You are on page 1of 5

Newton's law of cooling

Newton's law of cooling states that the rate of heat loss of a body is directly proportional to the difference in the temperatures
between the body and its surroundings provided the temperature difference is small and the nature of radiating surface remains same.
As such, it is equivalent to a statement that the heat transfer coefficient, which mediates between heat losses and temperature
differences, is a constant. This condition is generally true in thermal conduction (where it is guaranteed by Fourier's law), but it is
often only approximately true in conditions of convective heat transfer, where a number of physical processes make effective heat
transfer coefficients somewhat dependent on temperature differences. Finally, in the case of heat transfer by thermal radiation,
Newton's law of cooling is not true.

Sir Isaac Newton did not originally state his law in the above form in 1701, when it was originally formulated. Rather, using today's
terms, Newton noted after some mathematical manipulation that the rate of temperature change of a body is proportional to the
difference in temperatures between the body and its surroundings. This final simplest version of the law given by Newton himself,
was partly due to confusion in Newton's time between the concepts of heat and temperature, which would not be fully disentangled
until much later.[1]

When stated in terms of temperature differences, Newton's law (with several further simplifying assumptions, such as a low Biot
number and temperature-independent heat capacity) results in a simple differential equation for temperature-difference as a function
of time. This equation has a solution that specifies a simple negative exponential rate of temperature-difference decrease, over time.
This characteristic time function for temperature-dif
ference behavior, is also associated with Newton's law of cooling.

Contents
Relationship to mechanism of cooling
Heat transfer version of the law
The Biot number
Rate-of-change of temperature-difference version of the law
Temperature function-of-time solution in terms of object heat capacity
See also
References
External links

Relationship to mechanism of cooling


Convection-cooling is sometimes called "Newton's law of cooling." This use is based on a work by Isaac Newton published
Philosophical Transactions, 1701.[2]
anonymously as "Scala graduum Caloris. Calorum Descriptiones & signa." in

In cases where the heat transfer coefficient is independent, or relatively independent, of the temperature difference between object
and environment, Newton's law is followed. This independence is sometimes the case, but is not guaranteed to be so. The heat
transfer coefficient is often relatively independent of temperature in purely conduction-type cooling, but becomes a function of the
temperature in classical natural convective heat transfer. In this case, Newton's law only approximates the result when the
temperature changes are relatively small. Newton himself realized this limitation.

A correction to Newton's law concerning larger temperature differentials was made in 1817 by Dulong and Petit.[3] (These men are
better-known for their formulation of theDulong–Petit law concerning the molar specific heat capacity of a crystal.)

Another situation with temperature-dependent transfer coef


ficient is radiative heat transfer, which also does not obey Newton's law.
Heat transfer version of the law
The heat-transfer version of Newton's law, which (as noted) requires a constant heat transfer coefficient, states that the rate of heat
loss of a body is proportional to the difference in temperatures between the body and its surroundings.

[4]
The rate of heat transfer in such circumstances is derived below:

Newton's cooling law in convection is a restatement of the dif


ferential equation given byFourier's law:

where

is the thermal energy (SI unit: joule)


is the heat transfer coefficient (assumed independent of T here) (SI unit: W/(m2 K))
is the heat transfer surface area (SI unit: m2)
is the temperature of the object's surface and interior (since these are the same in this
approximation) (SI unit: K)
is the temperature of the environment; i.e. the temperature suitably far from the surface
(SI unit: K)
is the time-dependent thermal gradient between environment and
object (SI unit: K)

The heat transfer coefficient h depends upon physical properties of the fluid and the physical situation in which convection occurs.
Therefore, a single usable heat transfer coefficient (one that does not vary significantly across the temperature-difference ranges
covered during cooling and heating) must be derived or found experimentally for every system that can be analyzed using the
presumption that Newton's law will hold.

