Sei sulla pagina 1di 30

Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -

Prof. Pagliano

Lecture Course
Advanced Building
Physics

Prof. Lorenzo Pagliano

Academic Year 2017/2018

Part B

⋅ Convection
⋅ Thermal Circuits for conduction and
convection,

⋅ Series and parallel configurations


⋅ Thermal Bridges

page 1
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

DRAFT

page 2
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

1 HEAT TRANSFER BY CONVECTION AND NEWTON’S POSTULATES ............... 4


2 thermal resistance: expression in the convective case ............................. 10
3 PLANE WALL WITH 3 OR MORE LAYERS, CONSTANT AND UNIFORM
TEMPERATURES OF THE FLUIDS, NO INTERNAL HEAT GENERATION .............. 13
4 La resistenza unitaria di contatto .......................................................... 21
5 HEAT FLOW IN PARALLEL .................................................................... 23
6 STEADY-STATE THERMAL TRANSMITTANCE AND ITS USE IN THE CASE OF
PARALLEL FLOWS .................................................................................... 26

page 3
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

1 HEAT TRANSFER BY CONVECTION AND


NEWTON’S POSTULATES

The natural or free convection is a phenomenon that occurs as a result of the fact that
if two points of a fluid are at different temperatures, they have also different density
and this influences the forces they are subjected, in the presence of a gravitational
field.
For example, if we heat a fluid from below, the less dense (higher temperature) part of
the fluid, subjected to hydrostatic pressure for the law of Archimedes, will move to
other areas at a lower temperature carrying with it its own internal energy. This
phenomenon occurs for example in a liquid in a container when it is heated by a flame:

Figure B.1: Liquid contained in a vessel heated by a flame

The region immediately above the flame warms up and the liquid decreases its density,
then undergoes a net force from the bottom up for the action of the gravitational field
on volumes of different density (Archimedes' principle).
The cooler liquid has a higher density and therefore moves down, taking the place of
the warmer liquid that tends to rise.
The phenomenon just described is called natural convection.
If the fluid motion is produced by a fan, pump, etc.., the phenomenon is called forced
convection.
In both cases, the energy transport called convection is related to the motion of
macroscopic portions of matter (carrying with it its own internal energy U), but also to
transport phenomena due to molecular collisions (which are always present wherever
there is matter.
page 4
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

On the contrary in the case of pure conduction (in a solid or a fluid at rest), where
there is no macroscospic motion and energy transport is solely due to molecular
collisions.

Figure B.2: Examples of natural convection

Consider a surface S of area A that separates a solid from a liquid. Let’s consider the
temperature of the liquid at a point sufficiently distant from the surface so that this
temperature can be considered not subject to influence from the wall, and let’ call it
T∞.
The temperature profile in the case of convective exchange is of the type shown in
Figure B.3. Near the wall the fluid temperature is equal to the wall surface. In the
immediate vicinity of the wall the fluid has a high thermal gradient, which diminishes
as one moves away from the wall.
Also, near the wall the fluid is stationary (no macroscopic velocity) due to friction, heat
is hence transferred by conduction through this very thin layer, while away from the
wall macroscopic portions fluid are in motion and thus the transport of energy occurs
via convection.

page 5
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

Figure B.3: convective temperature profile for exchange on a wall

The above discussion shows that the prediction of the amount of energy exchanged
between the surface and the fluid requires a detailed understanding of transport
phenomena of mass collision energy, and molecular and turbulent fluid flow and the
boundary layer. All the complexities involved in such an analysis can be merged (or if
you will, apparently hidden) in a single parameter by introducing the law or Newton's
postulate (introduced by Newton in 1701) which postulates that the heat flow by
convection is proportional to the area of solid-fluid and lapped by the temperature
difference between surface and undisturbed fluid:

Φconv = h (TS − T∞ ) A [W ] Newton’s Postulate

Where the coefficient of proportionality h is indicated by the name of convective


heat transfer coefficient.
In fact, you may also consider the above equation written as a definition of h
implied:

Φconv ⎡ W ⎤
h ≡ ⎢ m2 K ⎥
(Ts − T∞ ) A ⎣ ⎦

Since Φ conv has the dimension of power [ W ] , it will have dimensions ⎡⎣ Wm −2 K −1 ⎤⎦ ;

page 6
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

Figure B.4: Temperature curve for a laminar flow over a surface

Ricordiamo che nello strato di fluido immediatamente adiacente alla superficie (e fermo
rispetto alla superficie a causa della viscosità del fluido) si avrà scambio puramente
conduttivo.

