A Student's Guide To Dimensional Analysis (Don S. Lemons)
A Student's Guide To Dimensional Analysis (Don S. Lemons)
A N A LY S I S
www.cambridge.org
10.1017/9781316676165
This publication is in copyright. Subject to statutory exception and to the provisions of relevant
collective licensing agreements, no reproduction of any part may take place without the written
permission of Cambridge University Press.
A catalogue record for this publication is available from the British Library.
Cambridge University Press has no responsibility for the persistence or accuracy of URLs for
external or third-party Internet Web sites referred to in this publication and does not guarantee
that any content on such Web sites is, or will remain, accurate or appropriate.
Contents
Preface
Acknowledgments
1 Introduction
1.1 Dimensional Homogeneity
1.2 Dimensionless Products
1.3 Dimensional Formulae
1.4 The Rayleigh Algorithm
1.5 The Buckingham π Theorem
1.6 The Number of Dimensions
1.7 The Number of Dimensionless Products
1.8 Example: Pressure of an Ideal Gas
1.9 A Mistake to Avoid
Essential Ideas
Problems
2 Mechanics
2.1 Kinematics and Dynamics
2.2 Effective Dimensions
2.3 Imposed Dimensions
2.4 Example: Hanging Spring-Mass System
2.5 Example: Hanging, Stretched Cable
2.6 Asymptotic Behavior
2.7 Example: Speed of Sound
2.8 Example: Side Window Buffeting
2.9 Example: Two-Body Orbits
Essential Ideas
Problems
3 Hydrodynamics
3.1 Fluid Variables
3.2 Example: Water Waves
3.3 Surface Tension
3.4 Example: Largest Water Drop
3.5 Viscosity
3.6 Example: Hydraulic Jump
3.7 Example: Equilibrium Pipe Flow
3.8 Example: Non-Equilibrium Pipe Flow
3.9 Scale Models
Essential Ideas
Problems
6 Quantum Physics
6.1 Planck’s Constant
6.2 Example: Blackbody Radiation
6.3 Example: Spectral Energy Density of Blackbody Radiation
6.4 Example: Bohr Model
6.5 Atomic Units
6.6 Example: Quantum Ideal Gas
6.7 Example: Quantized Radiation from an Accelerating Charge
Essential Ideas
Problems
7 Dimensional Cosmology
7.1 The Fundamental Dimensional Constants
7.2 Eddington-Dirac Number
7.3 Example: Gravitational-Electrostatic Mass
7.4 Example: Classical Electron Radius
7.5 Classical Scales
7.6 Fine Structure Constant
7.7 The 15 Scales
7.8 Planck Scale
7.9 Example: Planck Mass
7.10 The 22 Quantities
Essential Ideas
Problems
A small dog could probably carry on his back two or three dogs of his
own size; but I believe that a horse could not carry even one of his own
size.
[1]
If Galileo had not thought dimensionally, he could not have made this
interesting argument.
By the time of Isaac Newton (1643–1727) scientists had begun to think
in terms of combinations of different dimensions. For instance, the
dimension of speed is length divided by time, the dimension of acceleration
is length divided by time squared, and, according to Newton’s second law,
the dimension of force is mass times length divided by time squared.
Newton regarded mass, length, and time as primary, fundamental
dimensions and combinations of these as secondary, derived ones. [2]
One of the first things a physics student learns is that one should not
add, subtract, equate, or compare quantities with different dimensions or
quantities with the same dimension and different units of measure. For
instance, one cannot add a mass to a length or, for that matter, 5 meters to 2
kilometers. This rule against what is sometimes called “adding apples and
oranges” means that every term that is added, subtracted, equated, or
compared in every valid equation or inequality must be of the same
dimension denominated in the same unit of measure. This is the principle of
dimensional homogeneity.
The principle of dimensional homogeneity is nothing new. Scientists
have long assumed that every term in every fully articulated equation that
accurately describes a physical state or process has the same dimension
denominated in the same unit of measure. However, it was not until 1822
that Joseph Fourier (1768–1830) expressed this principle in a way that
allowed important consequences to be derived from it. [3]
Symmetry under Change of Units
Behind the principle of dimensional homogeneity is a symmetry principle.
