Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Laws and Models: Science, Engineering and Technology: Hall, Carl W. "Frontmatter" Boca Raton: CRC Press LLC, 2000

Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Hall, Carl W.

"Frontmatter"
Laws and Models: Science, Engineering and Technology
Boca Raton: CRC Press LLC,2000
PREFACE
As distinguished from the laws of society, physical laws have been developed to describe or
represent the response of a subject (person, elemental particle, earth, moon, stars) in a certain
natural (chemical, biological, physical) or a controlled environment, and expressed in such
a way that they can be used to predict the response. The law that is general or covers a wide
variety of situations is useful in describing the parameters of less general and more specific
conditions. Laws are identified according to the phenomena and/or the person(s) who first
stated them, or for whom the relationships were named. Where possible, this book identifies
the statement of the law with the name(s) of the person(s) expressing the law, or to person(s)
whom colleagues wish to honor. It is not a simple matter to identify the person after whom
a law is named. Which Bernoulli? Which Curie? Was it Wentzel or Wenzel? Was it Fisher
or Fischer? And when two or more names are used to identify a law—was one person a
student of the other? Often, little biographical information can be found on these early
investigators after whom laws are named, unless they continued in the field. In the early
years, there was a penchant for secrecy of one’s findings, which might become known only
after several years, because they appear in a foreign-language journal or are included the
scientific notes of a deceased person. Sometimes, the earlier persons identifying a phenom-
enon or relationship are recognized by having a law named after them, and sometimes not.
It is not unusual for a law to be named after the person who publicized it rather than the
individual who expressed it earlier.
My collection of laws began 40 years ago. Laws were commonly espoused in mathematics,
physics, and chemistry books, and I found them intriguing and useful. I began to see laws
as primarily developed by scientists as a method of describing the results of their work. Many
of these laws, such as those of Kepler, Newton, Bernoulli, etc., are fundamental to scientists
and engineers. However, many relationships of the natural world with which the engineer
worked did not fit the accepted definition of laws identified by scientists at the time. There
began to appear in the literature very useful relationships in the physical and chemical worlds,
and belatedly in the biological world, described as models, identified by some writers as laws,
that were not as general as the previous concept of laws. A special effort has been made in
this book to include laws and models pertaining to medical, physiological, and psychological
laws in fields that are becoming important in engineering. In the literature, one can find
numerous examples of laws described as axioms, canons, effects, principles, rules, theorems,
or theories, where these can stand alone. It is interesting to note that a law is often described
in terms of a principle, a theorem, or an effect. Those not incorporated under laws have been
identified as models. One can think of the world being represented by laws as described by
scientists and models as used by others, particularly engineers and technologists. It is common
that a particular relationship is called a law by one author and a rule, or some other term, by
another author. I have included these various designations (axiom, effect, postulate, principle,
rule, theorem, etc.), in addition to laws as represented in the literature. No attempt has been
made to adjudicate which designation is most appropriate. Dimensionless groups, such as
named after Reynolds or Mach, and called numbers, form a very important category of
models, of which at least 100 have been included. These numbers form an important category
of relating variables. In this collection, I have identified as models those scientific relationships
not given the designation of laws. The laws and models are listed in strict alphabetical order,
omitting apostrophes, brackets, commas, and dashes.
Empirical laws are frequently used to describe relationships, in contrast to theoretical laws.
Empirical laws, such as the laws of Kepler, are frequently explained by theoretical laws but

