Vous êtes sur la page 1sur 28

Article

pubs.acs.org/IECR

An Evaluation of the Usual Simplifying Assumptions


Stuart W. Churchill*
Department of Chemical and Biomolecular Engineering, University of Pennsylvania, 311 A Towne Bldg., 220 South 33rd St.,
Philadelphia, Pennsylvania 19104, United States
ABSTRACT: Some of the simplifying assumptions that underlie the characteristic concepts of chemical engineering are
identied and their impact is examined. Many of them are found to be obsolete. They remain in textbooks and perhaps in
computer packages out of inertia and, in some instances, out of misdirected respect for those who conceived them. It is
concluded that students, teachers, and industrial practitioners should question the continued validity of simplifying assumptions
and of the viability of the concepts and expressions that incorporate them.

1. INTRODUCTION
The formulation and adoption of useful concepts, and, in many
instances, ones that are unique to chemical engineering, have
helped it to ourish as an academic subject and as a profession.
The broadest and most notable concepts are illustrated by the
unit operations, the unit processes, transport phenomena, the
rate processes, and more-specic ones by the equilibrium stage,
the perfectly mixed reactor, the Hougen and Watson models,
the Ergun equation, and the Colburn j-factor. These concepts
have resulted from observations in plant operations and in the
laboratory as well as from theoretical analyses and ashes of
insight. All in all, they constitute an essential element of both
education and practice in chemical engineering.
In most instances, the formulation and generalization of each
of these concepts have been dependent upon one or more
ingenious idealizations or simplications. Over the course of
time, the idealizations and simplications that are implicit in a
particular concept often become lumped together and known
as the usual simplifying assumptions. They are thereafter
accepted without much question by students, teachers, and
industrial practitioners, even though advances in analysis or in
computer hardware and software may now allow their
elimination or provide the basis for their replacement. The
quantitative error arising from these idealizations and
simplications, if recognized, is often compensated for or
partially compensated for in industrial practice by a correction
factor or an eciency. That expedient may be acceptable in
practice in the short run, but the possibility of its elimination or
improvement by improved modeling should always be kept in
mind. A fudge factor or eciency is never an acceptable
substitute in textbooks or the classroom for an understanding
of the cause of inaccurate predictions.
Newtons fourth rule of scientic reasoning, which has stood
the test of time, can be expressed as Propositions collected
from observation of phenomena should be viewed as accurate
or very nearly so until contradicted by other phenomena. This
rule not only identies the source of new concepts and their
utility, even if approximate, but also notes that they should be
abandoned if and when proven false.
The criticality of a simplifying assumption is demonstrated
decisively by the history of the attempts to predict the speed of
sound in a gas. In deriving an expression for its prediction, early
2012 American Chemical Society

scientists (including Newton) made the seemingly reasonable


simplifying assumption of an isothermal process and obtained
dp 1/2 RT 1/2

ua =
M
d T

(1)

whereas the correct simplifying assumption is an isotropic


process, which leads to
dp 1/2 RT 1/2

ua =
M
d S

(2)

where is the heat capacity ratio. As an aside, Newton cleverly


fudged his experimental data, which actually agree closely with
the yet unknown eq 2, to provide support for the erroneous
isothermal expression. Fortunately, he disclosed the false
reasoning behind his fudging.
Because of the ubiquitous inertia in academia and in
industrial practice, the discovery that an idealization or a
simplication can be eliminated and that a concept can thereby
be improved or replaced by a better one often does not
immediately become common knowledge or precipitate
corrective action. In all of the arts, sciences, and technologies,
the practitioners, including teachers, are prone to resist change
and to cling to familiar concepts long after they have become
outworn or superseded. Anderson1 has noted that textbooks
and handbooks in chemical engineering ordinarily have a longer
shelf life between revisions than in other elds of engineering,
and even upon revision may not be brought up to date in every
respect. One easily implemented expedient is for a teacher to
call attention to errors and out-dated concepts and idealizations
in textbook that they assign, and to issue brief corrections and
updates in print or electronically. That procedure is often
resisted by teachers, because it imposes a burden to relearn and
replace familiar concepts with which they are comfortable by
new and unfamiliar ones that may stretch their mathematical
Special Issue: L. T. Fan Festschrift
Received:
Revised:
Accepted:
Published:
230

March 23, 2012


September 27, 2012
September 27, 2012
September 27, 2012
dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

ones. When I was fresh out of college and working in a


petroleum renery, I observed that operators, who were highschool graduates, had great skill in bringing a fractionating
column on-line and zeroing in on the optimal separation, even
though their concept of distillation was typied by the phrase
the reux knocks back the heavies. On the other hand, I
realized that the engineer in charge understood what was going
on inside that black box and was thereby able to choose the
optimal reux ratio and feed tray with little or no trial and error.
Before turning to several of the specic topics that comprise
chemical engineering and its practice, such as thermodynamics,
separations, uid ow, heat transfer, and reactor design,
attention is focused on general techniques such as dimensional
analysis and correlation that directly evoke simplifying
assumptions.

and computational skills. Students should be pleased to be


brought to the frontier of their eld by such an action by a
teacher but they are often irritated rather than pleased by the
revelation that they have been asked to acquire and depend
upon a textbook that is out of date or even outright wrong in
some sections. New concepts that originate in academia may
come to the attention of industrial practitioners only through
new recruits or by slipping in unnoticed in updated computer
packages.
If and when corrective action in the form of replacement or
updating of a concept is undertaken, the result is often quite a
surprise and sometimes a very benecial one by virtue insights
or discoveries.
Outdated concepts are relatively easy to identify in textbooks
and handbooks but not so readily in computer packages.
Advances in theory and in improvements in computer
hardware, software, and algorithms are not the only ones that
permit or warrant the elimination of idealizations that were
once thought necessary or acceptable, but attention herein will
focus on these two.
Concepts may serve a useful role by themselves in terms of
understanding. On the other hand, if they are to be applied in a
quantitative sense, they may need to be supplemented by
graphical, tabular, or algebraic correlations. These supplementary correlations are also subject to erroneous idealizations and
to obsolescence.
The objective of this manuscript is to identify some of the
idealizations and simplications that are no longer necessary
and some of the accordingly outdated concepts, practices, and
correlations, and insofar as possible, suggest suitable replacements. The process of identication herein is carried out most
conveniently in terms of specic examples, although some
general principles emerge.
Some concepts that have been dismissed prematurely are also
identied as well as some idealizations and simplications that
are now classied as false or obsolete but have historical
signicance in that they led to valid concepts that might not
otherwise have been discovered.
It is obviously not feasible to discuss the idealizations and
simplications in all aspects of chemical engineering, and the
chosen topics are only illustrative. Preference is given to those
that are well-known by all chemical engineers and thereby
require minimal description. The specic illustrations are
preferentially drawn from those limited areas of chemical
engineering in which I have some experience, and a number
include my own work because of rst-hand familiarity with the
details.
An identication of idealizations and simplications and an
evaluation of their validity should be included in every
publication of research ndings, but such identications are
often given short shrift and minimal attention relative to the
new ndings. Authors of books in chemical engineering should
accept the responsibility of identication and evaluation of
idealizations and simplications so the unjustied ones do not
continue to be taught and used unwittingly.
One of the reviewers of this article raised two worthy
questions. If an expression produces acceptable predictions why
should you care about the details and why should you tamper
with it? The answer is that new methodologies and more exact
expressions may produce signicantly improved designs and
greater production, and that an understanding of the
shortcomings of present methodologies and expressions is a
necessary prelude to the choice and substitution of improved

2. DIMENSIONAL ANALYSIS
Rayleigh,2 in 1915, began his denitive publication on
dimensional analysis with the following statement: I have
often been impressed by the scanty attention paid even by
original workers in physics to the great principle of similitude. It
happens not infrequently that results in the form of laws are
put forward as novelties on the basis of elaborate experiments,
which might have been predicted a priori after a few minutes
consideration. The phrase a few minutes consideration may
be a fair description for Rayleigh but constitutes an
exaggeration of the capabilities of we ordinary mortals.
By and large, chemical engineers have heeded Rayleighs
advice and applied dimensional analysis. In particular, the
solutions, correlative equations, and graphical correlations for
transport are almost always expressed in terms of dimensionless
groupings of variables with dimensions. It might appear
unnecessary to review this topic, which most chemical
engineers presume they understand, but my experience
indicates that such presumed understanding is often incomplete
and/or faulty. A brief exposition of some of the false
simplications and assumptions associated with dimensional
analysis follows.
2.1. Dimensional Analysis of a List of Variables.
Rayleigh in the aforementioned publication got dimensional
analysis, as applied to a listing of variables, exactly right. It is my
personal opinion that we would be better o if all subsequent
contributions were simply ignored (except possibly as bad
examples).
He started by postulating an expression in the form of a
power series of a term composed of the product of powers of all
of the primitive variables. He then focused his attention on only
one of these terms and determined these powers insofar as they
are constrained by the conservation of dimensions, including
time. A typical solution might be
Nu = ARe nPr m + BRe 2nPr 2m + ...

(3)

Rayleigh realized that this result does not mean that the Nusselt
number (Nu) is proportional to the product of a power of the
Reynolds number (Re) and a power of the Prandtl number
(Pr), such as

Nu = ARe nPr m

(4)

Rather, it merely indicates that Nu is some unknown arbitrary


function of Re and Pr only, namely,

Nu = {Re , Pr }
231

(5)

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

straightforward dimensional analysis, the following result for


the velocity distribution in fully developed turbulent ow in a
round tube:

Unfortunately, many chemical engineers, including academics,


fail to make this distinction and then compound the error by
plotting experimental data for Nu vs Re and/or Pr on loglog
coordinates to determine values of A, n, and m. I confess to
wasting a signicant amount of time in my younger days trying
to rationalize the prevalent value of n = 0.8 as the power of Re
in expressions with the form of eq 4 as a theoretically based
value of 4/5. I eventually realized that powers of dimensionless
groups other than plus and minus unity only occur in
asymptotes, and that values such as 0.8 are rounded-o artifacts
of the choice of some particular range of the variable (here, Re)
and have no theoretical signicance. I have yet to discover any
exceptions to that conclusion but also have yet to nd or derive
a formal proof or disproof. Accordingly, pending the discovery
of an exception or the derivation of a disproof, I propose the
total avoidance and elimination of correlating equations in the
form of empirical power functions and, of course, of their
products, from manuscripts, journals, and new or revised
textbooks.
That false concept, as represented by eq 4, is legacy of an
analysis by Nusselt3 in 1909, in the very investigation that
dened the dimensionless group named in his honor. It is a
candidate for the most pernicious concept in the history of heat
transfer. Rayleigh knew better at the time.
The methodology of Rayleigh includes one, often-ignored,
special proviso. If two variables occur only as a product, such as,
for example, wcp in the model for a heat exchanger, they are to
be treated as a single variable.
2.2. The Minimal Set of Dimensionless Variables.
Hellums and Churchill4 devised a methodology that identies
the minimal set of dimensionless variables that are required to
describe the behavior represented by a mathematical model
consisting of one or more dierential and/or algebraic
equations and the associated initial conditions (if the behavior
is time-dependent) and boundary conditions. If a similarity
transformation is possible, this methodology identies it. The
procedure is somewhat tedious but so straightforward that
White and Churchill5 wrote a now outdated computer program
for its complete execution, including the reduction of a partial
dierential to an ordinary one in the event of the identication
of a similarity transformation. The speculative elimination of
each questionable variable and/or term should always be tested
when using the method of Hellums and Churchill.
2.3. Speculative Dimensionless Analysis. Of course, the
methodology of Rayleigh cannot correct a wrong, incomplete,
or redundant listing of variables. This is a classical illustration of
garbage in, garbage out. However, the very real diculty in
choosing variables can be turned to an advantage. Churchill6
suggested that the process of dimensional analysis be
considered to be a speculation and that the process be repeated
with questionable variables added or deleted one at a time or
even two at a time, with or without a rationale. The results are
then compared with experimental data or numerically
computed values to identify the dimensionless groups and
asymptotes that provide the best representation. This
repetitious process requires a signicant amount of eort but
the reward may be the elimination of an unnecessary parameter,
and that is a huge advance. Dimensional analysis with
alternative variables of equal validity may lead to results of
dierent utility.
As an example, Prandtl,7 who pioneered throughout his
career in the use of speculative dimensional analysis and in
particular in turbulent ow, in obtained in 1926, by

y( )1/2 a( ) 1/2
1/2
w
w

,
u =

(6)

He then expressed eq 6 in a new, compact notation that has


remained the standard to this day, namely,
u+ = {y+ , a+}

(7)

He thereupon speculated that velocity distribution near the wall


might not be dependent signicantly on the radius of the tube.
Eliminating the dimensionless group that includes the variable a
produces what is now known as the universal law of the wall:

u+ = {y+ }

(8)

The validity of eq 8 as an asymptote for y 0 and as a good


approximation for much of the cross-section in a round tube
has been amply conrmed by experimental measurements. The
elimination of a+ as a variable in the region near the wall has
become a usual simplifying assumption, and one that has
retained its viability to the present day. The term universal
signies the subsequent observation that eq 8 applies to
unconned as well as conned ow and to most if not all
geometrical congurations
The derivation of eq 8 illustrates the criticality of the choice
between alternative variables in the process of speculative
dimensional analysis. Had Prandtl chosen P or even dP/dx,
which are valid alternatives to w, as the dependent variable, he
could not have derived this law by this process. My guess is that
he tested many variables and combinations thereof before
arriving at the ultimate productive choice.
With this success in hand, Prandtl logically turned his
attention to the other extremethe region near the centerlineand speculated that the velocity gradient there might be
dependent primarily on the turbulence and negligibly on the
viscosity. The consequent routine dimensional analysis for du/
dy without the viscosity as a variable led to
+

y+
du +
+
+ =
dy
a

(9)

and the formal indenite integration of eq 9 from the centerline


to an arbitrary nearby location led to
uc+ u+ = {1} {y+ /a+} = {y+ /a+}

(10)

which is known as the universal law of the center. The term


u+c u+ is known as the velocity defect. The neglect of the
viscosity as a variable near the center of a round tube is now
recognized as a usual simplifying assumption, but it should not
be overlooked that it is a valid one only in terms of du+/dy+ and
only for conned ows with symmetry.
Millikan,8 a few years later in 1938, noted that both the law
of the wall and the law of the center were reasonable
approximations for experimental data in the region intermediate to the wall and the centerline, and speculated that
there might be some region of y/a in which their predictions
were essentially equal. With great mathematical insight, he
recognized that only one expression conformed functionally to
that requirement, namely
u+ = A + B ln{y+ }
232

(11)

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

often is a repository of idealizations and simplifying


assumptions.
Although some generic guidelines exist for evaluation of the
validity and applicability of particular asymptotes, most of the
choices involve details that arise from the nuances of physical
behavior rather from mathematical or structural considerations.
The general constraints include the following:

and its counterpart


uc+ u+ = B ln{a+/y+ }

(12)

Equations 11 and 12 are known as the laws of the turbulent


core.
The importance of eq 7, in another respect, should not be
overlooked, because of the developments relative to the velocity
distribution that followed from it, namely, eqs 812. It can be
inferred from eq 7 that integration of the velocity over the
cross-section results in
um+ = (2/f )1/2 = {a+}

u+m

(1) The asymptotes must both be free of singularities.


(2) The asymptotes must both be upper bounds or must
both be lower bounds.
(3) The asymptotes must intersect once and only once.
(4) Limiting values of zero and innity are not directly
applicable as asymptotes.
Flexibility often exists in the choice among asymptotes that
meet these requirements. Correlations in the form of the CUE
are becoming commonplace, if not yet the norm, in uid
mechanics and convective heat transfer. However, the
applicability of the CUE is not limited to those two topics.
Correlations in this form have been devised for applications as
diverse as the pressure drop in ow through a packed bed,
binary vaporliquid equilibrium, the rate of enzymatic
reactions, the rate of a human running on a track, and the
approximate representation of mathematical functions such as
erfc{x}. Two asides are irresistible if not essential to the
objective herein. First, the algebraic representation of Fermats
last theorem is a special case of eq 14, and second the Danish
polymath, Piet Hein, made a career out its applications to
architecture and design (see Gardner11).
The exponent p in eq 14 was conceived to be arbitrary and,
as already noted, a protocol was devised for its evaluation.
However, the predictions of eq 14 are so insensitive to the
numerical value of p that an integer or the ratio of two integers
is ordinarily chosen for convenience. In a few instances, the
combining exponent has been found to correspond to a
theoretical solution. In most of those, such as the Ergun (or
Forchheimer) equation (see Figures 3 and 4 in ref 10) or
Ohms law, the theoretically determined combining exponent is
1 or 1. As an aside and an example of a misidentied
simplifying assumption, the two terms (asymptotes) of which
the Ergun equation is comprised are often cited as representing
laminar and turbulent ow but the latter one actually represents
inertial ow.
The one notable exception, with regard to a theoretically
derived combining exponent other than 1 or 1, is that for
assisting forced and free convection. A value of 3 was
determined from experimental data (see Figures 2 and 3 in
12
) and subsequently discovered to correspond to theoretical
solutions rst derived by Kitaura and Tanaka13 and then
independently by Ruckenstein.14 The solution of Kitaura and
Tanaka was subsequently generalized for multiple mechanisms
of assisting convection (for example, translation, rotation, and
vibration).
The utilization of eq 14 for the formulation of correlating
equations constitutes an inconspicuous revolution in chemical
engineering in that it is gradually replacing power functions,
which, as noted, are always in functional error when applied
over a range of an independent variable or parameter.
Expressions in the form of eq 14 are approximations by virtue
of interpolation and empirical by virtue of the arbitrary
combining exponent (the particular power-mean). On the
other hand, their predictions are not only insensitive to the

(13)
+

The dependence of
on a is perhaps the most important
member of the usual simplifying assumptions for turbulent ow
in a smooth round tube. Equation 13, does not preclude
relationships in the form of f = {Re} because Re = 2a+u+m, but
the latter prove to be implicit and require iterative solution
whereas those in the form of eq 13 do not.

3. CORRELATION
Correlations provide an essential resource for the design of
chemical plants and the analysis of chemical processing. Fame
and credit often accrue to those who devise useful correlations
rather than to those who obtain the experimental data upon
which they are based. From the earliest days of chemical
engineering, these correlations have taken the form of graphical
representations or of empirical algebraic equations representing
straight lines or simple curves. That practice has been utilized
for both physicalchemical properties and for the rates of the
individual processes. Although theoretical concepts have yet to
replace correlations on a broad scale, they have had an everincreasing role in chemical engineering in two respects. First, by
providing a substitute for experimental data in the form of
values obtained by means of the numerical solution of a
theoretical model, and second by providing guidance for the
construction of correlating equations. The rst role is outside
the scope of this manuscript, but several examples of the second
role follow.
3.1. The ChurchillUsagi Equation (CUE). In 1972,
Churchill and Usagi9 proposed the following canonical
equation for correlation:
y{x} = [y0 {x} p + y{x} p]1/ p

(14)

In 1974, they explored its extended usage for more than one
independent variable and/or three or more regimes.10 Equation
14 may be noted to constitute the pth power-mean of the two
limiting asymptotes, y0{x} and y{x} for vanishingly small and
unlimited large values of x, respectively, and thereby to
interpolate between them. This means of correlation was
utilized earlier by others; our contribution was the recognition
of its potential generality and the codication of its
construction, including a methodology for the choice of a
numerical value for the arbitrary exponent. The choice of p as
the symbol for the combining asymptote is a deliberate attempt
to avoid confusion with n, which is commonly used to
symbolize powers of particular variables or dimensionless
groups of variables.
The asymptotes are the critical element of the CUE. They
may either or both incorporate a theoretically based dependence on parameters and secondary variables, as well as being
based on the primary independent variable. Their selection
provides the opportunity, if not the necessity, for ingenuity, and
233

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

For the ow of an idealized uid of constant density and


constant heat capacity through a round tube, it is dened as the
integral of the radial temperature distribution weighted by the
radial velocity distribution; thus,

numerical value of that exponent but, in most instances,


remarkably accurate.