Formulas and correlations are available in many references to calculate heat transfer coefficients for typical configurations and fluids.
For laminar flows, the heat transfer coefficient is rather low compared to turbulent flows; this is due to turbulent flows having a
thinner stagnant fluid film layer on the heat transfer surface.[5] However, note that Newton's law breaks down if the flows should
transition between laminar or turbulent flow, since this will change the heat transfer coefficient h which is assumed constant in
solving the equation.

The Biot number


The Biot number, a dimensionless quantity, is defined for a body as:

where:

h = film coefficient or heat transfer coefficient or convective heat transfer coefficient


LC = characteristic length, which is commonly defined as the volume of the body divided by the surface area of the

body, such that

kb = Thermal conductivity of the body


The physical significance of Biot number can be understood by imagining the heat flow from a hot metal sphere suddenly immersed
in a pool, to the surrounding fluid. The heat flow experiences two resistances: the first at the surface of the sphere, and the second
within the solid metal (which is influenced by both the size and composition of the sphere). The ratio of these resistances is the
dimensionless Biot number.
If the thermal resistance of the fluid/sphere interface exceeds that thermal resistance offered by the interior of the metal sphere, the
Biot number will be less than one. For systems where it is much less than one, the interior of the sphere may be presumed always to
have the same temperature, although this temperature may be changing, as heat passes into the sphere from the surface. The equation
to describe this change in (relatively uniform) temperature inside the object, is the simple exponential one described in Newton's law
of cooling expressed in terms of temperature difference (see below).

In contrast, the metal sphere may be large, causing the characteristic length to increase to the point that the Biot number is larger than
one. In this case, thermal gradients within the sphere become important, even though the sphere material is a good conductor.
Equivalently, if the sphere is made of a thermally insulating (poorly conductive) material, such as wood or styrofoam, the interior
resistance to heat flow will exceed that of the fluid/sphere boundary, even with a much smaller sphere. In this case, again, the Biot
number will be greater than one.

Values of the Biot number smaller than 0.1 imply that the heat conduction inside the body is much faster than the heat convection
away from its surface, and temperature gradients are negligible inside of it. This can indicate the applicability (or inapplicability) of
certain methods of solving transient heat transfer problems. For example, a Biot number less than 0.1 typically indicates less than 5%
error will be present when assuming a lumped-capacitance model of transient heat transfer (also called lumped system analysis).[6]
Typically, this type of analysis leads to simple exponential heating or cooling behavior ("Newtonian" cooling or heating) since the
amount of thermal energy (loosely, amount of "heat") in the body is directly proportional to its temperature, which in turn determines
the rate of heat transfer into or out of it. This leads to a simple first-order differential equation which describes heat transfer in these
systems.

Having a Biot number smaller than 0.1 labels a substance as "thermally thin," and temperature can be assumed to be constant
throughout the material's volume. The opposite is also true: A Biot number greater than 0.1 (a "thermally thick" substance) indicates
that one cannot make this assumption, and more complicated heat transfer equations for "transient heat conduction" will be required
to describe the time-varying and non-spatially-uniform temperature field within the material body. Analytic methods for handling
these problems, which may exist for simple geometric shapes and uniform material thermal conductivity, are described in the article
on the heat equation.

Rate-of-change of temperature-difference version of the law


As noted in the section above, accurate formulation for temperatures may require analysis based on changing heat transfer
coefficients at different temperatures, a situation frequently found in free-convection situations, and which precludes accurate use of
Newton's law.

As noted above, Newton's law behavior when stated in terms of temperature change in the body, also requires that internal heat
conduction within the object be large in comparison to the loss/gain of heat by surface transfer (conduction and/or convection), which
is the condition where the Biot number is less than about 0.1. This allows the presumption of a single "temperature" inside the body
(as a function of time) to make sense, as otherwise the body would have many different temperatures inside it, at any one time. This
single temperature will generally change exponentially
, as time progresses (see below).