Si potrà dunque calcolare il flusso di energia uscente dalla superficie (normalmente alla
superficie, dunque nella direzione z nel disegno) che attraversa questo strato fluido
fermo, usando il postulato di Fourier per la conduzione:

∂T
Φ = − λf A
∂z
Dove λ f è la conduttività del fluido 1, e risulta dunque:

∂T ∂T
− λf A − λf
Φ ∂z ∂z
h ≡ = =
(Ts − T∞ ) A (Ts − T∞ ) A (Ts − T∞ )

Per poter calcolare h è dunque necessario ricavare il profilo di temperatura nel fluido
∂T
adiacente la superficie (cioè la funzione T(z) da cui ricavare ), e siccome questo è
∂z
influenzato dal campo di moto del fluido (velocità in ogni punto), è in definitiva
necessario risolvere il problema fluido-meccanico completo, che consiste nel
determinare le tre componenti della velocità, la pressione, temperatura e densità del
fluido, ognuna dipendente da posizione e tempo.
Le equazioni che governano il fenomeno sono:

1
E NON del solido. Stiamo calcolando il flusso per conduzione in uno straterello di FLUIDO FERMO.

page 7
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

• le equazioni del moto del fluido (derivanti dalla legge di Newton del moto
! !
F = ma ) costituiscono tre equazioni differenziali (una per ogni delle tre direzioni
ortogonali),
• la conservazione della massa e dell’energia forniscono altre due equazioni
differenziali,
• e l’equazione di stato del fluido fornisce una sesta equazione (algebrica).

La risoluzione completa del problema richiede la soluzione del sistema costituito da


queste sei equazioni in sei incognite, con le appropriate condizioni al contorno.
Questo può essere fatto in modo analitico solo in alcuni casi di geometria semplice,
In generale occorre ricorrere a delle semplificazioni, determinando h per via empirica,
correlandolo ad altre variabile e determinando sperimentalmente i coefficienti che
compaiono nelle correlazioni, in diverse condizioni di geometria e moto.

So the coefficient h is not a thermophysical properties (i.e, a size that is not


dependent only on the physical properties of the fluid, such as density ρ, dynamic
viscosity µ (see below), specific heat at constant pressure c p , thermal conductivity λ),
it depends on a large number of factors (different from the physical properties of the
fluid) including:
• the difference in temperature between the bodies;
• in case of forced convection from the undisturbed fluid velocity w , or the
coefficient of cubic expansion β of the fluid in case of natural convection;
• the geometry of the heat exchange configuration that can sometimes require
the specification of a geometrical parameter (for example, the diameter of a
pipe, the distance between two plates ,...).
To understand intuitively the fact that h varies with the geometric configuration, as
mentioned above, all else being equal, consider the example given in Figure B.5
representing a room in a building.

Figure B.5: Hot plate on the lower temperature T1 (a) and top (b) with respect to air with T2 <T1.

page 8
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

If we suppose the floor to be warmer than air convection will happen easily if the plate
is located at the bottom since the warm fluid (hence with lower density) will tend to
rise in the gravitational field and will find no opposition from the walls/floor/ceiling.
Convection will start with more difficulty if the hot surface, at equal temperatures and
conditions of the fluid, is placed in a horizontal position but higher than the fluid (such
as a hot ceiling) because in this case the lower-density portions of fluid will be in
a higher position than the colder, high-density ones that are at the bottom. Movements
of people in the room, heat sources in the room, asymmetries,.. will anayway get
movement started but with little disturbance in the second situation the heat transfer
by convection will be lower in the second situation. Note however that the hot
plate/surface also exchanges with the environment by radiation, hence its overall
effectiveness of heat transfer should also consider this aspect, but for now here we are
discussing only the mechanism of convection.
From this example should be obvious that the phenomenon of convection depends on
the geometry and therefore it is not possible to determine the convection coefficient
given ONLY the thermo physical parameters of the fluid and surface temperatures: one
must specify also the geometry of exchange (in our example: in which position is the
hot surface).