4. CONCEPTUAL CONCEPTS AND VARIABLES


The use of conceptual concepts and variables in chemical
engineering is so pervasive that it may be forgotten that they
are arbitrary and subject to limitationsthe usual simplifying
assumptions. Several examples follow.
4.1. The Heat-Transfer Coecient. The heat-transfer
coecient is perhaps the best example of a conceptual variable.
Newton,15 in 1704, noted that, in free convection from a heated
immersed body (a horizontal cylinder), the heat ux is
proportional to the surface area and to the dierence in
temperature between the body and the surrounding air. This is
perhaps the greatest single conceptual advance in the history of
heat transfer and related subjects in that it allows a reduction in
the number of quantities required to describe the process from
four to one. That one remaining compound variable is now
known as the heat-transfer coecient, and, in algebraic
notation, Newtons concept may be expressed as
h = Q /A(Ts T)

Tm

r 2
ur d
a

ur r 2
Tr d
um a

(17)

This quantity can, in principle, be determined by means of a


direct but disruptive experiment, namely, diverting the entire
stream through a stirred vessel and measuring its exiting
temperature, thus leading to the now-obsolete alternative name:
the mixing-cup temperature.
The mixed-mean concentration of the species A in a uid
stream of constant density can similarly be expressed as
CAm

ur r 2
CAr d
um a

(18)

These three mixed-mean quantities are well-known, but it is


important to note the usual simplifying assumptions in their
denition. They may also be dened and applied for variable
physical properties.
4.3. The Equivalent Thickness for Pure Conduction.
This quantity, as dened by e = h/k = jw/kT, has repeatedly
been proposed as an alternative to the heat-transfer coecient,
but, with one notable exception; it has always proven to be
inferior in terms of correlation, generalization, and insight, and,
but for that exception, might have been relegated to the ashbin
of failed concepts. That exception was its use by Langmuir16 in
1912 in the context of an analysis of the heat loss by free
convection from the lament of a partially evacuated electrical
light bulb. He utilized an equivalent thickness for thermal
conduction to derive an approximate expression for the eect of
the curvature of the cylindrical lament, vis-a-vis, a vertical at
plate, on the rate of heat transfer by free convection. The
resulting expression, which, after nearly a century, has not been
improved upon, and which is based on the applicability of the
log-mean area for conduction across a cylindrical layer, is

(15)

The attribution of the concept to Newton has been disputed by


some because eq 15 and the coecient h itself are later
expositions.
This concept, although usually encompassing some degree of
approximation and scorned by a few iconoclasts, has remained
in active use for over 300 years. The few outright failures are
generally a consequence of the invalidity of the simplifying
assumptions in a particular application rather than in general.
One such failure is illustrated in Section 10.11.
Eventually, the concept of a heat-transfer coecient was
adapted for forced convection in tubular ow by replacing the
free-stream temperature with the mixed-mean temperature of
the uid. With the passage of time, the concept of a heattransfer coecient has also been adopted and adapted for
equivalent coecients for ow and mass transferfor example,
as the friction factor, the drag coecient, the orice coecient,
and the mass-transfer coecient. These compound variables
have been utilized by chemical engineers for more than one
hundred years, and they remain invaluable and irreplaceable.
Most of the technical database in heat transfer is compiled in
terms of these coecients.
4.2. Lumped-Parameter Models. The adaptation of the
concept of a heat-transfer coecient for heat exchange between
a stream of uid and a wall in terms of the mixed-mean
temperature may have been the origin of the concept of a
lumped parameter. Although models based on lumped
parameters are scorned by some elitists, they are the workhorse
of process design and a few of them merit identication here.
The mixed-mean velocity appears throughout the literature
of chemical engineering, because of the pervasive use of tubular
ow in processing chemicals and petroleum. For constant
density, it is dened as
um

Nu =

ln 1 +

2
Nuf

(19)

Here, Nuf symbolizes the correlating equation for laminar free


convection from a vertical at plate. It should be noted that the
eective thickness is not present in the nal expression. This
relationship has also been found to be uniquely useful as an
approximation for both free and forced convection from a
horizontal cylinder as the Grashof and Reynolds number,
respectively, approach zero, and is readily adapted for the
region of the entrance in laminar tubular ow and for mass
transfer. When applied to a spherical layer, this concept invokes
the geometric-mean area and leads to the exact asymptote for a
decreasing Grashof number or Reynolds number, namely, Nu =
2.
As a consequence of eq 19, the equivalent thickness is a
candidate for the most useful of simplifying assumptions.
4.4. Fully Developed Flow. Full development is invariably
a postulate, whether noted or not, in theoretical expressions
and correlating equations for both the laminar and turbulent
regimes of ow in a round tube. A regime of development of
the ow actually exists in all instances and depends critically on
the geometrical congurations before and at the inlet. The

(16)

The mixed-mean velocity is equal to v/A, where v is the


volumetric rate of ow and A is the cross-sectional area, and the
integration is unnecessary insofar as v is known.
The mixed-mean temperature has also proven to be a very
useful quantity in chemical engineering for the same reason.
234

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

approximation at best. Fortunately, this is a conservative


approximation, in that the heat-transfer coecient is highest at
the point of onset of heating or cooling. Most analytical and
numerical solutions for thermal convection are for fully
developed convection and thereby for uniform heating or
uniform wall temperature, because of the relative simplicity of
the behavior. The former condition can be approximated in
practice by electrical-resistance heating of the tube wall or by
equal-counter-enthalpic ow, and the latter condition is
accomplished by subjecting the outer surface of the tube to a
boiling uid for heating or to a condensing uid for cooling.
Constant viscosity, as well as density and heat capacity, are
implied by the concept of fully developed thermal convection.
One classical solution for developing convection in fully
developed ow is that of Graetz18 for uniform wall temperature.
It is in the form of the sum of an innite series of terms
consisting of eigen coecients and eigen functions, but reduces
to one term for fully developed convection. As an aside,
numerical integration has now become more accurate and less
time-consuming than utilizing the Graetz solution or its
counterparts for a uniform heat ux density or fully developed
turbulent ow.
It is obviously desirable to recognize the existence of the
simplifying assumptions identied here when applying a closedform solution, a numerical solution, or a correlating equation
for convective heat transfer.
4.6. Lower Limiting Values for Convection. The lower
limiting value of 2 for the Nusselt number (Nu) for free or
forced convection from the outer surface of a sphere as Re or
Gr approach zero is also useful as a test of experimental data;
lower values are presumably in error. All other nite bodies
have a greater limiting value.
The exact values of 48/11 and 3.567 for Nu in fully
developed convection in fully developed ow in a round with a
uniform heat ux density and a uniform temperature on the
wall, respectively, serve similar roles as limiting and test values.
4.7. The Log-Mean and Mixed-Log-Mean Temperature Dierence. These two quantities are applicable with
some constraints for countercurrent and concurrent ow in
heat exchangers, as well as the correction factor for deviations
from these mean temperature dierences in other congurations of ow. Those constraints, although too specialized to
justify description here, should not be overlooked.

pressure gradient, which is of primary concern in bulk


transport, and the radial velocity distribution, which is of
primary concern in forced convection and chemical conversions, vary greatly near the inlet. The postulate of full
development is so ubiquitous, not because it is always an
acceptable approximation, but because it results in such a great
simplication in the modeling. If the density changes
signicantly with length, because of the pressure drop, or if
the density and/or viscosity change signicantly with temperature, because of heat transfer or an energetic chemical
reaction, fully developed ow may not be attained before the
uid exits from the tubing. The number of tube diameters in
which the velocity prole and pressure gradient dier
signicantly from that for full development is dicult to
generalize, other than that it is much shorter in turbulent ow
than in laminar ow.
Whether or not the development of the ow is important
should be made on a case-by-case basis. There are three
possible choices: neglect the existence of the regime of
development, estimate the eects of its neglect, or take the
development into account rigorously. The rst choice is the
common one, but the second choice is the practical one, in
most instances. The third choice, that is computation of the
development, is discussed in the subsequent section on uid
ow.
4.5. Fully Developed Convection. The concept of fully
developed convection in tubular ow is as useful and as
pervasive as that of fully developed ow, but is much more
subtle and its initial formulation required more ingenuity. Fully
developed convection diers for fully developed or developing
ow. It was belatedly generalized by Seban and Shimazaki,17
who codied its dependence on the thermal boundary
condition at the wall.
If a uniform heat ux density is imposed on the wall of a tube
through which a uid is passing, the mixed-mean temperature
of the uid must increase linearly with axial distance insofar as
the density and heat capacity can be considered to be invariant.
It follows that the temperature of the wall must thereafter also
increase linearly at the same rate, and that (T T0)/(Tm T0),
(Tw T)/(Tw Tm) and the heat-transfer coecient must
approach asymptotic values. Fully developed convection for
uniform heating is thus dened by the near-attainment of
asymptotic values for (T T0)/(Tm T0), (Tw T)/(Tw
Tm), and the heat-transfer coecient. It follows that T/x
Tw/x Tm/x dTm/dx, which allows simplication of
the dierential energy balance.
A uniform wall temperature also results in an approach to an
asymptotic value for the heat-transfer coecient and for (Tw
T)/(Tw Tm) but not for (T T0)/(Tm T0). Seban and
Shimazaki recognized that the attainment of an asymptotic
value of the rst of these quantities implied that its derivative
can be equated to zero. They thereby formulated
T T T
T
= m w

x
x Tw Tm

5. THERMODYNAMICS
The thermodynamics that is taught in chemical engineering
constitutes not only the most important element in the
curriculum but also diers greatly from that taught in chemistry,
physics, and the other branches of engineering. The dierences
with respect to that taught in chemistry and physics have arisen
because chemical and petroleum processing take place primarily
in steady ow through round tubes or stirred vessels, and are
most expediently described in Eulerian rather than Lagrangian
coordinates. The dierences, with respect to that taught in the
other branches of engineering, have arisen because of the
involvement of chemical engineers with a broader range of
materials, including gases other than air and water vapor, gases
under vacuum and high pressure, liquids other than water,
including non-Newtonian ones, and two-phase dispersions,
including bubbles in liquids, droplets in both gases and liquids,
and solids in uidized and packed beds.
As graduate students in chemical engineering, several of us
selected, collectively, as a brave cultural adventure, an advanced

(20)

The two terms on the right-hand side of eq 20 are both


independent of the radius, and therefore T/x can be replaced
by dT/dx. Substitution of this expression in the dierential
energy balance results in considerable simplication.
Developing convection occurs in almost all practical
applications but is often overlooked, and fully developed
convection is applied for the entire length of a tube, which is an
235

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

of the limitations of the state of the art, and the reason for the
concept of a tray eciency.
This example illustrates the usefulness of simplifying
assumptions in a conceptual sense, even after they become
recognized as false, but it also illustrates the desirability of
identifying and acknowledging their limitations.
An implicit simplifying assumption of the McCabe-Thiele
concept is that it is strictly applicable only for binary mixtures.
Accordingly, caution is in order in the instance of a third
component. An example is the presence of a trace of propane in
the feed to a deisobutanizer. The propane accumulates in the
uppermost trays and the maximum ratio of isobutane to normal
butane occurs several trays below.

graduate course on thermodynamics in physics taught by David


M. Dennison, a former student of Niels Bohr. One of us asked,
on behalf of all, why he consistently referred to the van der
Waals equation as the real gas equation, although it was only
a mechanistic approximation. The graduate students in physics,
who constituted the vast majority of the class, were openly
irritated at our temerity in challenging their idol, but he took
the question good naturedly and seriously, and conceded that
that van der Waals equation was not exact. He went on to say
that he was astounded that chemical engineers did not nd it
exact enough for all practical purposes. As an aside, the van der
Waals equation is favored by physicists, because many of its
predictions are correct qualitatively, and physicists are less
concerned with quantitative predictions than are chemical
engineers. The purpose of this anecdote is to demonstrate that
a usual simplifying assumption that is acceptable in physics
may not be acceptable in chemical engineering.

7. FLUID MECHANICS
Chemical engineers share an interest in uid mechanics with
aerospace, mechanical, civil, environmental, petroleum, and
nuclear engineers, and, to a limited extent, physicists. However,
in almost all instances, the interests of chemical engineers dier
from the others in terms of geometry and by virtue of the
inclusion of reactions and a broader range of uids. Hence, the
illustrative topics herein are not necessarily applicable to those
other elds.
7.1. Laminar Flow in a Round Tube. The concept of the
friction factor, which originated at least as early as 1855 and
which herein is represented by the version introduced by
Fanning,20 namely f = 2w/umx2. It is not, as it sometimes
called by students, a fudge factor, but rather the
dimensionless ratio of well-dened measurable quantities. For
laminar ow in a round tube, the exact theoretical expression
known as Poiseuilles or Hagens law, namely,

6. SEPARATIONS
The separation of uids into their molecular components has
largely been appropriated by chemical engineers, and thereby is
chosen as the subject of the rst special eld herein.
The McCabe-Thiele Method. Most chemical engineers
rst encounter the phrase the usual simplifying assumptions
as an undergraduate student in a class in separations and in the
context of the graphical method for the calculation of the
required number of equilibrium stages in the partial separation
of the components of a binary mixture of miscible liquids in a
column with discrete trays. Each of the usual simplifying
assumptions, including, but not limited to, perfect mixing of the
liquid on each tray, perfect mixing of the vapor between the
trays, a xed molar rate of ow of liquid across each tray above
the feed tray and another xed rate below it, and the attainment
of the equilibrium composition in both the gaseous and liquid
phases leaving each tray, were necessary for the truly ingenious
formulation conceived by two graduate students (see the work
of McCabe and Theile19). Because of its relative simplicity, and
because it provides great insight, this methodology is still taught
to students and utilized by practitioners, even though all of the
simplifying assumptions have been proven to be crude
approximations at best.
In order to retain the merits of simplicity and insight of the
McCabeThiele methodology while recognizing its possible
inaccuracy, a fudge factor, called the eciency, was
introduced. The number of ideal stages is divided by this
eciency to estimate the number of real trays that are
required to produce the same separation. An entire literature of
irrational expressions was gradually devised to estimate this
eciency. Soon after my graduation with a BSE, I was entrusted
with the design and the supervision of the operation of
fractionating columns of large diameter and many trays in a
petroleum renery. I was stunned when the analysis of the
samples of the liquid collected from the overow of some of the
trays revealed the tray eciency to be greater than 100%. The
primary explanation was of course the variation of the
composition of the liquid across the tray, which results in a
greater rate of mass transfer than that corresponding to the
highly idealized model.
A generalized methodology to predict the behavior in a
distillation column in terms of uid mechanics and mass
transfer still does not exist, primarily because of widely varying
hardware and dimensions. The best solution, with respect to
the classroom, appears to make sure that the students are aware

f = 16/Re

(21)

has been found to provide adequate predictions under most


circumstances.
The principal idealizations that were utilized in the derivation
of eq 21 (the usual simplifying assumptions) are the postulates
of fully developed ow, of Newtons second law of motion
(namely, that the acceleration of a mass generates a force), of
Newtons law of viscosity (namely, that a velocity gradient
creates a shear stress), of invariant density and viscosity, and of
a Reynolds number, Re = 2aum/, of <1800. Only the
idealization of fully development is examined here.
The development of the ow in the laminar regime in a
round tube requires a considerable length that is dependent, as
already mentioned, on the geometrical conguration at the
entrance, and is signicant in many applications. The prediction
of the development has now become feasible for undergraduate
students and practicing engineers and should be considered as a
possibility. Developing laminar ow in a round tube is
described by momentum and mass balances in the form of
partial dierential equations in radius and axial distance. The
numerical solution of these coupled equations can be carried
out by means of standard, well-developed nite-dierence or
nite-element methods. One approach is to write and execute
ones own computer program with the guidance of textbooks in
uid mechanics and/or computer science. The other is to
utilize a computer package such as FluentTR that produces
solutions for this behavior. Self-programming has the advantage
of allowing for special eects such as heat transfer and chemical
reactions that may not be encompassed by the computer
package.
236

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

7.2. Annular Mixing in Laminar Flow. A simple


extension of developing laminar ow in a tube illustrates the
surprises that often occur and the consequences when the
limitations of the usual simplifying assumptions are overlooked.
Seider and Churchill21 investigated, both experimentally and
computationally, the merging of jetting streams of fully
developed laminar ow from the end of a round tube and
from the annulus formed by an outer concentric tube that
continued beyond the termination of the inner one. Their
choice of this geometrical conguration was prompted by an
interest in the interdiusion and reaction of traces of dissolved
species in the two streams. They expected the two streams to
merge smoothly, and the parabolic velocity distribution in the
inner tube and the nearly parabolic velocity distribution in the
annulus to undergo a smooth and rapid transition into a single
parabolic one.
Fortuitously, they chose an unsteady-state algorithm to solve
the nite-dierence model. Their numerical solution converged
to the expected steady state at low rates of ow but became
unstable and eectively insolvable at higher ones, as a result of
the wake formed at their point of intersection at the end of the
inner tube.
An outer glass tube was chosen for the experiments in order
to permit observation of the velocity distribution and the
process of mixing by means of the introduction of dye into the
inner stream. The expected behavior was observed at very low
rates of ow; however, at higher ones, the wake created at the
end of the inner tube, whose wall of course had a nite
thickness, resulted in an instability that could be triggered by
any disturbance such as someone walking in the laboratory or a
miniscule bubble of air on the outer wall. The ow was then
observed to become strongly turbulent immediately beyond the
point of merger, and a rapid process of almost perfect of mixing
to proceed across the tube. Further downstream, the
uctuations phased out and the expected parabolic velocity
was attained.
The unstated simplifying assumptions that proved to be false
as the rate of ow was increased were the zero thickness of the
inner tube wall and the existence of a steady state. The
compensating benet was the discovery of a practical means of
mixing two miscible streams of reactants.
7.3. The Orice Coecient. The volumetric rate of ow of
a uid through piping, which is an essential measurement in
chemical and petroleum processing, is often determined by
means of a sharp-edged oricea plate with a central
opening that has its minimum diameter upstream and an
outwardly curved expansion in approximation of the downstream half of a Venturi tube. The orice and associated
hardware, as supplied by a vendor, are accompanied by a table
and/or curve that indicates the rate of ow as a function of the
reading of a pressure drop or its equivalent. If an industrial
practitioner trusts the reading and the corresponding value, the
underlying simplifying assumptions as described here may be of
no concern.
On the other hand, the theoretical basis for this means of
measurement, which perhaps evokes simplifying assumptions
than any other illustration in this manuscript, is discussed in
every textbook on uid ow. Even so, most of those discussions
in textbooks fall short.
The momentum and force balance for a jet of decreasing
diameter in and beyond the upstream (sharp) edge of the
orice ow of a uid of constant density can be written as

u1 = C1

2(P1 P2)
[1 (D2 /D1)4 ]

(22)

The simplifying assumptions made in deriving eq 22 include


constant density, fully developed ow upstream and downstream, and inertial ow (negligible viscous shear) in the region
of the orice. The error due to a variable density, such as for a
gas, may be reduced by using a mean density. A general rule of
thumb, based on practical experience, is that the orice should
be at least 50 pipe diameters downstream from and 10
upstream from any disturbance, such as the tting of a ange. If
this approximation is applied and these conditions are met, the
value of the orice coecient C1 is approximately unity.
The diameter of the jet at the downstream point of
measurement is generally unknown, so D2 is replaced by D0
and C1 by C0, and then the dependence on D0/D1 dropped
converting eq 22 to
u1 = C0

2(P1 P2)
2(P1 P2)
C0
4

[1 (D0 /D1) ]

(23)

According to Lamb,22 a theoretical value of C0, corresponding


to the minimum diameter (or vena contracta) for a ow
bounded by a free streamline for the limiting case of D0/D1
0 was rst determined by Kirchho in 1869 to be /( + 2)
0.611. Although the coecient C0 depends on Re, and, as
implied, on D0/D1 as well, the reduced form of eq 23 with a
coecient of 0.611 is found in some textbooks without further
explanation. Presumably, a coecient that accounts for all the
idealizations is used in practice.
The functionality of eqs 22 and 23 is essentially exact and, in
principle, it should be now be possible to compute values of C0
by means of numerical integration of the partial dierential
equations of conservation without any idealizations. That does
not seem to have been done, but approximate values for
asymptotically large values of Re0 that have been determined
from theoretical considerations (see ref 23) are listed in Table
1.
Table 1. Theoretically Predicted Values of the Orice
Coecienta

A0/A1

C0

0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9

0.5848
0.5966
0.6048
0.6103
0.6130
0.6124
0.6075
0.5940
0.5740
0.5258

Data taken from ref 23.

