Similitude and Approximation Theory
Similitude and Approximation Theory
Stephen J. Kline
Similitude and
Approximation Theory
With 27 Illustrations
Springer-Verlag
Berlin Heidelberg New York Tokyo
Stephen J. Kline
Mechanical Engineering Department
Stanford University
Stanford, CA 94305
U.S.A.
9 8 76 5 4 3 2 1
usually yields information about all problems in the class studied. Fre-
quently, a fractional analysis is used to provide the approximate equation
from which an approximate analytical or numerical solution can be
obtained. Thus fractional analysis and numerical analysis are primarily
complementary rather than overlapping methodologies.
As almost every undergraduate in engineering and science is taught
today, one method for performing a fractional analysis is the procedure
called dimensional analysis. The meaning of the term fractional analysis
in the present connotation is very much in keeping with the spirit of the
discussion of dimensional analysis given in the first book on the topic by
P. W. Bridgman. 6 t However, at least six good books on dimensional
analysis are now available in English, and' they cover almost every facet
and view of the subject thoroughly. Consequently, only a relatively
brief resume is given in this work, with references to more extensive treat-
ments where pertinent. In addition to the resume of dimensional analy-
sis, this work covers two other methods of fractional analysis that seem
to have been at least somewhat lleglected in the literature; these are called
respectively the method of similitude and fractional analysis from the gOl'ern-
ing equations. Neither of these methods is original with the author.
However, some extensions and additions to each are included in this work.
The first chapter covers a more detailed discussion of the philosophy
and uses of fractional analysis and classifies the various types of problems
usually treated by such methods. The second chapter summarizes dimen-
sional analysis. The third chapter contains a broader discussion of the
method of similitude than auy other known to the author; it also includes
an introduction to the use of governing equations with examples employing
algebraic and illtegral equations. The fourth chapter covers fractional
analysis of differential equations with their boundary conditions. It
includes not only discussion of conventional problems of dimensional analy-
sis and modeling, but also the bases of approximation theory, construction
of estimates, a brief introduction to the use of the boundary-layer concept
and expansion methods for treating singular behavior, and, finally, exteu-
sioll of the concept of similarity by use of similarity variables and absorption
of parameters. The final chapter contains a comparison of the various
methods.
The level of discussion is intended to be appropriate for a current
senior or first-year graduate student in engineering or science in the United
States. The work is intended for use both by such students and by
workers in science and engineering who must often deal with problems for
which complete answers cannot be found. A working knowledge of under-
graduate mathematics is presumed, and a modicum of familiarity with
partial differential equations is required for Chap. 4.
t Superscripts refer to the list of rererenc.es at the end of the book.
Preface xiii
Stephen J. Kline
Acknowledgments
Chapter 1 Introduction 1
Chapter 2 Dimensional Analysis and the l'i l'heorem
Units and Dimensions 8
2-1 Units and Dimensions II
2-2 Types of Quantities Appearing in Physical Equations 9
a. Primary and Secondary Quantities 10
b. Physical Constants and Independent Dimensions 12
c. Nondimensional Quantities H
2-3 Dimensional Homogeneity of Physical Equations 15
2-4 Statement and Use of the Pi Theorem 16
2-5 Rationale of the Pi Theorem 22
2-6 Huntley's Addition 2:3
2-7 Examples of Application of Dimensional Analysis 24
2-8 Summary 31
xvii
xviii Contents
Spring constant k
Weight of mass m
-r-
Static equilibrium
position
x
.-1 ______ _
FIG. 1.1
appropriate analogue techniques, that is, with items 6, 7, and 8 of the list
of uses, and also with the physics (as opposed to the mathematics) of
item 5.
While items 6, 7, and 8 in the list of uses of dimensional analysis
above appear to be separate, they are in fact very closely related. All of
them center on the question, "What are the pertinent parameters of the
problem?" If the answer to this question is known, the governing inde-
pendent parameters and the model laws usually follow with little difficulty.
The establishment of analogue techniques requires some additional infor-
mation, but, as will be shown in Chaps. 3 and 4, this information can be
found by the same steps that are used to establish the independent param-
eters needed from the governing equations.
The foregoing discussion can be summarized by noting that the uses
of dimensional analysis can be grouped into two categories: (1) establish-
ment of the governing parameters and (2) effective management of the
parameters. These two main categories can be broken down to include
all the items in the list of uses on page 2 as follows:
1. Establishment and study of the governing parameters. This is central
to the following uses:
6. Establishment of model laws and similitude relations
7. Determination of independent parameters
8. Construction of physical and mathematical analogue techniques
II. Effective management of the parameters. This includes two sub-
groups:
A. Simple manipulative and checking processes including,
1. Unit checking
2. Checking algebra
3. Systematic conversion of units
B. Rearrangement of the parameters (usually into nondimensional
form) to provide:
4. Reduction of the number of independent parameters
5. Generalization and correlation of results
We have already noted that fractional analysis is primarily concerned
with group I; it takes group IIA entirely for granted; and it usually uses
group IIB more or less automatically. However, it is pertinent to
examine why this is so. A given problem in dimensional analysis, when
used as a technique of fractional analysis, has two primary parts: (1)
finding the parameters of concern; (2) manipulating these parameters into
the desired form. These parts correspond to the two main groupings
above, and they have been arranged in the order in which the problem
must be solved. Manifestly, it is impossible to make meaningful manipu-
lations on the parameters of a given problem until these parameters are
6 Similitude and Approximation Theory
In the case of length, the datum is taken to be zero; the unit of measure is
the inch as established by comparison with a standard inch; and the rule
for interpolation is given by means of fractions which are marked on
dividing machines. These operations regarding length are so well known
to us that we usually take them for granted.
A dimension is the qualitative concept or idea of the characteristic
measured by a given unit. Thus, in the example of the preceding para-
graph, the dimension is length. The unit employed to make the qualita-
tive idea of length quantitative may be an inch, a meter, or a mile.
In the early decades of this century, an argument still persisted in
the literature concerning whether dimensions (and units) were funda-
mental or relative in character. Bridgman 6 discusses this matter in much
detail and shows that the only tenable position is that dimensions (and
units) are relative quantities. In particular, they must depend on the
specific operational procedures employed in the measuring process. If
these operational procedures are altered not only the size but also the
type of dimensions and units needed will, in general, change. Bridgman
also established the fact that there is no unique "best" or "fundamental"
set of dimensions or units. Thus we could choose to measure geometric
sizes using area, instead of length, as the unit. In a given problem this
might be either more or less convenient, but it would be equally correct.
We might also choose to replace temperature by color in measurements on
very "black" hot bodies, but again this would not alter our final results,
provided we use a consistent and proper measuring process.
Bridgman also showed what was meant by the words proper measuring
process. In particular, Bridgman points out that the operational proce-
dures used must be such that the physical equations are satisfied no
matter what choice of units is made. This will be true, provided that the
operational procedure specified requires that if the unit of measure, item
2 above, is halved, then the number of units found is doubled. Or, more
generally, that the number of units measured is inversely proportional to
the unit of measure. Thus 1 foot equals 12 inches and the length of a
bar in feet is one-twelfth the length of the same bar in inches. This is a
fact which most of us would regard as common sense because of our
experience with physical systems, and it causes no difficulties in practice.
We shall therefore assume that this condition is fulfilled by all unit sys-
tems discussed in the remainder of this treatment.
In terms of the brief discussion just given, we can differentiate four types
of characteristics of systems that enter into physical equations as follows:
10 Similitude and Approximation Theory
1. primary quantities
2. secondary quantities
3. physical constants
4. nondimensional quantities
And we read; length has the dimensions of L, area has the dimensions of
L2, etc. Also we would read the equations which follow as "mass has the
units of pounds," or "mass is expressed in pounds."
pounds force A F A lb j
length A L A ft
time A t A sec
mass AM A Ibm
We now solve Newton's Second Law for gc and obtain:
c. Nondimensional Quantities
EOMoTO ~ 1
E ~ energy
M~mass
cp =
L.(ah)
aT p
c. L. G~).
u = internal thermal energy
h = enthalpy
T = temperature
Since any finite quantity raised to the zero power is unity, we usually say
for brevity that a nondimensional group has the units of 1. N ondimen-
sional quantities play a central role in all of the methods of fractional
analysis.
As noted by Bridgman 6 and others, not all correct equations are dimen-
sionally homogeneous. For example, one might choose to analyze a
macroscopic system of fixed mass in the absence of relativity effects. For
such a system the First Law of Thermodynamics can always be written:
Q=LlE+W (2.3)
F = ma (2.4)
F - ma = (toE + W) - Q (2.5)
t In most of the literature of dimensional analysis the term variable has been
used for this purpose instead of parameter. However, the word variable has a dis-
tinctly different, almost totally contrary, meaning as used in most governing equa-
tions of science and engineering. Since later in this work we will consider governing
equations in some detail, clarity demands the use of another term. As we will see,
this use of the term parameter is consistent with the definition of Chap. 1 and the
usage in governing equations in Chaps. 3 and 4.
Dimensional Analysis and the Pi Theorem Units and Dimensions 17
Pi Theoremt
possible form any more than Eq. (2.10) does. This is an important point,
and we shall return to it later on. Let us now turn again to the matter of
the conditions for use of the pi theorem and the underlying reasoning upon
which it rests.
With this example in mind, we now enumerate the conditions which
should be fulfilled in use of the pi theorem:
1. The list of dimensional parameters must contain all of the param-
eters of physical significance including all independent parameters
and one dependent parameter.
2. The nondimensional pi's as finally composed should contain, at
least once, each of the parameters in the original list.
S. The list of dimensions used to compose the physical parameters
must be independent, or else provision must be made to compen-
sate for the redundancy.
In many treatments, the foregoing requirements are not set forth
explicitly, and it is therefore desirable to discuss them briefly. Since the
method makes no provision for introduction of further parameters at any
stage beyond the original listing, the original list must contain all param-
eters of importance. Any parameters that are omitted from the original
list will be omitted from the solution, and such an omission is a clear error
in analysis. It is also desirable, for convenience, to have the dependent
parameter appear in only one group, but this is a matter of convenience,
not necessity.
As just stated, the final nondimensional form cannot contain more
parameters than the original list, but there is also the possibility that it
may contain fewer. However, if the nondimensional form contains fewer
parameters than the original list, it must imply that either the parameters
which appeared in the original list, but not in the final nondimensional
form, are of significance in the problem or they are not. If they are of
significance, they should be in the final nondimensional form. If they are
not of significance, they should not have been in the original list of physical
quantities, because this almost always increases the number of pi's unnec-
essarily and thus unduly complicates the solution. In other words,
inclusion of unnecessary parameters in the original list is not erroneous
(as is omission of pertinent parameters), but it does give added pi's.
These added pi's tend to defeat one of the basic purposes of the analysis, to
reduce the number of independent pi's to the lowest possible value.
The requirement concerning independence of the dimensions also is
necessary. The simplest way to show this is to examine what would have
happened had we used a redundant set of dimensions in the example above.
Suppose that we had decided to use all of mass, length, time, and force in
the analysis. Since Newton's Second Law of Motion does apply to this
Dimensional Analysis and the Pi Theorem Units and Dimensions 21
The method of utilizing the pi theorem in Example 2.1 and those which
follow has several names, none of which is universally accepted. These
Dimensional Analysis and the Pi Theorem Units and Dimensions 25
L
'11"1 = M
L
'11"2 =N
£2
'11"3 = at = Fourier number
where
a = ~ = thermal diffusivity
pCp
This answers the problem initially set as follows. In order to have similar
temperature behavior, it is necessary and sufficient to have geometric
similarity and the same Fourier number. However, we want to hold the
temperature above a given level for some known time. If we denote this
known cooking time by to and the time required for the center of the body
to reach the required temperature as indicated by a given value of '11"3 (the
Fourier number) as tf! then for a given body the solution is:
tmin = tf + to = tf + constant
And tf is found from the condition L2/a tf = constant. We can then con-
clude that two geometrically similar bodies will have similar temperature
versus time behavior if
L2
- = constant
a
_T
h out
6Tc -Tcout-Tcin
6Tmax·Tho -Tco
In In
6Th -Tho -Th
In out
FIG. 2.1
28 Similitude and Approximation Theory
It seems relatively clear from the sketch of Fig. 2.1 that the param-
eters involved in this problem can be written as:
(2.13)
This list contains nine parameters. The dimensions of these can all be
written in terms of four independent dimensions m, L, t, and T. The four
parameters U, A, Cp " and oTh will not form a nondimensional group.
Thus by the pi theorem we would conclude that if the list of Eq. (2.13) is
correct, there should be four independent pi's and one dependent pi.
However, the answer to this problem is well known; it is two independent
groups and one dependent (see for example Kays and London,23 or
McAdams,32 where the complete solution to this problem is given in closed
form and plotted on graphs). Obviously, there is something wrong with
our solution. Actually, there are two things wrong.
The first difficulty is that Eq. (2.13) contains not one but two depend-
ent parameters. Although it is not immediately evident, a careful study
of the dependency relations among the parameters reveals that only two
temperature differences should be included in our list. That is, given any
one of the three temperature differences listed and all of the other param-
eters in Eq. (2.13), the other two temperature differences are both fixed.
So we proceed to strike out oTc and we obtain:
(2.14)
ing equations. Not only have we checked that the list contains four
primaries that will not form a nondimensional group, but also the excep-
tion is in the wrong direction to be explained by linear dependence among
the primary dimension equations. Specifically, if the set of equations is
not linearly independent, then the actual number of groups needed is
greater than that predicted by Buckingham's original statement of the pi
theorem. But in this case the actual number of groups required is less
than that predicted by the pi theorem in the solution above.
Thus we have found an exception to the pi theorem that seems to be
inexplicable in terms of dimensional analysis alone. It then remains to
be seen if other methods can explain the source of this exception and, if so,
to examine whether it is reasonable to expect an able and experienced
worker to have found the source of the trouble using only the procedures
of dimensional analysis.
Example 2.4. We now approach a practical problem, by the Buck-
ingham method, where the answer is not altogether known. Let us con-
sider the correlation of the performance of centrifugal compressors. This
is a more complex problem than any of the previous examples, in fact,
unlike the previous cases we cannot write a single set of differential equa-
tions and boundary conditions governing the complete performance of all
of the class of systems we hope to correlate.
We begin again in the conventional fashion, that is, by attempting
on intuitive grounds to establish a list of parameters. We suspect that.
the following parameters might be of concern in this problem: flow rate Q,
pressure ratio PT, speed N, impeller inlet radius r;, impeller outlet radius To,
length characterizing the diffuser L, and the viscosity J.l, density p, speed
of sound a, and specific-heat ratio 'Y of the working fluid.
In the functional form we have:
given is far from complete. As a first step toward making the problem
manageable, we might therefore restrict ourselves to consideration of
geometrically similar machines. Such machines can be characterized by
one length, say D, and hence the number of groups apparently needed is
reduced by two. Thus we could formulate an answer to the reduced
problem of homologous units as:
11"1 = pr
Q
11"2 = ND3
11"3 = Re = Reynolds number = ~~
11"4 = M = Mach number = - Q2
Da
11"6 = 'Y
beyond the scope and purpose of the present volume. The foregoing
discussion is sufficient to show that dimensional analysis alone is of some
utility in this problem, but that it must be very consid,erably supple-
mented before a practical answer can be found. In particular, the inabil-
ity of dimensional analysis to answer questions of the sort, "When can a
given nondimensional group be neglected?" stands out clearly in this
problem.
Exalllple 2.5. Having given several examples in which various
shortcomings of dimensional analysis as a method of fractional analysis
have become evident, it is useful to work a problem where the method
succeeds particularly well, in order to illustrate the type of problem that
can be handled. Accordingly we consider small motions of a simple pen-
dulum, as shown in Fig. 2.2.
Let us take as the problem the determination of the parameter(s)
governing the frequency of the natural or free oscillation of the pendulum.
If we consider what parameters we would use to write the equations, we
would set down a list of pertinent parameters as:
f(w,L,m,g) = 0
where
w = natural frequency
L = length of pendulum
m = mass of pendulum
g = local acceleration of gravity
---1---/
I, ,/'
FIG. 2.2 mg
32 Similitude and Approximation Theory
Since there is only one group, it can at most be a constant. And we can
therefore write
C·g
w2 =--
L
This, of course, is the entire answer to the problem, lacking only the value
of the constant C. Thus we see that in this simple problem dimensional
analysis works exceptionally well. Indeed, it demonstrates that the
period must be independent of the mass of the pendulum, and it also
yields the explicit form of the solution.
To round out our examples by dimensional analysis we now consider
a more complex problem of mechanics.
Example 2.6. t Consider a more complex mechanical system con-
sisting of a simple beam in flexure with given loading and end conditions.
We hope to find a model by which we can determine the dynamic or
vibration characteristics of any beam by tests of a smaller beam or model.
In this problem, we suspect from experience that the vibration fre-
quency w will be a function of the following parameters:
w = w(E,p"p,Fj,ai,g)
where
E = modulus of elasticity
p. = Poisson's ratio
t This example is based on the excellent discussion of Prof. D. E. Hudson 18 and
a suggestion from Prof. S. H. Crandall.
Dimensional Analysis and the Pi Theorem Units and Dimensions 33
p = mass density
F; =j conditions sufficient to fix the loading, say q in number
ai = i lengths sufficient to fix the geometry of the beam, say 11. in
number
g = gravity constant, weight per unit mass
w2al (pga l )
-g- = f IF' IL, a;, {3i (2.15)
The conditions (2.16) demand that the materials employed have the same
Poisson's ratio, and that complete geometric and loading similarity be
strictly maintained. In general, this will represent such restrictive con-
ditions that modeling on this basis would be of little practical value.
In an attempt to overcome this difficulty we might examine a more
restricted class of problem. Let us assume that the body under study is
a simple rectangular beam of length L, depth h, and width b; let us assume
further that it is a simple cantilever beam, and the only load is the beam
mass p per unit length. Accordingly, the list of parameters is
w = w(E,p"p,b,h,L,g)
34 Similitude and Approximation Theory
w2L
g
= f (pgL, p.
E 'LL
!!.,!!:.) (2.17)
(2.18)
There are still too many conditions for manageable model studies, and
thus even in this relatively simple problem a correlation of workable
simplicity is not achieved by use of the pi theorem alone. From the point
of view of fractional analysis, we should like to be able to find a solution
involving fewer conditions than Eq. (2.18). This might be distorted or
approximate model laws covering the problem, it might be a grouping of
parameters in improved form, or it might be a statement of the exact con-
ditions under which some of the pi's in Eq. (2.17) can be neglected. These
are essentially the same questions we encountered in Examples 2.2 and
2.4, but they are framed slightly differently, because we are seeking a
solution in terms of a model law rather than a correlation. Again we will
defer further discussion of these questions until other types of solutions
have been developed.