Table B.1: Ranges of values for the convective heat transfer coefficient in W/m2K.

Air in natural convection 1-20


in forced convection 40-250
Water in natural convection 250-750
in forced convection 1000-12000
in boiling 1800-45000

page 9
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

2 THERMAL RESISTANCE : EXPRESSION IN THE CONVECTIVE

CASE

We can write:

⋅ (TS − T∞ )
Φ = h (TS − T∞ ) A =
1
[W ]
hA

Applying also in the case of thermal convection the general definition given in the
previous chapters of “thermal resistance” ℜ :

T1 − T2 ⎡ K ⎤
ℜ≡ ⎢W ⎥
Φ ⎣ ⎦
which implies:
T1 − T2
Φ=

we obtain the expression of thermal resistance in the particular case of convective
heat exchange:
1 ⎡K ⎤
ℜconv = ⎢W ⎥
hA ⎣ ⎦

Consider now an infinitesimal volume element dV, which is cut from the mathematical
surface S that separates a solid from a fluid, and suppose that the geometry and
boundary conditions are such that the temperature varies along the x axis only. Please
note that we do NOT assume stationary conditions in this discussion.

page 10
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

Figure B.6: infinitesimal element dV

We can write the energy balance for the control volume dV

dE
= ∑ φ IN + ∑ Π
! ⎡⎣W ⎤⎦
dt

If we suppose to make the control volume dV smaller and smaller, in the limit very
close to zero (mathematically: dV —> 0 ), also
• the energy production per unit time will tend to zero, since there is no volume
where to generate energy,
• the mass contained in the control volume will tend to zero, hence the energy
associated with that mass will tend to zero (and being constantly equal to zero,
its change per unit time will be zero).
Hence:

dE
= ∑ φ IN + ∑ Π
! ⇒ ∑φ IN
=0 ⎡⎣W ⎤⎦
dt

In case the heat flow is directed exactly and only in one direction (which is the case if
the geometry AND boundary conditions are such to ensure such a mono-dimensional
energy transfer as we supposed at the beginning), then the flow passes through the
vertical surfaces bounding dV and NOT across the horizontal surfaces.
Therefore, we can say that the flow of energy that enters the volume dV (in the limit:
dV —> 0 ) from the left (due to the phenomenon of conduction) must be equal in size
(and opposite in sign due to the convention “flow IN” adopted in the energy balance)
to the flow of energy that exits the volume dV to the right (due to convection).
Considering the modules of the flows they should have the same value, hence:

page 11
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

Φcondutt = Φconvett ⎡⎣W ⎤⎦ On the surface separating solid from fluid

We emphasize that this result does not require the hypothesis of stationary consitions:
dE/dt is zero, NOT because all the derivatives with respect to time are zero (which is
the case in stationary condition) but because in our case the energy E is constantly
zero since in the control volume there is no mass which can have energy associated
with it (in the limit : dV —> 0 which we are adopting ion order to represent the pure
geometrical surface separating the solid from the fluid).

So if the position the origin x=0 of the axis x in such a way that x = L is the
coordinate of the interface (mathematical surface without thickness) between fluid and
solid
Then at x = L the following relation will hold:

∂T
−λ
∂x
(
A = h A T ( L ) − T∞ ) ⎡⎣W ⎤⎦
x =L

• This holds both in stationary and non-stationary conditions (in the latter case,
both conductive and convective heat flows vary over time, but at every moment
they are equal to each other in module),

This relationship belongs to the family of boundary conditions called the third kind (or
Neumann).

page 12
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

3 PLANE WALL WITH 3 OR MORE LAYERS,


CONSTANT AND UNIFORM TEMPERATURES OF
THE FLUIDS, NO INTERNAL HEAT GENERATION
Consider now the case of a plane solid (e.g. a wall of a building) made up of several
layers of different materials (e.g plaster, bricks, insulation panels, ...). Let’s analyse
a simple case, that of a wall :

1. consisting of only two layers of homogeneous solid

2. being surrounded by two fluids, each having a constant (in time) and uniform
(in space) temperature at a certain distance from the wall

3. the initial transient is finished and the solid has reached steady state conditions
(which is possible due to Hp 2)

