Vous êtes sur la page 1sur 7

Fundamentals of Steady Flow thermodynamics Malcolm J.

McPherson

Chapter 3. Fundamentals of steady flow thermodynamics

3.1. INTRODUCTION ...........................................................................................1


3.2. PROPERTIES OF STATE, WORK AND HEAT .............................................2
3.2.1 Thermodynamic properties. State of a system....................................................................... 2
3.2.2 Work and heat ........................................................................................................................ 3
3.3. SOME BASIC RELATIONSHIPS ...................................................................4
3.3.1. Gas laws and gas constants.................................................................................................. 4
3.3.2. Internal Energy and the First Law of Thermodynamics......................................................... 7
3.3.3. Enthalpy and the Steady Flow Energy Equation ................................................................... 8
3.3.4. Specific heats and their relationship to gas constant .......................................................... 10
3.3.5. The Second Law of Thermodynamics................................................................................. 13
3.4. FRICTIONAL FLOW ....................................................................................14
3.4.1. The effects of friction in flow processes .............................................................................. 14
3.4.2. Entropy ................................................................................................................................ 16
3.4.3. The adiabatic and isentropic processes .............................................................................. 19
3.4.4. Availability............................................................................................................................ 20
3.5 THERMODYNAMIC DIAGRAMS..................................................................23
3.5.1. Ideal isothermal (constant temperature) compression. ....................................................... 24
3.5.2. Isentropic (constant entropy) compression ......................................................................... 26
3.5.3. Polytropic compression ....................................................................................................... 28

Further Reading ................................................................................................34

____________________________________________________________________________

3.1. INTRODUCTION

The previous chapter emphasized the behaviour of incompressible fluids in motion. Accordingly, the
analyses were based on the mechanisms of fluid dynamics. In expanding these to encompass
compressible fluids and to take account of thermal effects, we enter the world of thermodynamics.

This subject divides into two major areas. Chemical and statistical thermodynamics are concerned
with reactions involving mass and energy exchanges at a molecular or atomic level, while physical
thermodynamics takes a macroscopic view of the behaviour of matter subjected to changes of
pressure, temperature and volume but not involving chemical reactions. Physical thermodynamics
subdivides further into the study of "closed" systems within each of which remains a fixed mass of
material such as a gas compressed within a cylinder, and "open" systems through which material
flows. A subsurface ventilation system is, of course, an open system with air continuously entering
and leaving the facility. In this chapter, we shall concentrate on open systems with one further
restriction - that the mass flow of air at any point in the system does not change with time. We may,
then, define our particular interest as one of steady flow physical thermodynamics.

Thermodynamics began to be developed as an engineering discipline after the invention of a


practicable steam engine by Thomas Newcomen in 1712. At that time, heat was conceived to be a
massless fluid named 'caloric' that had the ability to flow from a hotter to a cooler body.
Improvements in the design and efficient operation of steam engines, made particularly by James
Watt (1736-1819) in Scotland, highlighted deficiencies in this concept. During the middle of the 19th
century the caloric theory was demolished by the work of James P. Joule (1818-1889) in England,

3- 1
Fundamentals of Steady Flow thermodynamics Malcolm J. McPherson

H.L.F. Helmholtz (1821-1894) and Rudolph J.E. Clausius (1822-1888) in Germany, and Lord
Kelvin (1824-1907) and J.C. Maxwell (1831-1879) of Scotland.

The application of thermodynamics to mine ventilation systems was heralded by the publication of a
watershed paper in 1943 by Frederick B. Hinsley (1900-1988). His work was motivated by
consistent deviations that were observed when mine ventilation surveys were analyzed using
incompressible flow theory, and by Hinsley's recognition of the similarity between plots of pressure
against specific volume constructed from measurements made in mine downcast and upcast shafts,
and indicator diagrams produced by compressed air or heat engines. The new thermodynamic
theory was particularly applicable to the deep and hot mines of South Africa. Mine ventilation
engineers of that country have contributed greatly to theoretical advances and practical utilization of
the more exact thermodynamic methods.

3.2. PROPERTIES OF STATE, WORK AND HEAT


3.2.1 Thermodynamic properties. State of a system.

In Chapter 2 we introduced the concepts of fluid density and pressure. In this chapter we shall
consider the further properties of temperature, internal energy, enthalpy and entropy. These will be
introduced in turn and where appropriate. For the moment, let us confine ourselves to temperature.

