chandrasekhar, S. (1961).
  48. ^ Planck, M. (1897/1903), p. viii.
  49. ^ Hintikka, J. (1988), p. 180.
  50. ^ Bailyn, M. (1994), pp. 65, 79.
  51. ^ Born, M.(1949), Lecture V.
  52. ^ Born, M. (1949), p. 44.
  53. ^ De Groot, S.R., Mazur, P. (1962), p. 30.
  54. ^ Denbigh, K.G. (1951), p. 56.
  55. ^ Fitts, D.D. (1962), p. 28.
  56. ^ Gyarmati, I. (1970), p. 68.
  57. ^ Kittel, C. Kroemer, H. (1980).
  58. ^ Bacon, F. (1620).
  59. ^ Partington, J.R. (1949), p. 131.
  60. ^ Partington, J.R. (1949), pp. 132–136.
  61. ^ Reif (1965), pp. 67–68
  62. ^ Maxwell J.C. (1872), p. 54.
  63. ^ Planck (1927), Chapter 3.
  64. ^ Bryan, G.H. (1907), p. 47.
  65. ^ Callen, H.B. (1985), Section 1-8.
  66. ^ Joule J.P. (1884).
  67. ^ Perrot, P. (1998).
  68. ^ Clark, J.O.E. (2004).
  69. ^ Halliday, David; Resnick, Robert (2013). Fundamentals of Physics. Wiley. p. 524.
  70. ^ Denbigh, K. (1981), p. 9.
  71. ^ Adkins, C.J. (1968/1983), p. 55.
  72. ^ Baierlein, R. (1999), p. 349.
  73. ^ Adkins, C.J. (1968/1983), p. 34.
  74. ^ Pippard, A.B. (1957/1966), p. 18.
  75. ^ Haase, R. (1971), p. 7.
  76. ^ Mach, E. (1900), section 5, pp. 48–49, section 22, pp. 60–61.
  77. ^ Truesdell, C. (1980).
  78. ^ Serrin, J. (1986), especially p. 6.
  79. ^ Truesdell, C. (1969), p. 6.
  80. ^ Lieb, E.H., Yngvason, J. (2003), p. 190.

Quotations

  1. ^ Denbigh states in a footnote that he is indebted to correspondence with Professor E.A. Guggenheim and with Professor N.K. Adam. From this, Denbigh concludes «It seems, however, that when a system is able to exchange both heat and matter with its environment, it is impossible to make an unambiguous distinction between energy transported as heat and by the migration of matter, without already assuming the existence of the ‘heat of transport’.» Denbigh K.G. (1951), p. 56.
  2. ^ «Heat must therefore consist of either living force or of attraction through space. In the former case we can conceive the constituent particles of heated bodies to be, either in whole or in part, in a state of motion. In the latter we may suppose the particles to be removed by the process of heating, so as to exert attraction through greater space. I am inclined to believe that both of these hypotheses will be found to hold good,—that in some instances, particularly in the case of sensible heat, or such as is indicated by the thermometer, heat will be found to consist in the living force of the particles of the bodies in which it is induced; whilst in others, particularly in the case of latent heat, the phenomena are produced by the separation of particle from particle, so as to cause them to attract one another through a greater space.» Joule, J.P. (1884).