4. there is no heat production

Figure B.7: Wall composed of two layers of different material

Let’s indicate with 1 and 5 the areas of fluid away from the wall (hence their
temperature being supposedly undisturbed by the wall), while 2, 3, 4 denote the
surfaces of separation of different materials. Note that the boundary conditions ensure
that the heat flow is one-dimensional (i.e. the undisturbed temperature in zone 5 is
ONE value of temperature, the same value at each point along y and z axis, which we
will call T 5 ; the same for temperature in zone 1, T 1 ).
In each of the two solid materials (due to stationary conditions, the absence of heat
generation and the one-dimensional combination of geometry and boundary
conditions), the function T(x) that we obtain by integrating the Fourier equation has a

page 13
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

linear type of Ti (x) = Ai*x + Bi , with different constants Ai and Bi in the two layers
(i= 1, 2);
In many practical cases the undisturbed fluid temperatures T1 and T5 are known but
not the intermediate temperatures.
It would therefore be very useful to be able to calculate the heat flow without going
through the step of calculating the values of temperature at intermediate points.

Figure B.8: Interface between two solid layers (to be done: change dV in
C.V.)

With this objective in mind we observe that any control volume C.V. of the type
described in the figure, based on the overall hypothesis made at the beginning (1 to 4)
is in stationary conditions and no energy is produced within it, hence the energy
balance:

dE
= ∑ φ IN + ∑ Π
 ⎡⎣W ⎤⎦
dt

will take the form:

⎡⎣W ⎤⎦
∑φ IN
=0

(We note that this result is independent of the position of the plane sections 1 and 2
that limit the control volume to the right and to the left).

This condition can be rewritten as:


Φ1 = Φ2 ⎡⎣W ⎤⎦

We can hence conclude that under the assumptions:

• stationary conditions,
• no internal generation

page 14
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

• flat mono-dimensional geometry and boundary conditions (hence mono-


dimensional heat flow)

the heat flow has the same value across all planes perpendicular to the propagation of
heat, however their location is chosen.
So under the hypothesis of flat mono-dimensional geometry and boundary conditions,
stationary condition and absence of energy production, we can conclude that:

Φ1CONV = ΦCOND
23 4
= Φ3COND = Φ5CONV ≡ Φ ⎡⎣W ⎤⎦

We can write the expressions of these heat flows as a function of thermal resistance,
(this is possible since we are in stationary conditions and the absence of internal
generation, and the fact that the problem has plane symmetry, see previous chapter
“A_conduction” for the definition and use of thermal resistances).

Φ1CONV =
(T − T ) = (T − T )
1 2 1 2

ℜ1 1
h1 A

T2 − T3 T2 − T3
Φ COND
23
= =
ℜ23 s23
λ23 A

T3 − T4 T3 − T4
Φ 34 = =
COND
ℜ 34 s34
λ34 A

T4 − T5 T4 − T5
Φ 5CONV = =
ℜ5 1
h5 A

Figure B.9: Heat Flow through a multilayer wall (to be done: Change Q in Φ and make clearer the location of
the flows drawing e.g. vertical sections with dotted lines)

We rewrite the four equations in order to isolate at their first member the changes in
temperature. Taking into account that, as we have just shown, the heat flow has the
same value through each plane section perpendicular to the x axis, we obtain:

page 15
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

T1 − T2 =Φ ℜ1
T2 − T3 = Φ ℜ23
T3 − T4 = Φ ℜ34
T4 − T5 = Φ ℜ5

Summing up the above equations member to member, the unknown intermediate


temperatures cancel out and we obtain T1 − T5 = Φ ⋅ (ℜ1 + ℜ23 + ℜ34 + ℜ5 ) , where the
unknown intermediate temperatures no longer appear.
Solving for the heat flow we obtain:

Φ=
(T − T )
1 5
⎡⎣W ⎤⎦
ℜ1 + ℜ23 + ℜ 34 + ℜ5

On the other hand we can write the heat flow between points 1 and 5, introducing a
thermal resistance between the two temperature nodes 1 and 5, accordinfg to the
general definition of thermal resistance:

Φ=
(T − T )
1 5
⎡⎣W ⎤⎦
ℜ15

CONCLUSION: It thus results that the thermal resistance between two points:

• separated by a certain number of solid and fluid layers, crossed by a one-


dimensional flow that goes through them in series (one layer after the other)

• in stationary (steady-state) conditions

• in the absence of internal heat generation

is equal to the sum of the resistances of the individual layers:

⎡K⎤
ℜ15 = ℜ1 + ℜ23 + ℜ 34 + ℜ5 ⎢W ⎥
⎣ ⎦

Where the explicit expression of the resistances of individual layers are (for flat
layers 2):

s ⎡K⎤
Conductive thermal resistance: ℜ COND ≡ ⎢W ⎥
λ A ⎣ ⎦

2
Here we have considered the case of plane geometry, but we will see later that similar conclusions can
be reached with other geometries, e.g. cylindrical, with an appropriate expression of the resistance of
individual layers.

page 16
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

1 ⎡K⎤
Convective thermal resistance: ℜ CONV ≡ ⎢W ⎥
hA ⎣ ⎦

Thus, the known characteristics of the layers, we calculate the resistance of each layer
and summing gives the total resistance between points 1 and 5.
1 s s 1 ⎡K⎤
ℜ15 = + 23 + 34 + ⎢W ⎥
h1 A λ23 A λ34 A h5 A ⎣ ⎦

ONLY FOR PLANE GEOMETRY , since the area A crossed by the flow is the same for
all sections it is possible to obtain a simplification by multiplying both members of the
above equation by A.
1 s23 s34 1 ⎡ m2 K ⎤
R15 = ℜ15 ⋅ A = + + + ⎢ ⎥
h1 λ23 λ34 h5 ⎣ W ⎦

We hence obtain the conclusion that the “unit resistance R ” between the extreme
point 1 and 5 (sometimes described as the unit resistance “seen between the two
temperature nodes T 1 and T 5 ”) is the sum of the unit resistances R through which
thermal energy flows in a series.
We should note that e.g. in a cylindrical geometry, the area crossed by the flow is
not the same for the different layers and hence the concept of unit resistance R is not
useful (unit resistances in series do not add up). In a non-flat geometry only the
concept of thermal resistance ℜ might be used if appropriate (i.e. given that the
various hypothesis: stationary conditions, no heat generation, mono-dimensional flow,
are met).

In conclusion, if one knows the geometrical and thermal features of the layers and the
outer temperatures T 1 and T 5 one can therefore calculate the heat flow through the
wall without the need to know T 2 , T 3 , T 4 . We note again that this heat flow is the same
through at each section. This would NOT be the case if one of the hypotheses would
not be met.

Φ =
(T − T )
1 5
⎡⎣W ⎤⎦
ℜTOT

Once one has calculated the heat flow it is possible, however, to derive the values of
the temperatures in any section. For example, since:

T1 − T2 = Φ i ℜ1

it follows: T2 = T1 − Φ i ℜ1

page 17
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

which we may write as:


T1 − T2 = Φ i ℜ1
We note that from the above expression follows that the difference in temperature
between two boundary surfaces of a certain layer is directly proportional to the
thermal resistance ℜ offered by that layer.
Since the value of the heat flow is the same in all layers and the temperature jump
is directly proportional to thermal resistance ℜ , it follows that, given the total
temperature difference ∆" = "$ − "& , this ∆" will distribute among the layers, the ones
with higher thermal resistance ℜ will show a higher temperature jump. E.g. in the wall
of a building in winter, the temperature will drop from 20°C to -10°C and the
insulating layer will be the one across which there will be the higher temperature drop.

T1 − T2 = Φ i ℜ1

Figure B.10: temperature as a function of position in the multilayer wall. (to do:REDRAW 1) in general,
best would be to assume realistic values for resistances and draw the exact temperature profile)

page 18
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

As an example of the implication of the above, replacing a single glazing unit with a
double glazing unit (DGU) and double glazing unit filled with a low-conductivity gas
(e.g. Argon), increases thermal resistance and has in winter conditions two main
thermal effects:
• reduction the heat flow between indoor and outdoor spaces
• increase the surface temperature of the glass surface facing indoor space (because
increasing the thermal resistance of the window increases the ∆" across it) .

A similar result is achieved in summer, in the reverse direction as for temperatures, if


one considers a window not subject to direct sun radiation (the case of presence of
sun radiation and how to protect the indoor from excessive solar gain will be treated in
a next chapter).