2-8 SUMMARY
The examples above show that in every instance the pi theorem gives
some useful information about the problem under study. In some simple
problems, such as that of the pendulum, it gives remarkably complete and
correct answers. The pi theorem also sets in particularly clear form the
enormous utility of the use of dimensionless groups in reducing the number
of independent parameters and correlating and generalizing solutions.
In addition to these useful properties, the pi theorem forms a partic-
ularly good framework for discussion of the nature of units, dimensions,
and related topics, although these more elementary topics have been
treated only very briefly here.
Dimensional Analysis and the Pi Theorem Units and Dimensions 35
3-1 INTRODUCTION
Probably one of the reasons that this method has lost favor in modern
times is that the latter half of the postulate is not broad enough to cover
all problems. Specifically, in many problems, knowledge about quantities
which do not depend on forces at all, such as heat transfer or electric
currents, may be sought. In fact, it would appear that the last half of
the postulate rests upon a purely mechanical view of the universe. Such
a view was prevalent in the nineteenth century, but is not in keeping with
the modern concepts of thermodynamics. Some broader foundation is
definitely needed.
One basis for such a broad foundation is given in Sec. 3-2b. But
before this is done, it will be helpful to study one example which can be
treated by use of force ratios alone in order to make clear the details of
the method.
FI pVL
-=- - =-
Fv /.I Fo E.
p
Reynolds Pressure Cauchy Weber Froude
number coefficient number number number
Fo E.L Fs S
Fv
Fv /.IV Fv /.IV
Stokes
number
Fo E. Fs S
Fs S FG = pLg
-=- Fo
Fo E.L Fo E.
Fa = pL2g
Fs
Fs S
The six forces most often encountered in These forces may be expressed dimen-
fluid flow: sionally:
Inertia forces L:::,. FI FI A pV2£2
Viscous forces L:::,. Fv Fv AILVL
Pressure forces L:::,. Fp Fp A /:1pL2
Compressive forces L:::,. Fo Fo A Es£2
Surface tension forces L:::,. Fs Fs ASL
Gravity forces L:::,. Fa - Fa ApL3g
where
p = mass density
L = length
V = velocity
/.I = viscosity
E. = isentropic bulk modulus of compression
S = coefficient of surface tension
g = local acceleration of gravity
Table 3.1, only six are apparently widespread enough in use to have
acquired generally accepted names.
The appearance in Table 3.1 of all of the most common correlating
groups of at least incompressible flow is, of course, not a coneidence.
These groups are widely used not only for historical reasons, but also
because the direct ratios of the governing forces express the correlating
groups in a particularly useful, simple, and readily interpreted form.
Indeed, although the Mach number is normally used for correlation pur-
poses, the quantity that almost always appears in the governing equations
is the square of Mach number, or the Cauchy number; and in Table 3.1 it
is the Cauchy number that appears. Far more often than not, the use of
M2 rather than M as a variable simplifies the equations, and historically
we would probably have been better off had M2 been adopted as the con-
ventional correlating group. However, the difference is not great and is
hardly worth discussion except to illustrate why the use of direct force
ratios so frequently yields, apparently fortuitously, a particularly useful
combination of parameters.
t Some readers may find the general remarks in this section regarding the rela-
tion between the governing equations and similitude hard to follow completely at
this reading since a number of different points are involved. If this occurs, it is
suggested that the reader continue through Chap. 4 and then read this section again;
many of the points discussed can be made entirely clear only in terms of examples.
42 Similitude and Approximation Theory
The existence of a set of equations and conditions which are the same
for two systems can be established on either of two bases. First, the two
systems can be physically of the same class, such as two tubes with fully
established flow. We then assume a priori that the governing equations
express some immutable laws of nature and will therefore always have the
same form provided only that we specify sufficient information about the
two systems. Second, we may know the governing equations from prior
experiments, so we can compare them directly whether or not the two
systems are the same. This second basis is broader and includes the first,
but it can be used only when the governing equations have been explicitly
developed.
In terms of this discussion we can now see that the classical method
of similitude based on force ratios employs a basis of the first type; it
implicitly assumes that some set of governing equations based on immu-
table laws of nature exists; it further assumes that the terms in these
equations can be expressed in terms of forces alone; and it then moves
directly to a solution of the problem by dealing with the ratio of forces to
guarantee that the parameters in the governing equations will have the
same values. Since it does not employ the governing equations explicitly,
in a sense it implies that, while they exist, they are not available in usable
explicit form. t
However, in the present state of physics, it is very rare that we
encounter a problem for which we cannot write at least a set of overall or
"black-box" governing equations based on the known macroscopic laws
of nature. This suggests that examination of some general list of gov-
erning equations should tell us what other types of parameters in addition
to forces must be fixed, if any, in order to guarantee equal values of the
nondimensional parameters in an appropriate set of governing equations
and conditions. No two workers would employ quite the same list of
fundamental equations, but a sufficient list for problems in continuum
analysis, that is, fluid flow, elasticity, classical electromagnetism, heat
transfer, and thermodynamics, is the following:
1. Conservation of Mass
2. Stoichiometric Principle (conservation of atoms, molecules, etc.)
3. Newton's Second Law
4. Equation of State (state principle)
5. First Law of Thermodynamics
6. Second Law of Thermodynamics
7. Rate Equation Theory (Fourier's law, Fick's law, Ohm's law, etc.)
8. Maxwell's Laws of Electromagnetism
t This remark applies in the same sense to all of dimensional analysis as sum-
marized in Chap. 2.
Similitude and Overall Equations 43
-------
=
at inside dt
Rate of Rate of energies --~
Rate of Rate of ener- Rate of deliv-
heat trans- entering with energy gies leaving ery of work
fer in mass including storage with mass in- excepting re-
reversible flow inside cluding reversi- versible flow
work of transport ble flow work work of trans-
of transport port
(3.1)
t If the reader is unfamiliar with this point of view, a more thorough discussion
is given in Ref. 24. It should also be noted that some recent authors, particularly in
rheology, have begun to call item 4 and item 7 taken together the constitutive relations;
this term is understood to include the stress strain and stress rate of strain relations.
It should be noted that these relations are not always known either theoretically or
empirically in their entirety. For simple substances, such as pure crystalline solids
and perfect gases, almost all the desired information can be calculated from kinetic
and statistical molecular models. In some other systems, most notably water, the
statistical theories are inadequate, but quite precise data is available over wide ranges.
In more complex and less studied substances, such as high polymers, very little is yet
known about the constitutive relations in either integrated or differential form.
For these reasons and perhaps others, workers in various fields will probably
want to modify the exact list of fundamentals to make it more suitable to their pur-
poses. Such modification is appropriate and a highly valuable study. However,
the list as given will be sufficient for the problems we will discuss in this volume and
for most problems in continuum analyses.
44 Similitude and Approximation Theory
where the summation signs imply addition of all inflows and outflows,
respectively, and
------ ------
(entire control (material inside (entire
surface) control surface) surface)
""'--" ""'--"
Forces due to Forces due to Forces Surface
transport of change of mo- due to forces;
momentum mentum inside fields shear and
across control control sur- such as pressure
surface face gravity
and mag-
netism
where
u = stress
A = area
V = volume
Equation (3.2) applies to any macroscopic control surface in the absence
of relativity effects. t
We now inspect Eqs. (3.1) and (3.2) to see whether they contain
partly the same or entirely independent information. Examination of the
momentum equation (3.2) shows that it contains the same forces that
occur in the mechanical terms of Eq. (3.1). Such mechanical terms occur
as the result of the action of forces of the sort enumerated in Table 3.1.
Thus it seems reasonable to state that so long as these forces are unaffected
by the other terms in the energy equation; the two types of action can be
viewed as independent, and problems which deal only with the force terms
can be solved by the use of Table 3.1 or other suitable tables of force
ratios alone. This is the case for incompressible flows, and it probably
accounts for the widespread use of the method of similitude based on
force ratios alone in problems of that kind. Such solutions date back at
least to Lord Rayleigh in the mid-nineteenth century. For more general
problems, however, Eq. (3.1) suggests that it will be necessary to introduce
other types of effects even if one is seeking only correlations of the forces,
since the other forms of interactions may then affect the force terms.
Apparently, if the heat flow, the thermal energy associated with mass
transport, or the thermal energy stored in the system are believed to
affect the variable one is seeking, then some appropriate energy ratios
must also be included in the pi groups formed. Thus, a possible tentative
conclusion is as follows. The force ratios, if complete, will account for
the mechanical-energy terms in the energy equation; this includes the
work term, the flow work, and any mechanical energies, such as those due
to gravity or kinetic energy (inertia), that are associated with mass trans-
port or motions inside the system. However, it is necessary to avoid
including some specific single effect, such as the hydrostatic pressure
change caused by motion in a gravity field, twice in a given solution.
Since the energy equation contains the mechanical terms, but the momen-
tum equations do not contain the thermal terms, the best place to check
the independence of various effects would seem to be the energy equation
(3.1) or its equivalent. However, in doing so it must be remembered
that it is not sufficient merely to lump all force effects into a work term
as is frequently possible in a conventional thermodynamic analysis. That
is, the mechanical-energy terms must be carefully reduced to their ele-
t Equation (3.2) reduces to the more conventional form when V; = V R if the
control volume is nonaccelerating and the reference of coordinates is fixed on the
control surface.
46 Similitude and Approximation Theory
ments by use of force diagrams, and a table, such as Table 3.1 for fluid
flow, must be employed to find the ratios of forces needed for similarity.
After this has been done, the results can be compared with the governing
overall energy equation to determine the totality of energy and force
ratios required to specify overall similarity of the two systems with regard
to both external interactions and detailed similarity of the energies and
forces inside the system.
This then provides a possible method of specifying the force and
energy ratios that will be required in order to insure complete similarity.
However, we must still ask, "Are there any other categories of quantities
or effects that can occur?" To answer this question we proceed to exam-
ine the remaining physical principles in the list given above, checking, in
each instance, to see if any new items must be added beyond the pertinent
force and energy ratios to specify solution of each equation and thus to
ensure similarity.
Taking the principles in the order of the list, we have first conserva-
tion of mass and of atoms. These introduce no new requirements, since
the equations of momentum and of energy, in the form given, both employ
the continuity equation and thus inherently include mass conservation.
The notion of the state principle does introduce a new possible set of
parameters, namely, those based on the properties of the system or its
surroundings. Dimensionless groups based on the ratios of properties do
play an important role in some problems. The concept of a functional
relation between the properties is independent of the other principles
involved, and if such an equation is needed to link the variables or parame-
ters occurring in the other equations, then one or more pi's involving
appropriate ratios of properties will in general be required. This point
will be made clearer by example below.
The next principle, the Second Law of Thermodynamics, does not
add any new parameters or variables to fractional-analysis problems.
Physically, this is due to the fact that the function of the Second Law, in
any analysis, is to prescribe the possible direction of real processes and to
locate steady-state or equilibrium situations in configurational problems.
This function does not add any new effects; it provides additional knowl-
edge about the energy terms already present. Mathematically, this can
be seen by noting that the variable introduced by the Second Law,
entropy, is fixed if the energy fluxes are fixed and the equation of state for
the system is prescribed.
Similarly, rate-equation theory and the laws governing electric,
electromagnetic, and gravity effects do not appear to introduce the need
for any new types of quantities. The rate equation merely specifies the
magnitude of some of the terms in the energy equation in terms of proper-
ties of the system and surroundings. The electrical and gravitational
Similitude and Overall Equations 47
equations are concerned with force and energy terms that appear in the
momentum and energy equations. If the force and energy equations are
complete, these terms will be included.
At this point, it is well to emphasize that the governing equations
employed are redundant and nonunique. They contain overlapping
information (for example, the force terms in the energy equation and the
mass-conservation concept in both energy and momentum equations).
They can be replaced, at least to the extent of rearranged form, by other
sets of governing equations that are equally valid. The precise equations
needed depend on the definition of the system to be studied; in many
simple problems there is a choice of governing equations to be used depend-
ing on the tastes of the analytical worker, because of the redundantinfor-
mation contained in the equations. t
The foregoing discussion and the sufficiency postulate for similarity
stated in Sec. 3-2b seem adequate to provide a firm foundation for simili-
tude methods. However, some questions regarding the independence of
the parameters employed will almost always arise. These questions can
be vexatious indeed, and they clearly depend on the amount of redundance
in the governing equations set down (or visualized). We can ignore the
redundancies and simply set down all the dimensionless parameters that
might possibly be included; such an answer will be correct, but it will
normally also be almost totally useless, since it will contain such a large
number of groups. N or is it possible, at least at present, to resolve these
redundancy problems in general form. To do so we would have to set
down the most general form of each of the governing equations explicitly,
and examine the relations among them. But this we cannot do, since
we are unable to write some of the principles, as for example, the state
principle, in anything but undefined functional form, and this is not suffi-
cient for the purpose. Moreover, the form of the equations depends on
the station of the observer in some cases. Nevertheless, it is quite possi-
ble to obtain some general results, as noted above, and it is much easier to
find specific results in any given case by examining the governing equa-
tions applicable to the particular problem.
But what is even more pertinent is that the use of the governing
equations in the fashion just discussed raises a logical paradox in regard
to the whole framework of fractional analysis as we have developed it up
to this point. In the resolution of this paradox lies the basis for consider-
able further improvements in methodology and understanding.
The entire discussion, up to this section, has been based on the pi
t This point seems to trouble many analysts, but the remarks made should clarify
it provided one accepts the idea that "nature is never inconsistent to itself" and thus
that the alternatives must give the same answer provided the solutions are properly
carried out.
48 Similitude and Approximation Theory
E F\ pc.T£2
qo F\ ALT q, F\ h,L2T q. F\ h£2T
• t
Conduction Radiation Convection Storage in solids
pL3C•
AL
-
WCp
h,£2
-WCp -h£2
WCp
--
wCpt
E, F\ wCpT
Stanton Storage in fluid
number or streams
NTU
h,L hL Llpe. LI
-X -
X tX
- - =-
Olt
qo F\ XLT
Biotnumber Fourier number Conduction
h pLc.
h,
- -h,t q, F\ h,£2T
Radiation
pc.L
-ht
q. F\ hL2T
Convection
Conduction = qo F\ ALT
Radiation = q, F\ dEL2Tf F\ h,L2T
Convection = q, F\ h£2T
.. pVc.T pL'c.T
Storage ill solIds = E, F\ - - = - - -
t t
Storage in fluid streams = E, F\ WCpT
where
v = volume h, = apparent convection coefficient
emissivity
E = for radiation
d = Stefan-Boltzmann constant X = th i d 'IffuSIVlty
..
01 = - erma
X = thermal conductivity of solid pC.
k = thermal conductivity of fluid NTU = number of thermal units; de-
w = mass flow rate fined as quantity shown. See
Cp , C. = specific heats at constant pres- Kays and London21
sure and constant volume p = density
L = characteristic length T = temperature
h = convection coefficient t = time
hL
The very common Nusselt number 1:::, k does not appear as a natural energy
ratio, but can be shown to be the product of an energy ratio, a force ratio, and
a property ratio.
Similitude and Overall Equations 51
Example 3.2. Using Table 3.2, let us first consider a purely thermal
conduction heat-transfer problem. We take again the cooking problem
of Example 2.2. Here we need to find a similarity rule governing the
cooking time of arbitrarily shaped solid bodies in an oven.
We note that all three modes of heat transfer may be important.
Radiation and convection occur at the surface of the body and conduction
occurs inside, as shown in Fig. 3.1. We now construct the simplest pos-
52 Similitude and Approximation Theory
I+----M---~
E. (energy stored
}--~~in~si~de~)-\--+--r--------~Y
FIG. 3.1
sible thermal circuitt to represent this problem (Fig. 3.2). Figure 3.2 is
not sufficient to solve the complete problem, but it will suffice to determine
what energy effects act independently.
We see that the sum of the radiation and convection heat transfer act
as a single energy effect in this problem, as is often the case, but that con-
duction is a separate effect. That is, the ratio of the radiation to the
convection is unimportant; only the sum of the two is pertinent, but
the ratio of conduction to convection plus radiation heat transfer is im-
t A thermal circuit is conventionally drawn using standard symbols for electrical
circuits which imply thermal elements by analogue.
hr
Rei Re2
hu
+
Ell E.2 E.3 Rc3
FIG. 3.2
Similitude and Overall Equations 53
L
1fa = M
L
1f4 =N
Thus we obtain the following answer. The cooking time will be the same
if all of 1ft, 1f2, 1fa, and 1f4 are fixed. Again we must interpret the time for
complete cooking in terms of tmin = tf + to as in Ex. 2.2. Clearly we
would also get the similarity rules for conduction.
By inspection of Table 3.2 we also see that increasing 1ft increases
the importance of the transient conduction effect and that increasing 1f2
increases the importance of the heat transfer to the surface. Thus we
can see that if 1ft is large and 1f2 is small, we would expect to have pri-
marily a conduction problem; while if 1ft is small and 1f2 is large, we would
have primarily a convection and radiation problem. Although we are
unable to make these conclusions quantitative by this method, they are
nevertheless useful information we were unable to achieve by use of the
pi theorem on this same problem in Chap. 2. For later reference it
should be noted that an added conclusion arises not so much from the
method itself as from the fact that the method allows us to use physical
information that we could not bring to bear within the framework of
the pi-theorem methods.
~ ________________________ ~~x
FIG. 3.3
Similitude and Overall Equations 55
of this type, we begin with some type of system diagram, in this case a
free body diagram of the forces.
In Fig. 3.3, it is seen that four forces affect the fall of the body:
gravity, drag, buoyancy, and inertia. The equation for the overall
dynamic equilibrium of the body is very simple; it is
(3.3)
where
Fg = gravity force
Fr = inertia force
FB = buoyant force (hydrostatic pressure forces)
F D = drag force
However, if the body is bluff, the drag force will have two causes, namely,
a pressure force and a viscous force. Thus we have for some cases at least
five forces; we will require four pi's, three independent and one dependent.
Three of these can be selected from Table 3.1; appropriate groups include:
Reynolds number, Froude number, t and drag coefficient. In addition
we need a group representing buoyancy force. This we can formulate
very simply as follows:
FB = PIV
Fg = PBV
FB _ PI
Fg - PB
where
PI = density of fluid
PB = density of body
V = volume of body
Thus we have found four pi's which will serve our purpose. However, if
instead of merely extracting similarity rules from these pi's, as we have
done previously, we choose to express them as an equation, we can find
further results.
We note that the Froude and Reynolds numbers and the drag coeffi-
cient all involve an inertia force. Thus we could produce them, or their
reciprocals, by the simple expedient of dividing Eq. (3.3) by Fr.