©2000 CRC Press LLC


are only approximately true or accurate and refer to directly observable objects and properties.
Theoretical laws, such as the Avogadro law, at least until the present, have not been open to
direct observation and measurement.
A distinction needs to be made between universal and statistical laws. A universal law
asserts a regularity that holds without exception, such as the law of conservation of charge
(the amount of electric charge in an isolated system is always preserved). Statistical laws
state that something happens in a certain percentage of cases—such as tossing a coin for
heads or tails. Many of the more recent laws, such as in quantum mechanics, are based on
statistical relationships. Statistical relationships are often difficult to express without mathe-
matical relationships that appear complicated. For this compilation, these are written in
general terms to convey the essence of the law rather than with detailed mathematical
expressions, which can be found in the sources cited.
Models are often used to represent variables in a relationship that otherwise cannot be
manipulated, or can be manipulated only with great difficulty or major expense. Thus, fields
of endeavor dealing with the atmosphere or geological formations or large complicated
systems can be modeled. The result could be that a model, perhaps due to its applicability,
at a later date earns enough acceptance of people in the field to be called a law. No attempt
has been made to distinguish among law, rule, theorem, model, etc., except as used or cited
by people in the field. Brief biographical information of the persons after whom the law or
model is named is provided, and sources of information are provided for the scholar who
wishes to pursue the subject. Cross-references help guide the reader to related laws and
models. An attempt has been made to minimize extremely complicated mathematical equa-
tions so that the reader can gain the sense of the law or model without knowing all the details.
Extensive sources are listed to assist the interested reader in learning more about a law or
model. In some cases, an entire book can be written on a specific law or model, so to expect
the subject to be completely covered here would be unrealistic.
With respect to mechanics or style, some decisions were made to streamline the manuscript,
and to make the presentations user friendly. The apostrophe was omitted from most of the
laws. It is a usual practice to omit ’s from names of laws named after two or more people,
so why not do the same for one person? This practice also made alphabetical ordering more
sensible and saved space. Diligence was exerted in attempting to find biographical data about
each person after whom or for whom a law or model was named. Perhaps a former graduate
student or laboratory associate was involved with the work but did not continue in the field
and was lost in the biographical data. Some of these may be recorded in foreign-language
biographies, and even though searched, several names could not be located. Some laws are
named after two or more people, unknown to each other, who separately found the same
relationship—the first may have originally expounded the law or model, and the second found
the same relationship or additional information to verify the original theory.
One is impressed by the breadth of activities of early scientists, such as von Baer; the
family connections, such as Bernoulli and the Curies; and the long time some early scientists
took to publish their work and the secrecy with which some people worked, thus denying
the original discoverers their rightful credit.

©2000 CRC Press LLC


LAWS AND MODELS:
AN INTRODUCTION
George A. Hazelrigg

From the dawn of recorded history, men and women have sought to manipulate their envi-
ronment in pursuit of a more comfortable, fulfilling, and safer life. The deliberate manipu-
lation or utilization of nature through the use of energy sources and natural resources, with
intended benefit, is engineering. The ancient stone nyrage of Sardina, which remain standing
today, attest that such feats of engineering date back over 4000 years. With time, our feats
have grown in both number and complexity, and today the process of engineering is conducted
by some 2 million engineers in the U.S.A. alone.
All feats of engineering, whether stone dwellings or space stations, require a particular
sequence of events. First, the engineer must understand the needs and wants of the society
or subgroup of society that is to be served. Second, the engineer must formulate concepts of
potential designs that might serve the designated needs and wants. Third, the engineer must
analyze the concepts to determine their functionality. Fourth, the engineer must optimize
selected candidate designs and choose a single preferred design. And fifth, the engineer must
design a production system to realize the selected design.
More recently, engineering does not stop even here but continues through the life cycle
of a design, even including disposal at the end of the product’s life. Throughout this process,
the goal of the engineer is to find a design, a production system, and a context for use of
designed products, processes, and services that can effectively and efficiently realize the goals
imposed by the underlying needs and wants of society. And to do this requires that the
engineer be capable of predicting the future—the future wants and needs of society, the future
context within which products will be used, and the performance of the as-yet-undesigned
products themselves. The engineer must, within reasonable bounds, be able to determine how
every proposed design would perform were it selected, and from this make rational choices
among design alternatives. This is quite a challenge and leads us into the forest of modeling.
In performing the design process, all available knowledge is drawn upon so as to evaluate
the various parameters of a functioning process or design—as might be represented by
mathematical, physical, biological, and chemical relationships that are often represented by
laws or models. At the same time, additional knowledge is sought, as needed or as perceived
as needed, to satisfy the design process.

A SYSTEM OF LOGIC
Beginning as early as the ancient Greeks, man sought to create a system of logic within
which better understanding and predictions could be facilitated. This system has become
known as mathematics. But, even to this day, we do not have a clear understanding of what
mathematics is and why it is such a powerful tool. Barrow (1991) proposes four views of
mathematics that are quite insightful.
The first view he calls formalism. This view defines mathematics to be “nothing more or
less than the set of all possible deductions from all possible sets of consistent axioms using
all possible rules of inference.” In this formalism, truth is defined as a statement that is
consistent with a set of axioms and the web of logical connections they imply. The ability