So improving the thermal performance of glazed surfaces in a building at the same


time saves energy and increases the comfort level in areas close to windows, which
further increase the usability of the space. In a building service industry where space
has a measurable economic value (e.g. commercial value and rent are proportional to
m 2 ), the payback time in the improvement of the frames should be assessed taking
into account the energy savings and the economic value of space where comfort has
been improved.
There are other improvements to be considered in the overall cost-benefit calculation,
e.g. the reduction of sound transmission, which will result by going from single to
double or triple glazing, and possibly reducing air infiltration through improvements of
the window frame).

page 19
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

Figure B.11: Single glazing (a), double glazing (b), double glazing with argon filled cavity (c) and their
surface temperatures of the inner face.

page 20
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

4 L A RESISTENZA UNITARIA DI CONTATTO

Nello studio della conduzione termica attraverso solidi multistrato, si è ipotizzato


all’interfaccia tra due strati in contatto tra loro un “contatto perfetto”, cioè senza
variazione di temperatura. Un contatto perfetto si ha con superfici perfettamente lisce
mentre in realtà anche superfici apparentemente lisce sono ruvide, con numerosi picchi
e cavità.
In pratica, quando due superfici sono pressate una contro l’altra, i picchi assicurano un
buon contatto, mentre le cavità costituiscono dei vuoti riempiti d’aria. Per la presenza
di cavità, che si comportano da isolante a causa della bassa conducibilità termica
dell’aria, l’interfaccia offre una certa resistenza alla trasmissione di calore per
conduzione, detta resistenza termica di contatto ℜc . I valori di ℜc determinati
sperimentalmente mostrano una notevole dispersione a causa della difficoltà nella
caratterizzazione delle superfici. Si può osservare, comunque, che la resistenza di
contatto diminuisce al diminuire della rugosità superficiale e all’aumentare della
pressione all’interfaccia. Si osservi, inoltre, che, nota la potenza termica Q trasmessa
attraverso un certo strato, la resistenza termica di contatto può essere determinata
misurando la differenza di temperatura delta T all’interfaccia e moltiplicandola per la
potenza termica Q . La maggior parte dei valori sperimentali della resistenza termica
unitaria di contatto è compresa tra 0,00001 e 0,001 m 2 K/W. La resistenza termica di
contatto può essere ridotta usando un liquido termicamente conduttivo, quale olio al
silicone, che viene applicato sulle superfici prima che vengano compresse una contro
l’altra. Questo accorgimento si adotta comunemente quando si incollano componenti
elettronici tipo transistori di potenza ai dissipatori.

page 21
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

Figura B.12: ingrandimento interfaccia superfici.

Come già detto non è solitamente possibile quantificare analiticamente in modo preciso
il valore della resistenza di contatto, di conseguenza si ricava caso per caso un suo
valore per via sperimentale.

In edilizia la resistenza di contatto data la sua scarsa influenza può essere trascurata.
Una muratura semplice di mattoni pieni 180x140x60, ad esempio, presenta una
resistenza termica di 0,18 m 2 K/W (da UNI 10355) che è molto maggiore del valore
massimo della resistenza di contatto, 0,001 m 2 K/W.

page 22
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

5 HEAT FLOW IN PARALLEL

It can be shown that even in the case of parallel flows the problem can be represented
via a model of heating circuits, under precise.

Consider, for example, a portion of a building where part of a wall is formed by bricks
and part of other materials with a different conductivity, and the surfaces at x = 0 and
x = L are at uniform temperatures (respectively T 1 , at x = 0 and T 2 at x = L).

Under certain conditions (in particular the hypothesis that the thermal conductivity are
not too different in different areas, as we will see) it can be assumed that the heat
flow occurs along the x axis and not along y and z directions, that can still be treated
as one-dimensional flow. This is equivalent to assuming that the contact interfaces
between the materials A, B and C behave as perfectly adiabatic surfaces.

Figure B.13: Wall with different materials in parallel with respect to the direction of the
heat flow and uniform temperatures on the surfaces x = 0 and x = L. (to do: eliminate
the temperature T B )

In this case the equivalent thermal circuit diagram becomes:

page 23
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

Figure B.14: Thermal circuit assuming uniform temperatures on the surfaces x = 0 and x = L.