Fg = 1 + FB + FD
FI Fr Fr
t Froude number is used in a different sense here than is usual for problems
restricted to systems on earth.
56 Similitude and Approximation Theory
Upon inserting
(3.4a)
. and no t·mg -F
On rearrangmg FB = PI
-
g PB
1 1
Np --1
NF(I-;;) NR
= - (3.4b)
where
FI
N F = Froude number = F
g
FI
NR = Reynolds number = F.
N p = pressure coefficient = :;
Here we must note that we have defined the pressure coefficient, the
Froude number, and the Reynolds number in the step between Eqs.
(3.4a) and (3.4b). That is, we define them to be the force ratios appearing
in Eq. (3.4a). This procedure is permissible since, as shown in Table
3.1, the form of the ratios is appropriate.
Equation (3.4b), although based on a very rudimentary analysis,
nevertheless provides considerably more information than could be
obtained from knowledge of the pi's alone.
In particular, we see immediately from Eq. (3.4b) that if the term
PI/ Pa is small compared to unity, it can be dropped from the solution,
since it will be insignificant. Similarly, we can establish from Eq. (3.4b)
over what ranges other parameters may be neglected. In an algebraic
equation, such comparisons of magnitude are obvious and always valid;
in a differential equation, the matter is more subtle, as we shall see in
Chap. 4.
At this point, we must be extremely careful about what we mean by
such terms as Reynolds number or Froude number. As already noted,
we can define a parameter which is the ratio of two forces in the problem,
say the inertia to viscous force. This ratio will have the form of a Reyn-
olds number, as shown in Table 3.1; but it is not necessarily the standard
form. That is, we cannot simultaneously define Reynolds number in
Similitude and Overall Equations 57
some arbitrary standard way, and also define it as the ratio of inertia to
viscous forces without checking to see that the two definitions agree for
the problem at hand. As a matter of fact, they often disagree. It then
follows that we can neglect a term which is small in the sense of the
example above, but it does not follow that a given force ratio is small
merely because a standard form of a parameter takes on a very small
(or large) value. Again, using the ratio of inertia to viscous force as an
example, it does not follow that this ratio will be small simply because
some arbitrary standard form of Reynolds number is large. Indeed,
some of the worst blunders in the history of fluid mechanics have been
based on a failure to make this distinction clearly. This particular
example (inertia and viscous forces) has become notorious in fluid mechan-
ics, but the difficulty is not isolated to that example or indeed to that
subject. The logical trap of the double definition for a single term exists
whenever we establish standard forms for the common parameters and
also try to attribute to these forms direct physical meaning for a very
wide range of problems. Since we do this in almost all technical fields as
they advance, it is important to be very clear about this point. That is,
we must never assume that we can judge magnitude on the basis of an
arbitrary standard form, no matter how useful that form has proved to
be in the past, until we have demonstrated that the form employed does
indeed give the pertinent energy, force, or property ratio for the particular
problem at hand.
In cases where the standard form and the form giving the pertinent
ratio of forces, energies, or properties are not the same, then the parameter
actually giving the ratio is almost always far more useful and relevant.
Moreover, if the reasoning regarding force, energy, and property ratios
above is correct, then it is always possible to form these more pertinent
ratios, provided we know enough about the problem in hand; in Chap.
4 we shall see in more detail what information is needed for this purpose.
With these remarks in view both the limitations and the uses of a table
of parameters, such as Table 3.1 or 3.2, can be further clarified. The
parameters of Tables 3.1 and 3.2 suggest the forms of parameters which
we should seek in a given problem. The tables thus can be used as a
checklist. They also form a starting place from which parameters that
do give more information about specific problems can be sought. But
we must always remember that the parameters of Tables 3.1 and 3.2 lack
specificity in the sense that the particular scales or values of physical
quantities best representing a given problem are not stated; we cannot
use the parameters as they stand in Table 3.1 or 3.2 for comparison of
magnitudes without further checking. Moreover, the establishment of
the proper scales is a research problem; sometimes it is an easy, almost
trivial task, but in some cases it is a difficult and subtle problem.
58 Similitude and Approximation Theory
f p dA + f pV2 dA = -; f p dV - p; - F:
(whole sur- (bottom (over
face, includ- surface) control
ing that volume)
of sphere)
(inside) (top
surface)
L,
z
Outflow
Streamline
Control on sides
surface
Inflow
FIG.3.3a
JpV dA = JpV dA + ~ Jp dV
(bottom (top (inside)
surface) surface)
Energy
The continuity equation contains only variables that also appear in the
momentum equation, but unfortunately the energy equation contains,
in addition to these variables, a new dependent variable, internal thermal
energy. t Thus we must find still one more equation linking internal
energy, density, and pressure. Such an equation is given by the equation
of state. If we are dealing with one pure Newtonian fluid substance,
then by suitable transformation of coordinates we can always find an
equation of state that has no more than two independent intensive prop-
erties and we may therefore write:
e = e(p,p)
To complete this discussion of the dependence of the variables on one
another, it is very helpful to use a more explicit equation of state. Let
us therefore assume that we are dealing with a medium that satisfies the
perfect-gas equation of state. For this type of substance we have the
very simple relation:
e = c.T
Also the enthalpy h is
h = e + pv = cpT
And thus for a perfect gas
p / p = pv = h - e = cp - cv = 'Y _ 1
e e e ~
where
cp
'Y=-
c.
1
v =-
p
de = T ds - pdv
which applies to any pure single fluid substance of the sort under dis-
cussion. More specifically, for the perfect gas the ratio of the net mechan-
ical energy of transport in a pressure field to the concomitant and unavoid-
able storage of thermal energy is precisely
as has just been shown. Thus we see again the crucial role of the explicit
dependency relations among the governing equations.
It is now possible to complete the fractional analysis by reasoning
as follows. The equations set forth above for momentum, energy, con-
tinuity, and the equation of state are sufficient to determine the dependent
variable sought. Thus if we can ensure that all of the variables and
parameters in the equations are fixed and are the same in two given cases,
then we can assert that similarity will be achieved on the basis of the
first grounds given in Sec. 3-2b, provided a unique solution to the equation
exists. This is true whether we can solve the equations or not, and in
this case guaranteeing fixed values for the variables is easier by at least
several orders of magnitude than solving the complete equations. In
fact we have already enumerated the forces needed and shown that the
equations also involve one property ratio which fixes certain significant
energy terms. We can proceed to obtain the desired fractional analysis
in either of two ways. We can simply set down the force ratios found in
the solution of the incompressible problem plus the Mach number and
the ratio of specific heats as a solution, or we can normalize the governing
overall equations and extract the pi groups that appear in the various
terms. The first method is easier and is in fact completed by the remarks
just made. The second method requires additional work, but will in
general yield more information. t Since it is quite complicated for the
case just given, we shall not complete such an analysis in this instance
but will give other examples below.
It is worth noting that the number of parameters in this problem is
unduly large, and the more complete analysis of the governing equations
to determine what terms might be neglected would certainly be in order
if we were seeking the most possible information about this problem rather
than attempting to illustrate methodology. Similar remarks apply to
t As we have just stressed, formal use of Tables 3.1 and 3.2 does not in general
give magnitudes.
62 Similitude and Approximation Theory
1 2
I I
I I
T7rDL I
I
I
« I
!pV2 dA !pV 2 dA
I:
1-1 2-2
PIA :1 I I
P2 A
I I
I I
I I
1 2
FIG. 3.4
f pV2 dA
1-1 - . . -
+ PIA'-v-"
=P 2A + T7r DL
'-v-" ___
+ f2 - 2 -pP. .dA-
FJ Fp Fp Fv FJ
If the flow is truly steady and fully established, we see immediately that the
two terms representing inertia forces cancel out, since p and V are the same
in any given element dA at the ends of the control surface. Thus we
obtain
(PI -
Fp
. po/A = ---
T7r
Fv
DL
This asserts that for a truly steady flow we should be able to obtain a
solution for the governing groups by dividing by F.; that is,
where a and (3 are appropriate length scales. This is indeed a proper and,
for many purposes, a very useful solution, t although it is not in the form
to which we are accustomed. It is readily verified from the actual known
complete solutiont as follows:
tJ.pj(LjD) b. 4f = 64 = 64/L
jpV2 = Re pVD
which, on rearranging, yields
4fRe = tJ.pj(LjD) pVD = tJ.p· D2 = 64
jp V2 /L /L VL
Note a = D, {3 = L; Table 3.1 does not give scales. Moreover, too-early
cancellation of lengths causes a loss of information; Huntley's addition
(Sec. 2-6) applies here.
This simplification will not work for the turbulent regime; the Reyn-
olds number appears to a different power in the expression for friction
factor, and the density will not cancel through the equation. N or would
we expect it to do so, since the inertia is important in the turbulent regime
inasmuch as it affects the fluctuating forces expressed by the Reynolds
stress§ even though the mean flow is steady.
Example 3.5. We now apply the same method to the problem of
the heat exchanger discussed in Example 2.3, where we found some inex-
plicable difficulties in application of the pi theorem.
Take again the same figure as in Example 2.3.
Transfer area -A;
surfa<;e conductance - U
aTe -Teout-Tein
6Tmax -Th . -Te .
In In
oTh-Th . -Th
In out
FIG. 3.5
t For example, it leads directly to the useful result that the Nusselt number is
constant in a fully established laminar flow in a tube.
t See any standard work on fluid mechanics, for example, Vennard.62
§ The relevant Reynolds stress is given as T = -pu'v', where ( )' indicates the
instantaneous deviation from the mean velocity and (-) indicates a time average.
64 Similitude. and Approximation Theory
We now set the overall rate and energy equations sufficient to deter-
mine the behavior without attempting to solve them. The energy equll-
tions are:
(3.5)
and
aq = (wcp)c aTc (3.6)
The overall rate equation is
(3.100)
means for providing compact solutions in graphical form for all possible
problems of this class. Thus derivation of the dimensionless parameters
from the governing equations yields, as we anticipated, not only a correct,
but also a particularly useful set of parameters.
The examples given are sufficient to demonstrate the utility of the method
of similitude as a means for finding governing parameters of the problem.
The construction of the relevant tables provides a most instructive means
for systematizing and improving correlation parameters. Although the
original developments of the methods were based on too narrow a founda-
tion, it has been shown that this basis can be broadened to include at
least any macroscopic system when the equations of state of the substances
are reasonably well understood and when there are no relativity or quan-
tum effects. Further generalization is probably possible, but has not
been attempted here.
An important weakness of the method of similitude is that it gives
only the standard forms of the parameters and does not tell us directly
the appropriate scales; thus without further analysis we cannot tell
whether or not these standard parameters indicate the magnitude of
terms in a specific problem.
In discussing the broadened basis for the method of similitude, we
were led directly to a more important conclusion about problems in frac-
tional analysis. In particular, both the discussion and the examples
given show the very considerable power of fractional analysis employing
the governing equations in some form. Not only is it useful to assume
the existence of immutable governing equations in order to provide a
broader and more complete basis for similitude, but actual use of the
governing equations, even in a very rudimentary form, leads immediately
to more information, to more useful parameters, and to more general
solutions to similitude problems. All of these points are evident in the
examples given above. These examples also show that the equations
are useful whether they are complete or incomplete and whether they are
algebraic, integral, differential, or combinations of all three. Examina-
tion of the examples given also suggests three important conclusions about
the direct use of governing equations for fractional analyses:
4-1 INTRODUCTION
malized form are the basis for almost all the procedures that follow. The
discussion accordingly begins with a detailed discussion of a standard
procedure for normalization, even though its full implications will not be
immediately obvious. Before we turn to discussion of this procedure one
more remark is necessary.
Some of the procedures that will be based on normalized equations
are rigorous and accurate; in many cases they are entirely rigorous and as
accurate as the description of physical reality provided by the governing
equations and conditions employed. Other procedures discussed are
definitely approximate and must be viewed as trial methods in which
results should be checked against data. An attempt is made throughout
to distinguish rigorous results from trial methods, that is, to state explic-
itly the degree of approximation in each case. However, no apology is
made or intended for the trial methods; all too often they are the only
recourse in difficult problems.
t Actually one makes the equation unit-free or unitless as noted in Sec. 2-2c,
but conventionally the term dimensionless is used; this practice will be followed here.
Governing Equations 71
As noted in Sec. 4-1, the precise choice of scales in step 1 is very important.
The method for choice of scales we will now suggest more or less automati-
cally provides a form in which all desired similitude information can be
obtained in those problems where one normalization is sufficient, and it
forms a very good starting point in those problems where ultimately
further transformation of variables is required. We will therefore use it
as a standard procedure; it will form the basis of almost all the remainder
of the discussion. This procedure is as follows:
The term approximate is used here in the engineering sense, that is,
the closest estimate which can be provided with reasonable effort. Usu-
ally it means within a factor of 2 or closer to the true value, and it would
not normally differ from the true value by as much as a factor of 10.
We now give an example to form the basis for further discussion and
to make the ideas clear.
Example 4.1. Illustration of Normalization. The system ana-
lyzed in Example 2.2 is shown again in Fig. 4.1. Once again the problem
is to determine a similarity rule for cooking such a solid body. The
governing differential equation for conduction inside the body is the well-
known "heat" equation.
at t = -0: T = Ti
at t = +0: X = OL }
y = O;M at ~he surface
z = ON
, T - Ta
72 Similitude and Approximation Theory
1'"---- M - - - ' ;
FIG. 4.1
(4.1b)
at l = -0: '1'=1
at l = +0: X=01}
__ 0' 1 T- = Ta/T; = 0 (since Ta has now been
~ : 0;1 taken as the reference temperature)
The steps carried out on the differential equation above are not modi-
fied in any essential way for integral or algebraic terms. It is only nec-
essary to note that the same steps can be carried out under an integral
sign provided we take proper account of changes in the limits of integra-
tion. The parameters can then be extracted, since they are constants.
and
It should be noted that although these pi's are composed from the system
scales, boundary conditions, and physical parameters, their form, i.e.,
the combinations appearing, is explicitly dictated by the structure of the
variables and parameters in the governing equation and boundary
conditions.
It is very important to understand thoroughly the meaning of varia-
tion in one or more of the variables as opposed to changes in values of the
parameters. To make this point entirely clear, we define a class of prob-
lems as follows. Any group of problems which obeys the same normalized
governing equations and boundary conditions will be called a class of
problems. It is specifically noted that this need not imply identical or
even geometrically similar systems; both physical details and the gross
nature of the systems can vary as long as they can be adequately repre-
t Parentheses are used for clarity in this first example only; they will not be
used in later equations.
t In this case the boundary value T. cancelled out, but this does not usually
happen. Such cancellation has, in fact, a useful physical meaning, which we will
discuss shortly.
Governing Equations 75
t Exceptions can occur if the boundary conditions are expressible only in terms
of implicit functions of the variables.
76 Similitude and Approximation Theory
Equation (4.1c) expresses only the nature of the solution and not its form.
We do not need to actually solve Eq. (4.1b) to obtain Eq. (4.1c). Never-
theless, we observe immediately from Eq. (4.1c) that the parameters of
t~e normalized governing equations are the same as the parameters of
the solution. Indeed, if the argument above is correct, this must be so
for all such equations. Thus we do not need to write the equivalent of
Eq. (4.1c). It is sufficient to normalize the governing equations and bound-
ary conditions and extract the dimensionless parameters by inspection. t If
the governing equations and conditions employed are complete and appropri-
ate, the parameters found must be a necessary and sufficient set for modeling
procedures. To clarify this assertion, consider its implications in the
present problem. Equation (4.1b) was purposely formulated so that it
has the same form for all problems of this class. Since the boundary
conditions in the normalized coordinates contain no parameters, the solu-
tion depends only on the parameters
and
and
t Many workers have used this or closely related procedures to solve specific
problems. The earliest formal treatment known to the author is that due to Ruark,44
who calls the procedure inspectianal analysis. Ruark's discussion has been expanded
by Birkhoff.6 However, Ruark gives only a small fraction of the many types of
results which can be achieved, and Birkhoff states the procedure is merely "sugges-
tive." Neither Ruark nor Birkhoff provide a discussion of when and how the pro-
cedure can be made rigorous, nor do they provide systematic procedures covering use
of the boundary conditions or consideration of the magnitude of terms.
78 Similitude and Approximation Theory
and
are the same for the two systems. If we denote the nondimensional
temperature at any arbitrary point in the body as Th we can then express
this result as
(4.2)
Since Eq. (4.2) holds for every point in the body, it is possible to state an
explicit answer based on the normalized equation for the similarity and
dimensional-analysis problems. To solve the dimensional-analysis prob-
lem, it is necessary only to read off the parameters from the normalized
governing equation and boundary conditions. Thus from Eq. (4.1b)
one finds the pi's to be
L2
11"1 = M2
£2
11"2 = N2
L2
1I"a =-
tcOl
To obtain the similarity rule, we again recall that the problem stated
that all points in the body were to be held above a certain minimum tem-
perature for a specified time to insure cooking. From Eqs. (4.1) and
(4.2) we can conclude that Tj for any two bodies of this type will reach the
same value when 11"1, 11"2, and 1I"a have the same values. Since the least
cooking will take place at the center of the body, we write Eq. (4.2) for
the center point. We again observe that for geometrically similar bodies
the specific requirement for total cooking time is that which yields a
given 1I"a plus a fixed cooking time to. Thus, as in Ex. (2.2), we obtain
tmin = tf + to; tf is found from the condition that 1I"a = constant and from
the values of Land 01 for the body. As already noted, bodies described
need not be totally geometrically similar. The pi's given are sufficient
for any problem in this class and hence for any body whose appropriate
thicknesses can be characterized by the dimensions L, M, and N alone.
Governing Equations 79
at t = -0: '1'=1
at t = +0: ~ : 0,1 } at the su~face
y - 0,1 '1' = Ta Td
Z = 0,1 Ti - Td
The solution will now depend on the parameter (Ta - Td)/(Ti - Td),
as well as on the three parameters found previously. Thus a simple shift
in datum complicates the solution to the similarity problem; only care-
fully chosen coordinates yield optimum answers. Usually there is a
physical motivation for such choices. In this instance, one measures
temperature from the equilibrium, or steady-state, value Ta. However,
there are no general rules, and in many cases it helps to try more than one
set of coordinates to see which set gives the most useful results.
In the example above we have shown that when a complete, appro-
priate set of governing equations and boundary conditions is known,
normalization can provide a rigorous solution to the canonical problem
of dimensional analysis or similitude. This procedure provides more
explicit forms of the parameters; it also provides a basis for one type of
distorted model and includes the definition of permissible distortions.
This by no means exhausts the possibilities inherent in study of the gov-
erning equations without explicit solution. However, before additional
procedures are discussed, it is important to make clearer what is meant by
complete and appropriate governing equations and boundary conditions.