The orice coecient actually rises rapidly with Re0 in the


laminar regime, goes through a maximum, and then decreases
and approaches the appropriate value in Table 1 (see, for
example, Tuve and Sprenkle24). Churchill and Usagi10
developed a correlating equation for that overall behavior in
the form of eq 14 as follows. From experimental data, they
devised, for the laminar regime (the asymptote for Re0 0),
the following purely empirical expression:
237

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research


C0 = 0.16Re01/2

Article

The validity of the NavierStokes equations has been


questioned, particularly by physicists who prefer a molecular,
and thereby statisticalmechanical model as a starting point for
uid mechanics. As an aside, when, as a student in a graduate
course in physics taught by George Uhlenbeck, I asked him
why he never mentioned the NavierStokes equations in his
discussions of uid mechanics, he replied that he was unsure of
their validity. Later, when I had become a colleague, he halfjokingly complained that my question as a graduate student had
caused him to waste two doctoral students in vain attempts to
derive the NavierStokes equations from a molecular model.
He subsequently described that result in the following
statement within a review on uid mechanics,30 Quantitatively,
some of the predictions from these equations surely deviate
from experiment, but the very remarkable fact remains that
qualitatively the NavierStokes equations always describe
physical phenomena sensibly. The mathematical reason for
this virtue of the Navier-Stokes equations is completely
mysterious to me. My own personal conclusion is that,
although these expressions may have uncertain antecedents and
validity, we have no real choice but to utilize them.
One important characteristic of the NavierStokes equations
is that they are ordinarily expressed in Eulerian coordinates,
that is with distance as the primary independent variable,
whereas the molecular model, as well all others in the hard
sciences such as physics and chemistry, are invariably expressed
in Lagrangian coordinates, that is with time as the primary
independent variable. The use of Eulerian coordinates is a great
advantage mathematically, because it simplies the treatment of
continuous processes while allowing the treatment of batch
processes. In particular, it expedites the treatment of the
molecular diusion of momentum, energy, and molecular
components, as well as of their transport by the turbulent
uctuations. The treatment of these eects in two or three
dimensions is awkward, if not eectively impossible, in
Lagrangian coordinates. Why the physicists and chemists sohandicap themselves is a mystery to me, but their unbending
preference for a molecular model rather than a continuum may
be one explanation. As subsequently noted in connection with
reaction engineering, the concepts of spacetime and space
velocity are unfortunate holdovers from the Lagrangian
formulations of chemists. In conclusion, the NavierStokes
model appears to be not only valid but preferable, except for
special conditions, such as a vacuum or a free surface in which
the molecular structure cannot be ignored.
The time averaging of the NavierStokes equations by
Reynolds is probably the greatest contribution to the modeling
of turbulent ow in its history, in that every viable solution for
tubular ow has utilized that form. Nevertheless, its validity
continues to be disputed. Reynolds actually spaced-averaged
the NavierStokes equations, and the validity of that almostuniversal adaptation also continues to be disputed.
The key to the representation of turbulent ow by Reynolds
was his decomposition of the components of the timedependent velocity and pressure into a time-averaged value
plus a uctuation. The time averaging thereby introduces
additional unknown dependent variables, such as uv, which
are known as the Reynolds stresses. The prediction of those
latter quantities, and this illustrated one in particular, has
provoked a vast amount of literature, of which only a few of the
key contributions that are most relevant to the main theme of
this manuscript are mentioned herein.

(24)

A. D. Baer25 subsequently called to my attention a theoretical


expression of Roscoe26 for external ows that can be
interpreted as follows for laminar ow through an orice:
C0 = (Re0/12 )1/2

(25)

Given that (1/12)1/2 = 0.1629..., this constitutes a remarkable


conrmation of both the functional and the numerical validity
(within 2%) of an expression based on conjecture and
experimentation, despite the many usual simplifying assumptions, and thereby justies its inclusion here. Equation 25 and
C0 = /( + 2) are not compatible as components of eq 14, in
that the former is an upper bound and the latter a lower one.
Churchill and Usagi10 accordingly tted experimental data for
the intervening regime with the purely empirical expression
C0 = 0.611 + 0.023 ln{29 000/Re0}

(26)

Equations 24 and 25, with a combining exponent of 7/3, based


on experimental data, result in the following expression for all
Re0:
C0 =

0.16Re01/2

1 +

0.16Re01/2
0.611 + 0.023 ln{29 000 / Re0}

7/3 3/7

(27)

Engineers in industrial practice may, as stated, nd this


discussion of the usual simplifying assumptions associated
with the use of an orice beyond their interests or needs, but it
provides the grounds for a productive learning experience and
better understanding by students.
7.4. Turbulent Flow in a Round Tube. Except for
chemical reactors, most processes in tubular ow occur in the
turbulent regime. Hence, an understanding of ow in this
regime is of great importance in chemical engineering, even if
much of the relevant work has been done in other branches of
engineering. The lengthy exploration that follows constitutes a
gold mine of simplifying assumptions, many of which are
shown to be of questionable validity.
Two ancient concepts serve as the starting point for almost
all the successful theoretical advances in the representation and
prediction of fully developed turbulent ow in a round tube.
The rst is the partial dierential model for the conservation of
mass and momentum known as the NavierStokes equations.27,28 The second is the time averaging of that set of
equations by Reynolds.29 Both the original and the timeaveraged equations incorporate simplifying assumptions of
which many, perhaps even most, users may not be aware.
The NavierStokes equations are partial dierential mass
and momentum balances. Their derivation implies that the uid
is continuous and isotropic, and thereby neglects its molecular
structure. These equations incorporate Newtons law of
viscosity, = (du/dy), and his second law, F = ma. The
higher-order, nonlinear derivatives in the NavierStokes
equations are the source of instability and thereby of the
uctuations in space and time. Given that all practical
predictions of turbulent ow in a round tube start with the
time-averaged form of the NavierStokes equation, these two
concepts and two laws could be considered the usual
simplied assumptions in this instance. First, their validity is
be examined and, then, that of the time averaging.
238

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

For stationary, fully developed turbulent ow of a uid with


invariant physical properties in a smooth round tube, the timeaveraged equation for the conservation of momentum in the
radial direction can be combined with an overall force
momentum balance to replace the pressure gradient by the
shear stress on the wall and thereby obtain
y

du
w1 =
uv

a
dy

Large eddy simulation (LES), which is a concept of


Schumann33 and others, was proposed to alleviate the
restriction of DNS to small Reynolds numbers by using DNS
with a relatively large grid size for the large eddies in the central
core, and the eddy viscosity, as predicted by the model, for
the region of small turbulent eddies near the wall.
Unfortunately, that usage evokes the very empirical and
inaccurate wall function that undermines the model.
7.6. Models for the Transport of Momentum by the
Turbulent Fluctuations. The discussion that immediately
follows is limited to fully developed ow in a round tube. An
expression for uv is needed to make eq 28 predictive. The
earliest and best-known approach is the representation of that
quantity by the eddy viscosity, which Boussinesq34 proposed as
an addition to the molecular viscosity, thereby converting eqs
28 and 30 to

(28)

This forcemomentum balance can also be expressed in terms


of the local shear stress in the axial direction, as
=

du
uv
dy

(29)
6

or, in the already-mentioned wall variables of Prandtl, as


1

y+
du +
=
+ (uv)+
dy +
a+

du
dy

(31)

t du+
y+

+ = 1 +
dy +
a

(32)

= ( + t )

(30)

It is important to note that uv is negative at all locations


within the uid stream in a round tube and that for
convenience (uv)+ = uv/w is arbitrarily dened so as to
be positive at all locations.
As contrasted with physicists, the simplifying assumptions
that are inherent in eqs 2830, namely, the validity of the
NavierStokes equations and time averaging, have not
generally been questioned by engineers; however, to this day,
the experimental data for uv are not of sucient accuracy
and scope to conrm or disprove the validity of these
expressions beyond any doubt. The solutions by direct
numerical simulation of the non-time-averaged equations that
have recently been carried out for a parallel-plate channel do
agree with those of the time-averaged ones but are too limited
in scope to constitute a general proof.
Ignoring the variation of the pressure with radius is a usual
simplifying assumption of which students should be aware,
because that variation is nite, although small, in all turbulent
ows, and of practical importance in a few special applications.
7.5. Numerical Solutions of the Primitive Equations of
Conservation. When direct numerical solution (DNS) of the
partial dierential equations for the conservation of momentum
in their primitive unaveraged form was rst proposed by Orszag
and Kells31 in 1980 for turbulent ows, I was skeptical about its
practicality, because of the supposition that the grid size would
need to be less than the scale of the turbulent eddies, which
ranges down to molecule dimensions. My reasoning was faulty;
the grid spacing only must be reduced to the point at which the
shear stress due to the turbulent uctuations become negligible,
relative to the viscous stress. When the results for parallel-plate
channels were attained by Kim et al.32 and others, beginning in
1987, I leapt to another false conclusion, namely, that such
solutions would revolutionize the treatment of turbulent ow
and obviate the need for and the use of closed-form modeling.
In the intervening 25 years, virtually no further progress has
been made in that respect. The promise of DNS has never
materialized, because the computational requirements proved
to be truly excessive, except for a parallel-plate channel at
Reynolds numbers barely above the minimum for fully
developed turbulence. Computations using DNS have made
one signicant contribution to tubular ow, namely, the
conrmation of the proportionality of the turbulent shear
stress near the wall to y3, and the provision of an approximate
value for the coecient thereof.

and
1

Although the eddy viscosity concept has been widely scorned


because of its heuristic, nonmechanistic origin, its validity in the
sense of freedom from functional failures such as singularities
has recently, after more than a century, been established.
However, in the absence of a means of predicting the eddy
viscosity itself from rst principles t/ merely serves as a
replacement for (uv)+ as a variable for correlation. One such
predictive expression for t is worthy of mention. The study of
the structure of the homogeneous turbulence that occurs far
from a surface by Kolmogorov,35 Prandtl,36 Batchelor,37 and
others led to the model, namely,

t = c 2/

(33)

Here, c is a theoretically based coecient but one of uncertain


numerical value. The kinetic energy of turbulence () and the
rate of dissipation () are described for predictive purposes by
special equations of conservation (see, for example, Launder
and Spalding38). The resulting very complex and heuristic
model is not described in detail herein, because its predictions
for fully developed ow are very poor in the critical region near
the wall, and it needs to be supplemented with a highly
empirical and inaccurate wall function. However, the
model, despite its defects, is, by default, the current model of
choice for developing turbulent ows.
The primary competitor of the eddy viscosity was the mixing
length of Prandtl,39 which results in
t = S2

du du
dy dy

(34)

and
1

+
y+
du +
du +
+ 2 du
+ =
+ + (S )
+
dy
dy dy +
a

(35)

The analysis that conrmed the conceptual validity of the eddy


diusivity revealed that the mixing length is singular at the
centerline and thereby fundamentally unsound. However, as
contrasted with the eddy viscosity, and before this fundamental
239

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

aw was discovered, expressions were proposed to predict the


mixing length. Prandtl40 suggested the expression

S = ky

The rightmost form of eq 39 reveals that the eect of the


turbulent uctuations is to reduce the velocity at each radius
relative to that for laminar ow in a simple and regular way.
The formal integration for the mixed-mean velocity, and the
thereby the friction factor, can similarly be integrated by parts,
as follows:

(36)

where k is a constant, and von Karman41 subsequently


postulated, as an improvement,

S=k

2 1/2
= um+ =
f

du/dy
d2u/dy 2

(37)

The velocity distribution resulting from eq 37 is not a


signicant improvement over that resulting from eq 36 and,
except for its imaginative elegance, would have long ago been
relegated to the ashbin of history. As an aside, I once asked him
how he had arrived at eq 37. He replied that it was the simplest
possible expression in terms of derivatives of the velocity that
has the correct net dimension. In retrospect, that strikes me as a
necessary but not a sucient condition. The empirical factor k
is called the von Karman constant and its value is usually taken
to be 0.400. Predictive expressions for t/ corresponding to
eqs 36 and 37 could be derived by eliminating du+/dy+ between
eqs 30 and 32, but this has apparently never been done, because
of the known deciencies of these expressions.
7.7. The Representation of Transport of Momentum
by Turbulent Fluctuations without a Model. Churchill and
Chan42 were contemplating the formulation of a an improved
correlating equation for t/, based on the combination of
asymptotes when it occurred to them that a correlating
equation for uv/w = (uv)+, might be a better variable in
that it would eliminate outright the need for a heuristic
concept, such as the eddy viscosity or the mixing length. Their
conjecture was validated by their success in devising simpler
and more general correlating equations, (see Churchill and
Chan43,44). (The chronological anomalies are an artifact of
dierent time spans between submission and publication.)
Soon thereafter, Churchill45 proposed an even better
dimensionless variable for correlation and design, namely,
(uv)++ (uv)/. The following are a few of the
unanticipated benets and bonuses that result from its choice.
First, it has simple, unambiguous, physical signicance as the
local fraction of the total shear stress due to the turbulent
uctuations. Second, it may be inferred from that physical
interpretation that (uv)++ is zero at the wall, cannot exceed
unity, and cannot be singular. Third, it turns out to be nite at
the axis, although that is not evident from its denition. Fourth,
in terms of (uv)++ the momentum balance becomes

y+
du +
1 + [1 (uv)++ ] = +
a
dy

f=

u dR2 =

a+
4

[1 (uv)++ ] dR4

(uv)++ dR4

(40)

16
1

Re[1 (uv)++ dR4]


0

(41)

Equation 41 expresses the behavior in more direct accord with


physical intuition and is seen to reduce to the Poiseuille
Hagen law, eq 21, in the absence of turbulent uctuations.
The expression of local and mixed-mean velocity in turbulent
ow in terms of deductions from those for laminar ow might
rationalized physically in retrospect, but such a structure is not
evident in the analogous expressions in terms of the eddy
viscosity and mixing length, and this deductive role for the
turbulence does not appear to have ever before been mentioned
in the literature. The analogous integrals of the eddy viscosity
and the mixing length can be reduced similarly, but such
simplications do not appear to have ever been implemented,
perhaps because of their greater complexity.
The elimination of du+/dy+ between eqs 30 and 32 results in
t

(uv)++
1 (uv)++

(42)
++

In the light of the analysis of the behavior of (uv) , eq 42


reveals that the often-reviled eddy viscosity is well-behaved in a
round tube and has a nite value at the axis. Although such a
nite value was previously conjectured to occur, it was never
before proven. Prompted by the interpretation of (uv)++, as
the fraction of the local shear stress due to turbulence, the eddy
diusivity ratio can be interpreted as the ratio of turbulent to
viscous transport. Boussinesq was either prescient or lucky in
his choice of a representation.
The corresponding elimination of du+/dy+ between eqs 30
and 35 results in

(38)

(S +)2 =

a
[1 (uv)++ ] dR2
2 R2
a+
a+ 1
(1 R2)
(uv)++ dR2
=
2
2 R2

u+ =

The rightmost term of eq 40 reveals that the eect of the


turbulent uctuations on the mean velocity is also a deduction,
this time from Poiseuilles law, which, for purposes of
integration, with respect to R4, indicates that the values of
(uv)++ near the wall are most heavily weighted and therefore
the most important, at least in this respect. The rst term on
the right-hand side of eq 40 is that for laminar ow, and the
second one is based on the eect of the turbulent uctuations.
The friction factor is increased because of the decrease in the
velocity for a xed value of a+. By virtue of the aforementioned
identity, Re = 2a+u+m = a+(8/f)1/2, eq 40 can be rewritten in
terms of f and Re as follows:

which can be integrated to obtain values of u+ as a function of


y+ and a+, and then u+ can be integrated over the cross section
to obtain values of u+m, as a function of a+, with only a
correlation for (uv)++ as a function of y+ and a+ as an input.
The formal integration of eq 38 for u+ can be expressed and
carried out, in parts, as follows:
+

a+
a+

4
4

(uv)++
++ 2
[1 (y/a)][1 (uv) ]

(43)

which reveals, perhaps denitively for the rst time, that the
mixing length, although otherwise well-behaved, is unbounded
at the centerline, which is a serious if not fatal defect.