If the boundary conditions are constant over time and there is no internal heat
generation, in each of the areas the one-dimensional flow (ie only in the direction of
the x) satisfies the equations, respectively:
T1 − T2 T1 − T2 T1 − T2
ΦA = ; ΦB = ; ΦC = [W]
ℜA ℜB ℜC
Under this assumptions, the total flow through the wall will be the sum of the three
streams "in parallel", and will hence be:
T1 − T2 T1 − T2 T1 − T2
Φ 1−2 = Φ A + Φ B + ΦC = + + =
ℜA ℜB ℜC

(
⎛ 1
= T1 − T2 ⎜ +)1
+
1 ⎞
=
(T − T )
1 2
=
(T − T )
1 2
⎡⎣W ⎤⎦

⎝ℜA ℜB ℜC⎠ 1 ℜ eq
⎛ 1 1 1 ⎞
⎜ℜ +ℜ +ℜ ⎟
⎝ A B C⎠

Having defined the equivalent resistance between the temperatures T 1 and T 2 in the
usual way, one obtains:
1
ℜeq =
⎛ 1 1 1 ⎞
⎜ + + ⎟
⎝ ℜA ℜB ℜC⎠
or

1 ⎛ 1 1 1 ⎞ ⎡K⎤
=⎜ + + ⎟ ⎢W ⎥
ℜ eq ⎝ ℜ A ℜ B ℜ C ⎠ ⎣ ⎦

That is the inverse of the equivalent resistance between the two isothermal surfaces T 1
and T 2 , is obtained by adding the inverse of the resistances that are seen in parallel
between the two temperatures T 1 and T 2.
If we define the “thermal conductance” as the inverse of thermal resistance ℜ and
denote it by G (sometimes denoted by C, but here we prefer to reserve the letter C to
denote the heat capacity), we can write:
page 24
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

1 ⎡W ⎤
GA ≡ ⎢K⎥
ℜA ⎣ ⎦

Φ 1−2 = Φ A + Φ B + ΦC = (T − T ) ( G
1 2 A
+ GB + GC ) = (T − T ) G
1 2

Therefore we can conclude that in cases of materials in parallel the thermal


conductances add up to the global conductance of the entire wall.
In this course we will try (for economy of language) to use mainly the thermal
resistances ℜ and only rarely conductances, in order to reduce the number of
variables to be used.

In cases where parts of the thermal circuit are in parallel and parts are in series one
should apply the rules seen before following a chose direction, e.g. from left to right.

Figure B.15: Example of series-parallel hybrid case (to do in (b): eliminate the middle line between the RF and
RG)

There are two different modelling choices as to the position of the adiabatic and
isothermal surfaces, which can lead to slightly different results. In the figure above the
choice has been made to consider the sections 1, 2, 3, 4, 5 as if they were each
isotherm. See exercises for more details.

page 25
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

6 STEADY-STATE THERMAL TRANSMITTANCE AND


ITS USE IN THE CASE OF PARALLEL FLOWS

In many cases it is usual to express the heat flow from the environment on the one side
of a body to the environment on the other side where the two environments have constant
temperatures T 1 and T 5 ,
in the following way:

(
Φ = U A T1 − T5 ) ⎡⎣W ⎤⎦

ϕ = U (T1 − T5 ) ⎡W ⎤
⎣⎢ m2 ⎦⎥

Where U is named “thermal transmittance” and is implicitly defined by the above


equations. We note that the concept of thermal transmittance (also called U-value) is
defined in stationary condition . We will discuss later how to describe and calculate
heat transfer under dynamic conditions.
The explicit definition of thermal transmittance is therefore:

Φ ⎡W ⎤
U≡
(
A T1 − T5 ) ⎣⎢ m2 K ⎦⎥

or
ϕ ⎡W ⎤
U ≡
(T1 − T5 ) ⎣⎢ m2 K ⎦⎥

ASHRAE gives the following definition, which coincides with the one just presented.
Note the slightly different terminology. The last sentence about film conductance
underlines that U is used in case where we consider also the surrounding
environments, where convection (and radiation) are the heat transfer mechanisms.