This problem is discussed in the following section.
80 Similitude and Approximation Theory
two solutions exist, then we must determine which solution is more stable
and what initial conditions are needed to guarantee similarity.
Again it is often desirable to rely on an expectation based on physical
knowledge. In this question of uniqueness we rely on the close relation
between the stability of physical systems and the uniqueness of mathe-
matical solutions. There are twO; cases of particular interest.
Suppose first that we can actually establish from mathematical con-
siderations that more than one solution is possible. Then we can usually
assume that the most stable solution will be the one found in nature.
Indeed, we can often purposely perturb the system to insure that this will
be so. If the actual solutions, or even their general form, is known, then
we also have a number of minimum principles, based primarily on the
Second Law of Thermodynamics, which can be used to show which solu-
tion is the most stable. However, in fractional analysis we do not usu-
ally know the solution, and we must then fall back on the expectation
that the most stable solution is the one which will almost always be found
in nature.
Second, consider cases in which we cannot establish uniqueness from
mathematical grQunds at all, that is, we cannot establisq ~ theorem to tell
us whether a unique solution is to be expected. However, if the equations
are known to be complete, we again expect that a solution does exist, and
we can then ask the necessary question directly of nature by an experi-
ment. That is, we observe directly whether two systems do exhibit the
same general type of behavior. This is particularly important when we
are comparing systems of the same class but of a different physical nature.
Usually, the amount of physical data needed is quite small, since gross
changes in the behavior normally occur between one solution and another.
Consider, for example, the question of phase stability of a pure fluid; it
takes only minimal observation to determine whether the liquid or gas
phase is more stable and actually occurs at a given temperature and pres-
sure. Thus a few crude observations are often sufficient. As we shall
see in Sec. 4-6, this same type of crude data is often of crucial importance
in approximation theory, for much the same reasons.
In summary, we can guarantee similar behavior of systems from the
governing equations only when the equations are physically complete
and mathematically appropriate. Since establishment of the physical
completeness is inductive, we can never be completely certain about it;
however, we can be certain to precisely the same degree that we can pre-
dict any solution at all about the class of problem discussed. In many
cases, the outcome can be predicted with a very high degree of assurance.
Indeed, all we are really saying here is that nothing is a dead certainty in
science; everything is subjected finally to the ultimate test of a check
against performance. In some cases, this check is more crucial than in
84 Similitude and Approximation Theory
others, since the degree of assurance which can be provided by past checks
and by the general theorems available from mathematics both vary from
one case to another.
In the cases where both the physics and the mathematics are well
known, we can predict model laws from normalized equations with as
much rigor as we can achieve in predicting anything. In cases where the
completeness of the equations is in more doubt or where the necessary
mathematical theorems are lacking, we can still seek such laws with the
expectation that they will usually be correct, but more checking will be
needed. In either event, if we use the best available equations, we should
expect to get the most reliable answers that can be furnished on the basis
of current knowledge. Like all expert knowledge, it may turn out to be
wrong once in a while, but the probability is in our favor compared to
other procedures.
Thus the sensible course of action seems to be to normalize enough
equations and also enough boundary conditions to provide a well-posed
mathematical problem. This procedure also provides several important
bonuses, as we shall see. It requires the analytical worker to consider just
what is needed to obtain a rigorous similarity or model law; this suggests
what additional information should be sought and also advises about the
urgency of a check. The procedure also almost automatically discloses
other information of importance about the physics and mathematics of
the problem and provides the basis for many other useful procedures in
fractional analysis. We turn next to the study of these other matters.
a. General Basis
b. Homogeneous Equations
Theorem 1
Theorem 2
If a given parameter appears to the same power in each term of the govern-
ing equations, and if this parameter does not appear in the boundary con-
ditions of the normalized equations in standard forIri, then the solution
for the dependent variable is independent of the given parameter, and
such a parameter can be excluded from the specification of similarity in
appropriate coordinates.
The discussion of Secs. 4-3 and 4-4 shows that we can employ the
governing equations and boundary conditions to find the governing
parameters and similarity rules. While we cannot guarantee that the
solutions found in this way always will be rigorous, they usually will be
in a particularly useful and relatively complete form, and they are the
best we can do with currently available knowledge. Moreover, these
processes make particularly clear the nature of model laws and similarity.
It is accordingly helpful to recapitulate the relations between model laws,
similarity, governing parameters, and analogues, keeping in mind these
processes relating to the governing equations.
Solving a similitude problem is virtually the same as finding the
governing parameters. The only distinction is the form of the two results.
Finding the governing parameters, in general, requires the following
approach.
The governing independent pi's in this problem are 11'2, 11'3, and 11'4. The
value of these pi's is sufficient to fix the value of the dependent group 11'1
at any given point in the domain of the solution.
In comparison, the similarity rule requires this sort of statement.
In order for the solution to two problems of the same class to be the same
(similar), it is sufficient that the value of all the independent pi's 11'2, 11'3,
and 11'4 be the same in the two problems.
In both cases, in order to prove the statement is correct, the conditions
enumerated in Sec. 4-3 must be satisfied. Moreover, it is relatively easy
18 Similitude and Approximation Theory
of Sec. 4-3 on rigor and assurance of solution are again directly applicable.
In addition, such model laws are not restricted to geometrically similar
systems or to other requirements of this type. If a means can be found
for ensuring that two systems will have the same normalized equations
and boundary conditions with the same value of the pi's which appear in
these equations, then similarity of behavior is assured up to the extent
indicated in Sec. 4-3.
b. An Alternative Procedure
where the subscript 1 stands for the first system, 2 for the second system,
and the A's are the scale factors. The DE for the first system is:
a2TI + a2Tl + a2TI _ 1 aTI
axi ay~ az~ - ~ at l
Formal substitution of the variables for the second system into the equa-
tions yields
AT a2T2 + AT a2T2 + AT a2T2 _ AT aT2
A; ax~ A~ ay~ A; az~ - aAI at;
dividing by AT/A; gives
a2T2 + A; a2T2 + A; a2T2 = A; aT2
ax~ A~ ay~ A; az~ aAt at2
t See, for example, RefB. 37, 18, and 20.
90 Similitude and Approximation Theory
(4.3a)
(4.3b)
(4.3c)
Scaling procedures usually stop at this point. However, the A's represent
the scale factors between system 1 and 2, and if we recall that the scales
for each should be measured in terms of the boundary conditions and
system size, then the Eqs. (4.3) result in
Thus Eqs. (4.3) give the same results as normalization but in less com-
plete form because the scale factors are left undefined. In order to trans-
late the results into a form in which they actually can be applied to a given
system, the scale factors must be determined. Moreover, the scale fac-
tors used must really characterize the system; if arbitrary scales are used,
even though they appear in the problem, all problems of the class may not
be represented in the same way. Under these conditions, there is no
assurance that similarity will really be achieved.
Another difficulty in examples of scaling procedures in the literature
is that in many cases such procedures have been applied to the governing
equations but not to the boundary conditions. As has been emphasized
repeatedly, the boundary conditions must also be brought to standard
size, and parameters sometimes occur in the boundary conditions in addi-
tion to those arising from the equations. t Serious errors in important
technical problems have occurred in at least one field due to the use of
scaling procedures on governing equations without concomitant study of
the changes occurring in the boundary conditions. It is, of course, pos-
sible to study the boundary conditions also using scaling procedures, but
the suggestion to do so is not implicit in the scaling procedure as it is in
the normalization suggested in Sec. 4-2.
t The importance of this point becomes even clearer in Example 4.7, where
more complex boundary conditions for this problem are examined.
Governing Equations 91
While the choice between the method of Sec. 4-2 and a test for invari-
ance of both the equation and boundary conditions is thus in part a matter
of preference, normalization brings out additional information on such
matters as cancellation of homogeneous parameters. It is also preferred
by the author, since it provides a form of the governing equations which
is immediately useful for several other purposes as shown below.
within the domain. Note that the parameters so formed will involve the
specific boundary conditions and system sizes, hence they will vary in
form from one problem to another. It cannot be expected that they will
always be in some conventional arbitrary fotm. However, when the
parameters found by this type of normalization do differ from the con-
ventional forms, they will tell us more about the problem in hand than
the conventional forms. Further examples as well as a discussion of the
validity of estimates of magnitude appear in Secs. 4-6 to 4-9.
Example 4.2. Consider again the cooking problem shown in Fig. 4.1.
Again we seek to model the cooking time after the body is placed in an oven.
It is required that all parts of the body be held above a given temperature
for a specified time. The initial temperature of the body is Ti throughout,
and the oven temperature is constant at Ta. We insert the body into the
oven at time t = O. The temperature at any point on the surface of the
body is denoted as T 8 • The governing differential equation for the tem-
perature distribution in the body is well known; it is the Fourier equation
of heat conduction
i)2T + a2T + a2T = ~ aT
ax 2 ay2 az 2 a at
Governing Equations 95
at t = -0: T = Ti
at t = +0:
X = O'L}
Y = O,M at the surface T = T.
z = O,N
But T. is itself unknown and must be determined from the rate equation
for convection from the oven air to the surface of the body. Since the
surface cannot store energy, the heat rate by convection to it must equal
the rate of conduction into the body. This condition gives
at t = -0: T= 1
at t = +0: T= 1
i = 0,1
jj = 0,1
Z = 0,1
We are assuming that Eq. (4.1b) was constructed so that each term
containing only variables in the normalized equation is approximately
unity. If this is true, then the magnitude of each term is given solely
by the nondimensional parameter. Moreover, since each term has a
physical meaning, and since we made the equation nondimensional by
division on the coefficient of one term, each parameter gives the magnitude
of the ratio of two important effects. Thus, for example, (L/ M)2 repre-
sents the ratio of net heat conduction in the y direction to net heat con-
duction in the x direction; we also note that L/ M raised to any power but
2 does not have this significance. Similarly the Fourier number £2/tca
which appears in Eq. (4.1b) represents the ratio of net heat conduction
in the x direction to storage of energy in the solid, and in the form set
gives the magnitude of this ratio to a first approximation.
If, as we also assumed in deriving Eq. (4.1b), the estimates of the
nondimensional variables, hence the nondimensional terms in the govern-
ing equation, are valid over the entire domain of interest, then only the
larger terms in the equation will be of concern. Consequently, we con-
sider systematic variations in the values of the pi's in the normalized
equation. First, suppose that L is small compared with both M and N
Governing Equations 97
(~ « 1 y (4.8a)
(t Y«1 (4.8b)
This result follows from Eq. (4.1b) and from the fact that the terms with
larger dimensions can be neglected, while those with equal dimensions will
not change the result. This answer is a very explicit, simple, and useful
similarity rule. It is a more complete answer than that obtained by
other methods in Chaps. 2 and 3. It provides a rigorous answer to the
question originally posed in Chap. 2 concerning whether the housewives'
cooking rule is correct. Clearly, the stated housewife's rule (20 minutes
per pound) is wrong, since it implies proportionality to the product of all
three characterizing length dimensions. This is dimensionally improper,
and it also ignores the crucial shape factors and exponents embodied in
the ratios (L/ M)2 and (L/N) 2. t
The result we have found so far holds only for the case where con-
duction is the controlling process; hl/2k» 1. This assumption implies
physically that the resistance to heat flow due to conduction is large com-
pared to that due to convection even for the shortest path to the center
l/2. It is quite possible that this condition is fulfilled for the larger lengths
but not the shortest. It is also fairly common that the value of h is dif-
ferent on the different sides owing to insulation or different contact con-
ditions. It is left as an exercise for the reader to examine the various
cases which can arise and to determine numerical criteria in terms of
governing parameters, showing when various terms can be neglected and
which cases are essentially one-dimensional when h takes on different
values on the various sides. The results have interesting implications
regarding the assumption of "one-dimensional flow."
estimates of these terms that apply near the singular point and also in
other regions. These two problems are often closely associated, but are
not identical; one can occur without the other.
If a given problem exhibits uniform behavior, and if we are concerned
with a domain of finite extent, then the procedures illustrated in Sec.
4-6a go through without difficulty and approximation theory is extremely
useful. If a problem displays nonuniform behavior, has singular terms,
or requires an infinite interval of integration, then we must expect mathe-
matical difficulties and will need additional special considerations. In
particular, we must always bear in mind possible nonuniform behavior of
the solution in the limit of very large and very small values of the parame-
ters, that.is, the behavior of the solution with one or more of the parame-
ters small may be entirely different from that with the same parameter set
identically zero. t Consequently, we must (1) discuss means for dis-
covering when uniform behavior can be assured, and (2) discuss what
steps can be taken when the behavior is nonuniform. These are both
complex problems, and in the present state of mathematical theory of
differential and integral equations it is not possible to give complete
answers. Some useful procedures, rules, and comments can nevertheless
be provided.
In general, previous discussion of these problems has been primarily
mathematical, and considerably less attention seems to have been given
to the use of physical information. In the present discussion, more
emphasis is placed on the use of physical information, for two reasons.
First, the present discussion is centered on fractional analysis and hence
is most concerned with those cases where complete mathematical theory
is not available. Second, the use of physical information, even in very
crude form, in conjunction with mathematical analyses of the present
type can be a very powerful tool and has in fact often been employed.
More explicit discussion of the places and ways where it should be used
(or sought) seems desirable.
t In earlier drafts of this work, a few of which are extant as dittoed notes, the
author attempted more exact general mathematical conditions. Unfortunately,
these attempts had loopholes, which were pointed out by friendly reviewers. In the
present discussion the conditions have been weakened to meet these difficulties; this
section of the earlier versions should not be used.
Governing Equations 103
:;;1===:;;=== -f-
I
I
7 I
I
I
I
I
I
I
I
I
I
OL-------------------L-------------__
o
x
FIG. 4.3
If Jlooks more or less like Fig. 4.2, then the second derivative is not larger
than
that is
where
L~
J = fm (1)
O~i;~l
104 Similitude and Approximation Theory
1~-----------t I
I
I
1 I
I
I
I
I
I
E I
I
I
OL-________________________ ~--------~~x
o
FIG. 4.4
(4.10b)
IdX2
I
d /1 : : : dj/ dx - 0 = .!.
2
E - 0 E2
In fact if djIdx changes very rapidly at the "corner," even this estimate
Governing Equations 105
or possibly
Hence for functions like those shown in Figs. 4.3 and 4.4 the smooth
estimates applicable to Fig. 4.2 get worse and worse as we go to deriva-
tives of higher and higher order.
We can summarize these results in the form of a rule.
Rule 3t
1. Uniform estimates of a function f and its derivatives up to order n can
be given over a finite region if the function and its derivatives up to
order n - 1 are continuous and differentiable and if the function is
smooth (that is, has a general nature like that of Fig. 4.2).
2. If also f is normalized to be U(l) and the variables Xi to run from 0 to 1,
then
r = au _ av
ay ax
r
In a potential flow the sum of the two terms giving is zero everywhere by
definition. In a logarithmic vortex, the flow is potential everywhere
except at the origin, as is readily shown from the appropriate expression
for the complex potential F(z)
F(z) = iA In z
Consider now a point arbitrarily near the origin; au/ay and av/ax become
very large. Eq. (4.lOb) can provide no estimate of either au/ay or
av / ax which will apply over any domain of the flow field that extends both
very close to and very far from the origin. Nevertheless, a uniform esti-
r
mate for the vorticity can be provided; indeed it is zero everywhere
except at the origin, and it can be neglected in a potential flow. Thus we
do not always need to have each derivative term small by itself.
A second case often occurs when we are interested, not in local effects,
but only in the behavior of some overall quantity. A term can be very
large at some point and not satisfy Eq. (4.lOb) in that neighborhood
without affecting the overall effect appreciably. An example of this
type of behavior occurs in the drag on a flat plate. Under the usual
assumptions, the equations indicate, fictitiously, an infinite skin friction
at the nose, but it is infinite only for a zero distance, and the limit behaves
in such a way that the overall drag is not appreciably affected by the
infinite value at the nose. That is, even though the equations predict
a skin friction at the nose in error by an infinite amount, the overall drag
for plates of finite length is given accurately by the same equations.
Even though continuity and differentiability are not always nec-
essary, there is this to say in justification for trying such assumptions in
continuum problems. The physics inherent in the problem often suggests
(but by no means proves) that we should expect such behavior. This
expectation is based on the fact that the nth order derivatives, those of
highest order in the differential equation, usually represent differences in
forces, energies, and properties; we expect such quantities to be continuous
in a continuum theory. Indeed, such principles as the Second Law of
Thermodynamics often tell us a good deal about the nature of the func-
tions we must expect at the outset. (At this point the reader might well
Governing Equations 107
look back at Sec. 4.6a, and ask himself if the Second Law does provide
the conditions required by Rule 3 for the cooking problem.) It is empha-
sized that this type of information is, however, merely suggestive and
not conclusive, since we can never be certain that the differential equa-
tion we have written is complete. We can expect a given type of physical
behavior with a very high degree of certainty, but there must always be
residual doubt that the equation we have written will predict this
behavior. In the case of the heat conduction equation this doubt is
minimal, and the estimates we found are quite useful. In such cases the
physical information provides strong motivation for expecting certain
types of behavior, hence assumption of such behavior followed by checking
is a highly appropriate way to proceed.
This remark by no means implies that the available mathematics
should not be used or that additional mathematical theorems should not
be sought. It means rather that the researcher cannot allow the absence
of formal mathematical proof to deter him from pursuing solutions and
that good hints for seeking such solutions can very often be found suc-
cessfully from the accumulated physical knowledge embodied in what we
have called the "completeness of the equations," in the associated physical
principles, and in physical data.
This attitude, which we might characterize by the phrase "forge
ahead and then check the results," is one we will adopt frequently in the
examples below. Its adoption is a characteristic difference between the
applied mathematician, who is primarily concerned with solving problems,
and the pure mathematician, who is mostly concerned with establishing
rigorous relationships within the framework of given conditions and defi-
nitions. We need to say a word also on what constitutes a check. It is
possible to check against data or against a more complete theory; that is,
a more complete equation. The check against data is always the ultimate
authority in science, and any check against more complete equations is
subject to uncertainty regarding how complete the equations really are,
as we have already stressed.
Finally, Rule 3 assumes some crude information about the form of the
answer. This is typical of approximation theory procedures, although
the point is frequently glossed over. Very often we have the necessary
data available, and their use is entirely proper, provided it is consistent
with the equations we are using as a mathematical model and it is under-
stood in the context of the remarks about completeness and checking.
Rule 3 gives us one way to establish estimates. It shows, in fact,
one reason why it is so often useful to normalize in the way suggested in
Sec. 4-1, since if the conditions of Rule 3 are satisfied (and perhaps in
other cases), this procedure will make the parameters show the estimate
of magnitude of the various terms. As we saw, these estimates, in general,
108 Similitude and Approximation Theory
decrease in accuracy as the order of the derivative increases, and all deriva-
tives beyond the first are only estimated to be less than or equal to 1.