(39)
240

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

Elimination of (uv)++ between eqs 32 and 35 results in


(S +)2 =

ones prove to be convenient and, because of the interrelationships, informative. A set of such correlating equations
devised by Churchill and Chan43,44 for (uv)+ (here, reexpressed in terms of (uv)++), u+, and u+m follows:

t / + (/t )2
1 (y / a)

(44)

0.7 y+ /10 3 8/7


(
)
[(uv)++ ]8/7 =
1 (y+ /a+)

This relationship between the two heuristic models for


transport of momentum by the turbulent uctuations has
apparently never appeared in print, perhaps because it has no
evident utility, except as a curiosity.
7.8. Correlating Equations for Fully Developed
Turbulent Flow. Progress in science and engineering does
not always follow a logical or straightforward path. Although
the following historical aside does not relate directly to what
follows, it is an invaluable illustration of the origin and
persistence of erroneous concepts and simplifying assumptions.
7.8.1. Power Laws. In 1903, Saph and Schoder,46 two
graduate students in civil engineering at Cornell University,
measured the rate of ow of water through vertical pipes of
various diameters and for various hydraulic heads and
correlated their data, in terms of
f = A + BRe n

2.5
4y+
2.5
+ exp + + 1 + +
a
a
y

+ 3

(u )

0.0791
Re1/4

(y+ )2

=
4
+
+

1 + y exp 1.75(y /10)

(49)

and

(45)

2 1/2
um+ =
f
= 1.989

47.6 2
a+
161.2

+ + + 2.5 ln
+
+
a
a
1 + 0.301eC

(50)

At rst glance, eqs 48, 49, and 50 are forbidding in detail, and
understandably provoke two questions: Is all that complexity
necessary? and Why are they expressed in terms of u+m and a+,
rather than f and Re ?. The answer to the rst question is that
their complexity is the price of the far-greater accuracy and
scope than for any previous ones in both a functional sense and
a numerical sense. (See, for example, Churchill et al.52) The
answer to the second question is that they can readily be
expressed in terms of f and Re by virtue of the identities (2/f)1/2
= u+m and Re = 2a+u+m, but they then appear to be even more
complex. Equation 50 must be solved iteratively for a specied
value of Re, but convergence is very rapid.
The derivation of eqs 48, 49, and 50 may be found in
Churchill and Chan43,44 or ref 53; the focus here is on the
idealizations and approximations. The quantity u+ is directly
related to (uv)++ through eq 39, and the quantity u+m to both
u+ and (uv)++ through eq 40. Therefore, the idealizations and
approximations incorporated in eqs 48, 49, and 50 are most
easily identied and evaluated by considering these three
equations as a whole. Equations 48, 49, and 50 each consist of
three basic terms (asymptotes), and eqs 48 and 49 incorporate
a combining exponent other than unity. The closed-form
integrations or dierentiations of the entire equations are
thereby either not feasible or lead to awkward expressions
whereas the asymptotes are readily integrated or dierentiated
as required.
7.8.1.1. Simplications Introduced in the Construction of
eq 49 for u+ = { y+, a+}. The construction of the correlating
equation for u+ is examined rst, because that process best
illustrates the origin of the simplications and idealizations that
pervade all three.
The Asymptote for y+ 0: The generally accepted asymptote
for u+ for y+ 0, namely,

(46)

48

Prandtl in 1926 made the understandable error of presuming


that the power of 1/4 had theoretical signicance, and utilized
dimensional considerations with great ingenuity to derive the
corresponding velocity distribution, namely,

y 1/7
u
=
a
uc

(48)

3
2

+ 3

9.025y+ 15 y+
10 y

+
+ 2.5 ln1 +

1 + 0.301eC+
4 a+
3 a+

They determined and tabulated slightly dierent values of A, B,


and n for dierent ranges of Re. In the literature of civil
engineering, they are credited with demonstrating that the
friction factor depends only on the Reynolds number (Re).
Although eq 45 might appear to have the form of eq 14, it does
not qualify for that format, because neither A nor BRen are valid
asymptotes.
In 1913, Blasius,47 who was a student of Prandtl, plotted
experimental data for the friction factor in a round pipe versus
Re in logarithmic coordinates and discovered that they,
including those of Saph and Schoder and those for several
dierent uids, could be represented reasonably well by a
straight line, corresponding to

f=

8/7

(47)

He did not recognize the functional failures of eq 47 at the


wall and centerline, or at least did not comment on them, but
on the basis of the poor representations of experimental data by
eqs 46 and 47 for large values of Re he correctly concluded that
the slope of 1/4 was an artifact of the low range of values of
Re in the experimental data available to Blasius. Accordingly,
another collaborator of his, Nikuradse,49 attempted to
resuscitate eq 47 by utilizing a variable power dependent on
Re, rather than 1/4, but the defect in the velocity distribution
persists for any power. Prandtl apparently overlooked the
generality of this failure, but he and Nikuradse soon abandoned
this approach in favor of analyses based on eq 7. Even so,
remnants of this awed concept still persist in some textbooks
and the recent literature (see Barenblatt50 and Cipra51).
7.8.1. Comprehensive Correlating Equations. It is evident
from eqs 39 and 40 that, in principle, a correlation for (uv)++
eliminates the need for ones for u+ and u+m. However, separate

u+ = y+ (y+ )4 + ...
241

(51)

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

is based on asymptotic expansions of the partial dierential


equations of conservation of momentum by Murphree54and
others for that limiting condition. Although experimental data
are limited, and direct numerical solutions are nonexistent for
round tubes, this asymptotic dependence is supported by both
experimental data and direct numerical solutions for parallelplate channels. As discussed subsequently, the equality of some
such results for round tubes and parallel-plate channels appears
to have strong support. Both the experimental data and the
numerically computed values of (uv)++ support the choice of
1.75 104 as a close approximation for the leading coecient
of eq 51, resulting in
u+ = y+ 0.000175(y+ )4

uc+ u+

y+
a+
++
=
[1 (uv) ]c 1 +
2
a

Here,
and [1 uv]c are the limiting values at the
centerline. By virtue of eq 56, the limiting value of 1 (uv)++
is 15/a+ and, thus, depends on a+. Such a dependence is
conrmed qualitatively by the aforementioned graphical
representation but not quantitatively, because of insucient
precision of the plotted values.
Path 3: Churchill and Chan43,44 subsequently discovered an
expression of Finley et al.56 in 1961 for fully developed
turbulent ow of a liquid down an inclined plane, namely,

(52)

u+ = A + B ln{y+ }
y + 2

y+
2y+
+ B + 1 + + 2C 3 +
a
a
a

(53)

(54)

An Asymptote for u+ for y+ a+: The asymptote in eq 49 for y+


a+ has many dierent origins but all roads lead to Rome.
Rather than an orderly and logical after-the-fact description of
its formulation, some of the paths, if not the details, are
described as an illustration of a typically untidy one.
Path 1: The rst step in the derivation was the search by
Churchill and Chan43,44 of a specic function for the right-hand
side of eq 14, the law of the center, that conformed to the
following three restrains at the centerline: a value equal to zero,
a derivative with respect to y+ equal to zero, and a nite value of
the eddy viscosity. The following appears to be the simplest:

uc+

y+
15
u =
1 +
2
a

u =

a+ 1

2 1+

(55)

y
a

tc

2
y+
a +
1 +
2 tc
a

(60)

Here, tc/ is the value at the centerline corresponding to


(uv)++
c .
Path 5: Taylor58 earlier still in 1921 predicted, on the basis of
the theory of homogeneous turbulence, that, near the
centerline, t should be proportional to a(w)1/2, and thereby
derived

A value of 15/2 for the arbitrary coecient E was found not


only to provide a good t for the experimental data but to
correspond to that of Reichardt55 in 1951 for a dierent
expression, thereby resulting in
uc+

(59)

and that the corresponding expression for u reduces to eq


56 as y+ a+. (In carrying out that limiting process, it is
essential to note that ln{a+/y+} 1 y+/a.) The latter two
terms of eq 59 can be considered to represent the wake near
the centerline, that is the excess over that predicted by eq 11 for
the turbulent core. The deviation due to the wake at the
centerline is, according to eq 49, (15/4) (10/3) = 5/12 =
0.417 and its maximum occurs at y+ = 0.75 a+, and is 45/64 =
0.703.
Path 4: Churchill and Chan53 also discovered subsequently
that, even earlier, Hinze,57 in 1959, had postulated correctly,
although contrary to Prandtl, Nikuradse, and other early
experimental investigators, that the eddy viscosity approaches a
nite value tc at the centerline. On that basis, he integrated eq
32 to obtain the following asymptotic expression for the region
near the centerline:
u+c

y+
+
+
uc u = E1 +
a

+ 2
+ 3
15 y
10 y
+

4 a
3a

u+ = 5.5 + 2.5 ln{y+ } +

+ 2

(y )
4
1 + y exp 1.75(y+ /10)

(58)

Equation 58 may be noted to reduce to eq 11 as y+ 0, and


Churchill and Chan found that the choice of C = 1/12, along
with the traditional choice A = 5.5 and B = 2.5, results in

However, eq 53 proves to fail to intersect the subsequently


derived asymptote for the central region, and Churchill and
Chan43,44 devised the following second-order approximation
that does:
u+ =

(57)

(uv)++
c

However, eq 52 is not directly applicable as a component of eq


14, because of the singularity at y+ = 0.000175(y+)4, namely at
y+ = 1/(0.000175)1/3 = 17.88. An obvious corrective
approximation that is free of that singularity is

y + 3
u+ = y+ exp 0.175

10

y+
a+
[1 (uv)++
]
1

c
a+
2

(56)

uc+

Path 2: Churchill45 noted from a plot by Churchill and


Chan43,44 of both experimental data and numerically computed
values that 1 (uv)++ approached a constant value as y+ a+,
thereby suggesting the following approximate integration of eq
38:

y+
1
u =
1 +
4
a

(61)

where tc/2a(w) is a constant. Equation 61 is basically


the same as eq 55. Thirty-three years later in 1954, Taylor59
determined a value of 0.04 for the coecient from
experimental measurements of turbulent diusion.
1/2

242

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

Path 6: Groenhof60 in 1969 surveyed the many experimental


determinations of the coecient and its analogues for heat
and mass transfer and found them to range from 0.031 to 0.04.
Even though his own experimental values were close to 0.04, he
concluded that a value of 0.032 is more representative. A
dependence of on Re (or a+) has been speculated by some,
but denite evidence is not available. The value of
corresponding to the factor 15/2 for eq 56 is 1/30 = 0.0333...
The Intermediate Asymptote for u+: The intermediate
asymptote for u+ in eq 49 is eq 11, with the aforementioned
values of A and B, namely,
u+ = 5.5 + 2.5 ln{y+ }

Likewise, dierentiation of eq 52 and substitution into eq 38


results in the following corresponding asymptote for (uv)++ as
y+ 0
(uv)++ =

(62)

This asymptote is singular at ln{y } = 5.5/2.5 or y+ =


exp(5.5/2.5) = 0.1108. This singularity can be avoided by
rewriting eq 62 as
(63)

and adding unity to the argument to obtain


u+ = 2.5 ln(1 + 9.025y+ )

(64)

The predictions of eq 64 are nite and positive for all values of


y+ and provide a close approximation for all values greater than
unity.
Equation 59, which can be interpreted as a power-mean of
unity for 5.5 + 2.5 ln{y+} and (15/4)(y+/a+)2 (10/3)(y+/
a+)3, may be modied correspondingly to become
+ 2
+ 3
15 y
10 y
u = 2.5 ln(1 + 9.025y ) +


4 a+
3a
+

um+ = A

3B
+ B ln{a+}
2

(68)

With the usual values of A = 5.5 and B = 2.5, eq 68 becomes

(65)

um+ = 1.75 + 2.5 ln{a+}

The Power-Mean of u+ for All Values of y+: The combination of


eqs 54 and 65, in the form of eq 14 with an exponent of 3,
nally results in eq 49. That combining exponent was chosen
based on many sets of experimental data for round tubes and
for parallel plates with b+ simply substituted for a+. As shown in
Figure 1 of Churchill and Chan,53 the data are well represented
by eq 49.
Looking back at eq 49, the velocity distribution is dominated
by the logarithmic term for the central region, but the term for
the region near the wall is critical because it strongly inuences
the radial transport of momentum, energy, and chemical
species, and the term for the wake is critical because it is the
region with the highest rate of ow. The variable e+C, which
represents the eect of roughness, is discussed subsequently as
a separate topic.
7.8.1.2. Simplications Introduced in the Construction of
eq 48 for (uv)++ = { y+, a+}. With eq 49 in hand, most of the
simplications, idealizations, and asymptotes have already been
identied, and the description of the construction of eq 48 for
(uv)++ is relatively simple. Dierentiation of eq 65, which
constitutes a combination of asymptotes whose simple addition
represents u+ in the central core and the wake, substitution of
that derivative in eq 38, and its rearrangement in terms of
(uv) ++ , followed by division by 1 (y = /a + ), and
approximation of 1 (2.5/y+) by exp{2.5/y+} results in the
following combined asymptote for the central core and the
wake:
2.5
4y+
2.5
(uv)++ = exp + + 1 + +
a
a
y

(67)

Equations 66 and 67 intersect, and their combination with an


exponent of 8/7 is based on the limited and scattered
experimental data of Wei and Willmarth61 and Eckelmann62
and the numerical values obtained by Rutledge and Sleicher63
and others by DNS, all for parallel-plate channels, result in eq
48. The absolute value is specied to avoid negative predictions
for y+ < 2.5/ln{a+/2.5}, which actually only occur for y+ 1.
The almost exact predictions are demonstrated in Figures 1 and
2 of Churchill and Chen.53 Equation 48 appears to be
somewhat simpler than eq 49, despite essentially the same
idealizations primarily, because lesser measures were required
to avoid singularities.
7.8.1.3. Simplications Introduced in the Construction of
eq 50 for u+m = (2/f)1/2 = { a+}. Equation 50 has many
functional predecessors, including eq 11 of Millikan,8 eq 45 of
Saph and Schoder,46 and eq 46 of Blasius,47 and many
predecessors for the arbitrary coecients and exponents that
they incorporate. Von Karman, Prandtl, and Nikuradse, on the
basis of the speculation that eq 11 provides an acceptable
representation for the entire cross-section, each integrated it to
obtain

u+ = 2.5 ln{9.025y+ }

0.7(y+ /10)3
1 (y+ /a+)

(69)

Although the decrease in the velocity relative to eq 11 in the


boundary layer and its increase in the wake counterbalance each
other to some extent, the real net eect is a reduction in u+m at
low Reynolds numbers and an increase at higher ones, as
predicted by eq 50. Representation of the eect of the viscous
boundary layer, in terms of reciprocals of a+, was conceived by
Churchill64 as a means of avoiding the singularities that would
have resulted from the direct use of eq 51. The values of the
coecientsnamely, 1.989, 161.2, and 47.6were chosen by
Churchill and Chan53 to t on the mean of the two sets of
values of u+m obtained by numerical integration, namely, those of
u+, using the predictions of eq 49, and those of (uv)++, using
the predictions of eq 48. The result is
2 1/2
47.6 2
161.2
+

+ 2.5 ln{a }
um+ = = 1.989
+
a+
a+
f
(70)

A correction for the eect of the viscous boundary layer is


missing from most, if not all, expressions in the current
literature for the friction factor, although its magnitude and its
variation with a+, and thereby with Re, are of signicance in
both a practical sense and an intrinsic sense. Instead, eq 11,
with arbitrary coecients determined from experimental data
and thereby accounting for the boundary layer as well as the
wake on the mean, has remained the expression of choice in
most textbooks and handbooks in chemical engineering. From
theoretical considerations, the coecient B in expressions for
u+ and u+m must have the same numerical value, but dierent

(66)
243

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

distribution and pressure drop in piping in which uniformly


sized and closely spaced particles had been glued on the
surface. He proposed the use of expressions with the form of
eqs 73 and 74 simply as speculative modications of eqs 11 and
62 rather than deriving them, and found that, with empirical
coecients, they were eective in representing his measurements for asymptotically large Reynolds numbers, although not
for lesser ones (see Figure 7 in ref 66). Because these
measurements for uniform roughness have apparently never
been duplicated, these correlating equations remain in the
literature 80 years later as the standard by default.
In 1938, Colebrook67 dened the ef fective (or natural)
roughness of real commercial piping (eC) as that which resulted
in the same friction factor (and mean velocity) at asymptotically large Reynolds numbers as that for uniform roughness.
For example, insofar as eq 74 is valid,

ones are often found in the counterparts of eqs 62 and 69,


because of their separate determinations.
7.8.2. Idealizations Introduced in Order To Represent the
Eects of Roughness. All piping has some imperfections in its
inner surface that are inherent in the material of construction
(such as steel, glass, ceramic, or concrete), the process of
manufacture (such as extrusion or casting), and the method of
joining (such as welding, threading, or anging). These
imperfections all disturb the ow to some extent. The
imperfections often become worse with time, because of
physical erosion, variations in physical stress, depositions, and
chemical corrosion. The concept of a smooth pipe is an
idealization that is never truly attained, although it may be
approached in a practical sense. In practice, the imperfections
are lumped together as roughness. Roughness has a small,
generally negligible impact in the laminar regime of ow in a
round tube but a signicant impact on the velocity distribution
and on the pressure gradient in the turbulent regime. It is a
prime example of a parameter that cannot be ignored but is
dicult to dene in a mathematical sense. The result is that the
usual simplifying assumptions associated with roughness are
based more on blind faith than on hard evidence.
The representation of roughness by a variable e with the
dimension of length, based on the same path used to derive eq
6, leads to
u+ =

{y , ay , ae }

C (2/f )1/2
eC
3
= exp

a
B
2

The substitution of a measured value of f for commercial piping


in eq 75 thus allows an evaluation of the eective roughness.
This ingenious concept, combined with the physical and
experimental concept of Nikuradse, represents a denitive step
toward reality in the choice and interpretation of variables for
the correlation and prediction of turbulent ow. In that context,
this concept is second only to time-averaging by Reynolds as an
intellectual advance.
Colebrook made an associated great stride in correlation and
prediction by proposing an expression for interpolation
between the friction factor for smooth piping at asymptotically
low Reynolds numbers and that for naturally rough pipe at
asymptotically large Reynolds numbers. In modern notation, it
can be expressed as

(71)

The speculation that the velocity is only weakly dependent on


the radius and primarily dependent on the roughness leads, in
turn, to the following law of the wall for rough piping:
y+
u+ =
e

(72)

Equation 10 remains valid as the law of the centerline for


rough as well as smooth piping, and the application of
Millikans speculation of a region of overlap leads to the
following law of the central core for rough piping:
y+
u+ = C + B ln +
e

2 1/2
e
1

= 2.5 ln c +
7.39a
2.25a+
f

2

f

= um+ = C

(73)

3B
a
+ B ln
2
e

{}

2 1/2
a+
= 1.75 + 2.5 ln

+
f
1 + 0.301eC

a+
= 1.75 + 2.5 ln
+
1 + 0.301(eC /a)a

(74)

It may be inferred that the eect of roughness is to decrease


both u+ and u+m to the extent of B ln{e+}. The usual choice for
the value of C is 8.5, and with the usual value of 2.5 for B, eq 74
becomes
2 1/2
a
= um+ = 4.75 + 2.5 ln
e
f

(77)

He did not provide any rationale for eq 77: he simply assertedg


that the two terms in the argument of the logarithm must be
additive.
Churchill68 interpreted eq 77 as an interpolation in the form
of eq 14 with a combining exponent n = 1 between early
versions of eqs 65 and 75. The latter expressions lead to

The coecient B in eq 73 is implied to be the same as that in


eq 11 for smooth piping, but C diers from A. Integration of u+
from eq 73 over the entire cross-section results in the following
approximate expression for the friction factor for rough piping:
1/2

(76)

(78)

Equation 78 may be updated as follows, based on eq 70:


2 1/2
47.6 2
161.2
+ +
= 1.989
+
a
a
f

a+
+ 2.5 ln
+
1 + 0.301(eC /a)a

(75)

65

Nikuradse in 1933, although, and perhaps because, he


recognized that the roughness of commercial piping was highly
variable and perhaps chaotic in amplitude and spacing, chose
uniform roughness as a parameter for correlation, and made
supporting experimental measurements of the velocity

(79)

In eq 77, following from eq 72, it may be inferred that, for


rough piping, eq 65 becomes
244

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

15 y+ 2
9.025y+
u+ = 2.5 ln1 +
+

1 + 0.301(eC /a)a+
4 a+

+ 3
10 y


3a

and
2 1/2
50 2
227
= = 3.30 + + +
a
a
f

+
1
a+
ln
+
+ {a }
0.436 1 + 0.301(eC /a)a

um+
(80)

Equation 80 poses the choice of a value for e+C or eC /a. In 1944,


Moody69 prepared a simplistic graph of values of eC/2a versus
pipe diameter with the type of piping (for example, steel, glass,
or tile) as a parameter. Although this plot or its equivalent has
been updated from time to time, considerable uncertainty exists
as to the reliability of those curves (or their alternative in the
form of a tabulation).
7.8.3. The Experimental Work of Zagarola. The sense that
a topic in science or engineering is nally well understood and
well in hand is invariably short-lived. A breakthrough or
contradiction soon emerges. The achievement represented by
eqs 4850 might have been expected to last long enough to
enter the textbooks and computer packages but a new set of
experimental measurements that diered signicantly from all
those upon which eqs 4850 are based soon emerged.
Zagarola,70 in his doctoral research on turbulent ow in the
Princeton superpipe, so-called because of its large diameter
(129 mm), obtained a coherent set of data that dier
signicantly from the generally accepted one of the past in
two respects: a smoother surface (obtained by super polishing)
and a higher range of Reynolds numbers. This work represents
a signicant contribution and, but for one omission, the failure
to measure the turbulent shear stress, namely, uv, and two
deciencies (the failure of the measurements to extend down to
a Reynolds number corresponding to the onset of fully
developed turbulence (a+ = 145 or Re = 4000), and the failure
of the measurements of the local time-mean velocity to extend
into the viscous sublayer (y+ < 10)), they might have become
the new standard. One important intrinsic contribution of this
investigation, which was not claimed by the author but was
cited by Churchill et al.,52 is the evidence that the miniscule
roughness of the tubing is a signicant factor, even at the
highest Reynolds number.
The measurements of Zagarola conform functionally to eqs
4850 but require tweaking of the coecients, as indicated by
the following expressions for (uv)++, u+, and u+m.