©2000 CRC Press LLC


of mathematics to describe or predict nature is purely coincidental, and we get little insight
as to why mathematics works.
The second view is inventionism. This view “regards mathematics as a purely human
invention…, it is a product of the human mind. We invent it, we use it, but we do not discover
it.” This view leads, indeed, to a concept of why mathematics works. In the evolution of the
species, those species whose brains could distinguish between past and future, and between
cause and effect, and that could make predictions in concert with nature, were more likely
to survive and reinforce the gene pool. Thus, in this view, natural selection led to a set of
species for which logic corresponds well with nature. And man is the culmination of this
evolution of logic.
The third view is the Platonic view. In this view, mathematics exists and man discovers
it. Indeed, the Platonic view holds that mathematics existed before the universe was created
and will exist after it ends. The Platonic view holds mathematics in much the same regard
as God. The physical world is inherently mathematical, and thus mathematical theories
describe natural phenomena. A logical consequence of this view is that all physical processes
can be modeled mathematically, and hence the universe could be modeled on a computer.
Furthermore, it would be impossible to distinguish between the simulation and the physical
universe.
The fourth view is called constructivism. This view was created at the end of the nineteenth
century as a result of the logical paradoxes of modern set theory and the strange Cantorian
properties of infinite sets. In this view, mathematics is defined to “include only those state-
ments that could be deduced in a finite sequence of step-by-step constructions starting from
the natural numbers, which were assumed as God-given and fundamental.” This view of
mathematics removes such familiar procedures as argument by contradiction, and thus reduces
the content of mathematics considerably.

PREDICTIONS
Today, in all logical activities, we predict the future using models, either explicitly or
implicitly. And all predictive models have their basis in mathematics. Indeed, it is necessary
that such be the case in order that resulting predictions be logically consistent with our
knowledge. So, we must ask, what is a model? For one thing, a model is an abstraction of
reality. Nature is simply too complex to understand in toto, and so we abstract from nature’s
reality those elements that are important to any given circumstance. A model then provides
a logical connection between the elements that we include in our abstraction. Thus, a model
is also a relationship between elements that clearly implies correlation, but it may also infer
cause and effect. It is only through models, and especially inferences of cause and effect,
that we gain an understanding of nature. And we validate our understanding by using models
of past phenomena to predict future events.
In engineering, we use models to combine disparate elements of knowledge and data to
make accurate predictions of future events. For example, an engineer might want to know
whether a bridge of a particular design will withstand a given load. It is, of course, possible
to guess at the answer directly. But models give us answers in which we have greater
confidence. They do this in a two-step process. First, a model enables us to disaggregate a
guess from a single quantity—for example, true or false, the bridge will stand—into a set of
parameters, each of which we must also guess, but for which we can make more precise
estimates. Then, in a second step, the model provides the framework within which we can
unify the data logically and consistently to obtain the desired result with, hopefully, less
uncertainty than we would obtain from a direct guess. For example, if we want to know the

©2000 CRC Press LLC


acceleration of a particle, the model a = F/m enables us to guess at F and m rather than a.
If it is possible to obtain accurate estimates of these parameters, we can use the model to
unify them into the desired result, namely an accurate estimate of a. In this view, the value
of a model lies in its ability to reduce uncertainty in the estimation of a variable.
This concept is illustrated in the figure below. One’s ability to estimate the parameter of
interest can be represented by a probability density function (pdf), or simply probability
function, on the original guess. The model disaggregates this guess into a set of parameters,
each of which in turn must be estimated. The idea is that it should be possible to estimate
each of these parameters with much less variance, that is, with much less peaked pdfs. Then
the model enables these estimates to be combined with a resulting estimate of the parameter
of interest that has a more peaked pdf.

A conceptual view of a model. Reprinted from Hazelrigg, G.A., Systems Engineering: An Approach to
Information-Based Design, Prentice Hall, 1996, used with permission.

FUNDAMENTAL ASSUMPTIONS IN MODELING


So far, we have identified that models provide understanding and that they enable improved
estimation. But what do models themselves depend on? It is important to recognize that
inherent in all models of physical phenomena is the assumption that the fundamental rules
of nature, which we call laws, are fixed across all space and time. It is our belief that, whenever
we obtain results that differ from those predicted by a model, the reason is (1) the model is
incorrect, (2) we have not included a sufficient breadth of elements in the model, (3) in
conducting the experiment, we did not isolate the system sufficiently from exogenous cor-
rupting factors, or (4) we erred in collecting the data or in computing the results. We rarely
challenge the underlying assumption of invariance of natural laws. Indeed, all physics is
based on the assumption of invariance, and the validation of all physical theories uses this
assumption; without it, rational prediction would be impossible. But we continually challenge

©2000 CRC Press LLC


the validity of statements that portend to represent the natural laws, and we validate physical
theories by repeated experimentation in temporally and spatially disparate cases.
The need to validate theories and their resulting models by spatially and temporally
disparate experiments imposes stringent conditions on new theories. In short, it says that they
must be connected to extant theories—they must explain physical phenomena within the
context of previous explanations. The conduct of a valid experiment demands that enough
be known about the environment that the experimenter be capable of isolating the experiment
from corrupting factors. For example, consider the demonstration that gravity acts on rocks
and feathers in the same way to produce acceleration. It was long felt that gravity acted
differently on different substances. But the understanding that air creates drag, which retards
the acceleration of the feather more than of the rock, led to the conduct of an experiment in
a vacuum. The corrupting factor was removed, and the experiment produced results that
confirmed the hypothesis. This understanding of corrupting factors is adequate only when
experiments are conducted on the margins of extant knowledge. As a result, theories tend to
build upon one another.