Reference to the temperature of substances is such an everyday occurrence that we seldom give
conscious thought to the foundations upon which we make such measurements. The most common
basis has been to take two fixed temperatures such as those of melting ice and boiling water at
standard atmospheric pressure, ascribe numerical values to those temperatures and to define a
scale between the two fixed points. Anders Celsius (1701-1744), a Swedish astronomer, chose to
give values of 0 and 100 to the temperatures of melting ice and boiling water respectively, and to
select a linear scale between the two. The choice of a linear scale is, in fact, quite arbitrary but leads
to a convenience of measurement and simpler relationships between temperature and other
thermodynamic properties. The scale thus defined is known as the Celsius (or Centigrade) scale .
The older Fahrenheit scale was named after Gabriel Fahrenheit (1686-1736), the German scientist
who first used a mercury-in-glass thermometer. Fahrenheit's two "fixed" but rather inexact points
were 0 for a mixture of salt, ice and water, and 96 for the average temperature of the human body. A
linear scale was then found to give Fahrenheit temperatures of 32 for melting ice and 212 for boiling
water. These were later chosen as the two fixed points for the Fahrenheit scale but, unfortunately,
the somewhat irrational numeric values were retained. In the SI system of units, temperatures are
most often related to degrees Celsius.

However, through a thermodynamic analysis, another scale of temperature can be defined that does
not depend upon the melting or boiling points of any substance. This is called the absolute or
thermodynamic temperature scale. N.L. Sadi Carnot (1796-1832), a French military engineer,
showed that a theoretical heat engine operating between fixed inlet and outlet temperatures
becomes more efficient as the difference between those two temperatures increases. Absolute zero
on the thermodynamic temperature scale is defined theoretically as that outlet temperature at which
an ideal heat engine operating between two fixed temperature reservoirs would become 100 per
cent efficient, i.e. operate without producing any reject heat. Absolute zero temperature is a
theoretical datum that can be approached but never quite attained. We can then choose any other
fixed point and interval to define a unit or degree on the absolute temperature scale. The SI system
of units employs the Celsius degree as the unit of temperature and retains 0 °C and 100 °C for
melting ice and boiling water. This gives absolute zero as -273.15 °C. Thermodynamic temperatures
quoted on the basis of absolute zero are always positive numbers and are measured in degrees
Kelvin (after Lord Kelvin). A difference of one degree Kelvin is equivalent to a difference of one
Celsius degree. Throughout this book, absolute temperatures are identified by the symbol T and
temperatures shown as t or θ denote degrees Celsius.

T = t(°C) + 273.15 K (3.1)

3- 2
Fundamentals of Steady Flow thermodynamics Malcolm J. McPherson

The Kelvin units, K, are normally shown without a degree (°) sign. Thermodynamic calculations
should always be conducted using the absolute temperature, T, in degrees Kelvin. However,
as a degree Kelvin is identical to a degree Celsius, temperature differences may be quoted in either
unit.

The state of any point within a system is defined by the thermodynamic properties of the fluid at that
point. If air is considered to be a pure substance of fixed composition then any two independent
properties are sufficient to define its thermodynamic state. In practice, the two properties are often
pressure and temperature as these can be measured directly. If the air is not of fixed composition
as, for example, in airways where evaporation or condensation of water occurs, then at least one
more property is required to define its thermodynamic state (Chapter 14).

The intensive or specific properties of state (quoted on the basis of unit mass) define completely the
thermodynamic state of any point within a system or subsystem and are independent of the
processes which led to the establishment of that state.

3.2.2 Work and heat

Both work and heat involve the transfer of energy. In SI units, the fundamental numerical
equivalence of the two is recognized by their being given the same units, Joules, where

1 Joule = 1 Newton x 1 metre. Nm

Work usually (but not necessarily) involves mechanical movement against a resisting force.
An equation used repeatedly in Chapter 2 was

Work done = force x distance

or dW = F dL Nm or J (3.2)

and is the basis for the definition of a Joule. Work may be added as mechanical energy from an
external source such as a fan or pump. Additionally, it was shown in Section 2.3.1. that "flow work",
Pv (J) must be done to introduce a plug of fluid into an open system. However, it is only at entry (or
exit) of the system that the flow work can be conceived as a measure of force x distance. Elsewhere
within the system the flow work is a point function. It is for this reason that some engineers prefer not
to describe it as a work term.