Equations 82 and 83 represent the data of Zagarola well within


their small scatter. The most striking dierence, with respect to
eqs 4850, is the value of 0.436 = 1/2.294 for the von Karman
constant, k, compared to 0.400 = 1/2.5. The value of 0.436 was
determined by Zagarola from his measurements for both the
velocity distribution and the pressure drop. The measured
values ranged from 0.41 to 0.45 and 0.436 and varied somewhat
with Reynolds number, and 0.436 was chosen as a mean, but a
creditable one. If it is correct, how could values of 0.400 or
thereabout prevail for some 94 years? A careful look back
reveals that the value of 0.400 or thereabout had its origin with
the data of Nikuradse,49 and with the very portion of the values
that are suspect of being penciled (see Miller71) in order to
conform to eq 36. Comparison of the tabulated values of the
velocity distribution in Nikuradse72 with those in Nikuradse49
reveals the adjustment displayed by Miller, and is a good
assignment for students in both utilizing the resources of their
library and in taking advantage of the treasures there in the
form of originals.
The coecients 14.88 and 3.30 in eqs 82 and 83,
respectively, are larger than their counterparts in eqs 49 and
50 and thereby compensate in part for 0.436, vis-a-vis 0.400.
Although renements of the experimental measurements,
numerical solutions, and theoretical structure for turbulent
ow are to be anticipated, eqs 8183 appear to represent
current state of the art. Why are expressions for f and u+ that
are known beyond question to be in serious error both
functionally and numerically not replaced by more recent and
soundly based ones such as eqs 82 and 83? The primary reason
is simply inertia. Another reason in the instance of the friction
factor is that the erroneous predictions are not viewed as being
important by many engineers in industrial practice. When
working in a petroleum renery, it came to my attention that
most of the piping was grossly oversized. The explanations I
received were these: (1) the chances of failure of piping due to
pressure, impact, corrosion, or bending increases as the
diameter is decreased; (2) an increase in the quantity being
processed invariably occurs over time, and one bottleneck is
thereby avoided; and (3) the marginal cost of the larger pipe
diameter is counterbalanced by the rst two factors. Piping
under vacuum is an exception to this rule of thumb. While
these reasons may provide an explanation for the use of
obsolete and inaccurate expressions for the friction factor in
practice, they do not excuse their retention in textbooks, the
classroom, and computer packages. The considerably greater
complexity of expressions such as eqs 8183 is not a serious
impediment when all calculations in process design are
performed on a computer.

+ 38/7
y
[(uv)++ ]8/7 = 0.7
10

6.95y+
1
1
+ exp

+
+ 1 +
0.436y 0.436a
a+

8/7

(81)

(u+)3

(y+ )2
=

+
1 + y exp 1.75

y+
10

( )

14.48y+
1

+
ln1 +
1 + 0.301(eC /a)a+
0.436

y + 2
y + 3

+ 6.824 + 5.314 +
a
a

(83)

8. HEAT TRANSFER
Chemical engineers were once leaders in research and
innovation in heat transfer. In recent decades, they have
ceded that role to mechanical engineers; nevertheless, heat
transfer remains pervasive in the processing of chemicals and
petroleum. The partial dierential equation for the conservation

(82)
245

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

of energy in fully developed ow without complicating eects is


linear in temperature and therefore easier to model, on the
whole, than uid ow. However, buoyancy, energetic and
nonequimolar reactions, and the dependence of physical
properties on temperature, pressure, and composition
introduce nonlinearities and thereby restrict the validity and
range of validity of simplifying assumptions and concepts.
8.1. The Failure of a Classical Solution for Thermal
Conduction. Thermal conduction is the most elementary
mode of heat transfer and might be presumed to be a closed
subject; however, two of the usual simplifying assumptions
made in its application to transient processes have recently
been shown to be totally wrong. Fouriers law for the rate of
thermal conduction, as it is normally expressed, namely,
j = k

dT
dx

The simple and correct explanation for the false predictions


of eq 86 is that the equations for the conservation of mass and
momentum must be solved simultaneously with that for energy
for a compressible medium, and all real materials including
liquids and solids are compressible to some degree. That
compressibility prevents the impossible behavior predicted by
eq 86. Numerical solutions of the equations of conservation by
Brown and Churchill75,76 avoid the anomaly and agree with
their experimental measurements of the very weak but nite,
decaying wave in pressure.
Raising the temperature of a solid surface impulsively is
revealed to raise the temperature of the adjacent gas. Its density
decreases and its pressure rises. That sudden increase in
pressure generates a slightly supersonic wave with a positive
amplitude in pressure and temperature. The temperature does
not rise ahead of the wavefront. A compressive wave traveling
at sonic velocity is also generated in a liquid or solid by a pulse
in temperature (as, for example, by an explosive) but the
amplitude is far less, because of their lesser change in density
with temperature.
The identication of the false assumption of incompressibility and the false concept of a second derivative in time has
several important consequences. First, Fouriers law is validated.
Second, eq 85 is revealed to be an approximation that, although
suciently accurate for most applications, fails in others. One
instance in which it is not valid is a familiar one, although not
generally recognized as such. A pressure wave, called thunder, is
generated by the rapid heating of air by a bolt of lighting. The
cylindrical region of the bolt serves the same role as an
impulsively heated at surface. The pressure in the adjacent
heated air rises almost instantaneously, in accordance with its
equation of state. The numerical solutions carried out by Brown
and Churchill illustrate how that phenomenon may be
predicted quantitatively. This tiny aspect of thermal conduction
may rarely be of importance in process design or operation, but
we were forced to take it into account in connection with
experiments in the Apollo spacecraft and found it applicable to
the laser pulses of the Space Defense Initiative. Even so,
textbooks and the eld of heat transfer should at least point out
this limitation in the applicability of eq 85.
Rayleigh77 understood the physics of the process described in
this illustration a century earlier, but he did not have access to
the precise instrumentation required for measurement of the
speed and amplitude of the pressure wave nor to the
computational hardware and software to carry out numerical
solutions. That led him to make some faulty assumptions in
order to simplify the description suciently to permit the
derivation of a solution in the form of an innite series. His
simplifying assumptions included the imposition and maintenance of an immaterial plane at high temperature, the
linearization of the equations of conservation, and, most
seriously, the postulate that the pressure wave had a sinusoidal
amplitude in pressure and temperature.
8.2. The Boussinesq Transformation. The study of
natural convection seems to have generated many simplifying
assumptions, a few of which are identied herein. The most
famous set of them is known as the Boussinesq transformation,
although that attribution has been questioned. They are almost
universally postulated to simplify the equations of conservation.
They include negligible viscous dissipation, and, most
importantly, invariant physical properties except for the density
in the buoyant (gravitational term), which is assumed to be
proportional to temperature and independent of pressure. The

(84)
73

does not appear explicitly in his 1822 book, but may readily
be deduced from it. Equation 84 has stood the test of time, but
some of its applications have not, and the usual simplifying
assumptions are at fault.
The application of eq 84 to unsteady-state conduction in one
dimension may be expressed as
c

T
T
=
k
t
x x

(85)

The well-known solution of eq 85 for a semi-innite region


with an invariant thermal conductivity, density, and specic
heat capacity, an initial temperature of T0, and a xed
temperature of TS imposed at x = 0 at t = 0 and maintained
thereafter is usually expressed as
x
T T0
= erfc
2
TS T0

kt

(86)

Equation 86 has long been recognized to be in error because it


predicts a nite change in temperature at all distances at all
times. Therefore, the speed of advance of a particular
temperature front must exceed the speed of light, contradicting
the theory of relativity. The most common and widely accepted
correction for this untenable result is the addition of a second
partial derivative in time, producing for invariant physical
properties,
k 2T
T
2T
+

c
=
k
t
uT2 t 2
x 2

(87)

Equation 87 is known as the hyperbolic equation of


conduction, with uT the velocity of a thermal wave usually
taken to be the acoustic velocity. An extensive literature exists
concerning priority for its derivation and the solutions thereof.
That model eliminates the innite rate of propagation but is
totally without validity. First, it violates the second law of
thermodynamics. Second, an acoustic wave implies a variation
of density with pressure. And third, a xed value of uT falsely
implies no period of induction due to inertia, no decay due to
viscosity, and no variation upon impact with a solid boundary.
The addition of the second derivative in time is generally
rationalized, presumably to imply his imprimatur, on the basis
of the rederivation of eqs 84 and 85 by Maxwell74 by means of
kinetic theory. However, a careful reading of that famous
publication reveals the opposite, namely, a proof by Maxwell
that there is no second-order derivative in time.
246

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

result is the replacement of g (p/x)/ by g(T T),


where
=

T P

the heat losses from water-heating solar collectors. Unfortunately, the vector potential does not provide a visualization of
three-dimensional (3D) uid motion equivalent to that of the
stream function in two-dimensional (2D) motion. As an
alternative, they conceived and constructed streaklines (uidparticle paths) from interpolated values of the discrete
components of the velocity as a substitute. They rst displayed
these streaklines as projected traces then isometrically and,
nally, dynamically.
The resulting visualizations of the pattern of circulation,
which they conrmed experimentally, proved to be very
eective in explaining the experimentally observed, radical
variations in the total heat ux with two angles of inclination,
dierent aspect ratios, nonuniform heating. and partial baes,
and also produced some conceptual surprises. For example, all
of the 3D streaklines for natural convection inside an enclosure
proved to be concentric, double, closed helices or degenerate
forms thereof, an observation that may have implications even
outside of uid mechanics. It may be noted that the particle
paths dened segregated roll cells similar to those observed by
Samuels and led to the development of a generalized
correlating equation for the overall rate of heat transfer.
The conception of streaklines and visual displays, which was
motivated by the desire to display the innovative 3D numerical
computations, was undoubtedly a greater contribution than the
computed thermal and uid dynamical results in themselves.
The critical element was the experimental visualization of the
streaklines and their comparison with the computed ones,
because their complexity initially was widely rejected as
improbable. This is a little-known simplifying assumption that
introduces no error.
8.6. Free Convection around a Heated or Cooled
Cylinder. At my behest, Dudley A. Saville attempted to carry
out a nite-dierence solution for laminar free convection from
an unbounded horizontal cylinder of innite length in
unbounded surroundings. He soon discovered that I had
chosen an ill-posed problem; the buoyant ow generated by the
heated cylinder nominally sets the whole universe in motion.
The assumption of the existence of a steady state solution for
thermal conduction from a cylinder of unbounded length to
unbounded surroundings is also false. A steady-state solution
does exist for thermal conduction from a cylinder of nite
length, but free convection results in a pattern of ow
associated with the ends of the cylinder and thereby poses a
yet-unresolved challenge for numerical analysis. The simplifying
assumption of a cylinder of innite length is an example of the
disastrous consequences that can result from a seemingly
reasonable idealization.
On the other hand, this failure led to a great leap forward.
Saville retreated to the thin-boundary-layer model and
discovered a modication of the integral transformation of
Gortler that led to a common solution for innitely long
horizontal cylinders and round-nosed vertical cylinders of fairly
general contour in the form of an innite series that converges
more rapidly than the classical solution of Blasius for a round
horizontal cylinder. (See ref 87.) This methodology was then
utilized to develop corresponding solutions for entire classes of
vertical and horizontal cylinders and for simultaneous heat and
mass transfer (see ref 88). The identication of the falsity of a
simplifying assumption led to results of greater importance than
those initially sought. As an aside, the simplifying assumptions
of thin boundary layer theory are worthy of a separate analysis,
but one is not pursued herein.

(88)

For an ideal gas, = 1/T. For a liquid, it is eectively constant.


These idealizations may be noted to imply that the maximum
dierence in temperature is small and that pressure changes
resulting from the motion are negligible. The approximations of
the Boussinesq transformation can be, and increasingly are,
avoided in numerical solutions, but at a signicant cost
computationally and with a loss of generality for the solution.
The idealized model and solutions incorporating the
Boussinesq transformation will and probably should continue
to appear in our textbooks and handbooks, because of their
reasonable accuracy and the insight that they provide, but the
simplifying assumptions should not be overlooked.
8.3. The False Transient. The vorticity/stream-function
model for natural convection has one serious drawback. The
parabolic partial dierential equations describing the conservation of momentum become elliptic and can no longer be
solved numerically by marching in time. Samuels and
Churchill78 solved that diculty by introducing a false transient
term that avoids the need for iteration, phases out, and does not
result in any residual error in the solution for the steady state.
8.4. An Improved Measure of Convergence. Humbert
Chu (see Chu and Churchill79,80) investigated the eect of the
location of a heating element in a living or working space. In
order to make the computational task feasible, he postulated
two-dimensionality, represented the heating element by a
horizontal isothermal strip of nite width and the cooling by an
opposing isothermal wall at a lower temperature. His
experimental work was limited to photographs of the paths of
particles of smoke, which were in excellent agreement with his
computed streamlines. He obtained surprising results of
practical interest and value both physically and numerically.
By plotting the steady rate of heat transfer, as represented by
the Nusselt number (Nu), versus the rate of circulation, as
represented by the maximum value of the stream function, he
discovered that two widely diering rates of circulation may be
achieved for the same rate of heating by the choice of the
vertical location of the heater, a result of obvious importance.
At rst, he appeared to have paid a penalty for the unorthodox
use of a nonconservative nite-dierence formulation (which
results in a dierence in the heat ux entering and leaving the
space), namely, hard-to-overcome protests from the reviewers.
However, this disputed formulation led to an important
discovery that more than counterbalances its algorithmic
deciencies. The dierence between the computed values of
the entering and exiting uxes phase, which phases out in the
limit of zero grid size, is a far-better measure of the rate of
convergence with grid size than conventional ones. This is a
novel and little-known simplifying assumption that is greatly
advantageous over the one that it replaces.
8.5. Three-Dimensional Streaklines. Aziz and Hellums81
were apparently the rst to conduct successful numerical
solutions for three-dimensional natural convection in an
enclosure. They did so for a cubical region by virtue of using
the vector potential, which was adapted by Hirasaki and
Hellums82 from its use in electromagnetics by Maxwell.
Ozoe and collaborators (see, for example, refs 8386), in
turn, adapted this methodology to characterize and minimize
247

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

laminar ow in a round tube with a uniform wall temperature


converges too slowly to be of practical value near the entrance.
Finite-dierence solutions are now feasible for that regime, but
not in 1928, when Leveque96 derived an asymptotic solution.
His solution is perhaps the best example in this manuscript of
the criticality of the idealizations and of the ingenuity to choose
the optimal ones. They consist of the neglect of longitudinal
diusion, the existence of such a thin thermal boundary that the
eect of curvature can be neglected and the velocity
distribution approximated by u = 4um y/a, that heat transfer
occurs only by thermal conduction, and the that the condition
of symmetry at the centerline replaced by T T0 as y .
Leveque recognized that these several idealizations made
possible the reduction of the partial dierential energy balance
to an ordinary dierential equation by means of an integral
similarity transform in the form of the combined independent
variable y(3umc/(xa))1/3. (Most likely, he used trial and error
to identify the idealizations that allowed for such a similarity
transformation.) He thereby reduced the partial derived a
solution for the resulting ordinary dierential equation in the
form of an integral, which he recognized as a tabulated function
known as the incomplete gamma f unction of 1/3 order. The
resulting expression for the Nusselt number Nu, dened in
terms of Tw T0 rather than Tw Tmx, is

8.7. The Boundary Condition for Free Convection on


a Vertical Plate. The almost-universal boundary condition
inherent in closed-form and numerical solutions for the free
convection generated by an isothermally heated vertical surface
include a plate of innite width and semi-innite height upward
immersed in a uid of innite extent, but with no motion below
the bottom edge. The solutions actually correspond to that of a
plate of innite extent vertically heated isothermally above and
cooled isothermally below. The absurdity of this common
simplifying assumption and its impact on the behavior is rarely
noted in the literature, although it should be.
8.8. Potential (Inviscid) Flow and the Boundary-Layer
Concept. The usual simplifying assumptions for external
forced convection was also formulated by Boussinesq,89 who, in
1905, derived solutions for inviscid ow over immersed bodies.
These solutions have since been shown by Galante and
Churchill90 to have practical value only as asymptotes for Pr
0, and in the thin-boundary-layer-model of Prandtl. The
mathematical formulation of Boussinesq and the resulting
uidmechanicalthermal predictions remain alive in some
textbooks and working expressions for heat transfer even today.
Their limitations should not go unremarked.
8.9. Integral Boundary-Layer Theory. When I was a
graduate student, the advanced textbooks on uid mechanics
and heat transfer included a section on integral-boundary-layer
theory. This methodology, which was apparently devised
independently by von Karman91 and Pohlhausen92 in 1921,
was based, for ow, on the postulate of an arbitrary velocity
distribution, thus allowing an analytical integration of the
momentum balance to obtain a closed-form expression for the
drag coecient. The equivalent process was applied for the
temperature distribution to obtain an expression for the heattransfer coecient. Because of the smoothing, almost any
distribution incorporating a sucient number of arbitrary
coecients could be forced to provide an approximation of
rst-order accuracy. Such a result was often misinterpreted as a
validation of the postulated form. One positive measure of
progress in the eld of transport is the disappearance of this
concept from modern books on uid mechanics and heat
transfer.
8.9. The Turbulent Prandtl Number. The eddy diusivity
for heat transfer, and thereby the turbulent Prandtl number,
have been treated by purists with even more scorn than the
eddy diusivity for momentum transfer. Based on the success in
modeling fully turbulent ow in terms of (uv)++, we started to
model convection in fully developed turbulent ow through a
round tube in terms of the fraction of the local heat ux density
due to the turbulent ux, namely, (Tv)++ = cTv/j but soon
recognized (see ref 93) that replacing (Tv)++ by Prt/Pr by
virtue of
Prt
(uv)++ [1 (T v)++ ]
=
Pr
(T v)++ [1 (uv)++ ]

Nux = 1.167Gz1/3

(90)

Here, Gz = wc/kx = RePr(a/2x). A very thin thermal boundary


layer relative to the one for ow implies that Pr . Despite
these idealizations or perhaps fortuitously owing to compensations, his solution has proven to be a valid and useful asymptote
with a greater range of applicability than could have even been
hoped for. Its predictions are within 1% only for Gz > 800 000
but within 5% for Gz > 40. It has been adapted for many other
conditions, one of which is noted subsequently. The current
context calls for the recognition of the many idealizations, an
appreciation of their choice by Leveque, and a recognition of
the resulting limitations. The most important limitation is the
requirement of a large value of Pe = Re Pr by virtue of a large
value of Pr, because Re is limited in magnitude to <1800 to
ensure laminar ow. Another concern is the slowly increasing
error with increasing distance from the entrance.
Laminar Condensation. In 1916, Nusselt97 derived the
following solution in closed form for the rate of condensation
in a laminar lm by the ingenious choice of simplifying
assumptions, namely, an isothermal vertical plate of length L
and temperature Tp, a pure saturated vapor at Tg, negligible
drag for the vapor, negligible heat capacity, inertia, and fully
developed (parabolic) nonrippling ow for the condensate:

1/4
3/4
S 3
= 4 = 0.9428
G 2 3
3 3
3
gS kS (Tg Tp) L

(89)

(91)

Here, G is the mass rate of condensation per unit breadth of the


plate and is the latent heat.
This solution can be interpreted as the ultimate example of
the unique utility of a conceptual variable. Those of you who
are scornful of such a purely theoretical result should ask
yourself how much experimentation with what degree of
accuracy would be necessary to produce such a relationship,
that is lead to the identication of the correct power
dependence of the rate of condensation on each of these
variables. Dimensional analysis would get you part way but not

was a winner in terms of simplicity and physical meaning,


namely, the ratio of fractions of the transport of momentum
due to turbulence and viscous shear to the same quantity for
the transport of energy. In fully developed turbulent ow, /w
is known, but its analogue in fully developed turbulent
convection, namely, j/jw, is not (see Churchill and Balzhiser94
and Heng et al.95).
8.10. The Leveque Solution. The classical series solution
of Graetz18 for developing convection in fully developed
248

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

to this degree of resolution. Theoretical and/or numerical


solutions have since been derived that avoid each of the
simplifying assumptions of Nusselt, and they should be used for
design. Even so, his solution remains a touchstone, both
because of the insight provided by its combination of variables
and because of the fairly accurate quantitative agreement of its
predictions with experimental data.
Some textbooks present this solution rearranged in terms of
the Nusselt number Nu (and thereby the heat-transfer
coecient) as a function of the Reynolds number Re, but
that is a misdirected eort because the rate of ow (the rate of
condensation), which is part of Re, is a dependent rather than
an independent variable.
The validity of the several simplifying assumptions that led to
eq 91 has been examined by Churchill.98 The simplication that
is most responsible for misprediction is the neglect of rippling.
Nusselt proposed this expression for water vapor; therefore, he
is not responsible if it is applied for other compounds for which
the neglect of the heat capacity of the vapor is not a good
approximation.

excised from the literature of chemical engineering.