UNITY ACROSS DISPARATE EVENTS


One goal of science, and indeed that of models, is to unify seemingly disparate events. This
is a noble goal, as it extends our predictive capabilities. Note, for example, how models
derived from Newton’s Laws unify events that span the universe in both scale and distance.
They provide that the motion of a feather floating on a pond and the motion of the planets
around the sun and, indeed, the motion of the galaxies are all governed by the same concept,
namely that the time rate of change of momentum of a body is proportional to the net force
exerted upon it. This notion of unification is extremely powerful. Not only does it broaden
our predictive capability well beyond the bounds of experience, it also provides a venue for
validation of the theory itself.
Models derive from two fundamentally different processes: first, from observation, but
also they derive from pure thought—mathematics. Our life process exposes us to many
phenomena: sunrise, sunset, warmth, cold, rain, sun, wind, lightning, and life itself. The mere
act of observing these natural phenomena in a methodical and careful manner provides us
with data from which we can formulate models. But our observations themselves are couched
in terms that can be thought of as models, some of which are so fundamental and well
accepted that they are recognizable as laws. For example, we speak of warm only in com-
parison to cold or hot. We imply a model of temperature. As we move to more sophisticated
measurements, models become indispensable. The measurement of a voltage implies the
existence of a model that defines voltage. At some level, all measurements—that is, all
gathering of numerical data—imply the existence of models, and the data themselves have
absolutely no meaning in the absence of their associated models. Thus, measurement or data-
taking and model building are intrinsically linked and utterly inseparable.
Yet, data themselves are not the basis from which some of the most important laws of
nature have been formulated. Consider Einstein’s formulation of the theory of relativity. A
postulated model of the nineteenth century was that the solar system was moving through
an ether. This model held that the speed of light should differ according to the direction in
which it travels. Within the context of this model, Michelson and Morley conducted their
famous measurements of the speed of light and determined that their hypothesis was incorrect.
Indeed, it appeared, within their ability to measure, that the speed of light is the same,
independent of direction. Einstein internalized this notion, accepting it as fact. But it is very
important to recognize that it was not established fact. He then went on to ask, what are the

©2000 CRC Press LLC


logical conclusions resulting from the assumption that the speed of light is independent of
direction? The theory of relativity is consistent with this assumption and the axioms of
mathematics. Thus, the theory emerged from pure, that is, mathematical thought.

THEORY AND DATA, DATA AND THEORY


A theory that derives from pure thought cannot stand, however, without the test of physical
data. And so such data must be obtained through the theory itself. A new theory makes
predictions that can be measured. Recognizing this, we can develop experiments to generate
the desired data. But there is often a complication. New theories tend to emerge on the margin,
near the limits of our ability to make measurements. This is because large discrepancies from
extant models are easily detected, and the faulty models can often be discarded and replaced
with more appropriate models. Thus, at least at the most fundamental levels, that is, at the
level of the fundamental laws of nature, the confirmation of new models demands measure-
ment at the limits of resolution with existing instruments. And so the advance of science with
the development of new theories is often retarded by the technology of scientific instruments.
But this is not always the case. Often, new theories emerge around new instruments, and
the technology of the instruments leads the science. A good example of this is the telescope,
which changed man’s interpretation of the universe, or the microscope, which led to the
microbial theory of disease, both by direct observation. New instruments often enable new
measurements, which lead to unexpected results. An example of this is the invention by Volta
of the voltaic cell. This cell enabled the separation of the elements and rapidly led to the
discovery of the elements. From this discovery came a whole new collection of models
governing chemistry.
In all cases, however, we see that the process of discovery is a process that proceeds on
the margins of extant knowledge. We build models on data and collect data on models. We
build instruments using models and interpret the readings that instruments give based on
models. And the most fundamental of these models we call laws.
And such is the subject of this book. Over the past 3,000 years, development and appli-
cation of physical laws has been a mainstay of the sciences, particularly the physical sciences,
and engineering. Beginning with Pythagoras’ Law of Strings, we have amassed a rich col-
lection of laws that govern and predict nature in almost all currently observable forms. This
book presents that collection in a unique format. Each law and model is given, along with
date credited, where available; keywords; authors or persons recognized, with a brief bio-
graphical sketch; sources of information; and cross-references.