Heat is transferred when an energy exchange takes place due to a temperature difference. When
two bodies of differing temperatures are placed in contact then heat will "flow" from the hotter to the
cooler body. (In fact, heat can be transferred by convection or radiation without physical contact.) It
was this concept of heat flowing that gave rise to the caloric theory. Our modern hypothesis is that
heat transfer involves the excitation of molecules in the receiving substance, increasing their internal
kinetic energy at the expense of those in the emitting substance.

Equation (3.2) showed that work can be described as the product of a driving potential (force) and
distance. It might be expected that there is an analogous relationship for heat, dq, involving the
driving potential of temperature and some other property. Such a relationship does, in fact, exist and
is quantified as

dq = T ds J (3.3)

The variable, s, is named entropy and is a property that will be discussed in more detail in Section
3.4.2.

3- 3
Fundamentals of Steady Flow thermodynamics Malcolm J. McPherson

It is important to realize that neither work nor heat is a property of a system. Contrary to popular
phraseology which still retains reminders of the old caloric theory, no system "contains" either heat
or work. The terms become meaningful only in the context of energy transfer across the boundaries
of the system. Furthermore, the magnitude of the transfer depends upon the process path or
particular circumstances existing at that time and place on the boundary. Hence, to be precise, the
quantities dW and dq should actually be denoted as inexact differentials δW and δq.

The rate at which energy transfers take place is commonly expressed in one of two ways. First, on
the basis of unit mass of the fluid, i.e. Joules per kilogram. This was the method used to dimension
the terms of the Bernoulli equation in Section 2.3.1. Secondly, an energy transfer may be described
with reference to time, Joules per second. This latter method produces the definition of Power.

dW dq J
Power = or (3.4)
time time s

where the unit, J/s is given the name Watt after the Scots engineer, James Watt.

Before embarking upon any analyses involving energy transfers, it is important to define a sign
convention. Many textbooks on thermodynamics have used the rather confusing convention that
heat transferred to a system is positive, but work transferred to the system is negative. This strange
irrationality has arisen from the historical development of physical thermodynamics being motivated
by the study of heat engines, these consuming heat energy (supplied in the form of a hot vapour or
burning fuel) and producing a mechanical work output. In subsurface ventilation engineering, work
input from fans is mechanical energy transferred to the air and, in most cases, heat is transferred
from the surrounding strata or machines - also to the air. Hence, in this engineering discipline it is
convenient as well as being mathematically consistent to regard all energy transfers to the air as
being positive, whether those energy transfers are work or heat. That is the sign convention utilized
throughout this book.

3.3. SOME BASIC RELATIONSHIPS


3.3.1. Gas laws and gas constants

An ideal gas is one in which the volume of the constituent molecules is zero and where there are no
inter-molecular forces. Although no real gas conforms exactly to that definition, the mixture of gases
that comprise air behaves in a manner that differs negligibly from an ideal gas within the ranges of
temperature and pressure found in subsurface ventilation engineering. Hence, the thermodynamic
analyses outlined in this chapter will assume ideal gas behavior.

Some twenty years before Isaac Newton's major works, Robert Boyle (1627-1691) developed a
vacuum pump and found, experimentally, that gas pressure, P, varied inversely with the volume, v,
of a closed system at constant temperature.
1
P ∝ (Boyle’s law) (3.5)
v
where ∝ means 'proportional to'.

In the following century and on the other side of the English Channel in France, Jacques A.C.
Charles (1746-1823) discovered, also experimentally, that

v ∝T (Charles' law) (3.6)

for constant pressure, where T = absolute temperature.

3- 4
Fundamentals of Steady Flow thermodynamics Malcolm J. McPherson

Combining Boyle's and Charles' laws gives

Pv ∝ T
where both temperature and pressure vary. Or, inserting a constant of proportionality, R',

Pv = R ' T J (3.7)

To make the equation more generally applicable, we can replace R' by mR where m is the mass of
gas, giving

Pv = m RT J (3.8)
v
or P = RT J/kg
m

But v/m is the volume of 1 kg, i.e. the specific volume of the gas, V (m3 per kg). Hence,

PV = RT J/kg (3.9)

This is known as the General Gas Law and R is the gas constant for that particular gas or mixture of
gases, having dimensions of J/(kg K). The specific volume, V, is simply the reciprocal of density,
1
V = m3/kg (3.10)
ρ
Hence, the general gas law can also be written as
P
= RT J/kg (3.11)
ρ
P kg
or ρ = giving an expression for the density of an ideal gas.
RT m3

As R is a constant for any perfect gas, it follows from equation (3.9) that the two end states of any
process involving an ideal gas are related by the equation