Hence we may keep terms we do not really need; however, we will not
throwaway any that should be kept.
There are ways in which the estimates of terms can be improved.
One of the commonest is direct use of physical information. This is
clearly in keeping with the philosophy we are using; however, such data
need to be much more detailed than the information needed for Rule 3.
Consequently, we use both procedures, that is, we use data when it is
available and we make appropriate assumptions about the gross nature of
the solution, subject to later checking, when it is not. In either case we
use the estimates to tell us which terms we can drop from the governing
equations to obtain simpler correlations or solutions where we cannot
handle the complete problem mathematically. With this in mind we
turn to the discussion of the relations between the magnitude of terms in
differential equations and in integrated results.
To illustrate the ideas simply, we consider first an ordinary differ-
ential equation. Take for example
T = T(x)
and the DE
T2 (~~) + CT = 0 (4. 11 a)
T!
L
'1'2 dT
dx
+ T CT = 0
a
or
- dT
T2 dx + 7rlT- = 0
where
LC
7rl = 'f2
a
Governing Equations 109
(4.llb)
(4.12)
-----------L
I
-2dT
T dx
FIG. 4.5
110 Similitude and Approximation Theory
FIG. 4.6
r<qs.n
since if the function g has an upper bound, say M 1, we can always define
g = .JL = 0(1)
Ml
These remarks are sufficient to show that we can handle very general
terms. If smooth estimates can be made, or if any other estimates can
be provided which are correct over the entire domain of integration, then
we can construct the pi's in such a way that they represent the magnitude
of the terms in an appropriate integrated relation.
Similar, but slightly different, procedures can be employed for partial
differential equations. In partial differential equations, it is necessary to
Governing Equations W
_Ii fx=o
~y=o _x fx=o T dd-xT +
2
x (- 2 11'"1 d- dx- dx- dy- = 0
dT) (4.14)
y
provided the boundary conditions are suitable. (We will explain this in
a moment.) Application of Eq. (4.10) to Eq. (4.14) yields an order of
magnitude
(4.15)
(x _ d-
J0 x X
+ 11'"1 J(x0 dT d-
dx x =
0
x2
2" + 11'"1 [T - T(O)] = 0
--
{Ii r [d~ +
i
JIi=O h=o dx
'11'1
dy avg
+
(dT) ]dXdY {Ii (i
JIi=O h=o
[~A(Y)
+ 'II'l~B(Y)] dxdy
where the functions ~ A(y) and ~ B(y) are the values of d'l'/ dx and d 'l' j dy
along the boundary x = 0 and (dTjdY)avg is an average value of d'l'jdy
integrated over x.
It is clear that further integrations will involve the magnitudes of
~A and ~B. These terms will influence the relation between the mag-
nitude of the term multiplied by '11'1 and the remaining term. Thus the
magnitudes of ~A and ~B in this case must also be made order unity if we
are to judge the magnitude of terms in the integrated solution from the
magnitude of the nondimensional parameters in the differential equation.
This is merely another way of saying that we must normalize sufficient
boundary-value data to specify a mathematically appropriate problem.
The gist of Eqs. (4.12) and (4.15) is that we can connect the
magnitudes of the terms in the differential equation with those in an
integrated solution, provided we can supply from some source the neces-
sary estimates of terms. These equations by no means indicate the way
we would actually go about integrating except in very special cases, but
they do provide a procedure for demonstrating that we can use the
magnitude of the pi's in the differential equations to characterize mag-
nitude of terms under proper conditions. Again these procedures do not
constitute a proof but only a path for exploration in specific cases subject
to later checking. Moreover, we have so far glossed over the most
difficult part of all-the actual provision of estimates in cases which are
not smooth; most of Secs. 4-7, 4-8, and 4-9 is devoted to examples of how
estimates are constructed in various cases, and these materials give only a
very brief introduction. However, once the framework is known, any
estimating procedure for a specific case can be fitted into it.
We summarize this situation with another rule.
Rule 4
d~x + CT d~ + kT2 X = 0
dt 2 m dt m
Suppose, further, that cT/m « 1; if we then drop the second term and
integrate, we obtain the solution
Comparison of Eqs. (4.20) and (4.21) shows that while Eq. (4.20) agrees
exactly at t = 0, it departs more and more from the complete solution
Eq. (4.21) as time increases.
Next consider, again fictitiously, that ({3r)2 » 1, and try to drop the
term d2xldt2• We obtain only
x = 0, x=O ( 4.22)
for all t
Using Eq. (4.22) we cannot even match the boundary conditions; we
have no solution at all unless the boundary conditions are trivial, that is
at t = 0: x=O dx = 0
dt
This type of difficulty often occurs when we try to drop the most
highly differentiated terms; in such cases we can no longer satisfy all the
boundary conditions and we can therefore obtain solutions only for
certain very limited values of the boundary data. This type of difficulty
can sometimes be resolved by various kinds of expansion in powers of the
parameters, or by a technique which we call boundary-layer theory.
These techniques are briefly introduced in Sees. 4-8 and 4-9.
It is emphasized that dropping terms from the equation for the simple
harmonic oscillator fails, not because approximation theory is wrong, but
because we have supplied erroneous estimates. We have defined x and t
in such a way that the terms d2xldt2 and x are approximately unity, but to
do so we had to use the free parameter r. Since we can vary kim,
superficially it would appear that we could make (klm)r2 anything we
please; this is not the case. Equation (4.19) is itself a relation between t,
x, and (klm)r2. If we are to have any oscillatory solution, Eq. (4.19)
demands that the two terms d 2xjdr2 and (klm)r 2x must be of the same
order. Physically, to maintain an oscillation we must have a dynamic
equilibrium between restoring and inertia forces. Since the equation is
complete, it shows this behavior also. That is, both physically and
mathematically it is not possible to control kim and r separately; as kim
becomes large r becomes small, and conversely. In fact, the relation
between kim and r2 is well known, and is seen directly from the complete
solution to be
or
Thus our assumptions about the magnitude of the parameter are wrong.
Governing Equations 117
We give this almost absurdly simple example to show the type of reason-
ing that has sometimes been used to give approximation theory a bad
name. The trouble is not in the theory, but rather in the adoption of bad
estimates.
Two additional comments are pertinent here. First, we have made
the difficulty regarding estimates very explicit by choosing to make i and
d2i/di2 approximately unity so that the magnitudes show in the parameter.
In most of the literature this is not done, so that the poor estimate relates
not to the parameter but to a term involving derivatives of the dependent
variable. If, as is also often done, an attempt is then made to apply
approximation theory based on the magnitude of the parameters without
regard to the magnitude of the terms in the variables, the procedure
hinges on the actual value of the derivatives. In effect one then merely
hopes that the terms in the variables are approximately unity in size,
and if possible checks this "hope" later on. It would seem preferable
to make the need for estimates of the terms in the variables explicit at the
beginning, although either procedure will work if one is clear about what is
going on.
Second, it is worth mentioning that this same line of argument-that
the whole of the two terms must be the same size if the equation is to be
satisfied and if solutions are to be of an expected type-can often be used to
make estimates of the magnitude of a free parameter such as T. We shall
utilize this procedure later on.
Still another difficulty, beyond those illustrated by use of the simple
harmonic oscillator, occurs when the governing equations or boundary
conditions are nonhomogeneous. By Theorem 1, the value of the depend-
ent variable plays a role which cannot be normalized out. It follows that
we may not be able to find a single estimate that will hold for all problems
of the class considered; that is, for all possible values of the parameters.
(Here we are not talking about nonuniform estimates within a single
problem, but rather about nonuniform estimates between one problem
and another.) This remark will come as no surprise to anyone familiar
with hydro mechanics or with problems in electromagnetic theory. The
solutions found in such cases are typically parameter dependent, that is,
gross differences occur in the overall solution between problems with
different values of the governing parameters, for example, the Reynolds
number. The moral is that when we use data to make estimates for
application of Rule 3 or 4, we must be sure that the data apply to the
range of values of the parameters of concern when the equations or bound-
ary conditions are nonhomogeneous.
N ow that we have given one useful example and also highlighted some
of the mathematical deadfalls and booby traps in approximation theory,
it is appropriate to summarize what we can do with it, and then turn to the
more constructive task of working additional examples.
118 Similitude and Approximation Theory
b, mass per unit length p. Assume that the cross section A may vary
along the beam. A sketch of the system is shown in Fig. 4.7a. Again we
seek a model law from which we can determine the frequency of the first
natural mode of vibration w. The governirig differential equation is well
known from the theory of elasticity. For example from Hudson,18 we
find the equation for the governing forces is
iJ ( i J 2y ) iJ2 y
iJx2 EI iJx2 + pA iJt2 = 0 (4.23a)
where
x = direction along the beam
y = direction of deflection
E = modulus of elasticity
I = moment of inertia about neutral axis (bh 3/12) for rectangular
beam)
t = time
p = mass density of beam
A = cross section area of beam
y..!-o>----- L ----~
Section A-A
(a)
~~------L------------~
(b)
+
y
FIG. 4.7
120 Similitude and Approximation Theory
y
Trace of point at arbitrary
section of beam
x-aL; O~a~l
FIG. 4.8
For a cantilever beam the boundary and initial conditions for a free
vibration can be taken as
at x = 0: y=O dy =0
dx
2
dy
at x = L: -- =0 ~(Eld2y) = 0 (4.23b)
dx 2 ax 2 ax
at t = 0: Yz=L = c}t (d Y) = Y_L =0
dt z=L
Moreover we expect an oscillation with a deflection curve which is smooth.
Indeed, we expect a general form like that shown in Fig. 4.8 for the first
mode. Note that we do not need the exact equation of the curve, only a
very general knowledge of its form. Vast numbers of results also assure
us that Eqs. (4.23a) and (4.23b) are complete and appropriate, provided
that the material is isotropic and the deflection small. Thus we can
expect to find an accurate and complete set of parameters by normalizing
these equations. If we consider a quarter cycle of vibration, we can make
the various derivative terms U(I) by defining dimensionless variables as
follows:
_ x
x=L
y='!tc}
l = wt
where c} is the maximum deflection at the end of the beam and w is the
reciprocal of the time required for a quarter cycle of vibration. A quarter
cycle is used in this case since the beam goes from rest to maximum velocity
(or returns) approximately each quarter cycle. Thus the estimate of
t Measured from position of static equilibrium.
Governing Equations 121
acceleration is
a2y = (dy/dt)t_1/W - (dy/dt)t_o = a/(l/w) - 0 = aW 2
at 2 l/w l/w
The estimate of a4y/(Jx 4 should be given reasonably well by Eq. (4.lOb),
since the curve of y versus x begins with zero deflection and slope and is
expected to be smooth, hence substitution of x and y as defined should give
(J4y/(Ji 4 ~ 1.
Substituting the nondimensional variables into Eqs. (4.23) yields
~ [EI (J4y
L4 (Jx 4
+ 2 (J(EI) (Jay + a (EI) (J2y] + aw 2A a y = 0
(Jx (Jx a
2
(Jx 2 (Jx 2 P
2
(Jl2
(4.23c)
(4.24a)
at x = 0: y=O ay = 0
(Jx
(J2y (Jay
at x = 1: (Jx2 = 0 ax 3 = 0 (4.24b)
at t = 0: Yx=l = 1 (a y) =0
(Jt x=l
Since the governing equations are homogeneous in y and the boundary
conditions parameter-free, 0 cancels from the equations as expected.
Equations (4.24) apply to the entire range of deflections where Eq. (4.23)
adequately represents the system.
From Eqs. (4.24) it is possible to find two types of results directly.
First, it is a simple matter to extract the one governing parameter
where
Equation (4.26) can be used to give the required model law and similitude
rule for vibration in this mode immediately. This is simply that the ratio
of acoustic velocity to VIR be constant. Thus the researcher can adjust
frequency, etc., to meet experimental needs and still obtain accurate
predictions. In this sense Eq. (4.26) provides a basis for distorted as well
as for normal models for beams of constant cross sectional properties.
If we treat w as a parameter whose value is unknown but desired, we
can extend the solution still further by resort to reasoning based on the
meaning of the normalized governing equation. The second term in
Eq. (4.24a) is a measure of the inertia forces resulting from acceleration of
the mass elements of the beam during vibration; the first term is a measure
of the elastic forces arising from the flexural bending of the beam. If a
vibration is to persist, these two forces must be of the same magnitude.
Since the terms in the variables in Eq. (4.24) have been made approxi-
mately unity by construction, the parameter must also be approximately
unity if both terms are to be the same magnitude, provided our estimates
were correct. For any given beam shape, this allows us to estimate the
value of w. We obtain
w='\};;w
lET rad/sec (4.27a)
The exact solution to this problem as given, for example, by Jacobsen and
Ayre 21 (page 80) for a rectangular cross section, is
4w = 3.5159 '\}PbhD
rET rad/sec
or
w = 0.879 '\JrET
PfihI) rad/sec (4.27b)
Hence the very simple estimate of Eq. (4.27a) is entirely correct in form
Governing Equations 123
iJ<!i) «1
iJ2(Ei) //1 (4. 28b)
iJi 2 "''''
EI
(EI)o ~ 1
If the conditions of Eq. (4.28b) are not fulfilled, then an additional
modeling condition must be fulfilled for similar behavior, namely,
[ EI(i)] [EI(i)]
(El)o m = (EI)o p
That is, the distribution of EI as a function of i must be the same
for model and prototype. Usually it will be possible to satisfy all three
of these conditions simultaneously, hence practical modeling can be
achieved in relatively complex cases. As in the simpler case described
by Eq. (4.23), strict geometric similarity is not required to obtain exact
modeling of the vibration behavior, and good approximations can be
obtained for a very wide range of conditions.
Example 4.4. Consider a cantilever beam of uniform cross section
with a large rigid flange welded to the end as shown in Fig. 4.9. Assume
124 Similitude and Approximation Theory
~~------------------~~===P1r ~~----~~x
z Beam
11
--H-
FIG. 4.9
the beam can vibrate up and down. Assume also that the mass can rotate
about the beam center, but that torsional rotation and vibration motion
in the z direction are constrained from occurring.
The equation for the beam vibration is again
i}4y i}2y
EI i}x 4 + pA iJt 2 = 0 (4.29a)
at x = 0: y=o dy = 0 (4.29b)
dx
i}2y JF i} 28F
at x = L: i}x 2 = EI i}t 2
(4.29c)
at t = 0: (4. 29d)
where
Y F = deflection of flange
8F = angular displacement of flange
J F = mass polar moment of inertia of flange
M F = mass of flange
I = section moment of beam, as in Example 4.3
E = Young's modulus of beam
flange measured from the position of static equilibrium with respect to the
beam. To complete the set of equations it is thus only necessary to add
the conditions of compatibility at x = L
fj=J!.
o
_ x
x=L
l = wt
We also define
- dfj
(}=-
di
Then
dY) 0
tan (} F = ( dx x=L ~ L
at t = 0: fj~=l = 1 ( dfj) = 0
dt ~=l
126 Similitude and Approximation Theory
But all of 11'1, 11'2, and 11'3 contain the parameter w2 which we will
usually want to be the dependent parameter. Thus to simplify we form
11'2 J FL Elw 2
71"4 = 71"1 = pAL· L3 Elw 2
But
pAL = mass of beam :!: MB
Also
J F = MFr~
where
r F = radius of gyration of flange
71"8 = Z
Since the equations are known to be well posed and to represent the
Governing Equations 127
physics very accurately for small deflections, we can expect that these
parameters should be an accurate, complete, and quite useful set.
This problem can also be solved with integral equations. The
reader who is familiar with the details of beam theory may find it useful to
verify the solution in the following way. Using energy methods write a
single integral equation governing the behavior of the system. Verify
that the transformation of this equation into nondimensional form which
satisfies the requirements of Rule 3 will lead to only parameters which are
fixed when 11"1, 11"5, 11"6 above are fixed.
FIG. 4.10
U OU + v OU __ - --
1 op
+ IIV' 2U
oX oy pox
and
ov ov lop
U -
oX
+ v -oy = - - -
P oy
+ IIV'2V
where
U = x component of velocity
v = y component of velocity
II = kinematic viscosity = 1-1/P
p = density
p = pressure
V'2 = Laplacian operator
Continuity must also be satisfied. For this case the continuity equation
can be written
P- P
= ---
t:..PL
where
L =length of object
o = undetermined length
U = x component of velocity far upstream
V = undetermined velocity
t:..P L = largest pressure difference between two points on the body;
P is to be measured from pressure far upstream as datum
y momentum:
au + L ~ aD = 0
ax U 0 afj
By construction au/ax is unity order except near the stagnation point,
which presents special problems. t We also want afJjag to be unity order
and we must satisfy continuity at each and every point; this is possible
only if
LV = 1 (4.30a)
Uo
We can guarantee this by defining V as
v~iu
=L (4.30b)
(4.31b)
u=O ii=O
at fj = 00: (4.31c)
u= 1 ii=O
71 = (~y
UL
72 = - /I = Reynolds number -~ ReL
APL
73 = p U2 = Euler number -6 Eu
Thus we see again that we must expect solutions that are parameter-
dependent. That is, we should anticipate the possibility of finding
different solutions, possibly even entirely different types of solution, for
different values of the independent nondimensional governing parameters.
This possibility indeed occurs, as is well verified experimentally in the
case of the flows governed by Eqs. (4.31) for a large variety of cases.
In this instance also we normally invert the variables. That is, the
mathematical form shows 0 as independent and U as dependent. How-
ever, this would be exceedingly awkward experimentally, and we normally
prefer to treat U as independent and 0 as dependent.
In this particular problem we are still seeking a choice of parameters
such that the terms in the variables of Eqs. (4.31) will be made unity
order; but since we are treating 0 as dependent, we may not be able to
select a single value that will achieve this for all possible values of Reynolds
and Euler numbers. This is in fact the case. Hence we examine the
nature of possible solutions and approximations when RCL is small com-
pared to 1 when it is approximately one, and when it is large compared to
1. Take first the case where the Reynolds number is small compared to 1.
If RCL is small compared to unity, then Eqs. (4.31) show that regard-
less of the size of (L/O)2 one of the two viscous terms multiplied by 1/RCL
must be large compared to unity.
The order of magnitude of the terms near the body is as indicated
below each term in the following equation
U au + ii au = Eu ap + _1_ [a u + (~)2 a u]
2 2
If the flow field is infinite, we cannot guarantee that Eq. (4.32a) will hold,
for the reasons discussed in Sec. 4-6b. Indeed, it frequently does not give
correct results for the flow at infinity.