©2000 CRC Press LLC


AUTHOR
Carl W. Hall, Ph. D., P.E., is a consultant, having retired in 1990 as deputy assistant
director, engineering, of the National Science Foundation. After three years of infantry service
in the U.S. Army, including the awards of CIB and the bronze star, in the ETO during WWII,
he resumed his academic studies. He earned his bachelor of science (summa cum laude)
degree in agricultural engineering at Ohio State University in 1948; his master of mechanical
engineering at the University of Delaware in 1950; and his doctorate at Michigan State
University in 1952. He taught at those institutions and at Washington State University, where
he was dean of engineering and professor of mechanical engineering, from 1970 to 1982.
He participated in the SMG program at Harvard University.
Dr. Hall has held numerous consulting and professional assignments with industry, foun-
dations, universities, the United Nations, and the World Bank. He has served as a national
officer and is a fellow of ABET, ASAE, ASME, and chaired the EAC (Engineering Accred-
itation Commission). He has also been active in NSPE, ASEE, AIBS, is a fellow of AAAS
and AIMBE, and has won numerous national and international awards. He is a member of
the National Academy of Engineering.
He is author or editor/coeditor of 28 books and has written more than 400 articles and
papers. His most recent book is The Age of Synthesis. His specialty areas are heat and mass
transfer as related to drying and energy.

©2000 CRC Press LLC


ACKNOWLEDGMENTS
Thanks are expressed to the following for their advice and assistance in the preparation
of the book: G. I. Barenblatt, Arthur Bergles, Clayton Crowe, Claudia Genuardi, Mildred
Hall, George Hazelrigg, Karla Kalasz (illustrations), A.S. Mujumdar, Edward A. Murphy III,
and Bob Ross, and many of the public and institutional libraries in the U.S., especially the
Arlington Public Library, George Washington University Library, and the Library of Congress.
Carl W. Hall

©2000 CRC Press LLC


DISCLAIMER
Information in this book has been obtained from numerous sources considered to be
reliable. However, the editor and publisher cannot guarantee the accuracy or completeness
of information. Due to the nature of the book, users of the information may need to go to
one or more of the sources of information cited in order to appropriately use it, depending
on the application under consideration. The editor and publisher cannot be responsible for
any errors, omissions, or damages arising out of the use of this information. The book is
provided as a means of supplying information without rendering engineering or other pro-
fessional services. An appropriate professional should use the information for the design,
construction, maintenance, and management of elements or systems.

©2000 CRC Press LLC


POSSIBLE FIRST
RECORDED PHYSICAL LAW
The first physical law expressed in mathematical terms could well be the Pythagorean Law
of Strings:

Pythagoras, c. 582–c. 500 B.C., Greek philosopher and mathematician, stated that
harmonic sounds were given by strings whose lengths were in simple numerical
ratios, such as 2:1, 3:2, 4:3, etc.
Gamow, G., 1961, 1988
Porter, R., 1994

©2000 CRC Press LLC


DEFINITIONS AND
DESCRIPTIONS
Laws and models can be expressed in a number of ways, as illustrated by the following
identifiers. An effect or theorem identified by one author could be called a law by another
author. A well known example is the Bernoulli theorem, called the law of large numbers.
Usage often determines the identification of the statement. For example, Newland Relation-
ship became known as a law; Delta function is not a function; Hamilton Principle is not a
principle; and many laws are axioms, such as the Laws of Conservation of Energy. Those
relationships not called laws are categorized as models in this book. Some examples of models
are presented here.

APHORISM—A concise statement of a principle, truth, or opinion; a maxim; an adage.