Assumption of rapid internal conduction also allows use of the so-called lumped capacitance model. In this model, the amount of
thermal energy in the body is calculated by assuming a constant heat capacity and thus thermal energy in the body is assumed to be a
linear function of the body's temperature.

Temperature function-of-time solution in terms of object heat capacity


Given that the body is treated as a lumped capacitance thermal energy reservoir with a (total thermal energy content) which is
proportional to (simple total heat capacity) and (the temperature of the body), then . It is expected that the system will
experience exponential decay in the temperature difference of body and surroundings as a function of time. This is proven in the
following sections:
From the definition of heat capacity comes the relation . Differentiating this equation with regard to time gives the
identity (valid so long as temperatures in the object are uniform at any given time): . This expression may be
used to replace in the first equation which begins this section, above. Then, if is the temperature of such a body at time
, and is the temperature of the environment around the body:

where is a positive constant characteristic of the system, which must be in units of , and is therefore sometimes
expressed in terms of a characteristic time constant given by: . Thus, in thermal systems,
. (The total heat capacity of a system may be further represented by its mass-specific heat capacity multiplied by
its mass , so that the time constant is also given by ).

The solution of this differential equation, by standard methods of integration and substitution of boundary conditions, gives:

If:

is defined as : where is the initial temperature difference


at time 0,

then the Newtonian solution is written as:

This same solution is more immediately apparent if the initial differential equation is written in terms of , as a single function
of time to be found, or "solved for."

See also
Thermal transmittance
List of thermal conductivities
Convection diffusion equation
R-value (insulation)
Heat pipe
Fick's law of diffusion
Relativistic heat conduction
Thermomass theory
Churchill–Bernstein equation
Fourier number
Biot number
False diffusion

References
1. History of Newton's cooling law(http://paginas.fisica.uson.mx/laura.yeomans/tc/Sci-Edu-Springer-2010.pdf)
Archived
(https://web.archive.org/web/20150614122639/http://paginas.fisica.uson.mx/laura.yeomans/tc/Sci-Edu-Springer-201
0.pdf) 2015-06-14 at the Wayback Machine.
2. 824 (https://books.google.com/books?id=x8NeAAAAcAAJ&pg=P A824#v=onepage&q&f=false)–829; ed. Joannes
Nichols, Isaaci Newtoni Opera quae exstant omnia, vol. 4 (1782), 403 (https://books.google.com/books?id=Dz2FzJq
aJMUC&pg=PA403)–407.
3. Whewell, William (1866).History of the Inductive Sciences from the Earliest to the Present imes.
T
4. Louis C. Burmeister, (1993) “Convective Heat Transfer”, 2nd ed. Publisher Wiley-Interscience, p 107ISBN 0-471-
57709-X, 9780471577096
5. http://www.engineersedge.com/heat_transfer/convection.htm Engineers Edge, 2009, “Convection Heat
Transfer”,Accessed 20/03/09
6. Frank Incropera; Theodore L. Bergman; David DeWitt; Adrienne S. Lavine (2007).
Fundamentals of Heat and Mass
Transfer (6th ed.). John Wiley & Sons. pp. 260–261. ISBN 978-0-471-45728-2.

See also:

Dehghani, F 2007, CHNG2801 – Conservation and rTansport Processes: Course Notes, University of Sydney ,
Sydney
John H Lienhard IV and John H Lienhard V
, 'A Heat Transfer Textbook', Third Edition, Phlogyston Press, Cambridge
Massachusetts [1]

External links
Heat conduction - Thermal-FluidsPedia
Newton's Law of Coolingby Jeff Bryant based on a program byStephen Wolfram, Wolfram Demonstrations Project.

Retrieved from "https://en.wikipedia.org/w/index.php?title=Newton%27s_law_of_cooling&oldid=866336777


"

This page was last edited on 29 October 2018, at 19:02(UTC).

Text is available under theCreative Commons Attribution-ShareAlike License ; additional terms may apply. By using this
site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of theWikimedia
Foundation, Inc., a non-profit organization.

You might also like