P1V1 PV J
= 2 2 = R (3.12)
T1 T2 kg K

Another feature of the gas constant, R, is that although it takes a different value for each gas, there
is a useful and simple relationship between the gas constants of all ideal gases. Avogadro's law
states that equal volumes of ideal gases at the same temperature and pressure contain the same
number of molecules. Applying these conditions to equation (3.8) for all ideal gases gives

mR = constant

Furthermore, if the same volume of each gas contains an equal number of molecules, it follows that
the mass, m, is proportional to the weight of a single molecule, that is, the molecular weight of the
gas, M. Then
MR = constant (3.13)

The product MR is a constant for all ideal gases and is called the universal gas constant, Ru.
In SI units, the value of Ru is 8314.36 J /K. The dimensions are sometimes defined as J/(kg mole K)
where one mole is the amount of gas contained in M kg, i.e. its molecular weight expressed in
kilograms. The gas constant for any ideal gas can now be found provided that its molecular weight is
known

3- 5
Fundamentals of Steady Flow thermodynamics Malcolm J. McPherson

8314.36 J
R = (3.14)
M kg K

For example, the equivalent molecular weight of dry air is 28.966, giving its gas constant as
8314.36 J
R = = 287.04 (3.15)
28.966 kg K

Example.
At the top of a mine downcast shaft the barometric pressure is 100 kPa and the air temperature is
18.0 °C. At the shaft bottom, the corresponding measurements are 110 kPa and 27.4 °C
respectively. The airflow measured at the shaft top is 200 m3/s. If the shaft is dry, determine
(a) the air densities at the shaft top and shaft bottom,
(b) the mass flow of air
(c) the volume flow of air at the shaft bottom.

Solution.
Using subscripts 1 and 2 for the top and bottom of the shaft respectively:-

P1 100 000 kg
(a) ρ1 = = = 1.1966
RT1 287.04 × (273.15 + 18) m3

In any calculation, the units of measurement must be converted to the basic SI unless ratios are
involved. Hence 100 kPa = 100 000 Pa.

P2 110 000 kg
ρ2 = = = 1.2751
RT2 287.04 × (273.15 + 27.4) m3

kg
(b) Mass flow M = Q1 ρ 1 = 200 × 1.1966 = 239.3
s

M 239.3 m3
(c) Q2 = = = 187.7
ρ2 1.2751 s

Example.
Calculate the volume of 100 kg of methane at a pressure of 75 kPa and a temperature of 42 °C

Solution.
The molecular weight of methane (CH4) is

12.01 + ( 4 × 1.008 ) = 16.04

8314.36 J
R (methane) = = 518.4 from equation (3.14)
16.04 kg K

Volume of 100 kg

mR T 100 × 518.4 × (273.15 + 42)


v = = = 217.8 m3 from equation (3.8)
P 75 000

3- 6
Fundamentals of Steady Flow thermodynamics Malcolm J. McPherson

3.3.2. Internal Energy and the First Law of Thermodynamics

Suppose we have 1 kg of gas in a closed container as shown in Figure 3.1. For simplicity, we shall
assume that the vessel is at rest with respect to the earth and is located on a base horizon. The gas
in the vessel has neither macro kinetic energy nor potential energy. However, the molecules of the
gas are in motion and possess a molecular or 'internal' kinetic energy. The term is usually shortened
to internal energy. In the fluid mechanics analyses of Chapter 2 we dealt only with mechanical
energy and there was no need to involve internal energy. However, if we are to study thermal effects
then we can no longer ignore this form of energy. We shall denote the specific (per kg) internal
energy as U J/kg.

Now suppose that by rotation of an impeller within the vessel, we add work δW to the closed system
and we also introduce an amount of heat δq. The gas in the vessel still has zero macro kinetic
energy and zero potential energy. The energy that has been added has simply caused an increase
in the internal energy.

J
dU = δW + δQ (3.16)
kg
The change in internal energy is determined only by the net energy that has been transferred
across the boundary and is independent of the form of that energy (work or heat) or the process path
of the energy transfer. Internal energy is, therefore, a thermodynamic property of state. Equation
(3.16) is sometimes known as the non-flow energy equation and is a statement of the First Law of
Thermodynamics. This equation also illustrates that the First Law is simply a quantified restatement
of the general law of Conservation of Energy

1 kg

Internal δW
Energy
U

δq
Figure 3.1. Added work and heat raise the internal energy of a closed system

3- 7

Vous aimerez peut-être aussi