“Thermal transmittance U is the heat flux 3 under steady-state conditions from the
environment on the one side of a body to the environment on the other side, per unit
temperature difference between the two environments, in W/(m 2 K). Thermal
transmittance is sometimes called the overall coefficient of heat transfer or U-factor .
Thermal transmittance includes surface film conductance .“

3
which we call heat flow rate in this course
page 26
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

The relationship of the steady state thermal transmittance U (or U-value) with the
thermal resistance and the unit resistance (when it is defined, i.e., in plane geometry)
can be derived by comparison with their implicit definition given by the equations

T1 − T5
Φ =
ℜ15
(
= U A T1 − T5 ) ⎡⎣W ⎤⎦

T1 − T5
ϕ=
R15
(
= U T1 − T5 ) ⎡W ⎤
⎣⎢ m2 ⎦⎥

From which:
1 1 ⎡W ⎤
U15 = =
R15 ℜ15 A ⎣⎢ m2 K ⎦⎥

Where (only in case of plane geometry):

1 s23 s34 1 ⎡ m2 K ⎤
R15 = ℜ15 A = + + + ⎢ ⎥
h1 λ23 λ34 h5 ⎣ W ⎦

In cases where it is possible to model the process by heat flows in parallel it is


convenient to make the calculations using the thermal transmittance. For example,
in the case of a window made of frame of thermal transmittance U f and area A f and a
glass of transmittance U g and area A g , if we assume one-dimensional flow between the
same two temperature nodes (indoor air temperature T i and outdoor air temperature
T e , if we are neglecting heat transfer by radiation. We will see in the following
chapters that when radiation cannot be neglected, the temperature to be considered
will be different and h will be defined and calculate by taking into account both
convection and radiation).

page 27
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

( ) ( ) (
Φ = U w ( Ag + A f ) Ti − Te =U g Ag Ti − Te + U f A f Ti − Te )
From which we obtain the expression of the thermal transmittance of the window U w :

U w ( Ag + A f ) = U g Ag + U f A f
And finally:
U g Ag + U f A f ⎡W ⎤
Uw =
Ag + A f ⎢⎣ m2 ⎥⎦

Note the range of values that can be realized


Window: from about 6 to about 1 W/m 2 K
Opaque structures: from about 1,5 to 0,1 W/m 2 K
So the heat flow going through the building envelope can be reduced by a factor 5 to
10. In case we act in a similar way also on ventilation losses, a reduction of energy
needs for heating and cooling of a factor 5 to 10 is possible.

from Wikipedia, to be checked and improved:


“Thermal transmittances of most walls and roofs can be calculated using ISO 6946, unless there
is metal bridging the insulation in which case it can be calculated using ISO 10211. For most ground
floors it can be calculated using ISO 13370. For most windows the thermal transmittance can be
calculated using ISO 10077 or ISO 15099. ISO 9869 describes how to measure the thermal
transmittance of a structure experimentally.

page 28
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

Typical thermal transmittance values for common building structures are as follows:
• single glazing: 5.7 W/m²K;
• single glazed windows, allowing for frames: 4.5 W/m²·K;
• double glazed windows, allowing for frames: 3.3 W/m²·K;
• double glazed windows with advanced coatings: 2.2 W/m²·K;
• double glazed windows with advanced coatings and frames: 1.2 W/m²·K;
• triple glazed windows, allowing for frames: 1.8 W/m²·K;
• triple glazed windows, with advanced coatings and frames: 0.8 W/m²·K;
• well-insulated roofs: 0.15 W/m²·K;
• poorly insulated roofs: 1.0 W/m²·K;
• well-insulated walls: 0.25 W/m²·K;
• poorly insulated walls: 1.5 W/m²·K;
• well-insulated floors: 0.2 W/m²·K;
• poorly insulated floors: 1.0 W/m²·K;
In practice the thermal transmittance is strongly affected by the quality of workmanship and if
insulation is fitted poorly, the thermal transmittance can be considerably higher than if insulation is
fitted well. “

From: Comfort and Energy Efficiency Recommendations for Net Zero Energy
Buildings
Sartori, J. Candanedo , S. Geier, R. Lollini, F. Garde, A. Athienitis, L. Pagliano

page 29
Parte B 2017/11/29 DRAFT Course of Advanced Building Physics -
Prof. Pagliano

page 30

Potrebbero piacerti anche