Similarly for confined flows or near the body, the y momentum
equation can be simplified for low ReL. It is also useful to multiply by
RCL; this yields for the two momentum equations
a2{i L 2 a2{i
L ApL ap
---+-+
2 --=0
J-L U ai
ai [)2 ail
(4.32b)
2
L ApL L2 ap a fj L2 a2fj
----+-+ - -=0
J-LU [)2 ay ai 2 [)2 ay2
Governing Equa.tions 133
If/u I"'"
iJi -
(iJU/iJY)y_8 -( iJU/iJy)y=O
8
O-U/8 U
~-o--82
FIG. 4.11
this definition of 0 does provide an estimate for d2il/dfP that is unity order
inside the boundary layer.
We must add that this normalization leads to approximate, rather
than exact, boundary conditions, since we replace
il = 1 at fj = 00 by il = 1 at fj = 1
This still does not settle the actual numerical value of o. However,
we observe from Eq. (4.31d) that if oiL = 1, then we would still drop all
the viscous terms and again be led back to the discrepancies of the Euler
equations. Some writers have accordingly argued that we must therefore
take
(L/ i3)2 = 1
ReL
or
f = ~R~L (4.33)
Thus we try dropping the term a2u/ax 2from the x momentum equation.
The y momentum equation has magnitudes as indicated
_aij _aij ap (L)2 1 a2ij (L)2 a2ij
u ax + v ay = - Eu ay B + ReL ax2+ B ay2
(1) (1) 0(1) ? »1 «1 1 1
p* = -'L
Ap.
where, again, the datum for p IS free stream pressure. t Then by
construction
ap*
~",,1
ag
The total term involving the pressure in the y momentum equation can
then be written
+Eu ap
ay
(f)2
0
= ~L Ap. ~p* (~)2
pU 2APL ay 0
",,1 »1
(~y «1
t The reader should note that the crux of this derivation lies in the distinction between
magnitudes in the x direction (L,tlPI) and y direction (8,tlps). See remarks on Huntley's
addition in Chap. 2.
136 Similitude and Approximation Theory
This follows from the fact that there is no other term in the equation
greater than unity, and there would be no way to create an equality,
unless t:.p, is at least this small. (t:.P6 could be smaller than this, but not
larger.) Consequently, to the accuracy of the boundary-layer approxi-
mation we can take
t:.P6 = 0
happen to be those very particular ones we would get from the solution
based on smooth estimates. The situation is like that found in the case of
the simple harmonic oscillator; we cannot match the boundary conditions
at all if we drop the derivatives of highest order. Here the particular
boundary conditions that would agree with smooth estimates are those of
the inviscid solution. Since this is not the case, a rapid change occurs
near the boundary, creating the boundary layer. This particular kind of
behavior frequently occurs when we try to drop terms involving the
highest order of derivatives in a given variable. It thus often pays to
assume the solution has the general form of Fig. 4.3 in such cases even
when data are lacking and then to check the result against the more
general equation. This idea is quite powerful, as we shall see in discussing
the method of zonal estimates in Sec. 4-8b. The basic concept, as we have
already stated, hinges on dividing the problem into a series of regions in
each of which we can supply uniform estimates. These regions need not
actually be at the boundary; for example, we often treat a shock wave in
the body of a fluid by similar approximations. Also we may need more
than two distinct regions.
A comparison of the procedure given above with the derivation of the
boundary-layer equations found in most texts will also show the advan-
tages of making terms in the variables <=::: unity. In particular, the great
improvement in knowledge embodied in the parameter (Ljo)2jReL com-
pared to ReL alone will be immediately evident to anyone familiar with
the solutions and correlations of boundary-layer theory. In connection
with the earlier remarks on force ratios (Sec. 4-5c) the particular Reynolds
number which gives the ratio of inertia to viscous forces for this type of
boundary layer is (Ljo)2jReL; it is therefore not surprising that it is a
peculiarly useful form.
In fluid mechanics, nonuniform behavior is very common indeed.
Since presently available mathematical methods are usually inadequate to
obtain solutions to the complete equations, use of data on overall flow
patterns for typical cases is extremely important. In this sense data are
even more crucial in fluid mechanics than in many other fields. As is well
illustrated by the discussion of Prandtl's boundary-layer equations, these
data need not be very detailed, but they must characterize the typical
overall flow pattern which actually occurs. Most of the really important
analytical errors in the history of fluid mechanics have arisen from an
implicit assumption of an overall flow pattern that does not actually occur
in nature for the values of the governing nondimensional parameters of
concern.
the similarity conditions for turbulent boundary layers. For the present
purpose the derivation provides one of the few available examples of the
use of intermediate derivatives and of the search for criteria in terms of a
priori undefined parameters. Despite the historical importance of the
result as a basis for construction of analytical forms of the mean velocity
profiles, the derivation has several shortcomings. However, these also
turn out to be instructive.
Define a stream function for two-dimensional, incompressible flow as
'1'1 = 'I'm + 'I'
where
'1'1 = total stream function
'I'm = stream function of mean flow
'I' = stream function of perturbation
Using these definitions, the Navier-Stokes equations for a two-dimen-
sional flow can be written ast
u::Z:um + a'l'
ay
v = 0 _ a'l'
ax
U = component of velocity in x direction
( )m = time mean
v = component of velocity in y direction
An equation for perturbation (fluctuating motion) only is then obtained
by expanding the original equation and substituting the conditions above.
This yields
a(V2'1')
at
+ (um + a'l')
ay ax
aV2'1' _ (a 2um + aV2'1') a'l' = II (V4'1' + asum)
ay2 ay ax ays
We next try to define nondimensional parameters which satisfy the
conditions of Rule 3 for the region which is not immediately adjacent to
the wall but slightly farther out, called the turbulent core or logarithmic
t See, for example, Schlichting,46 p. 58ff.
Governing Equations 139
(4.35)
and
Bl» 1 (4.36b)
p
and as before
(OUm/OY)o and (02U m /oy2)o are constant parameters dependent on the mean
velocity profile, not variable in size, in this treatment.]
_ OV2~ _ l(02Um/oy2)o o~ + B OV2~ a~
y ax (aum/ay)o ax l(aum/ay)o ax ay
B a~ aV2~ = 0
l(aum/ay)o ax ay
This equation requires that similarity between two different situations
can exist only if the two dimensionless parameters
B
and -O-:l(-=-au-m--;-/a-:-y-=-)0 = 11' 2
are the same in the two cases. If we are looking for a general sort of
similarity law, that is, one which holds for all problems of this class, then
it must follow that such a law can exist only if both of these parameters
are constant in all flows of this type. If we have normalized the equations
properly, this must be so, since if either 11'1 or 11'2 could vary, then some
cases would exist in which similarity would not occur. Thus to the
order of the approximation made, necessary conditions for a possible
similarity law can be stated as follows:
_ l(a 2um/ay2)o _ C _
11'1 - (aum/ay)o - 1 - 1£
(4.37a)
and
(4.37b)
We want B such that a~/ax = 0(1) and a~/ay = 0(1) for all cases. We
therefore set B2 such that Tip = B 20(1). Thus we can define B2 as:
B = v* = ~~
Thus we see that a definition for B that will satisfy the conditions for
Rule 4 is the familiar "friction velocity" of hydraulics. Adopting this
142 Similitude and Approximation Theory
Toven = To sin ~
T
+ Ta (4.38)
t- = -t > 1 (4.40a)
te
Also from Sec. 4-6a, te ~ l2/ Ot• Inserting this into Eq. (4.40a) gives
t» !::.Ot (4.40b)
Since we are not now interested in the early transient period, we assume
Eq. (4.40b) is satisfied and examine only the periodic steady-state
behavior.
Since the Biot number is large, the boundary conditions for this
portion of the problem are
T. = Tsurfac. = To sin ~
T
where all T's are measured from T" as datum. For simplicity, let us also
assume the body is roughly spherical, so that L ""'-' M ""'-' N ""'-' l; this
simplifies, but does not restrict, the analysis in this case. Equation (4.39)
thus becomes
T- • = T. . :;:-e t-)
To = sm (t
Thus, as before, we obtain a Fourier number l2/t eOt, but we also find the
added nondimensional parameter te/T from the boundary conditions.
We now consider the effect of variations in the parameter te/T.
Suppose first that
~«1
T
l'he solution of Sec. 4-6a shows that the entire body will reach a tempera-
ture within 1 percent of any newly applied surface temperature in a time
Governing Equations 145
~»1
T
Then we observe that the fluctuations are too rapid for the body as a
whole to follow. The typical depth of penetration of the temperature
effect, which we will call ~, can be estimated directly from Eq. (4.40b) by
reasoning as follows. The transient effect is important only where the
Fourier number l21ta is ~ 1. Hence, if we want to find the depth to
which the temperature change is roughly 1 percent of that applied (T.),
then we must find the section where l2jta ~ 3;t the value of l at this point
gives us~. Since we know the time involved (time for one-quarter
cycle = ; T ), we can solve for the length ~
(4.40c)
t The use of 3te gives an estimate, based on exponential behavior, that is appro-
priate for linear equations.
146 Similitude and Approximation Theory
pared with the cylinder wall thickness allows operation with gas tempera-
tures far higher than those which can be utilized in systems where steady
high temperatures are necessary, such as gas turbine engines.
This example shows clearly that (1) a change in boundary conditions
alone can entirely alter the important parameters and (2) a boundary-
layer problem can occur even in linear, homogeneous equations. How-
ever, we still have the advantage of homogeneity, in that the system
response can be correlated independent of the magnitude of To when
measured in T coordinates. We can also still employ the linear property
of the governing equation to argue as follows. Superposition allows con-
sideration of the temperature wave To cos (t/T) as one component of a
Fourier analysis of a more complex wave shape. Both the simple and
more complex waves can be considered as departures from an average or
backbone curve which gives the steady-state solution. This wave is a
function solely of the mean inputs of heat, that is, temperature dis-
tribution, applied to the surface over a long time. In this way, a very
clear picture of virtually all of the important features of heat conduction
can be found from the differential equation and the boundary conditions
without need for solution. Moreover, such results are readily formulated
for bodies of very complex shapes; time-consuming and costly detailed
calculations can in this way often be avoided.
b. Zonal Estimates
2
E-d u + (1 - X-2) U- + U-2 = 0
di 2
at i = -1: u=o
at x = +1; u=o
Suppose that the variables u and x have already been normalized. Sup-
pose further that we are interested in a particular problem where we know
that 0 < E «1. Under these conditions boundary-layer behavior is a
possibility. On a trial basis we can assume that the solution is smooth
except near i = + 1 and x = -1, where we may have a boundary layer.
We look first for a solution applicable in the interior (away from the
boundaries) which is assumed to be smooth. Dropping the E d2u/di 2
term we have
------~~----------~-----------+~------~~%
We want to keep a2u/a~2 in the boundary layer and at the corner, since
we anticipate that the second derivative must be large near the corner
(see discussion relating to Figs. 4.2 to 4.4 on pages 102 to 105).
Accordingly we set i3 = -j; this gives
a2u
ajp + Ei~(2 - Ei~)u + u2 = 1
This equation is not easy to work with. For convenience we look for a
solution u which blends directly into uo• We define a new function w(~)
as
u = uo(x) w(~) +
Substituting this into the differential equation, we obtain
a2 -
E1+ 2/l a~~ + E-/l~(2 - E-/l~)W + 2u w+ w2
o
+ [E-/l~(2 - E-/l~)uo + u~ - 1] = 0
But by construction of Uo the three terms set off inside the brackets at the
end are zero. Also we want to retain a2w/a~2 for the reasons already
stated; so again we take f3 = --t. This gives
a2 -
a~~ + 2u w+ w2+ 2E!~W -
o E~2w = 0
Since E « 1, we drop the last two terms. The resulting equation inte-
grates once immediately on multiplication by aw/a~. This gives
1 !l-)2 + 2 - - + ~-3 - C
"2 ( uw
aij uow 3 -2
1
Since for large ~ we want to match smoothly to u., we require that both w
and aw/a~ approach zero. If this is to hold, we can have a solution for
large ij only if C t = O. This application of an extra boundary condition
(aw/a~ = 0 for a large ~) is typical of the matching required at the edges
of zones in boundary-layer methods. Moreover, we only need a solution
for wnear x = 1. In this region u = 1 + H where H involves terms of
order Et and higher. We can show this by inserting the definition of ~
into the solution for uo•
~ = (1 + x)E-i
x= - 1
~E!
x2 = ~2E - 2~ei 1 +
2uo = ± V 4 + (1 - X2)2 - (1 - X2)
= ± v'r-:4-+-:--7(2;o-,~-'et;-_-~--=-2e7) _ (2~ei - ~2E)
The last term has only terms in e1 and higher; the lowest power of E from
150 Similitude and Approximation Theory
the term with the radical is E!. Since we have neglected terms of this size
in finding uo, it is consistent to use
2uo = ± y4 Uo = ±1
in the equation for w. Moreover (o2wjofi2)2 is necessarily >0, and if
Uo = + 1, then for fi = 0, w = -1. But then the equation for wbecomes
-!( +value) + 2( -1)2 - i ,e 0
That is, there is no way to satisfy the differential equation for wand also
the boundary condition at fi = 0 simultaneously if Uo = + 1. Accord-
ingly we set U = -1. The boundary conditions for wthen are
O
at fi = 0: u=o
at fi ~ 00: u~uo
w= 3 [1 - tanh 2 (~ + A ) ]
where
A = arctanh ~i
This gives the boundary layer near x = -1. Since the solution is sym-
metric, we can write the boundary-layer expression near x = +
1 by
inspection as
where
r= (x - I)E!
And a solution for u is given in three pieces as
near -1 ~ x ~ -1 + r:
u= -1 +w
for - 1 + r ~ x ~ 1 - r:
u = -uo
1- r ~ x ~ 1:
u = -1 + ji
For this solution to be meaningful we must show that there is a 0 « 1 for
which w ~ 1, that is, fi» 1. If we can do this, the mathematics will
display boundary-layer behavior and we will have constructed a good
solution for the values of E which give us the properties of 0 and fi just
Governing Equations 151
Uo - 1 + alijEt + a2ij2E + .
and
Moreover
a. Poincare's Expansion
(4.41a)
t- = -t
T
d2x
dl2
+mx=0kT2 -
(4.41c)
at l = 0: x=O d::dt = 1
As in Sec. 4.6c, we seek an approximate solution under the fictitious
assumption that kT2/m ->, O. As before we find on integrating d2X/dl2
twice
x= Cl +c l 2
and
d~2- d2-® d2-® d2-® d2-®
_ _X_+ ~+ 2_X_+ 3~+ (4.42c)
dl2 - dl2 ~ dl2 ~ dl2 E dl2
(4.44)
We must now formulate the boundary conditions for the unknown func-
tion xeD. By construction
(4.42a)
and
(4.42b)
But we have already satisfied the initial condition for x@ and (dx/dt)z-o
by construction of x@. Thus we must take
xeD(O) = 0 x®(O) = 0 etc. (4.46a)
(a-at
~eD )
x=o-
-0 (adt-®)
~
x=o-
-0 etc. (4.46b)
Addition of the conditions (4.46a) and (4.46b) will then satisfy the original
conditions (4.41b). On setting Eqs. (4.46) into Eq. (4.45), we obtain
C4 = C3 = O. Hence
fa
xeD = - -
3!
By continuation of the same processes, we find
x® = +-t'
5!
17
x® =
7!
156 Similitude and Approximation Theory
d~ = 0
dt
And we cannot satisfy the boundary conditions. If we attempt a
Poincare-type expansion, we set
x = x® + I'x<D + 1' 2x<D + . . .
dx _ dx® dx<D 2 dx® .
dt -Tt+I'Tt+1' Tt+
d2x d2x® d2x<D d2x<D
dt2 = dt2 + I' dt2 + 1'2 dt2 +
Governing Equations 157
but
so
b. Lighthill's Expansion
(4.48a)
at t= 1: x= 1 (4.48b)
(4.49)
or d(xt) = 0
dt
The solution so obtained is
xt = Cl = 1 (4.50)
where 1 is found from Eq. (4.48b). Equation (4.50) does not exhibit
the proper behavior as E ~ O. Hence we try the Poincare expansion.
Again let
x = x® + EX<D + E X<D + 2
then
dx _ dx® dx<D 2 dx®
dt - dt +E dt +E dt +
Substituting into Eq. (4.48a) gives
. ( dX®
(t + EX® + E2X<D + E3X® + ) dt + E dx® dX®)
dt + E dt
2
+ x® + EX<D + X® + . . .
E2 = 0
with boundary conditions
at t = 1: x® = 1 etc.
Solution for x® from the terms without E gives
d -®
t ~t + x® = 0
160 Similitude and Approximation Theory
or, as before,
lJJl = 1
-to' _ 1
x'" - ""t
But this solution not only has the wrong values for small E it also has an
undesired singularity at 1 = O. So we try the solution for lSJ. We
obtain on equating to zero terms in E to the first power
d -® d-<D
x® ~l +x ~l +x<D=O
Solving by inserting the solution for x® and d;; and by evaluating con-
stants from the boundary conditions gives
x<D= -~(1-~)
Thus x<D has an even worse singularity at 1 = O. Computation of higher
order terms only makes matters still worse, as the reader can verify for
himself.
When artificial and undesirable singularities of this sort appear, we
try next the expansion method of Lighthill. The essence of the method
is to expand not only the dependent variable but also the independent
variable in powers of the small parameter. To do this we define a new
variable :r which replaces 1; we then seek a form for :r which will eliminate
the singularity and hence make the problem uniformly valid in:r. We
proceed as follows. Let
x = x®(:r) + EX<D(:r) + d!)(:r) + .
and
1 = :r + El<D(:r) + E2l®(:r) + ...
Substituting these expansions into Eq. (4.48a), we obtain
- _ dx d:r _
(t + EX) d:r dl + X = 0
or
- _ dx _ d:r
(t + EX) d:r + X dl = 0
and writing terms only to first order, we have
[:r + El<D(T) + ... + EX®(:r) + EXQ)(:r) + ...J
2
d-® d-<D ]
[ ::r + E ::r + . .. + [x® + EX<D + ...J
dl<D
[1 + E d:r +
Governing Equations 161
- dd1' +x
-0
T
x -(!) _
-
0
(4.51)
But the variable is now 1', not t, and l' is still subject to choice, that is, we
have not yet determined its form. Equating to zero terms in E to first
power gives
-@ d-0 dl@
- dx
T d1'
+ x-@-(t-@+-0)x-0
- - x d1' - x d1'
t-",
o 7 ( 1--
=-
2
1)
1'2
Then at this order of approximation
_ 1
x =-:: (4.53a)
T
and
-t=1'+E 7 ( 1-",f2
2 1) (4.53b)
(4.53c)
c. WKBJ Expansion
However, this does not lead to converging series. If instead one trans-
forms the equation by setting
R= cU(x)
eters are encountered. For details, the reader should see the references
above.