ASSERTION—A belief or an idea; subject to proof or verification. Mayer Assertion
(regarding energy).
AXIOM, MAXIM—Self-evident or accepted as self-evident; Newton’s first two laws
originated as axioms; axiom of choice in statistics; a basic proposition in a formal
system presented without proof. Wolff described axioms as indemonstrable, theo-
retical, and universal, thus belonging to the synthetic. See POSTULATE. Design
A.; Grinnell A.; Grice M.; Peano’s A.; Playfair A.; Rheology A.
CANON—General rule or generally accepted basic principle. Molar Extinction C.;
Morgan (Lloyd) C.
CONCEPT—An idea. Quark C.
CONJECTURE—To infer from incomplete information or evidence. Kepler C.
CONVENTION—A belief widely accepted by long usage.
COROLLARIES—Propositions that follow with little proof from another already
proven.
CRITERION—A principle against which a proposition can be treated or tested. Chau-
venet C.; Lawson C. (for fusion); Routh C.
CURVE—A line that represents an equation or relationship of variables. Linear Growth
C.; Lorenz C.; Pareto C.; Wunderlich C.
DIMENSIONLESS NUMBER or GROUP—A ratio of two or more numbers with
units (dimensions), taken with respect to physically significant units, that provide a
meaningful number without units.
DOCTRINE, DOGMA—Idea put forth as true, not necessarily supported by evidence.
DUALISM—A view that, in some areas of concern, there are two fundamental entities,
such as mind and body, good and evil, true and false, and plus and minus. Analysis
and synthesis can be considered a dualistic approach.
EFFECT—A response to a stimulus. Abney E.; Aschoff E.; Auger E.; Barkhausen E.;
Bauschinger E.; Becquerel E.; Bohr E.; Compton E.; DeHaas-Van Alphen E.; Debye
E.; Doppler E.; Edison E.; Ettingshausen E.; Fenn E.; Forbush E.; Gudden-Pohl E.;
Gunn E. (electron); Hall E.; Joule-Thomson E.; Kerr E.; Kelvin E.; Leduc E.; Magnus
E.; Majorama E.; Mass-Law E.; Meissner E.; Mossbauer E.; Paschen-Black E.;
Peltier E.; Photoelectric E.; Piezoelectric E.; Piobert E.; Purdy E.; Purkinje E.;
Seebeck E.; Sewall Wright E.; Skin E.; Stark E; Thomson E.; Voigt E.; Zeeman E.
EQUATION—A statement of equality between two expressions. Airy E.; Arrhenius
E.; Clapeyron E.(rate reaction); Bragg E.; Deutsch-Anderson E.; Dittus-Boelter E.;
Gibbs-Helmholtz E.; Gibbs-Thomson E.; Hamilton E. (position of a body); Lange-

©2000 CRC Press LLC


Hicks E.; Lorenz Force E.; Navier-Stokes E.; Nernst E.; Nutting-Scott Blair E.;
Quadratic E.; Stark-Einstein E.; Van der Pol E.; Van der Waals E.
FORMULA, FORMULATION—A general principle stated in mathematical termi-
nology. Balmer Series F.; Cauchy F.; Clausius F.; Leibnitz F.; Nyquist F.; Planck F.;
Shannon F.
FUNCTION—A mathematical relationship. Dirac Delta F.; Gibb F.; Goedel’s F. (a
parallel to theorem).
HYPOTHESIS—A statement of conjecture subject to proof. A hypothesis is that
inertial forces arise from the gravitational interaction between the accelerated body
and mass of the distant stars. Avogadro H.; Gaia H.; Helmholtz-Lamb H.; Starling H.
INEQUALITY—A statement of inequality between expressions. Camp-Meidell I. is
called a law by E. B. Wilson.
LAW—A general principle that represents all cases where properly applied. DeMorgan
L.; Einstein L.; First L. of Thermodynamics, et al; Ohm L.
MAXIM—See AXIOM
METHOD—A procedure or order to be followed. Runge-Kutta M.
MODEL—Tentative representation for testing a phenomenon (structure, equation,
relationship, etc.). Burger M.; Charney M. (of Navier-Stokes Equation); Covering-
Law M.; Dirac M.; Linear M.; Maxwell M.; Newton-Kelvin M.; Sommerfeld M.
(metal); Turing M.; Viscoelastic M.
NUMBER—See DIMENSIONLESS NUMBER or GROUP.
PARADOX—A true statement that appears to be contrary to other statements or
opinions. Russell P.
PARAMETER—An arbitrary constant or equation or other expression that can be
used for comparison. Coriolis P.; Hydrodynamic P.; Magnetic force P.
PHENOMENON—An observable fact or event. Bezold-Brucke Ph.; Gunn Ph.; Loeb
Ph.; Magnetoelastic Ph.; Relaxation Ph.; X-ray Ph.
POSTULATE—An observation or projection of a relationship based on insight or
intuition that becomes a hypothesis if proved to be true; often used synonymously
with axiom; but Wolff described postulates as demonstrable, practical and particular,
thus belonging to the analytic. Avogadro P.; Hill P.; Koch P. (bacteriology).
PRINCIPLE—A basic truth, law, or assumption. Babinet P.; Bernoulli P.; Berthelot-
Thomsen P.; Blackman P.; Boltzmann Ionic P.; D’Alembert P.; Exclusion P. (Pauli);
Franck-Condon P.; Gilbert P.; Hardy-Weinberg P.; Homeopathy P.; Huygen P.;
Lagrange P.; Least Action P.; Le Chatelier P.; Limiting Factor P.; Lister P.; Maupertius
P.; Quantum P.; Relativity P.; Saint-Venant P.; Uncertainty P. (Heisenberg); Van’t
Hoff P.; Venturi P.
PROBLEM—A question proposed for examination or solution. First Digit P. (Ben-
ford).
PROOF—Expressions dealing with mathematical symbolic logic. Goedel P.
PROPOSITION—A statement proposed to represent a relationship. See RELATION-
SHIP; STATEMENT.
RELATION(SHIP)—A thought advanced as true. Bohr R.; Gibb R.; Koschmieder R.;
Newland R. (that became known as the Law of Octaves); Ostwald-Freundlich R.
RULE—Standard procedure for solving a class of problems. Abegg R.; Antonow R.;
Area, R. for; Bergmann R.; Blanc R.; Blondel R.; Constant Proportions, R. of;
Ecogeographic R.; Fleming R.; Geiger and Nuttall R.; Gibb Phase R.; Grimm-
Sommerfeld R.; Guldberg R.; Fajans R.; Jordan R.; Hahnemann R.; Helmholtz-
Thomson R.; Hesse R.; Hildebrand R.; Kramer R.; Lewis-Randall R.; Signs, R. of