Bibliography of cited references

  • Adkins, C.J. (1968/1983). Equilibrium Thermodynamics, (1st edition 1968), third edition 1983, Cambridge University Press, Cambridge UK, ISBN 0-521-25445-0.
  • Atkins, P., de Paula, J. (1978/2010). Physical Chemistry, (first edition 1978), ninth edition 2010, Oxford University Press, Oxford UK, ISBN 978-0-19-954337-3.
  • Bacon, F. (1620). Novum Organum Scientiarum, translated by Devey, J., P.F. Collier & Son, New York, 1902.
  • Baierlein, R. (1999). Thermal Physics. Cambridge University Press. ISBN 978-0-521-65838-6.
  • Bailyn, M. (1994). A Survey of Thermodynamics, American Institute of Physics Press, New York, ISBN 0-88318-797-3.
  • Born, M. (1949). Natural Philosophy of Cause and Chance, Oxford University Press, London.
  • Bryan, G.H. (1907). Thermodynamics. An Introductory Treatise dealing mainly with First Principles and their Direct Applications, B.G. Teubner, Leipzig.
  • Buchdahl, H.A. (1966). The Concepts of Classical Thermodynamics, Cambridge University Press, Cambridge UK.
  • Callen, H.B. (1960/1985). Thermodynamics and an Introduction to Thermostatistics, (1st edition 1960) 2nd edition 1985, Wiley, New York, ISBN 0-471-86256-8.
  • Carathéodory, C. (1909). «Untersuchungen über die Grundlagen der Thermodynamik». Mathematische Annalen. 67 (3): 355–386. doi:10.1007/BF01450409. S2CID 118230148. A translation may be found here. A mostly reliable translation is to be found at Kestin, J. (1976). The Second Law of Thermodynamics, Dowden, Hutchinson & Ross, Stroudsburg PA.
  • Chandrasekhar, S. (1961). Hydrodynamic and Hydromagnetic Stability, Oxford University Press, Oxford UK.
  • Clark, J.O.E. (2004). The Essential Dictionary of Science. Barnes & Noble Books. ISBN 978-0-7607-4616-5.
  • Clausius, R. (1854). Annalen der Physik (Poggendoff’s Annalen), Dec. 1854, vol. xciii. p. 481; translated in the Journal de Mathematiques, vol. xx. Paris, 1855, and in the Philosophical Magazine, August 1856, s. 4. vol. xii, p. 81.
  • Clausius, R. (1865/1867). The Mechanical Theory of Heat – with its Applications to the Steam Engine and to Physical Properties of Bodies, London: John van Voorst. 1867. Also the second edition translated into English by W.R. Browne (1879) here and here.
  • De Groot, S.R., Mazur, P. (1962). Non-equilibrium Thermodynamics, North-Holland, Amsterdam. Reprinted (1984), Dover Publications Inc., New York, ISBN 0486647412.
  • Denbigh, K. (1955/1981). The Principles of Chemical Equilibrium, Cambridge University Press, Cambridge ISBN 0-521-23682-7.
  • Greven, A., Keller, G., Warnecke (editors) (2003). Entropy, Princeton University Press, Princeton NJ, ISBN 0-691-11338-6.
  • Guggenheim, E.A. (1967) [1949], Thermodynamics. An Advanced Treatment for Chemists and Physicists (fifth ed.), Amsterdam: North-Holland Publishing Company.
  • Jensen, W.B. (2010). «Why Are q and Q Used to Symbolize Heat?» (PDF). J. Chem. Educ. 87 (11): 1142. Bibcode:2010JChEd..87.1142J. doi:10.1021/ed100769d. Archived from the original (PDF) on 2 April 2015. Retrieved 23 March 2015.
  • J.P. Joule (1884), The Scientific Papers of James Prescott Joule, The Physical Society of London, p. 274, Lecture on Matter, Living Force, and Heat. 5 and 12 May 1847.
  • Kittel, C. Kroemer, H. (1980). Thermal Physics, second edition, W.H. Freeman, San Francisco, ISBN 0-7167-1088-9.
  • Kondepudi, D. (2008), Introduction to Modern Thermodynamics, Chichester UK: Wiley, ISBN 978-0-470-01598-8
  • Kondepudi, D., Prigogine, I. (1998). Modern Thermodynamics: From Heat Engines to Dissipative Structures, John Wiley & Sons, Chichester, ISBN 0-471-97393-9.
  • Landau, L., Lifshitz, E.M. (1958/1969). Statistical Physics, volume 5 of Course of Theoretical Physics, translated from the Russian by J.B. Sykes, M.J. Kearsley, Pergamon, Oxford.
  • Lebon, G., Jou, D., Casas-Vázquez, J. (2008). Understanding Non-equilibrium Thermodynamics: Foundations, Applications, Frontiers, Springer-Verlag, Berlin, e-ISBN 978-3-540-74252-4.
  • Lieb, E.H., Yngvason, J. (2003). The Entropy of Classical Thermodynamics, Chapter 8 of Entropy, Greven, A., Keller, G., Warnecke (editors) (2003).
  • Maxwell, J.C. (1871), Theory of Heat (first ed.), London: Longmans, Green and Co.
  • Partington, J.R. (1949), An Advanced Treatise on Physical Chemistry., vol. 1, Fundamental Principles. The Properties of Gases, London: Longmans, Green and Co.
  • Perrot, Pierre (1998). A to Z of Thermodynamics. Oxford University Press. ISBN 978-0-19-856552-9.
  • Pippard, A.B. (1957/1966). Elements of Classical Thermodynamics for Advanced Students of Physics, original publication 1957, reprint 1966, Cambridge University Press, Cambridge.
  • Planck, M., (1897/1903). Treatise on Thermodynamics, translated by A. Ogg, first English edition, Longmans, Green and Co., London.
  • Planck. M. (1914). The Theory of Heat Radiation, a translation by Masius, M. of the second German edition, P. Blakiston’s Son & Co., Philadelphia.
  • Planck, M., (1923/1927). Treatise on Thermodynamics, translated by A. Ogg, third English edition, Longmans, Green and Co., London.
  • Reif, F. (1965). Fundamentals of Statistical and Thermal Physics. New York: McGraw-Hlll, Inc.
  • Shavit, A., Gutfinger, C. (1995). Thermodynamics. From Concepts to Applications, Prentice Hall, London, ISBN 0-13-288267-1.
  • Truesdell, C. (1969). Rational Thermodynamics: a Course of Lectures on Selected Topics, McGraw-Hill Book Company, New York.
  • Truesdell, C. (1980). The Tragicomical History of Thermodynamics 1822–1854, Springer, New York, ISBN 0-387-90403-4.

Further bibliography

  • Beretta, G.P.; E.P. Gyftopoulos (1990). «What is heat?» (PDF). Education in Thermodynamics and Energy Systems. AES. 20.
  • Gyftopoulos, E.P., & Beretta, G.P. (1991). Thermodynamics: foundations and applications. (Dover Publications)
  • Hatsopoulos, G.N., & Keenan, J.H. (1981). Principles of general thermodynamics. RE Krieger Publishing Company.

External links

  • Heat on In Our Time at the BBC
  • Plasma heat at 2 gigakelvins – Article about extremely high temperature generated by scientists (Foxnews.com)
  • Correlations for Convective Heat Transfer – ChE Online Resources

Понравилась статья? Поделить с друзьями:
  • A word for try hard
  • A word for true to life
  • A word for to never give up
  • A word for to keep up with
  • A word for to hold on tightly