1. There is only one solution possible for the dependent variable for
all problems of the class considered in the new coordinates.
2. This solution must hold for all values of both the variables and the
parameters of the original equation.
at t = 0: x = 0 dx = 0
dt
We normalize this equation by use of the variables found previously
_ x
X=-
o
- t
t = -
T
Governing Equations 165
U(1)
t Note t+ also contains the previous type of similarity, since for fixed time
behavior we require k and m or kim = constant.
Governing Equations 167
/
/
O__~------~~--~r--------------r--------~t+
/ t+=T..jk/m
/
---r-
k
Static equilibrium
position
r----'----,
FIG. 4.12 x
168 Similitude and Approximation Theory
at t 0: dx = 0
dt
:3
d2x
dtH
+ x__- Po (fJ) +
«lAk cos yk/m t
(4.58)
dx
at t+ = 0: X= 1 dt+ = 0
we must start with a condition which makes the initial conditions parame-
ter-free. This is, of course, true for entirely undamped systems, although
it is unnecessary for real systems at t+» 1, since any finite damping
results in a steady-state condition independent of the initial conditions
after a long time. In this case, if we had taken x = 0 ~ OA at t = 0,
then the initial conditions would have read
dx
at t+ = 0: dt+ = 0
Such forms are not usually plotted, since all real systems have damping.
It is interesting to note again in this example the effect of Theorem 1.
The homogeneous equation (4.56) for the free vibration in natural coordi-
nates is entirely parameter-free so the solution does not depend on the
value of the parameters. The solution to the nonhomogeneous equation
(4.58) for the forced vibration does depend on the value of the parame-
ters; however, both cases are linear.
where
- =y- x
y l = tw x=-
Oinit L
at x = 0: y=O (ly = 0
(lx
(l2jj (lay
at x = 1 : (li2 = 0 (lxa = 0 (4.24b)
at t= 0: iix=l = 1 Cii)
(It ;:=1
=0
where
L beam length
=
w = reciprocal of time for one-quarter cycle
r = radius of gyration of beam section
e = acoustic velocity in beam
An appropriate t+ is then
x+ = /£2w x = /w Lx _ _ x_ (4.60)
'\j re Vrc L - vre/w
Hence we observe the natural unit of length for free transverse vibrations
as Vre/w. However, this produces a parameter in the boundary condi-
tion at x = 1, since x+ then equals £2w/re. Moreover, w is in general
unknown initially, so the coordinates of Eq. (4.59) would appear to be a
better choice than those of Eq. (4.60). Inserting Eq. (4.59) into (4.28a),
we have
(4.61)
and the boundary conditions (4.24b) are unaltered. From Eq. (4.61) all
the results of Example 4.3 are easily achieved by reasoning as follows.
For a free vibration to occur, the terms in Eq. (4.24a) must be of the same
order. For smooth vibration curves, which are expected on physical
grounds, iJYJ/iJx 4 = U(l) and fj = U(l); it follows that f must run 0 ~ 1
over a one-quarter cycle. This leads directly to the estimate found pre-
viously as Eq. (4.27a), namely,
w =
[ET
'\}PAJ) =
rrc
'VV rad/sec (4.27a)
specify similar vibration behavior for any value of the parameters. Equa-
tion (4.24a) shows that two beams will have similar behavior for the same
values of x, t, and Vw/rc. But Eq. (4.61) shows the same result can be
achieved by requiring constant values of x and t+. Hence we need not
require constant Vw/rc if we adopt V/rc as the unit for time measure-
ment. This, in turn, suggests correlation of beam vibrations with a time
coordinate t+. This correlation is more powerful than necessary for the
free vibration alone. However, it is useful for forced vibrations. Con-
sider a uniformly distributed forcing function on a cantilever beam of
FIb per unit length. The governing equation is then
a4y a2y
EI ax4 + pA iii! = Fo cos (3t (4.62)
(4.63a)
at x= 1: (4.63b)
at l = 0:
And
{3V {3 forcing frequency
- ---
rc w natural frequency
FoL4
EI A length
--_~ V~
Flow
FIG. 4.13
( aq;) _
ay 1/=0 -
U(ddxY)
1 body
(4.64b)
(4.65)
Let
- = -cp-
cp
U1L
_ x
x=L
- y
Y=L
where
L = body length
Then Eq. (4.64a) becomes
a2~ I a2~
ax2 + I - M2 afp = 0 (4.66a)
( a~)
ay 1/=0
L (acp)
= U lL ay 11=0 =
(ddxY) body
(4.66b)
(ay+a~ ) 11+=0 = VI -
I
M2 ay
(a~)
ii=O
I
= VI - M2 dx
(dY)
body
(4.70)
d~) =
(dx tf' (x) (4.72)
body
(4.73)
We further define
t+" = t
VI-M2
Then Eq. (4.73) becomes
(aya~) 1/+=0
= t+f'(x) (4.74a)
Thus two bodies, say a and b, will have the same governing equation and
boundary conditions in ip, x, y+ coordinates if we adjust their thickness
factors to give the same t+". That is, the required rule for constancy of
(Cphody is
~.
t+ = constant = --;===::~ = --;=:::::::::;~ (4.74b)
VI - Ma 2 VI - Mb 2
where Ma and Mil are the free-stream Mach numbers of bodies a and b.
Thus to obtain constant (Cp)body in ip, x coordinates, we must adjust the
thickness of the body according to Eq. (4.74b), but retain the same family
or shape, that is, the same f(x).
Several points are of interest here. We observe that a different
normalization of ip would not affect Eq. (4.69), since it is homogeneous
in ip; it might, however, affect the expression for Cp• For example, if one
defines
tp* = B~ = Bip
U1L
where B is an arbitrary nondimensional constant, Eq. (4.69) is unaltered
and the boundary condition (4.70) becomes
(aip:)
ay 1/+... 0
a~),,+=0 = Bt+f(x)
= B ( ay (4.75)
Since we are here concerned only with constancy of the boundary condi-
tions for varying values of M, Eq. (4.75) meets the requirement. How-
176 Similitude and Approximation Theory
(4.76)
The constant B is thus a free parameter, that is, we can give it any con-
venient value. Thus we have not used up all the choices available to us.
By proper selection of B it is possible to derive all the usual similarity
rules for compressible flow under these conditions. These rules are called
the Prandtl-Glauert rules and Gothert's rule. These relations are sum-
marized in recent texts on aerodynamics along with examples of applica-
tion; a particularly good discussion is given by Liepmann and Roshko. 30
The equation as given above would result in imaginary values of t+
for supersonic flow (M2 > 1). However, restudy of the equations shows
that it is only necessary to insert ViI - M21 for VI - M2 in order to
make the rules correct for both subsonic and supersonic flow. t These
rules do not hold for M '" 1, however, since the Eq. (4.64a) is not com-
plete for such cases. Transonic flow is discussed in the next example.
Liepmann and Roshko 30 develop results identical to those above by
using two sets of coordinates, one represents a body a and the other body
b. They then demonstrate that the equations for body a will yield those
of body b, provided the coordinates are selected to satisfy the equations
for y+ and t+, that is, Eqs. (4.68) and 4.74b). The reader may want to
verify that either procedure is equivalent to stating an invariance prop-
erty of the equations and boundary conditions in ;p, X, y+ coordinates
with boundary conditions stated in terms of t+.
It is noted that the rule for similarity in the flow is not the same as
that for the boundary conditions. Similar points in the flow field are
given by constant values of y+, that is, by
y+ = VI - M2 y = constant (4.68)
Similar boundary conditions are maintained by constant values of t+,
that is, by
t
t+ = = constant (4.74b)
VI- M2
Thus, for example, as M increases from zero in subsonic flow, we retain
similarity by using a "narrower" body according to Eq. (4.74b), but the
effect on the flow field becomes "wider" according to Eq. (4.68).
Finally, it is again emphasized that similarity rules can be found
directly upon establishment of natural coordinates for the equations and
t Note, however, one does not compare a subsonic body with a supersonic one
or conversely.
Governing Equations 177
(4.77)
where all symbols are as in Example 4.11 and'Y = ratio of specific heats.
The boundary conditions can also be taken as those of Example 4.11.
Again, we seek a rule which preserves (Cphody. Since Eq. (4.77) is
not homogeneous in cp, we must expect to lose the free parameter B of
Example 4.11; indeed to find a similarity rule we must take a definite
value for B. Accordingly, we normalize using the variables
* Bcp
cp = UL
_ x
x=L
- y
Y=I
Equation (4.77) becomes
UL i)2cp* 1 UL i)2cp* ('Y + 1)M2 U2£2 i)cp* i)2cp*
BL2 i)x 2 +1- M2 BL2 i)fP = 1 - M2 UDlfi i)x i)x 2
Dividing by U / BL yields
(J2cp* 1 i)2cp* ('Y + 1)M2 1 i)cp* i)2cp*
i)i 2 +1- M2 i)y2 = 1 - M2 B i)x i)x 2
(4.78)
(4.79)
then
* _ ('Y + I)M2 IP
IP - I - M2 UL
The boundary conditions at infinity are again unaltered by the trans-
formation of variables. The boundary condition
(:;)y=o = U (~~)bOdY
becomes
or
iJy+
*)
( iJ IP
1/+=0 = VI -
B M2 (ddxY) body =
(dY)
('Y + I)M2
(1 - M2)! dx body
d~)
( dx body
= t fUe) (4.81)
Then
ay+
*)
( iJ IP
1/+=0
= ('Y + I)M2 t . f' (x)
(1 - M2)t
We can achieve a similarity rule by requiring
+ _ ('Y + 1)M2t _
t - (1 _ M2)1 - constant (4.82a)
so that
(4.83)
y Velocity
coordinates
'-------~u
Space
coordinates
l
'------~x ~~~-~~------------~~
FIG. 4.14
best made clear by example, we defer further general remarks to the end
of this section.
Example 4.13. The Suddenly Accelerating Flat Plate. Con-
sider a thin plate of infinite length initially at rest in the coordinates shown
by Fig. 4.14. The plate is immersed in an incompressible, viscous fluid of
infinite extent, which is also at rest. At time t = 0 the plate is accelerated
to velocity U. The motion of the fluid layers for times greater than
t = 0 is sought. For this case all but two of the terms in the governing
momentum equation are zero, so the governing equation reduces to
au a2u (4.85a)
iii = p ay2
Then
T'Y = pTY"
where primed quantities denote differentiation of the function with respect
to its argument. Upon division by TY we obtain
T' Y"
T = Jl y = constant = -A
T = cle- Xt
Y = C2 cos (~ Y) + Ca sin Y) G
Thus the expression for the velocity is
where
TJ = TJ(y,t)
The resulting equation would be second-order in the new independent
variable TJ; it would thus require two boundary conditions on .", unlike
the original equation where three boundary conditions in y and tare
required.
This reduction in number of boundary conditions requires that two
of the original boundary conditions be related in a specific way if we are
to obtain an exact solution of the transformed equation. In particular
it requires
u[.,,(a,y)] = u[.,,(t,{3)] (4.87)
Condition (4.87) follows from the fact that, if it were not true, it would
then be impossible to satisfy all three of the original boundary conditions
in terms of only two boundary conditions on .".
Thus two of the boundary conditions of the problem as originally
set can be written as
at t = a: u = u[.,,(a,y)]
(4.88)
at y = (3: u = u[.,,(t,{3)]
where
where a is a constant and nand m are real numbers which may either be
greater or less than zero; their exact values are to be determined.
184 Similitude and Approximation Theory
In order to keep our new variables distinct from the old set, we choose
a second new variable ~ to be equal to t, the old independent variable
appearing in the denominatort of 1]. Next we require the resulting equa-
tion, after the transformation of variables t, y ----> ~,1], to be separable into
a function of ~ alone and a function of 1] alone.
By the chain rule of calculus we can write au/ay in terms of the new
coordinates ~,1] as follows
(4.89)
t This is the simplest choice for r from the standpoint of the mathematics involved
in the transformation (because the highest order derivative of u with respect to t is
less than that of u with respect to y). Whether this choice for r works or not must
be found by carrying through the analysis; if it does not work, the only conclusion
obtained is that the assumed transformation does not yield a similarity solution.
Governing Equations 185
Examination of Eq. (4.90) now shows that it will separate into a function
of ( alone on one side and function of 7J alone on the other if (I-2m =
constant,t that is, 2m - 1 = 0 or m =!. Thus we find the required
separation variable is 7J = ay/vt. Again defining A as the separation
constant, Eq. (4.90) becomes, on insertion of m = t
-g' t = -TJ f'-
g 21
+ /la 1"1 = A
2 - (4.91)
Since b, CI, and 1(0) are all constants, we have the condition that A = O.
The function g is then determined; g = constant = CI. Equation (4.91)
can then be written as
(4.92)
Y
TJ = 2 V~
at TJ = 0: 1= 1 (4.93b)
at TJ = ao: 1=0
t This requirement that the coefficients containing t (or I) have zero exponent
is the general condition required to fix m. See, for example, Yang56 and Hansen.u
186 Similitude and Approximation Theory
erf x _ =
y2
1
y211'-z
J e-
z t'/2 dt
In this case, the investigation of similarity properties has led us all the
way to a complete solution of the problem. Equation (4.S5a) can also
be solved in similarity variables in somewhat more general form without
any assumption of initial condition. One can then seek the initial con-
ditions for which the solution can be shown to hold (details are given by
Abbott and Kline 1). One employs the more general transformation of
variables
ay
71 = 'Y(t)
where 'Y is an unknown function. We will not give the details here, but
instead present two other examples and then turn to a discussion of the
meaning of such solutions, their similarity properties, and some general
remarks about solutions of this type.
au au a2u
U-+V-=II- (4.94)
ax ay ay2
(4.95)
at y = 0: u=o V=o
at 00' u = U
Governing Equations 187
ay {Old vars.
1/ = y(x)' and ~ = x :
Since a, b, I'( 00), and Uo are all constants, the necessary relation is
g/'Y = constant = C2. Solving the two relations "'(u' = Cl and g/'Y = C2
simultaneously gives
and
Evaluation of the term "'(g' - "'('U shows it is identically zero. Hence the
separation constant A is zero, and Eq. (4.99) becomes
I'" + c1b
va
ff" = 0 (4.100)
'fI -
-
Y I +
U
\j v(x
xo)
(4.103)
(4.104)
(4.105)
knowing c, and Ca. The conclusion, as stated above, remains the same: another rela-
tion between 'Y and g is required.
190 Similitude and Approximation Theory
2p fo u dy =
00 2 constant (4.107)
---
Velocity-O
--
y
........-
........-
//
/"'"
•
/
x
----- ---
........................
FIG. 4.15
Governing Equations 191
then becomes
, , ,ff"
'Yg - 'Y g = 'Yg f'2 + b"a f'"1'2 (4.109)
This is the same equation which was found for the Blasius problem, Eq.
(4.99), since the same partial differential equations apply-only the
boundary conditions are different. The boundary conditions correspond-
ing to Eq. (4.106) are
at." = 0: bg'f(O) =0
at." = 0: ba2 JL f" (0) = 0
'Y2 (4.110)
at." = 00: baflf'(oo) =0
'Y
For Eq. (4.109) to be separated, it is necessary that
'Yg' = constant = Cl
A second relation between 'Y and g can be found by considering Eq. (4.107)
(see footnote, p. 188), which can be written as
10" u dy = constant = C2 (4.111)
r" a2b2 'Y2g2 f'2(.,,) !a d." = b2a g2'Y Jr"0 f'2(.,,) d."
J0
= C2
The solutions to these equations can be found in closed form and are
given, for example, by Pai. 38
The solution to this problem shows how the integral boundary-condi-
tion equation (4.111) is used to obtain the second relation between 'Y and
g necessary to determine the proper similarity variable from the family of
possible variables found by considering only the partial differential equa-
tion (4.108). In this solution, Xo represents the potential core or starting
length, and could be determined from appropriate data. t
We now turn to some general remarks about similarity variables.
First, we discuss the physical meaning. In all three cases studied we have
reduced an equation in two independent variables to one with a single
independent variable. In Examples 4.13 and 4.14 we found a solution
dependent on the new coordinate Tf alone. However, in Example 4.15 we
found the solution for the stream function dependent on the new coordi-
nate Tf and also on x, since the function g(x) is not a constant in this case,
and by construction
'IF = ag(x) . J(Tf)
Thus there are at least two distinct cases. Actually, it is instructive to
consider three classes. In all cases we have sought solutions in terms of
new coordinates purposely constructed so that separation of variables is
achieved in the conventional sense. Using coordinates x,y initially, new
coordinates r,Tf as above, and a dependent variable 'IF,
1. The most general separated product form is
d 1!-constant-cl
_ _+---,1!-constant-c 2
FIG. 4.16
-- --
y
1 //,/'""'
yax 2/ 3
u/um-constant
---
FIG. 4.17
The second reason that one cannot find all possible variables readily
is best illustrated by considering how one seeks such variables. If we
set a well-posed problem, then we write down the complete boundary con-
dition, and, by definition of such problems, only one solution exists.
Under these circumstances we can find only one or no similarity variables
under a given transformation, since otherwise a contradiction to the
existence of a unique solution occurs (barring, of course, functionally
dependent transformations, such as from 'YJ to 'YJ2). However, if we want to
study similarity variables of a given equation with the view toward estab-
lishing a number of similarity properties and/or solutions, then we must
omit some or all of the boundary conditions. It may then be possible to
find an infinite number of different similarity variables for anyone assumed
transformation of variables. This is the case for the boundary-layer
equation under the transformation 'YJ = ayh(x) using a product solution.
This point can be seen from the footnote of page 188; it shows that the
boundary-layer equations are satisfied by an 'YJ of the form
ay
'YJ = (ax + b)p
where p is an arbitrary exponent related to the separation constants Cl
and Ca of the required differential equations for 'Y and (J. The relation is
1
P = -,---,--,---:-
1 +
(cI/ca)
There are thus an infinite number of forms of 'YJ which will separate Eq.
(4.108). The equations in 'Y and g are relatively simple in the case of the
boundary-layer equations, but they can become quite involved.
It follows from the above remarks that the specific form of a suitable
similarity variable depends as much on the boundary conditions as on the
equations. In all three of the examples above, the form of the similarity
variable is really picked out from the boundary conditions-either at the
outset by hints, or later on explicitly. Indeed, the possibility of finding a
solution, a correlation, or a similarity property in terms of a similarity
variable for a given differential equation depends essentially on the bound-
ary conditions. This can be seen from condition (4.87). Consideration
of the integrations needed also shows that if the boundary conditions
cannot be expressed in terms of the similarity variable 'YJ alone, then one
of the original coordinates x or y must appear in the limits of integration
and a solution in terms of 'YJ alone cannot be obtained. A solution in
terms of 'YJ and a weighting function g(x) may be possible, but this appears
highly unlikely. A more promising approach has been adopted by H.