©2000 CRC Press LLC


(algebraic); Matthiessen R.; Maxwell R.; Richards R.: Simpson R.; Species R.;
Thomson-Nernst R.; Trapezoidal R.; Weddle R.; Young R.
STATEMENT—May be a definition or relationship. Mach S.
TAUTOLOGY—A compound proposition that is true regardless of values assigned,
such as “A or not A.”
THEOREM—An idea demonstrably true or proven proposition; a well formed formula
in a given logic system. Bayes T.; Chebyshev T., called the Law of Large Numbers
T.; Earnshaw T.; Goedel (Godel) T.; Lami T.; Last T. (Fermat); Pythagoras T.;
Roberts-Chebyshev T.
THEORY—System of circumstances leading to explanation of a phenomenon. Arrhe-
nius-Ostwald T.; Bar T.; Bayes-Laplace T.; Bernoulli Kinetic T.; Buckingham T.;
Gravitation, T. of; Helmholtz Accommodation T. (eye); Helmholtz T.(color); Lewis
T.; Recapitulation T.; Ptolemy T.; Quantum T.; Relativity T. (Einstein); Usanovich
T.; Young-Helmholtz T.
THESIS—A proposition forming ideas into expressible form to be communicated to
others, usually written but can be spoken or subject of a composition. Church T.

©2000 CRC Press LLC


LIST OF ILLUSTRATIONS
Energy consumed by mammals and birds
All or none
Simpson rule
Trapezoidal rule
Change of applied magnetic field
Loading and unloading to plastic strain
Triple point for water
Force between particles
Bulk flow
Carnot cycle
Impulsive motion
Color wheel
Conjugacy
Control volume
Cosine
Einthoven (EKG, ECG)
Elasticity
Exponential Failure
Fermat
Diffraction
Friction forces
Gerber
Movement of glacier ice
Gompertz
Hall effect
Hertz contact
Hooke vs. Newton
Increasing returns
Isotherms
Wall pressure
Lanchester
Lens law
Machines
Reciprocal deflections
Indentation
Mohr circle
Gaussian distribution and normal distribution
Force AC is equivalent to AB + BC
Pareto
Poisson ratio
Power function: hyperbola
Power function: parabola
Logarithmic power curves
Pythagorean
Quadrants
Reflection of light at a plane surface

©2000 CRC Press LLC


Reynolds number
Stiffness panels
Stress law of fracture
Survival
Tangents
Thermocouple
Toothed gearing
Transonic theory
Change in temperature when crushed ice is heated
Wall and velocity defect
Viscoelastic material models
Wagner
Wave motion
Yoda power law
Young-Helmholtz
Young modulus

©2000 CRC Press LLC


ABBREVIATIONS
AAAS American Association for Advancement of Science
ACS American Chemical Society
AGI American Geological Institute
AIChE American Institute of Chemical Engineers
AIME American Institute of Mining, Metallurgical and Petroleum Engineers
AIP American Institute of Physics
AMA actual mechanical advantage
API American Petroleum Institute
ASCE American Society of Civil Engineers
ASME American Society of Mechanical Engineers
BET Brunauer-Emmet-Teller
DOD Department of Defense (U.S.)
DOE Department of Energy (U.S.)
e.g. for example
eng., engr. engineer, engineering
erf error function
esp. especially
est. estimate(d)
etc. and so forth
exp(x) (x) is the exponent of e
ff. and following
i.e. that is
ISO International Standards Organization
LH latent heat
ln logarithm to the base e
log logarithm to the base 10
MTBF mean time before failure (same as)
MTTF mean time to failure
n.a. not available
n.d. no date; or undated
NAE National Academy of Engineering (U.S.)
NAS National Academy of Sciences (U.S.)
NASA National Aeronautics and Space Administration (U.S.)
NBS National Bureau of Standards (presently NIST)
NIST National Institute of Standards and Technology (U.S.)
NIH National Institutes of Health (U.S.)
NSF National Science Foundation (U.S.)
NTP normal temperature and pressure (25O C at 1 atm.)
NUC National Union Catalogue
NYPL New York Public Library
TMA theoretical mechanical advantage