Perfect radial mixing may be retained as a simplifying
assumption but only in the sense of an upper bound for
the conversion or lower bound for the length.
(2) The use of space velocity or space time as the
independent variable, which is equivalent to postulating
plug ow. It should also be excised.
(3) Reaction in fully developed ow with no radial mixing,
which is the counterpart of plug ow or perfect radial
mixing. This simplifying assumption has value only as a
lower bound for the conversion or an upper bound for
the length. It implies initiation of the reaction over the
cross-section at the inlet. That inlet condition does not
appear to be possible except for a photochemical
reaction. The introduction of two streams, each with
one of the reactants, necessitates consideration of the
conversion during the process of mixing.
(4) Reaction-rate mechanisms in terms of concentrations,
which violates thermodynamic considerations. Correction of existing expressions is a formidable task because
the conditions under which the rate constants were
determined must be known.
(5) A power of temperature as a factor of the Arrhenius
equation. This correction is made necessary by virtue of
the expression of reaction rates in terms of concentration.
(6) Reaction-rate mechanisms expressed in terms of nonintegral powers of the concentrations (or activities),
which violates collision theory.
(7) The postulate of local equilibrium or of a pseudostationary state for the concentration of the free-radical
reactants.
(8) Isothermal or adiabatic conditions, which cannot can be
achieved with an energetic gas-phase reaction.
(9) The use of a heat-transfer coecient without correction
for the inuence of an energetic reaction.
(10) The failure to distinguish between the rate of change and
the rate of reaction.
(11) The failure to use free-radical mechanisms for processes
such as combustion, even though many of them have
been codied quantitatively.
(12) The failure to recognize the concept of stiness, which is
often encountered in the computation of the initiation of
reactions proceeding by free-radical mechanisms.
(13) The failure to note the one idealization that is always
justiable in a practical sense, namely, that of negligible
diusion in the longitudinal direction.
The rst textbook on reaction engineering that avoids these
false simplifying assumptions will make all prior ones instantly
and totally obsolete. The changeover will be accepted without a
blink by students, because they do not need to unlearn the false
concepts, and they will nd the new ones to be consistent with
their preparation in thermodynamics and transport. Some
mismatch in impedance can be expected when the rst wave of
so-liberated students reach industry, but the new concepts will
prevail, because the resulting designs will be more accurate.
9.2. Combustion. Combustion should logically be a branch
of reaction engineering but has been ceded to other brands of
engineering and science. This exclusion may stem in part
because it is rarely carried out in tubular ow or in stirred
vessels as are most reactions in chemical processing.
Combustion has been one of the primary elds of my research
but, uniquely and exceptionally, in tubular ow.

9. REACTION ENGINEERING
Reaction engineering is a unique glory of chemical engineering
but it has become encumbered with a great many false concepts
and unnecessary idealizations. Some of these concepts and
idealizations have been carried over from physical chemistry
and, in particular, from the expression of chemical conversions
in Lagrangian terms; however, most are a consequence of
simplications made in order to obtain closed-form solutions.
Reaction engineering is one of the most mathematically
oriented and mathematically advanced of all the courses in
the curriculum of chemical engineering, but instead of utilizing
advanced mathematics to solve more-complex and more
physically realistic models, most of the attention has been
focused on simpler and less realistic ones in order to abet the
mathematics. I have recently reviewed this topic in another
context (see ref 99), so only a skeletal description is presented
here. Furthermore, in the interests of brevity, attention is
limited to homogeneous reactions in tubular ow.
9.1. Usual Simplifying Assumptions That Are Invalid,
Misapplied, or Overlooked. These include, but are not
limited to, the following:
(1) Plug ow, which occurs physically only when a semisolid
such as ice cream is pushed through a tube with a solid
plunger. The same solution is obtained by the physically
conceivable asymptotic condition of perfect radial
mixing. Plug ow is often cited as a reasonable
approximation that is <12% in error, but that assessment
is based on the operational behavior, that is on the
change in conversion for a xed length. The error in
design, that is in the required length for a xed
conversion, may be as great as 100%. Furthermore, the
error is nonconservative in that the predicted conversion
is too great and the predicted length is too short.
Turbulent ow is often implied to approach plug ow
but it does not. Also, tubular reactors are rarely operated
in the turbulent regime (see Churchill and Pfeerle100),
because, for most reactions, the required length would
then be excessive. Most tubular gas-phase reactors
double as heat exchangers but that supplementary
process is excluded by the postulate of plug ow. The
usual simplifying assumption of plug ow should be
249

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

As a separate eld, combustion has acquired some highly


questionable simplifying assumptions. The rst of these is the
frequent use of global models for the rate of reaction.
Combustion always proceeds by means of free-radical reactions.
Solids and liquids do not burn: they must rst be pyrolyzed
(thermally decomposed), and the products of decomposition
must rst be vaporized. This process is often overlooked or
misunderstood, because the thermal energy produced by the
reaction is almost independent of the mechanism of reaction.
The application of this mechanism made it possible to predict,
almost to the minute, the time delays before collapse of each of
the twin towers of the World Trade Center (see Watkin et
al.101) This achievement is a classical example of the criticality
of the choice of idealizations.
The concept of a pseudo-stationary state, as applied by J. A.
Christiansen102 to solid-state-catalyzed reactions, is one of the
greatest contributions in the history of chemistry and chemical
engineering, in that it explained the previously puzzling
observations of fractional and bilinear mechanisms. However,
it is inapplicable for the free radicals by means of which
combustion proceeds, because of the infrequency of the
collisions between the molecules and free radicals (see Pfeerle
and Churchill103). This is a classical example of the application
of an ingenious concept outside its eld of validity.

transport of energy and momentum was eected wholly by


eddies moving from the bulk of uid to the wall, and, hence,
that the mass rates must be equal, and that, as a consequence,

h
= w2
cum
um

(92)

which may be expressed in dimensionless symbols as


St =

f
2

(93)

or as
Nu =

RePrf
2

(94)

Equations 9294 predict that the rate of convective transfer of


heat, as represented by h, St, or Nu, is equal to the rate of
convective transfer of momentum, as represented by w, f, or
RePrf, respectively. Equation 92 has the advantage of being the
most explicit but eq 94 is in more common usage, by virtue of
familiarity with Nu. The practical utility of eqs 93 and 94 is that
they serve to predict St and/or Nu from a correlating equation
for f. Although the intent of Reynolds was to devise a general
expression for prediction, the modern role of his analogy is as
an asymptote for Pr = Prt or an approximate one for Pr 1.
The Reynolds analogy is free of explicit empiricism, which is a
real merit, but it is inexact, because of the simplistic mechanism
of transport upon which it is based.
The analogy of Colburn,109 also known as the Colburn jfactor correlation, was one of many attempts to improve upon
the Reynolds analogy. It still appears in many textbooks and
perhaps even computer packages, although its functionally is in
error in every conceivable respect, and its numerical predictions
are as much as 40% in error in the very range of conditions in
which it was formulated (see Churchill and Zajic110). It lives on
out of a mixture of inertia and respect for Colburn. The process
of its derivation is worth examining if only to reveal how such
an erroneous expression can come into existence. It is probably
safe to say that few, if any, of the present or past users of the
Colburn analogy have had a complete knowledge of the details
of its derivation and, hence, of the underlying postulates,
idealizations, and simplications. A jaded reader might respond,
Who cares if it works?, but it does not work and is thereby a
pointed illustration of why a knowledge of these details is
important to practitioners as well as teachers.
The rst step is the derivation of the Colburn analogy is the
construction by Drew et al.111 in 1932 of the following
correlating equation for the Fanning friction factor for smooth
pipe for Re > 3000:

10. ANALOGIES
Many of the analogies in common use by chemical engineers
were rst conceived in other branches of engineering. Even so,
analogies appear to have a more essential role in chemical
engineering, perhaps because of the breadth of interests and the
involvement with processes in which ow, heat transfer, mass
transfer, and chemical reactions occur simultaneously and/or
interactively. Analogies may be considered as a form of
correlation in that they purport to allow the prediction of some
quantity by virtue of a correlation or theoretical relationship for
some other quantity. They may thereby eliminate the need to
obtain experimental data or carry out numerical solutions for
the analogous process.
The dierential equations of conservation for momentum,
energy, and species display some commonalities that suggest
analogous behavior but they also display fundamental dierences, and many of the analogies now in common use sprung
from some ingenious insight or from the observation of
commonalities in experimental data, numerically computed
values, or correlations thereof.
Attention is rst given to the classical analogies, then to some
that are less familiar, and nally to some well-known
relationships that may not be recognized as being analogies.
10.1. Analogies between Heat and Momentum
Transfer in Turbulent Flow. More analogies have been
proposed between the transport of energy and momentum in
fully developed turbulent ow in a round tube than for any
other pair of processes, and they may even exceed in number
and usage the correlating equations devised for heat transfer
alone. The current author recently reviewed the derivation and
validity of the most common analogies in this category,104 and
only the more important of his conclusions need be mentioned
here. The analogies of Reynolds,105 Prandtl,106 Thomas and
Fan,107 and Churchill108 for energy and momentum transfer are
representative of the heuristic and mechanistic ones. Their
predictions are unreliable functionally and numerically.
Of these four analogies, only that of Reynolds retains more
than historical interest. In 1874, he speculated that the

f
0.0625
= 0.0007 +
2
Re 0.32

(95)

The structure of this awkward expression is attributed to


Lees,112 although it was used earlier by Saph and Schoder,44
who are mentioned in the article. Colburn then discovered and
demonstrated that the following near-analogue of eq 95:
j StPr 2/3 = 0.0007 +

0.065
Re 0.32

(96)

provided a fair representation on the mean of the somewhatscattered data for heat transfer. (As an aside, reading a centuryold learned discussion of the state of the art of this behavior
compensates for the eort of nding it, despite a badly bungled
250

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

structure is independent of the expression for Prt/Pr. That


result, in turn, prompted Yu et al.116 to redo Reichardts
derivation completely with turbulent transport represented by
the left-most term of eq 48. The result is

reference.) In eq 96, Colburn invokes the Stanton number (St),


as well as introducing a new variable called the j-factor, which
represents both of the two presumed analogues, in this case f/2
and StPr2/3 . He then asserts that eq 96 can be approximated by

j StPr 2/3 =

0.023
Re 0.2

Pr 1
Pr 2/3 1
1
= t
+ 1 t
Pr Nu
Nu Pr Nu1

(97)

and then re-expressed in the old manner as


jRe =

Nu
= 0.023Re 0.8
Pr1/3

1/2

with Nu1 = Ref/2 and Nu = 0.7343 (f/2) Re Pr . The latter


coecient corresponds to 0.7001/3 33/2/2, where 0.700 is the
value determined by direct numerical simulation (DNS) for the
coecient of (y+)3 in the asymptotic expansion of Murphree53
for u+ near the wall. Both eqs 100 and 101 are free of any
explicit empiricism, which, as noted in connection with the
Reynolds analogy is a merit but not a guarantee of exactness.
Because it was apparent from its structure that eq 101 could be
valid only for Pr Prt, Churchill and Zajic117 devised a
complementary analogy for Pr Prt. It incorporates some
empiricism and is not reproduced here; however, in
combination with eq 101, it may be seen in Figure 3 of
Churchill and Zajic117 to represent the most reliable computed
and experimental values for all conditions better than any prior
expressions. That comparison illustrates the potential benet
from not only identifying and evaluating the usual simplifying
equations but replacing them with better ones.
The use of eq 101 for numerical prediction requires an
expression for Prt/Pr. Neither a theoretical expression nor a
reliable correlating equation currently exists (see ref 118), but
the predictions of eq 101 and its counterpart for Pr 0 are
insensitive in that respect.
10.2. An Analogy for the Eddy Conductivity. In 1960,
Abbrecht and Churchill119 determined the eddy conductivity
experimentally in a developing temperature eld and thereby
conrmed their conjecture that the uctuations in temperature
were generated by those in the velocity, and thereby
independent of the temperature eld and of the thermal
boundary conditions. It follows that Prt/Pr has the same
independence. The limited experimental data and numerical
solutions by DNS appear to conrm this conjecture.
10.3. The Analogy between Heat and Mass Transfer.
The second most commonly applied analogy in chemical
engineering is that between energy transfer and component
transfer. This analogy is based on the apparent congruence of
the equations of conservation of energy and species, and is
ordinarily implemented simply by substituting the Sherwood
number (Sh) and the Schmidt number (Sc) for the Nusselt
number (Nu) and the Prandtl number (Pr), respectively, in a
solution or a correlation for heat transfer. The earliest such
usage appears to have been that by Chilton and Colburn,120 as
an extension of the Colburn analogy. It would appear to have
great utility, because the experimental data and correlations for
heat transfer are more accurate and more extensive than those
for mass transfer. Unfortunately, it has failed critical tests such
as the precise electrochemical measurements of Shaw and
Hanratty,121 and the DNS computations of Papavassilou and
Hanratty.122 The supercial similarity is spoiled by several
factors, principally the bulk motion generated by nonequimolar
mass transfer, the dependence of Sc on composition, and the
diering dependences of Sc and Pr on temperature and
pressure. Also, the highest value of Pr for ordinary uids is less
than the lowest value of Sc, so no current experimental data for
Nu and Sh directly overlap. Despite the disproofs by Hanratty
and others, this concept persists without warnings of its

(98)

which he characterizes as quite comparable with the following


expression, recommended by McAdams113 for heating:
Nu
= 0.0225Re 0.8
Pr 0.4

(99)

He notes that the exponent of 1/3 for Pr in eqs 96, 97, and 98
was chosen as more or less of an average of the values ranging
from 0.3 to 0.4, proposed by prior investigators. That choice
happened to be the subsequently derived theoretical value for
the hypothetical limiting case of Pr and Re ;
however, as a xed power, it is grossly in error for values of Pr
approaching unity for which, as already noted, Nu is nearly
proportional to Pr. The dependence of Nu on Re is
subsequently revealed herein to undergo a gradual shift from
proportionality to Ref/2 to proportionality to Re(f/2)1/2 as Re
increases within the turbulent regime. The functionality of the
Colburn analogy was doomed to fail, because of the choice of a
xed power of Re as well as of a xed power of Pr. It survived as
a standby in process design because its aura prompted veteran
designers to use it, even though they needed to correct its
predictions with their own fudge factor. As an aside, in the mid1940s, when I worked in process design and took part in the
choice of which heat exchanger to purchase, we compiled a list
of the safety factors in applied by various vendors as
deciphered from their bids. As a direct consequence, the most
accurate design was not necessarily the winner; if a gross
overdesign was competitive pricewise, we chose it because the
excess capacity would both constitute a safety factor and allow
for a possible future increase in the rate of processing.
The analogy of Reichardt114 for moderate and large values of
Pr and a uniform wall temperature is worth examining, because
it has a dierent and more theoretically based origin than any of
the prior ones and because it served as a stepping stone for the
currently most accurate one. He started by taking the ratio of
expressions for the radial velocity distribution in terms of the
eddy viscosity and its analogue for the radial temperature
gradient, in terms of the eddy conductivity, and thereby
eliminating the distance from the wall as a variable. He then
made four ingenious but questionable simplications, which
allowed him to integrate formally the resulting relationship
between the temperature and the velocity. His nal result
included several denite integrals for which he devised
empirical correlating equations.
Churchill et al.115 improved upon one of Reichardts
simplications and noted that the revised analogy could be
expressed in the following simplied form:
Pr 1

Pr 1
1
= t
+ 1 t
Nu Pr Nu1
Pr Nu

(101)
1/3

(100)

Here, Nu1 and Nu are the asymptotes for Pr = Prt and Pr


, respectively. This is a great step forward because the
251

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

intentional sacrice to obtain commonality and insight. As an


aside, the term in brackets, after substitution for F, K, and ,
may be identied as the Rayleigh number in the case of free
convection, but the counterparts for the other three gravitational processes do not appear to have accepted names.
The corresponding analogy for the turbulent regime is

shortcomings in many textbooks, and perhaps in some


computer packages.
10.4. The Analogy between Electrical and Thermal
Conduction. The analogy between electrical and thermal
conduction is simply presumed by most engineers to be exact
for the steady state. However, Ludwig Boltzmann was unsure
and proposed an experimental investigation of its validity as the
subject of the doctoral research of Lise Meitner, who is better
known for her unrequited discovery of nuclear ssion by means
of slow neutrons (see Rayner-Canham and Rayner-Canham123). She conrmed the validity of the presumption.
This analogy has few active applications but two are of great
importance, namely Ohms law of conduction through
resistances in series and parallel and Maxwells law of
conduction through dispersions. They were both conceived in
electrical terms but their analogues in chemical and thermal
processing have proven to be of great utility. Maxwells law has
been simplied and generalized for ordered dispersions and
generalized as an approximation for packed and uidized beds
(see ref 124).
10.5. The Analogy of MacLeod. Rothfus and Monrad125
determined the conditions that force the velocity distribution in
fully developed laminar ow in round tubes and parallel-plate
channels to be congruent. Remarkably, these conditions are
exactly equivalent to the expression of the results for round
tubes, in terms of a+ and y+, and those for parallel-plate
channels in terms of y+ and b+. That analogy is intriguing but of
little direct practical importance. However, it led to an
important result. Rothfus and Monrad credit a doctoral
student, Alexander MacLeod,126 with the conjecture that this
congruence might carry over to turbulent ow. A plot of
experimental data for u+c versus a+ and b+ by Whan and
Rothfus127 (or see Figure 5 in ref 66) conrmed that conjecture
and also demonstrated that it does not apply to the regime of
transition. This discovery of congruence is of great importance,
not only because it allows experimental data for the two
geometries to be used interchangeably, but even more
importantly, because it allows the numerical values obtained
for parallel-plate channels by DNS to be applied for round
tubes. Furthermore, from the analogy of MacLeod it can be
inferred that Prt/Pr is identical in a round tube and a parallelplate channel in terms of a+ and b+ as conjectured by Abbrecht
and Churchill.119
10.6. An Analogy between Dierent Buoyant Processes. In 1954, Emmons128 recognized that laminar free
convection, lm condensation, lm boiling, and lm melting are
all controlled by thermal conduction across a lm moving at a
velocity controlled by gravity, and derived the following
generalized solution for the mean heat-transfer coecient:
FL3 1/4
hmL
=

k
KkT

FL3 1/3
hmL
= A

k
KkT

(103)

Here, A is an empirical coecient. Emmons values of F, K, and


convert this expression to the accepted form for turbulent
free convection, but the limited and very scattered experimental
data for lm condensation, lm boiling, and lm melting in the
turbulent regime do not appear to conrm the predicted
independence from L. In a qualitative if not a quantitative
sense, the analogy of Emmons represents one of the broadest
generalities in all of transport.
10.7. The Equivalent Diameter. Perhaps the most widely
used analogy in practice is that of the equivalent diameter. This
concept allows correlations for ow and/or heat transfer in one
geometry, usually a round pipe, to be utilized as an
approximation for other geometries. The concept requires the
choice of some arbitrary methodology to determine the
equivalent diameter. The most common choice is the hydraulic
diameter, which is dened as four times the cross-sectional area
divided by the wetted perimeter. The hydraulic diameter results
in an overprediction of the friction factor for laminar ow in a
parallel-plate channel by 50% but a lesser error in turbulent
ow and in annuli and rectangular ducts. The laminar-equivalent
diameterthat is, the expression that produces the exact value
for laminar owis somewhat more accurate but requires the
availability of a solution or correlation for the latter.
10.8. The Radiative Heat-Transfer Coecient. If
thermal radiation occurs in series or parallel with thermal
conduction, forced convection, and/or free convection, it is
convenient to linearize its dependence on temperature by
means of the following expression in order to allow the
application of Ohms laws, which are restricted to processes
that are linear in the potential:
hR