Gortler.16 Gortler has employed separation coordinates to reduce bound-
ary-layer problems in viscous fluid flow to series expansions in just one
196 Similitude and Approximation Theory
the same class and can be found from appropriate normalization of the
governing equations and boundary conditions. The second type is
equivalent to a reduction in the number of independent variables and is
based on constant values of the new variables. It describes a purely
internal similitude relating two points both inside a single system of the
class considered. The third type is based on constant values of a natural
coordinate which combines parameters with variables to eliminate one or
more of the parameters. It is equivalent to measurement of behavior in
units based on the nature of the system itself. The third type of simi-
larity often contains the first type and can be employed to "swap-off"
changes in value of parameters for changes in value of the original
variables. It thus provides a rigorous basis for distorted models and
allows construction of similarity properties which may be impossible from
consideration of the parameters alone. Understanding these relations
and the concomitant processes in the governing equations provides a more
general and more unified picture of similitude than can be achieved from
study of the parameters alone.
All three types of similarity can be derived from the normalized
governing equations and conditions without need for solution.
When the governing equations and conditions are normalized sys-
tematically, a great deal of information can be obtained. This includes
establishment of a set of governing dimensionless parameters which is as
rigorous and complete as the governing equations and conditions.
Governing parameters found in this way are almost invariably par-
ticularly useful and they can be made to express directly the ratio of
magnitudes of the important effects in the problem if sufficient informa-
tion is available. However, when this is done one does not usually obtain
the standard form of the parameters.
Use of information on magnitudes allows derivation of approximate
equations and yields approximate solutions and similarity properties. It
thus provides a numerical criterion for when one or more parameters can
be dropped from a given correlation or may become unimportant. How-
ever, all procedures based on magnitude are more subtle and prone to
difficulty than the similarity procedures based on the complete equations
and conditions. A sharp distinction should be made between the simi-
larity conditions which can be made as rigorous and complete as the
knowledge of the equations and conditions on the one hand, and the
approximation-theory procedures for which only relatively weak Ilrules"
can be given on the other hand. It is typical of approximation-theory
procedures that many hints about trial procedures are used as well as
some information about the general nature of the expected solution.
This philosophy, that is, using available information and hints, and
then checking the answer against the complete equation and the physical
198 Similitude and Approximation Theory
in several problems, but do not seem to have been widely used in other
fields.
In fields such as hydrodynamics, where the complete basic equations
are intractable to exact solutions, the various methods just enumerated
provide the primary basis for calculations. It is probably fair to say that
all but a handful of the existing solutions in fluid mechanics depend on
one or more of the techniques just discussed. It follows that clear insight
into such fields of study and even understanding of the current literature
are to a large degree contingent on clear understanding of these processes.
A review of the procedures above shows that all the similitude results
demonstrated or indicated can be viewed in either of two equivalent ways.
First, they can be considered as a means of describing two or more dis-
tinct situations by a single set of equations and conditions. Second, they
can be viewed as an invariance under some form of coordinate trans-
formation. Thus we can say that the external similitude of dimensional
analysis can be expressed by reducing the equations and conditions for a
specified point in each of two distinct problems in a single class to a single
normalized equation with constant values of the parameters and variables.
We can also view the same property as invariance of the equations for a
given point in the system under transformation from the parameters of
one system to the parameters of the other in the normalized equation.
Thus it can be considered either in terms of reduction of all problems of a
given class to a single equation or as constancy in the value of the govern-
ing parameters under transformation. These two processes are mathe-
matically equivalent. If substitution of the values of one problem gives
the identical result in the normalized equation as that of the other, then
by construction both problems can be reduced to the same normalized
equation; the difference is only one of viewpoint.
Similarly, the symmetries due to self-similitude embodied in the
similarity variables of Sec. 4-lOb can be viewed either as invariance
under transformation of certain coordinates (as Morgan has shown) or as
a reduction of all problems of the class with a certain type of boundary
condition to a single simpler equation in terms of the same coordinates.
Finally, the natural coordinates discussed in Sec 4-lOa show that problems
with different values of the parameters can be brought to the same equa-
tion if we can find transformations to new variables which make the
equations and boundary conditions parameter-free. This too can be
viewed as invariance under coordinate transformation with the given
boundary conditions in the natural coordinates.
200 Similitude and Approximation Theory
c. Final Remarks
attempt has been made to illustrate the various major types. Moreover,
there are obvious gaps in the mathematical foundations which one hopes
professional mathematicians will fill as time goes on. These include not
only existence and uniqueness theorems for specific classes of equations
and boundary conditions but also additional theorems for approximation-
theory procedures and more adequate rules for anticipating when non-
uniform behavior is to be expected as well as its type and probable loca-
tion. There are also many places where extensions are possible in view
of the framework of the subject exhibited by this discussion. Some
obvious examples would seem to be applications of the technique of
natural coordinates to additional problems as a basis for distorted models,
extensions of the investigations of solution of nonlinear partial differential
equations by use of separation coordinates for more general boundary
conditions than have heretofore been examined, t more systematic
investigation of other transformation functions for construction of both
separation and similarity variables, and attempts to apply simultaneously
the boundary-layer idea with the uniformization idea as a more powerful
means for dealing with troublesome singUlarities. Finally, the rapidly
developing theory of integral equations and integral transformations,
almost entirely omitted here, can be used with good effect in the context
of fractional analysis. The possibilities for systematization and clari-
fication inherent in procedures based on the governing equations by no
means appear to be exhausted. On the contrary, only a beginning has
been made.
We turn now to a comparison of the available methods in the final
chapter.
t Since the first draft of this work, one example of this procedure has been pub-
lished, see Gortler,16 pp. 4-187.
5 Summary and Comparison
of Methods
5-1 INTRODUCTION
a. The Pi Theorem
rems on the equations are not known. Here we are considering cases where the
equations are complete and appropriate. Sedov 46 and Smith48 have also stated that
the conditions can be made rigorous, but did not elaborate.
t See Examples 4.11 and 4.12 in Sec. 4-10.
206 Similitude and Approximation Theory
a. Power
b. Rigor
Essentially the same comparison exists for rigor as for power. Use
of the differential equations can be made as rigorous as the equations and
the state of mathematical and physical knowledge about them. On the
other hand, exceptions occur to the pi theorem method even when the list
of quantities is correct. In fact, it seems fair to say that the establish-
ment of the list of quantities for solution by the pi theorem and the list of
forces and energies employed in the method of similitude both represent
arguments to plausibility rather than rigorous procedures. As such they
are particularly liable to error and difficult to check without additional
procedures. Undoubtedly, some of the governing differential equations
we use are incorrect in some of the problems we now believe they cover.
However, this is a matter that has at least had the careful attention of
many workers and is thus subject to a minimum of suspicion. Moreover,
it is a matter on which the accumulated empirical evidence of the scientific
method can be brought to bear increasingly as time passes; it is subject to
continuous and controlled improvement.
c. Accuracy
Using the pi theorem, the bulk of the class was never able to solve
problems in fractional analysis properly. On problems where the answer
was not known to the class in advance or readily found in the literature,
the percentage of correct t answers achieved in finding a list of governing
pi's was in the range of 10 to 30 percent in all classes. Using similar
standards, the percentage of correct answers employing the method of
similitude was 25 to 60 percent, and using the governing differential
equations, 70 to 90 percent. Results found by students from the pi
theorem invariably showed very large scatter in the number of groups
found. For example, in a problem where external similarity can be
specified by use of two independent nondimensional groups, typical
answers would be received with one to eight or ten groups. Moreover,
plots of frequency of occurence of the number of groups found by the
students failed to reveal a cluster about some single number. The plots
were typically irregular, thus suggesting basic inability to cope with the
problems by such a technique rather than a consistent error of some sort.
It is possible that the writer's bias affected these results, but this is con-
sidered unlikely for three reasons: (1) after early instances of such results,
every attempt was made to eliminate the author's opinion from the
remarks'on the pi theorem for two consecutive years; no noticeable change
in results occurred; (2) the majority of the students had used only the pi
method in prior classes, and some of them invariably were initially biased
in its favor and said so; (3) similar results have been reported to the
author by instructors at several other universities; unfortunately no
statistics from these other sources are available.
Further discussion of the possible reasons for these results on accuracy
may be useful in order to illuminate comparisons regarding the rationale
of the various methods. On the surface the methods of similitude and
dimensional analysis appear accurate in the main, even though some
troubling exceptions occur now and then. However, study of the many
published examples from which this appearance stems shows that most of
these examples are not new solutions to research problems, but are merely
formalizations of previously known information from experiment or from
more powerful theory. Study of the literature reveals relatively few new
problems in fractional analysis actually worked for the first time by the
pi theorem. Moreover, as already mentioned, the real difficulty in
employing the pi theorem for fractional analysis does not appear to be the
exceptions which now and then occur, but is instead the failure of the
method to provide any means for direct inclusion of the full physical
information available. The impact of this deficiency on accuracy can be
seen in part by examining the types of questions which the analytical
t The word "correct" here is used to mean obtaining the minimum number of
pi's including the proper physical quantities in any form whatsoever.
210 Similitude and Approximation Theory
worker must face in using the pi theorem for fractional analysis. Within
the rationale of the pi theorem, one proceeds to seek directly a list of
secondary quantities representing the parameters. The analytical
worker is therefore forced to ask, "What are the physical quantities of
importance in this problem?" Thus in Example 2.2 (laminar flow in a
pipe) he must ask, "Does the viscosity act independently from density in
this problem, or should only their ratio, the kinematic viscosity, be
employed in the list of physical quantities?" In Example 2.3 (on heat-
exchanger analysis) the question that arises is, "Does the specific heat act
in this problem independently of the flow rate of the fluid, or can a single
product or ratio be used to represent them?" As has been shown by
examples, this type of question does not hinge on independence in the
sense of ability to vary the value of one of these physical quantities and
hold the other constant, but depends instead on the specific form in which
the parameters combine in a particular problem. Consequently these
questions can be answered with assurance only when the form of the relations,
that is, the equations governing behavior, are known.
Thus inside the framework of analysis using the pi theorem there are
two choices. The first is to operate under the assumption that nothing is
known about the form of the functional relation among the variables and
parameters. This was until recentlyt and probably still is the accepted
view in the English-language literature. If nothing is known about the
form of the relation among the variables and parameters, then it must
follow that it is unreasonable to expect even an able worker to answer
questions like those posed in the preceding paragraph correctly a large
percentage of the time. Under this first assumption, using the pi theorem
alone, the answers to such questions lie outside of the framework as well as
the details of the knowledge available; they are essentially "unknowable."
The reader can check his own reactions on this point by comparing
the solutions to the beam-vibration problem achieved in Examples 2.6
and 4.3, respectively. He should ask himself, "In reading Example 2.6,
did I at that point discover that Poisson's ratio was irrelevant to the
oscillation of simple transverse beams with small deflections?t Did I see
how to reformulate the five groups of Example 2.6 into the one more power-
ful group of Example 4.3?" Finally, in reading Example 2.6, "Was I
aware of the significance of altering the boundary conditions from homo-
geneous to nonhomogeneous on the required method of correlation?"
(Compare Example 4.4.) The tests with classes in several instances used
t Since the preparation of the first draft of this work, the treatise of Sedov 46 has
been made available in English; it clearly espouses a different view.
t The reader's indulgence is asked for this small booby-trap regarding Poisson's
ratio; experience suggests such a device is necessary to counterbalance the effects of
20-20 hindsight.
Summary and Comparison of Methods 211
viewing this same point is the following. As many writers have very
properly stressed, the mathematical procedures of dimensional analysis
alone yield very little; a clear understanding and use of physical behavior
in the analysis are essential. Direct introduction of the governing equa-
tions and boundary conditions is one method for supplying as much of
this physical information as is available in the general literature.
d. Simplicity
e. Input Information
Indeed, the general rule seems to be that the less we know about the
mathematical models for a given system the simpler the correlation
methods we are forced to use. Thus in highly complex turbo machinery
we can use ideas of geometric similarity and force ratios with governing
overall equations relatively successfully. In highly irreversible processes
we can only state certain limited mathematical symmetry properties. In
individual biological organisms we can make statements like "birth
processes of mammals are generally similar." Finally in the interactions
we call human affairs we still must rely, unfortunately, on the art of
politics as embodied in the often fallible judgment of individual statesmen.
When the equations and boundary conditions are known, but exist-
ence theorems guaranteeing a unique solution are not available, it is still
almost always profitable to employ procedures based on the equations.
A number of such examples have been given in Chap. 4; they show that
more information is obtained, a better basis provided for direct use of
physical data, and sometimes even information on missing data or theory
is revealed. The governing equations of viscous fluid flow fall in this
class. Indeed, since these equations are inherently nonhomogeneous,
since few exact solutions have been found, and since parameter-dependent
solutions dependent on stability considerations are so common, it is not
surprising that the main trend of advance in fluid mechanics in this
century has been based on various processes relating to fractional analyses
of the differential equations. Thus we find the bases of the whole field of
boundary-layer analysis, the methods employed in testing high-speed
aircraft, the construction of similarity solutions, and a number of other
results and techniques all emerging as different aspects of fractional
analysis of the governing differential equations and boundary conditions.
These methods and solutions are clearly useful even though mathematical
uniqueness still cannot be proved.
It is again emphasized that direct data must playa distinct role when
the equations are nonhomogeneous and the complete solution cannot be
established. To proceed effectively in such cases requires some knowl-
edge, or a shrewd guess, about the overall form of the solution.
the first group. Derivation of similarity rules and model laws from
governing equations fall into the second group. Approximation theory
procedures overlap, and provide some relations between, the procedures
of the second and third groups.
Moreover, as time progresses, we will develop complete equations for
more situations, and we will solve the easy problems in more and more
fields. Under these conditions the procedures of approximation theory
and more rigorous similitude processes must necessarily tend to become
more and more important.
Since we will never reach the situation where complete and appropri-
ate governing equations can be written for all problems of concern, we will
always need the more qualitative procedures of the pi theorem and the
method of similitude. It is well to bear in mind, however, that they are
qualitative in applying them in research problems where the situation is
not really understood in advance. This is one reason why the method of
similitude is useful as a semi-independent check on pi theorem procedures.
It also implies that any result found by these methods should be viewed
with a healthy scientific scepticism until empirical evidence sufficient to
justify results is in hand.
b. Implications in Teaching
The author is well aware that the conclusions expressed in Sec. 5.4a
are not in accord with those held by many writers and teachers in this
area. A number of authors have stated that the pi theorem is the begin-
ning and end of dimensional analysis and some of these authors have used
the term dimensional analysis to mean something quite close to what has
been called fractional analysis in the present volume. The vast majority
of published examples of fractional analysis employ the pi theorem, and
very little has been written on the methodology of fractional analysis
based on use of the governing equations. Some authors have even gone
to the extreme of using the pi theorem to "check" results obtained from
governing equations. t In many current undergraduate courses the pi
theorem is presented and the other methods are not. This is a con-
siderable weight of history and opinion. Nevertheless, if the conclusions
above are correct, or even partially correct, then the English-language
literature particularly has been relying too much on the method of
dimensional analysis alone and this method has often been pushed beyond
its useful limits. A reconsideration of the utility of the various methods
in both published works and undergraduate courses would seem to be in
order.
t In the author's opinion this is analogous to calibrating a micrometer with a
yardstick.
216 Similitude and Approximation Theory
Since the present volume is concerned with the development and use
of fractional analysis as a tool for the analytical worker in science and
engineering, it is appropriate to make a few closing remarks in an attempt
to assess what remains to be accomplished.
Dimensional Analysis
Since the late nineteenth century this method seems to have been
largely neglected. As is shown by the examples, it is a very simple
technique. In addition, it provides a good basis for systematizing and
utilizing physical intuition, and it provides a useful cross-check on answers
obtained from dimensional analysis. It would certainly seem profitable
for workers in various fields to prepare additional reference material of the
sort given in Tables 4.1 and 4.2 to make the method more readily appli-
cable to their specific specialties. Such tables not only serve as an aid in
Summary and Comparison of Methods 217
d. Final Remark
generality for the class of problems under study. We can state this as a
nonuniqueness theorem:
It may be possible to obtain a simpler and more general similitude prop-
erty for any given problem if we are shrewd enough to find it.
Apparently this theorem will be with us for some time, since it seems
unlikely that we shall be able to be more precise until considerably more
general information is available on governing equations, appropriate
boundary conditions, existence and uniqueness theorems for nonhomo-
geneous and nonlinear partial differential equations, and the transforma-
tion properties of differential equations. Thus our non uniqueness
theorem presents a challenge to the research engineer, the scientist, and
the mathematician. The challenge to the research engineer and scien-
tist is to find new coordinates which provide increased simplicity and
generality in various classes of problems. The challenge to the mathe-
matician is to increase the theoretical foundations for approximation
theory, for boundary layer and expansion methods, and for the trans-
formation properties of complex sets of differential equations.
References
50. Van Driest, E.: On Dimensional Analysis and the Presentation of Data in
Fluid-Flow Problems, J. Appl. Meek., Trans. ASME, 13: A-34 (1946).
51. Van Dyke, M.: Higher Approximations in Boundary Layer Theory, J.
Fluid Meek., part 1,14: 161(1962); part II, 14: 481 (1962); part III, 19:
145 (1964).
52. Vennard, J. K.: "Elementary Fluid Mechanics," 2d ed., John Wiley &
Sons, Inc., New York. 1948.
53. von Karman, T.: Mechanical Similitude and Turbulence, NACA TM 611
(1931).
54. Wecker, N. S., and W. D. Hayes: Self-Similar Solutions, AFOSR TN.
60-894, Air Force Office of Scientific Research (1960).
55. Wilson, E. B.: "Introduction to Scientific Research," McGraw-Hill Book
Company, New York, 1952.
56. Yang, K. T.: Possible Similarity Solutions for Laminar Free Convection on
Vertical Plates and Cylinders, J. Appl. Meek., 27(2): (June, 1960).
Name Index
-
Jacobsen, L. S., 122, 168
den Hartog, J. P., 168 Jeffreys, B. S., 152, 153, 162
Duncan, W. J., 4, 10 Jeffreys, H., 152, 153, 162
226 Similitude and Approximation Theory
McAdams, W. H., 28
Michal, A. D., 196 Van Driest, E. R., 18, 23, 28, 35
Morgan, A. J. A., vii, 180, 194, 196, 199 Van Dyke, M., 152, 163
Morse, P. M., 153 Vennard, J. K, 36, 63