©2000 CRC Press LLC


PHYSICAL CONSTANTS
(IN SI UNITS)
acceleration constant or Newton constant of gravitation, G
6.67259 × 10–11 m3kg–1s–2
acceleration in free fall due to gravity (standard value), g
9.80655 ms–2
Arrhenius—see universal gas constant, R

atomic mass, u (unified) (see Weizacker mass law)


1.6605402 × 10–27 kg
Avogadro constant, NA
6.0221367 × 1023 mol–1
Bohr magneton, eh/2me, µ B or β
9.274025 × 10–24 JT–1 (T = telsa)
Bohr radius, ao
0.529177249 × 10–10 m
Boltzmann constant, k (R/NA), or B
1.38066 × 10–23 JK–1
classical electron radius, re
2.81794092 × 10–15 m
Compton wavelength, λC
2.42631 × 10–12 m
Curie constant, C
Nµ 2/3k
where N = number of parametric ions per unit volume,
µ = magnetic moment of each ion, k = Boltzmann constant

e = 2.71828…, the base of the natural logarithm system, a transcendental irrational number

electron radius, re
2.81794 × 10–15 m
electron rest mass, me
9.1093897 × 10–31 kg
electron volt, eV
1.60218 × 10–19 J
elementary charge, e
1.60218 × 10–19 C (C = coulomb)
Faraday constant, F
9.6485309 × 104 Cmol–1 (C = coulomb)
gas constant—see molar gas constant

gravitational constant, GN
6.67259 × 10–11 m3kg–1s–2
Lorentz constant, L (some references use Lo)
2.45 × 10–8 watt-ohmK–2

©2000 CRC Press LLC


Loschmidt number, NL
2.686763 × 1025 m–3 mole/L
magnetic constant, µ o
4π × 10–7 Hm–1
mass of alpha particle, m
6.56 × 10–27 kg
molar gas constant, R
8.314510 Jmol–1K–1
molar volume (ideal gas), RT/ρ
22.41410 Lmol–1 or 22.414 m3mol–1 or 2.241383 × 10–3 m3mol–1
muon mass, mµ
1.8835327 × 10–23 kg
neutron rest mass, mr
1.674929 × 10–27 kg
Newtonian constant of gravitation, G
6.67259 × 10–11 m3kg–1s–2 or Nm2kg–2
Number of molecules per cc of gas at NTP, N
2.705 × 1019/cm3 (or /L)
pi, π, ratio of circumference of circle to its diameter
3.14159… (a transcendental irrational number)
Planck constant, h
6.6260755 × 10–34 Js
proton rest mass, mp
1.6726231 × 10–2 7kg
R—see universal gas constant

radius of an electron, a
2.8 × 10–16 m
Rydberg constant, R∞
10.9737315 × 106 m–1
standard atmosphere, Po
101325 Pa
Stefan-Boltzmann constant, σ
5.67051 × 10–8 Wm–2K–4
universal gas constant, R, or molar gas constant
8.314510 Jmol–1K–1
velocity of light in a vacuum, c
2.997925 × 108 ms–1
Wien displacement law constant, b
2.897756 × 10–3 mK
zero, Celsius scale, To
273.15 K
Sources
Kaye, G. W. C. and Laby, T. H. 1995; Lide, David R. 1994 (75th ed.). CRC Handbook of
Chemistry and Physics. CRC Press, Boca Raton, FL.
Besancon, Robert M. 1974 (2nd ed.); The Encyclopedia of Physics. Van Nostrand Reinhold
Co., New York, NY, 1067 p.
Lerner, R. G. and Trigg, G. L. 1991. Encyclopedia of Physics. Addison-Wesley Publishing
Co., Reading, MA, 1408 p.

©2000 CRC Press LLC


CONTENTS
Laws and Models: An Introduction
George A. Hazelrigg
Possible First Recorded Physical Law
Definitions and Descriptions
List of Illustrations
Abbreviations
Physical Constants (in SI Units)
Laws and Models
Sources and References

©2000 CRC Press LLC

You might also like