(Ts4 Ts4)
2
)(Ts + T) 4Tm3
= (Ts4 + T
Ts T
(104)

The exact value of Tm can be calculated for each specied set of


values of Ts and T; however, since the variation of the
absolute value is ordinarily constrained, some arbitrary, xed,
mean value is usually employed as an approximation. A
radiative thermal conductivity may also be dened but it is less
useful.
10.9. The Eddy Viscosity and Mixing Length. These
two models, which were discussed in Section 7.6, could be
interpreted as analogies. The eddy viscosity was obviously
conceived by Boussinesq as an analogue of the viscosity, and
the mixing-length concept was put forward by Prandtl as an
analogue of the mean-free-path of gaseous molecules. The
critique of the mixing-length model as a very poor analogy by
Bird et al.129 may be justied, but this concept nevertheless has
proved, as mentioned in Section 7.8, to have validity, if not
utility, as a model for turbulent transport.
10.10. The Analogy Based on the Leveque Solution.
Martin130,131 in 2002 devised (and, in 2005, revised and
extended) an analogy between heat transfer and the pressure

(102)

Here, F is the gravitational force per unit volume of the lm, K


a numerical factor characterizing the viscous shear, and the
increase in the mass rate of ow of the lm per unit of heat
transfer. (Equation 102 is of course related to eq 91.) He
tabulated values of the factors F, K, and for each of the four
processes, based on qualitative considerations. One merit of
this analogy is that it allows a common plot of experimental
data for all four processes. More-accurate coecients have been
derived for each of the four individual processes, so the
inaccuracy of the tabulated values of Emmons represents an
252

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

obtained by mean of a numerical solutionbut that


compromise of standards should be acknowledged. Engineers
in industry are employed under the presumption that they stay
aware of advances in technology and that they can be counted
upon to utilize the best methodologies and information. If they
fail to live up to that standard, they are susceptible to
replacement by computer packages and nontechnical key
punchers.
Experimental measurements are the ultimate test of the
validity of simplifying assumptions, and a necessary one in cases
of geometric complexity. However, measurements that fulll
that need in both scope and reliability rarely exist. The everexpanding scope and accuracy of numerically computed
solutions ll that void to some extent, but they are rarely the
ultimate test.
A familiarity in depth with the recent literature on a topic
may be sucient to identify the shortcoming of a classical
expression in a qualitative sensefor example, the failure of the
derivative with respect to the radius, to go to zero at the
centerline of a round tube. However, an extended investigation
may be necessary to evaluate the consequences of some
shortcoming in a quantitative sensefor example, the eect of
an overly simplied velocity distribution on the mixed-mean
chemical conversion in a tubular reactor. The latter form of
evaluation would, in the case of many of the simplications
identied herein, require a separate dedicated investigation.
The preference of the reviewer (and of myself) for evaluation
of the error due to simplifying assumptions is a comparison of
the predictions with experimental measurements, but that is
rarely possible, because of the paucity of the latter. In 1959, in
connection with a 50th Anniversary Meeting of the AIChE,
Robert R. White and I were asked to review the experimental
foundations of chemical engineering. Our nding was that, to a
rst-order approximation, there are not any. The actual
written conclusion in the ensuing publication (White and
Churchill134) was The eld of chemical engineering is still
open for precise and careful experimentation on relatively
simple systems to provide the foundations of experimental
evidence, which by and large are still missing. In the process of
the current investigation, I was forced to conclude that such
fundamental measurements are still missing in many if not most
classic topics, despite presumed advances in instrumentation.
Such measurements have apparently not been carried out,
because they are expensive and tedious, compared to
theoretical work, and unrewarding, compared to exploratory
research. It is amazing how much faith many engineers have in
a structure that has been erected on such a weak or nonexistent
foundation. Numerical solutions are a possible standard for
evaluation of simplifying assumptions, but only if you believe
that the partial dierential equations of conservation are exact
for the condition of interest. That being said, such evaluations
already exist in a few cases, as illustrated by the following
examples.
Cleland and Wilhelm135 utilized both experimental measurements and nite-dierence calculations to evaluate the error in
the traditional idealized model for a tubular reactor. Churchill
and Yu136 utilized nite-dierence computations to evaluate the
errors in the prediction of the chemical conversion for gasphase reactions, because of the postulate of plug ow, adiabatic
ow, or isothermal ow. The concept of plug ow is misleading
in a physical sense and a very poor idealization in terms of
modeling in that it may be in error by as much as 50%;
furthermore, it is bad engineering in that it underpredicts the

drop based on the Leveque solution (see Section 8.10), for


plate-type exchangers with cross-corregated channels of the
chevron type, for packed beds, for tube bundles with crossedrod matrices and many other spacewise periodic arrangements,
and for external ows. This analogy is an amazing generalization but it is not an asymptote. Instead, it is an as empirical
approximation, and its validity and its functional accuracy need
to be examined by the user, case by case. It would appear to
have utility as a rst-order approximation in situations in which
the complexity of the geometry yet precludes numerical
solutions.
10.11. An Analogy between Chemical Reaction and
Convection. Although energetic chemical reactions have been
known for over 40 years to enhance or attenuate convection
strongly this important aspect of behavior does not appear to
have found its way into any textbooks or handbooks on heat
transfer or mass transfer. My encounter with this phenomenon
began in 1972 when the observed heat-transfer coecient in a
ceramic tube in which a ame was stabilized by thermal
radiation appeared to be an order of magnitude greater than the
prediction of the standard models. That led to the formulation
of a model in the form of an analogy between convection and
the rate of an energetic homogeneous reaction (see ref 132).
That model, together with experiments and essentially exact
numerical solutions to test it (see Yu and Churchill133),
revealed that the enhancement of the heat ux density at the
walland, thereby, the heat-transfer coecientare due to
the greater conversion in the slow moving uid near the wall
than on the mean, Because of that aspect of the behavior, they
independently considered the possibility of adaptation of the
solution of Leveque or at least his linearization of the velocity
prole but found it unnecessary, in view of the success of the
numerical solutions.

11. EVALUATION OF THE SIGNIFICANCE AND


MAGNITUDE OF THE ERROR DUE TO
SIMPLIFYING ASSUMPTIONS
The criticisms, comments, and suggestions of the reviewers
usually result in improvements in my manuscripts. This section
has been added in response to a reviewer who speculated that
my motive was academic purity and asked for a demonstration
of the impact of the simplifying assumptions that I identied as
false, obsolete, or unnecessary, and preferably by means of
experimental measurements rather than by numerical solutions.
That criticism enhanced this investigation by prompting the
following additional considerations.
Some of the usual simplifying assumptions result in serious
errors in a practical as well an intrinsic sense, and their
identication and elimination constitute an important advance
in a commercial sense. Examples cited herein include (1) plug
ow, which may result in the underdesign of a chemical reactor
by 50%, and (2) the Colburn analogy, which may result in the
underdesign of a heat exchanger by 40%.
Second, engineers are generally presumed to have a
professional obligation to utilize the best available technical
resources and methodologies, and to restrict their choices to
those that are scientically sound and have achieved general
acceptance. This investigation has identied many expressions
in the permanent literature of chemical engineering that no
longer meet those standards, because of new resultseither
theoretical or experimental. Compromises may be excusable in
the interests of conveniencefor example, the use of a
correlating equation, rather than the slightly more exact results
253

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

analysis and in methods of numerical solution. Unfortunately,


the failure to re-examine them regularly has, in many instances,
prevented the elimination or replacement in books, in the
classroom, in computer packages, and in industrial practice of
the obsolete concepts and relationships as well as idealizations
and simplications that are no longer necessary. Realistically,
textbooks can never be kept wholly up to date. New concepts
and improved idealizations are being conceived and false ones
identied while a book or a new edition is being printed, but
updates can readily be distributed to students and discussed in
the classroom.
It is important in a professional sense that industrial
practitioners of chemical engineering, as well as teachers and
students, have a knowledge of the origins as well as the
limitations and shortcomings of the formulas and computer
packages that they use. They should welcome improvements,
rather than resist them.
The examination of several particular and diverse examples
herein is not purported to serve as a thorough housecleaning
but rather as an illustration of what needs to be done routinely
in every nook and cranny of chemical engineering and the
associated sciences and technologies. Every teacher and book
writer should accept this responsibility, and every engineer has
an obligation to remain up to date by means of regular, critical
examinations of their own technical storage bin.
The qualitative and quantitative impacts of retention of some
of the obsolete simplications were reviewed in the previous
section, in response to a reviewer. Some important ones not
mentioned there are listed here, but they also are merely a
representative selection.
(1) Graphical and tabular correlations are passe in the age of
computers.
(2) Dimensional analysis of a list of variables identies a set
of dimensionless groups that are sucient to describe the
behavior. It does not produce an algebraic expression in
the form of products of powers of those groups.
(3) Correlating equations in the form of a single power
function or a product of power functions are
fundamentally in error and their replacement should be
sought. Correlating equations in the form of the powermean of asymptotes represent a promising alternative.
(4) The avoidance of a correlating equation that requires an
iterative solution, such as eq 50 for a specied Reynolds
number Re, ceased to be a valid excuse long ago when
programmable computers became generally available, but
it is still often cited today.
(5) The usual simplifying assumptions that led to the
introduction of that terminology back in the 1940s are
still invoked by the concept of an ideal stage and the
methodology of the McCabeTheile diagram. The ratio
of number of ideal trays to physical ones should be
identied as a measure of the failure of the concept of an
ideal stage. Realistic modeling of distillation in a tray
column is not yet feasible in a generalized sense and may
never be, because it depends on the details of the
hardware. However, qualitative description of the
process, in terms of uid ow, mixing, and mass transfer
is invaluable in that it explains the failure of the idealstage concept and points the way to improvement.
(6) The eddy viscosity and mixing length are inferior to the
local fraction of the shear stress due to turbulence as

required length of a reactor for a given conversion. The


concepts of isothermal and adiabatic reactions are misleading,
because, for most gas-phase reactions, the former is
unattainable and the latter is impractical. The error due to
the neglect of the region of development of the velocity prole
and/or the region of mixing of two streams of reactants may be
quite signicant, because the rate of reaction is highest there.
Just as with the case of the ideal stage in distillation, the error
associated with developing ow is dicult to generalize. Fully
developed ow is a conservative idealization in this application
but complete mixing is not.
Beginning in 1961, both numerical computations and
experimental measurements have been used to evaluate the
considerable enhancement or attenuation of the heat-transfer
coecient in tubular ow quantitatively, because of an
energetic chemical reaction. (See, for example, Yu and
Churchill.133) It is reckless to ignore this eect.
Expressions for the error resulting from each of the
simplifying assumptions invoked in the derivation of eq 17
for lm condensation have been devised by Churchill.137 The
failure of condensers to maintain the temperature of electronic
gear in the distant early warning line (DEW), because of the
neglect of the heat capacity of the liquid (a usual simplifying
assumption), led to his investigation.
Figure 6 of Pfeerle and Churchill103 displays the numerically
computed error resulting from the postulate of a pseudosteady-state concentration for the free radicals in a ame.
The error due to eq 68 (with ones favorite coecients)
relative to eq 50 for parametric values of a+, and that due to eq
62 (with ones favorite coecients) relative to eq 49 can readily
be calculated, so there is no need for tabulations or graphs.
Figure 3 of Churchill and Zajic117 displays the considerable
fractional error in the prediction of Nu for fully developed
turbulent ow in a round tube by all prior analogies. As already
mentioned, the Colburn analogy results in the greatest error.
The analogy of McLeod for fully developed turbulent ow
and fully developed laminar ow was conrmed by means of
experimental measurements, as described in Section 10.5. From
that agreement, it may be inferred that the radial distributions
of the time-mean velocity and of the fraction of the shear stress
due to turbulence local results obtained for a parallel-plate
channel by means of DNS are applicable for a round tube.
Although limited to a hypothetical condition at low Re values,
the results obtained by DNS agree almost exactly with those
obtained by numerical solution of the time-averaged equations
of conservation for the same conditions and thereby support
the validity of the partial dierential model obtained by timeaveraging.

12. SUMMARY AND CONCLUSIONS


A representative set of the concepts and expressions that
characterize both education and practice in chemical engineering has been examined with the objective of identifying the
presence and impact of the idealizations and simplications
made in their formulation.
Most of the idealizations and simplications were in
themselves visionary and ingenious at the time when they
were rst made. They were essential to the development of
most of the characteristic and enduring concepts of chemical
engineering. As the usual simplifying assumptions, they
rightfully became part of the lore of chemical engineering.
Some still retain validity, at least as asymptotes, but many have
been reduced to historical artifacts by virtue of progress in
254

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

(7)
(8)
(9)
(10)
(11)
(12)
(13)

Article

(4) Hellums, J. D.; Churchill, S. W. Simplification of the


Mathematical Description of Boundary and Initial Value Problems.
AIChE J. 1964, 10, 110.
(5) White, C. W., III; Churchill, S. W. A Computer Program for
Dimensional Analysis and the Indentification of Similarity Transformations. Proc. Second World Congress Chem. Eng. 1981, I, 331.
(6) Churchill, S. W. A New Approach to Teaching Dimensional
Analysis. Chem. Eng. Educ. 1997, 31 (Spring), 118.
(7) Prandtl, L. U ber die ausgebildete Turbulenz. Verhdl. 2. Kongr.
Techn. Mech., Zurich 1926, 62.
(8) Millikan, C. B. A Critical Discussion of Turbulent Flows in
Channels and Circular Tubes. Proc. Fifth Int. Congr. Appl. Mech.
(Cambridge, MA) 1938, 386.
(9) Churchill, S. W.; Usagi, R. A General Expression for the
Correlation of Rates of Transfer and Other Phenomena. AIChE J.
1972, 18, 1121.
(10) Churchill, S. W.; Usagi, R. A Standardized Procedure for the
Production of Correlations in the Form of a Common Empirical
Equation. Ind. Eng. Chem. Fundam. 1974, 13, 39.
(11) Gardner, M. Mathematical Carnival. In A New Round-Up of
Tantilizers and Puzzles from Scientic American; Vintage: New York,
1977; p 240.
(12) Churchill, S. W. A Comprehensive Correlating Equation for
Laminar, Assisting, Forced and Free Convection. AIChE J. 1977, 23,
10.
(13) Kitaura, Y.; Tanaka, H. (title in Jpn). Kagaku Kogaku 1964, 19,
793.
(14) Ruckenstein, E.; Rajagopalan, R. A Simple Algebraic Method of
Obtaining Heat or Mass Transfer Coefficients under Mixed
Convection. Chem. Eng. Commun. 1980, 4, 15.
(15) Newton, I. Scala graduum caloris. Philos. Trans. R. Soc. (London)
1701, 22, 824.
(16) Langmuir, I. Convection and Conduction of Heat in Gases.
Phys. Rev. 1912, 34, 401.
(17) Seban, R. A.; Shimazaki, T. T. Heat Transfer to a Fluid Flowing
Turbulently in a Smooth Tube with Walls at Constant Temperature.
Trans. ASME 1951, 73, 803.
(18) Graetz, L. U ber die Warmeleitungsfahigkeit von Flussigkeiten.
Ann. Phys. Chem. 1883, 18, 79.
(19) McCabe, W. L.; Theile, E. E. Graphical Design of Fractionating
Columns. Ind. Eng. Chem. 1925, 17, 605.
(20) Fanning, J. T. A Practical Treatise on Hydraulics and Water
Supply Engineering; Van Nostrand: New York, 1877.
(21) Seider, W. D.; Churchill, S. W. Confined Jet Mixing in the
Entrance of a Tubular Reactor. AIChE J. 1971, 17, 704.
(22) Lamb, H. Hydrodynamics; 6th Ed., 1st American Ed.; Dover:
New York, 1945; p 99.
(23) Churchill, S. W. The Art of Correlation. Ind. Eng. Chem. Res.
2000, 39, 1850.
(24) Tuve, G. L.; Sprenkle, R. E. Orifice Discharge Coefficients for
Viscous Fluids. Instruments 1933, 6, 201.
(25) Baer, A. D. Private communication.
(26) Roscoe, R. The flow of viscous fluids around plane articles.
Philos. Mag. 1949, 40, 338.
(27) Navier, C.-L. Memoire sur les du mouvement des fluides. Mem.
Acad. R. Sci. 1833, 6, 389.
(28) Stokes, C. G. On the Theories of Internal Friction of Fluids in
Motion and of the Equilibrium and Motion of Elastic Solids. Trans.
Cambridge Philos. Soc. 1845, 8, 287.
(29) Reynolds., O. On the Dynamical Theory of Incompressible
Viscous Fluids and the Determination of the Criterion. Philos. Trans. R.
Soc.(London) 1895, 186, 123.
(30) Uhlenbeck, G. E. Some Notes on the Relation between Fluid
Mechanics and Statistical Physics. Ann. Rev. Fluid Mech. 1980, 12, 1.
(31) Orszag, S. D.; Kells., L. C. Transition to Turbulence in Plane
Poiseuille and Plane Couette Flow. J. Fluid Mech. 1980, 96, 159.
(32) Kim, J.: Moin, P.: Moser, R. Turbulence Statistics in Fully
Developed Channel Flow at Low Reynolds Numbers. J. Fluid Mech.
1987, 177, 133.

variables and should be eliminated from the literature,


except possibly as historical artifacts.
The model has value only as a crude approximation
for developing ow.
Direct numerical simulation (DNS) has failed to live up
to its great initial promise, and large eddy simulation has
failed as a supplement thereof.
Despite its questionable basis, the equivalent thickness of
Langmuir has proven to produce useful approximations
for limiting behavior and for the eect of curvature.
The Colburn analogy is invalid functionally and in error
numerically. It should be excised except as a historical
artifact.
The analogy between heat and mass transfer should be
identied as disproven.
The hyperbolic equation of conduction should be
identied as false. Its specious origin has some historical
value as a bad example.
Reaction-rate mechanisms, in terms of concentrations,
are contradictory to thermodynamic principles. That
contradiction is apparent in the temperature dependence
and the equilibrium constant. The elimination of this
false concept is a long-term project but should be
initiated now by the editors of journals that publish
articles in reaction engineering.

12. APOLOGIA
The identication herein of false or outdated concepts,
idealizations, and simplication is not meant as a criticism of
the pioneers in the technology of chemical engineering and of
the associated branches of engineering and science who
originated them. They were risk-takers and did not have the
benet of the experimental data, numerical solutions, and
analyses that are now available to us. We see further because we
stand on their shoulders. Newton, Rayleigh, Reynolds, Prandtl,
von Karman, McCabe, Theile, and Colburn are among my
idols. In that spirit, I took the liberty of even identifying a nowobsolete contribution of L. T. Fan, in whose honor this
manuscript was prepared, as well as several of my own false
starts.
I have two excuses (not apologies) for the choice of many of
the idealizations and simplications from or related to the work
and experiences of myself and my collaborators for illustrative
examples. First, they provide a rich source of examples of false
starts and uncertainties of which I am all too aware. Second,
they allow me to avoid embarrassing fellow chemical engineers
by revealing their missteps.

AUTHOR INFORMATION

Corresponding Author

*Tel.: 215-898-5579. Fax: 216-573-2093. E-mail: churchil@


seas.upenn.edu.
Notes

The authors declare no competing nancial interest.

REFERENCES

(1) Anderson, T. J. Chemical Processing of Electrons and Holes.


Chem. Eng. Educ. 1990, 26, 1 (Winter), 3.
(2) Lord Rayleigh (J. W. Strutt). The Principle of Similitude. Nature
1915, 95, 66.
(3) Nusselt, W. Der Warmeubergang in Rohrleitungen. Waermeuebergang Mitt. Forsch.-Arb. Ing.-Wes. 1909, 89, 1750.
255

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

(33) Schumann, U. Subgrid Scale Model for Finite Simulations of


Turbulent Flow in Plane Channels and Annuli. J. Comput. Phys. 1976,
18 (4), 376.
(34) Boussinesq, J. Essai sur la theorie des eaux courantes. Mem.
presents divers savants Acad. Sci. Inst. France 1877, 23, 1.
(35) Kolmogorov, A. N. Equations of Motion of an Incompressible
Turbulent Fluid. Izv. Nauk SSSR Ser. Phys. VI 1942, (12), 56.
(36) Prandtl, L. U ber ein neues Formelsystem fu r die ausgebildete
Turbulenz. Nachr. Ges. Wiss. Gottingen Math.-Phys. Klasse 1945, 6.
(37) Batchelor, G. K. The Theory of Homogeneous Turbulence;
Cambridge University Press: Oxford, U.K., 1953.
(38) Launder, B. E.: Spalding, D. B. Mathematical Models of
Turbulence; Academic Press: London, 1972.
(39) Prandtl, L. Bericht uber Untersuchungen zur ausgebildete
Turbulenz. Z. Angew. Math. Mech. 1925, 6, 136.
(40) Prandtl, L. Neuere Ergebnisse der Turbulenz. VDI-Z. 1933, 77,
105.
(41) von Karman, Th. Mechnische A hnlichkeit und Turbulenz. Proc.
3rd Int. Congress Appl. Math. 1930, 1, 85.
(42) Churchill, S. W.; Chan, C. Turbulent Flow in Channels in
Terms of Local Turbulent Shear and Normal Stresses. AIChE J. 41
1995, 42, 2513.
(43) Churchill, S. W.; Chan, C. Improved correlating equations for
the friction factor for fully developed turbulent flow in round tubes
and identical parallel plates, both smooth and rough. Ind. Eng. Chem.
Res. 1994, 33, 2016.
(44) Churchill, S. W.; Chan, C. Theoretically Based Correlating
Equations for the Local Characteristics of Fully Turbulent Flow in
Round Tubes and Between Parallel Plates. Ind. Eng. Chem. Res. 1995,
34, 1332.
(45) Churchill, S. W. New Simplified Models and Formulations for
Turbulent Flow and Convection. AIChE J. 1997, 42, 1125.
(46) Saph, A. V.: Schoder, E. H. An Experimental Study of the
Resistance to the Flow of Water in Pipes. Trans. Am. Soc. Civ. Eng.
1903, 51, Paper 944.
(47) Blasius, H. Das A hnlichkeitsgesetz bei Reibungsvorgangen im
Flussigkeiten. VDIForschungsh. 1913, 131.
(48) Prandtl, L. U ber den Reibungswiderstand stromender Luft, Ergebn,
Aerodyn; Versanst Gottingen III Lieferung., Munchen-Berlin, 1927; 1.
(49) Nikuradse, J. Gesetzmassigkeiten der turbulenten Stromungen
in glatten Rohren. VDIForschungsh. 1932, 356.
(50) (a) Barenblatt, G. I. Scaling Laws for Fully Developed Turbulent
Shear Flows. Part I. Basic Hypotheses and Analyses. J. Fluid Mech.
1993, 248, 513. (b) Barenblatt, G. I. Scaling Laws for Fully Developed
Turbulent Shear Flows. Part II. Processing and Experimental Data. J.
Fluid Mech. 1993, 248, 521.
(51) (a) Cipra, B. A New Theory of Turbulence Causes a Stir Among
Experts. Science 1996, 272 (May 17), 951. (b) Cipra, B. Sci. Lett. 272
(Aug. 22).
(52) Churchill, S. W.; Shinoda, M.; Arai, N. An Appraisal of
Resources and Methodologies for Prediction of Flow and Convection
in Channels. Thermal Sci. Eng. 2002, 10 (2), 1.
(53) Churchill, S. W. The Prediction of Turbulent Flow and
Convection in a Round Tube. Adv. Heat Transfer 2000, 34, 255.
(54) Murphree, E. V. Relation between Heat Transfer and Fluid
Friction. Ind. Eng. Chem. 1932, 24, 726.
(55) Reichardt, H. Vollstandige Darstellung der turbulenten
Geschwindigkeitsverteilung in glatten Leitungen. Z. Angew. Math.
Mech. 1951, 31, 201.
(56) Finley, P. J.; Khoo, C. P.; Chin, J. P. Velocity Measurements in a
Thin Turbulent Water Layer. Houille Blanche 1966, 21, 713.
(57) Hinze, J. O. Turbulence; McGrawHill: New York, 1959.
(58) Taylor, G. I. Diffusion with Continuous Movements. Proc.
London Math. Soc. 1921, 24A, 196.
(59) Taylor, G. I. The Dispersion of Matter in Turbulent Flow
Through a Pipe. Proc. R. Soc., Ser. A 1954, A223, 446.
(60) Groenhof, H. C. Eddy Diffusion in the Central Region of
Turbulent Flows in Pipes and Between Parallel Plates. Chem. Eng. Sci.
1970, 25, 1005.

(61) Wei, T.; Willmarth, W. W. Reynolds-Number Effects on the


Structure of a Turbulent Channel Flow. J. Fluid. Mech. 1980, 204, 57.
(62) Eckelmann, H. The Structure of the Viscous Sublayer and the
Adjacent Wall Region in Turbulent Channel Flow. J. Fluid Mech. 1974,
65, 439.
(63) Rutledge, J.; Sleicher, C. A. Direct Simulation of Turbulent Flow
and Heat Transfer in a Channel. Part I. Smooth Walls. Int. J. Num.
Methods Fluids. 1993, 16, 1051.
(64) Churchill, S. W. Turbulent FlowsThe Practical Use of Theory.
Book in preparation.
(65) Nikuradse, J. Stromungsgesetze in rauhen Rohren. VDI
Forschungsh. 1933, 361.
(66) Churchill, S. W. The Art of Correlation. Ind. Eng. Chem. Res.
2000, 39, 1850.
(67) Colebrook, C. F. Turbulent Flow in Pipes with Particular
Reference to the Transition Region Between Smooth and Rough Pipe
Laws. J. Inst. Civ. Eng. 19381939, 11, 133.
(68) Churchill, S. W. Empirical Expressions for the Shear Stress in
Turbulent Flow in Commercial Pipe. AIChE J. 1973, 19, 395.
(69) Moody, L. F. Friction Factor for Pipe Flow. Trans. ASME 1944,
66, 671.
(70) Zagarola, M. V. Mean-ow scaling of turbulent pipe ow. Ph.D.
Thesis, Princeton University, Princeton, NJ, 1996.
(71) Miller, B. The Laminar Film Hypothesis. Trans. ASME 1949, 71,
457.
(72) Nikuradse, J. Widerstandgesetz und Geschwindigkeitsverteilung von
turbulenten Wasserstromunden in glatten und rauhen Rohren. Verh. 3rd
Int. Kongr. Technol. Mech.: Stockholm, 1930; p S.239.
(73) Fourier, J. B. J. Theorie Analytique De La Chaleur; GauthierVillars: Paris, 1882.
(74) Maxwell, J. C. On the Dynamical Theory of Gases. Philos. Mag.
1867, 157, 49.
(75) Brown, M. A.; Churchill, S. W. Experimental Measurements of
Pressure Waves Generated in a Confined Gas by Impulsive Heating of
a Surface. AIChE J. 1995, 41, 205.
(76) Brown, M. A.; Churchill, S. W. Finite-Difference Computation
of the Wave Motion Generated in a Gas by a Rapid Increases in the
Bounding Temperature. Comput. Chem. Eng. 1999, 23, 357.
(77) Lord Rayleigh (Strutt, J. W.) On the Conduction of Heat in a
Spherical Mass of Air Conned by Walls at Constant Temperature.
Philos. Mag. 1899, 47, 314.
(78) Samuels, M. R.; Churchill, S. W. Stability of a Fluid in a
Rectangular Chamber Heated from Below. AIChE J. 1967, 13, 77.
(79) Chu, H. H.-S.; Churchill, S. W. The Effect of Heater Size,
Location, Aspect Ratio and Boundary Conditions on Two-Dimensional, Laminar Natural Convection in Rectangular Channels. J. Heat
Transfer, Trans. ASME 1976, 98C (194), 513.
(80) Chu, H. H.-S.; Churchill, S. W. The Development and Testing
of a Numerical Method for Computation of Laminar Natural
Convection in Enclosures. Comput. Chem. Eng. 1977, 1, 103.
(81) Aziz, K.; Hellums, J. D. Numerical Solution of the ThreeDimensional Equations of Motion for Laminar Convection. Phys.
Fluids 1967, 10, 314.
(82) Hirasaki, G. J.; Hellums, J. D. A General Formulation of the
Boundary Conditions on the Vector Potential in Three-Dimensional
Hydrodynamics. Quart. Appl. Math. 1968, 26, 342.
(83) Ozoe, H.; Kazumitsu, Y.; Sayama, H.; Churchill, S. W. ThreeDimensional, Numerical Analysis of Laminar Natural Convection in a
Confined Fluid Heated from Below. J. Heat Transfer, Trans. ASME
1976, 98C (202), 519.
(84) Ozoe, H.; Yamamoto, K.; Churchill, S. W. Three-Dimensional
Numerical Analysis of Natural Convection in an Inclined Channel with
a Square Cross-Section. AIChE J. 1979, 25, 709.
(85) Ozoe, H.; Sato, N.; Churchill, S. W. Experimental Confirmation
of Computed, Three Dimensional Streaklines for Natural Convection
in an Inclined Rectangular Enclosure Computed. Kagaku Kogaku
Ronbunshu 1979, 5, 19 (English transl., Int. Chem. Eng. 1979, 19, 454).
(86) Yamamoto, K.; Hiroyuki, O.; Chao, P. K.-B.; Churchill, S. W.
The Computation and Dynamic Display of Three-Dimensional
256

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Industrial & Engineering Chemistry Research

Article

translation: The Principles of Turbulent Heat Transfer, National


Advisory Committee for Aeronautics, Technical Memorandum TM 1408,
Washington, DC, 1957).
(115) Churchill, S. W.; Shinoda, M.; Arai, N. A New Concept of
Correlation for Turbulent Convection. Thermal Sci. Eng. 2000, 8 (4),
49.
(116) Bo, Y. U.; Ozoe, H.; Churchill, S. W. The Characteristics of
Fully Developed Turbulent Convection in a Round Tube. Chem. Eng.
Sci. 2001, 56, 1781.
(117) Churchill, S. W.; Zajic, S. C. Prediction of Fully Developed
Convection with Minimal Explicit Empiricism. AIChE J. 2002, 48, 927.
(118) Churchill, S. W. A Reinterpretation of the Turbulent Prandtl
Number. Ind. Eng. Chem. Res. 2002, 41, 6393.
(119) Abbrecht, P. H.; Churchill, S. W. The Thermal Entrance
Region in Fully Developed Turbulent Flow. AIChE J. 1960, 6, 268.
(120) Chilton, T. H.; Colburn, A. P. Mass Transfer (Absorption)
Coefficients. Ind. Eng. Chem. 1934, 26, 1183.
(121) Shaw, D. A.; Hanratty, T. J. Turbulent Mass Transfer to a Wall
for Large Schmidt Numbers. AIChE J. 1978, 23, 28.
(122) Papavassiliou, D. V.; Hanratty, T. J. Transport of a Passive
Scalar in a Turbulent Channel Flow. Int. J. Heat Mass Transfer 1997,
40, 1303.
(123) Rayner-Canham, M. F.: Rayner-Canham, G. W. A Devotion to
Their SciencePioneer Women in Radioactivity; Chemical Heritage
Foundation Press: Philadelphia, PA, 1997.
(124) Churchill, S. W. The Thermal Conductivity of Dispersions and
Packed BedsAn Illustration of the Unexploited Potential of Limiting
Solutions for Correlation. In Advances in Transport Processes, Vol. IV;
Mujumdar, A. S., Mashelkar, R. A., Eds.; Wiley Eastern: New Delhi,
1986; pp 274293.
(125) Rothfus, R. R.; Monrad, C. C. Correlation of Turbulent
Velocities for Tubes and Parallel Plates. Ind. Eng. Chem. 1955, 47,
1144.
(126) MacLeod, A. C. Liquid turbulence in a gas-liquid absorption
system. Ph.D. Thesis, Carnegie Mellon University, Pittsburgh, PA,
1951.
(127) Whan, G. A.; Rothfus, R. R. Characteristics of Transition Flow
between Parallel Plates. AIChE J. 1959, 5, 204.
(128) Emmons, H. W. Natural Convection Heat Transfer
Correlation. In Studies in Mathematics and Mechanics Presented to
Richard von Mises by Friends, Colleague and Pupils; Academic Press:
New York, 1954.
(129) Bird, R. B.: Stewart, W. E.: E. N. Lightfoot. Transport
Phenomena; John Wiley & Sons: New York, 1960.
(130) Martin, H. The Generalized Leveque Equation and Its Practical
Use to Predict Heat and Mass Transfer from the Pressure Drop. Chem.
Eng. Sci. 2002, 57, 3217.
(131) Martin, H. The Leveque Analogy or How to Predict Heat and
Mass Transfer from Friction. In Proceedings of the 4th International
Conference on Heat Transfer, Fluid Mechanics and Thermodynamics,
Cairo, Egypt 2005; Paper K2.
(132) Churchill, S. W. An Analogy between Reaction and Heat
Transfer. AIChE J. 2006, 52, 3645.
(133) Yu, Bo.; Churchill, S. W. The Extraordinary Effects of
Energetic Reactions on the Nusselt Number in a Round Tube with a
Uniform Heat Flux Density on the Wall. Int. J. Heat Mass Transfer
2008, 51, 2990.
(134) White, R. R.; Churchill, S. W. Experimental Foundations of
Chemical Engineering. AIChE J. 1959, 5, 353.
(135) Cleland, F. A.; Wilhelm, R. H. Diffusion and Reaction in
Viscous-Flow Tubular Reactor. AIChE J. 1956, 2, 48.
(136) Churchill, S. W.; Yu, Bo. Effects of Transport on Reaction in
Homogeneous Tubular Flow. Ind. Eng. Chem. Res. 2006, 45, 8583.
(137) Churchill, S. W. Laminar Film Condensation. Int. J. Heat Mass
Transfer 1986, 29, 1219.

Streaklines for Natural Convection in Enclosures. Comput. Chem. Eng.


1982, 6, 161.
(87) Saville, D. A.; Churchill, S. W. Convection in Boundary Layers
Near Horizontal Cylinders and Vertical Axisymmetric Bodies. J. Fluid
Mech. 1967, 29, 39.
(88) Saville, D. A.; Churchill, S. W. Simultaneous Heat and Mass
Transfer in Free Convection Boundary Layers. AIChE J. 1970, 16, 268.
(89) Boussinesq, J.: Theorie analytique de la chaleur, mise en
harmonique avec la Thermodynamique ey avec la Theorie de la lumiere,
Vol. II; Gauthier-Villars: Paris, 1903.
(90) Galante, S. R.; Churchill, S. W. The Applicability of Solutions
for Convection in Potential Flow. Adv. Heat Transfer 1990, 20, 353.
(91) von Karman, Th. Laminare und turbulenten Reibung. Z. Angew.
Math. Mech. 1921, 1, 233.
(92) Pohlhausen, K. Zur naherungsweisen Integration der Differentialgleichen denr laminaren Reibungschicht. Z. Angew. Math. Mech.
1921, 1, 252.
(93) Churchill, S. W. New Simplified Models and Formulations for
Turbulent Flow and Convection. AIChE J. 1997, 42, 1125.
(94) Churchill, S. W.; Balzhiser, R. E. The Radial Heat Flux. Chem.
Eng. Prog. Symp. Ser. 1959, 55 (29), 127.
(95) Heng, Ly; Christina, Chan; Churchill, S. W. Essentially Exact
Characteristics of Turbulent Convection in a Round Tube. Chem. Eng.
J. 1998, 71, 163.
(96) Leveque, J. Le lois de la transmission de la chaleur par
convection. Ann. Mines 1928, 13, 201, 305, 381.
(97) Nusselt, W. Die Oberflachenkondensation des Wasserdamphes.
VDI Z. 1916, 60 (541), 569.
(98) Churchill, S. W. Laminar Film Condensation. Int. J. Heat Mass
Transfer 1986, 29, 1219.
(99) Churchill, S. W. The State of the Art of Education in Reaction
Engineering. Ind. Eng. Chem. Res. 2011, 50, 8806.
(100) Churchill, S. W.; Pfefferle, L. D. The Refractory Tube Burner
as an Ideal Stationary Chemical Reactor. Ind. Chem. Eng. Symp. Ser.
1984, No. 87, 279.
(101) Watkin, S.: Mihalik, A.: Churchill, S. W.: Godshalk, J. B.
Chemical Guidelines for Safety in High-Rise Buildings based on
Chemical Process Principles. In The 2007 AIChE Annual Meeting
Conference Proceedings on CD-ROM.
(102) Christiansen, J. A. On the Reaction between Hydrogen and
Bromine. Mat.-Fys. Medd-Kgl Danske Vidensk. Selskab. 1919, 1, 1.
(103) Pfefferle, L. D.; Churchill, S. W. NOx Production from the
Combustion of Ethane Doped with Ammonia in a Thermally
Stabilized Burner. Combust. Sci. Technol. 1986, 49, 235.
(104) Churchill., S. W. Critique of the Classical Analogies for Heat,
Mass, and Momentum Transfer. Ind. Eng. Chem. Res. 1997, 36, 3866.
(105) Reynolds, O. On the Extent and Action of the Heating Surface
of Steam Boilers. Proc. Lit. Soc. Manchester 1874, 14, 7.
(106) Prandtl, L. Eine Beziehung zwischen Warmeaustausch und
Stromungswiderstand der Flussigkeiten. Phys. Z. 1910, 11, 1072.
(107) Thomas, L. C.; Fan, L. T. Heat and Momentum Transfer
Analogy for Incompressible Boundary Layer Flow. Int. J. Heat Mass
Transfer 1971, 14, 715.
(108) Churchill, S. W. Comprehensive Correlating Equations for
Heat, Mass and Momentum Transfer in Fully Developed Flow in
Smooth Tubes. Ind. Eng. Chem. Fundam. 1977, 16, 109.
(109) Colburn, A.P. A Method of Correlating Forced Convection
Heat Transfer Data and a Comparison with Fluid Friction, Trans. Am.
Inst. Chem. Eng. 1933, 29, 174, 219, 220, 232, 249.
(110) Churchill, S. W.; Zajic, S. C. Prediction of Fully Developed
Convection with Minimal Explicit Empiricism. AIChE J. 2002, 48, 927.
(111) Drew, T. B.; Koo, L. C.; McAdams, W. H. The Friction Factor
for Clean Round Pipes. Trans. Am. Inst. Chem. Eng. 1932, 28, 174.
(112) Lees, C. H. On the Flow of Viscous Fluids through Smooth
Circular Pipes. Proc. R. Soc. London 1915, 91, (623, Nov. 2), 46.
(113) McAdams, W. H. Heat Transmission; McGrawHill: New
York, 1933; p 86.
(114) Reichardt, H. Die Grundlagen des turbulenten Warmeuberganges. Arch. ges. Warmetechn. 1951, No. No. 6/7, 129 (English
257

dx.doi.org/10.1021/ie300773e | Ind. Eng. Chem. Res. 2013, 52, 230257

Vous aimerez peut-être aussi