Symmetry principles tell us that something remains the same as something
else is changed. Here the something that remains the same is the form of the
equation or inequality and the things that are changed are the units in which
the dimensions of its terms are expressed. Thus, if we change the unit of
length from meters to kilometers and the form of the equation does not
change, that equation observes this particular symmetry and is, at least in
this regard, dimensionally homogeneous. If we change all of its units and
the form of the equation does not change, this equation is fully
dimensionally homogeneous.
An equation can be useful without being dimensionally homogeneous.
For instance,
s = 4.9t2
(1.1)
correctly describes the downward displacement s
denominated in meters of an object falling freely from rest for a period of
time t denominated in seconds. Yet (1.1) is not invariant with respect to
changes in its units of measure. Compare (1.1) with
(1.2)
(1.3)
where y − yo is the object’s displacement from its initial position yo, υyo is
its initial velocity, g is the magnitude of the gravitational acceleration, and
our coordinate system is oriented so that y becomes more negative as the
object falls and time advances. Equation (1.3) observes the principle of
dimensional homogeneity. For the dimension of y − yo is length; the
dimensions of υyo and t are, respectively, length/time and time so that the
dimension of their product υyot is length; and the dimension of gt2 is
length/time2 multiplied by time2 or, again, length. Furthermore, (1.3)
contains no dimensional constants masquerading as dimensionless numbers
– as does (1.1).
One consequence of the dimensional homogeneity of (1.3) is that
dividing each of its terms by gt2 produces an equation,
(1.4)
(1.5)
(1.6)
I have often been impressed with the scanty attention paid even by
original workers to the great principle of similitude. It happens not
infrequently that results in the form of “laws” are put forth as novelties
on the basis of elaborate experiments, which might have been
predicted a priori after a few minutes’ consideration.
[4]
(1.7)
T: α − 2δ = 0,
(1.8a)
M: β = 0,
(1.8b)
and
L: γ + δ = 0.
(1.8c)
The three equations (1.8a), (1.8b), and (1.8c) constrain the
four unknowns, α, β, γ, and δ, to a family of solutions β = 0, γ = − α/2, and
δ = α/2 parameterized by α. [The symbols T, M, and L preceding equations
(1.8) identify the source of each constraint.] Therefore, (Δtg1/2/R1/2)α is
dimensionless for any α, which means that Δtg1/2/R1/2, as well as θ, is
dimensionless.
Once we know the dimensionless products that can be formed out of
the model’s dimensional variables and constants, we know they must be
related to one another by an undetermined function, that is, in this case
expressed by
(1.9)
Np = NV − ND.
(1.10)
Statement (1.10) is the most common expression of the
π theorem.
1.6 The Number of Dimensions
Most dimensional analysts adopt M, L, and T as dimensions appropriate for
mechanical processes and states. We did so in describing the marble on the
interior surface of a cone. In that case, ND = 3. Furthermore, since Δt, m, R,
g, and θ describe the marble’s motion, NV = 5. Therefore, according to
(1.10) NP = NV − ND, 2(=5 − 3) complete and independent dimensionless
products should be produced. By applying the Rayleigh algorithm we find
these to be Δtg1/2/R1/2 and θ. The set Δtg1/2/R1/2 and θ is complete because
every possible dimensionless product of Δt, m, R, g, and θ can be expressed
as a product of some power of Δt2R/g times some power of θ. And its
members are independent because Δt2R/g and θ are not powers of each
other.
But the minimum number of dimensions needed to express the
dimensional formulae of the NV dimensional variables and constants is not
always 3, as it is in this example. Neither is the identity of the minimum
number of dimensions necessarily M, L, and T – as they often are in
mechanical problems. Rather, the dimensions required are, in
Buckingham’s words, the “arbitrary fundamental units [dimensions] needed
as a basis for the absolute system.” [7] And only when we can ensure that
ND is the minimum number of dimensions needed can we depend on
NP = NV − ND to be observed. Otherwise, NP = NV − ND remains a mere
“rule of thumb” – often observed but sometimes not. We will learn how to
recognize the minimum number of dimensions in Section 2.2.
1.7 The Number of Dimensionless
Products
Note that the more dimensionless products, π1, π2, … πNP, produced, the
less determined the state or process described by the dimensional model.
After all, if only one product π1 is produced, the result sought assumes a
form f(π1) = 0 whose solution π1 = C is in terms of a single undetermined
dimensionless number C. However, if two dimensionless products, π1 and
π2, are produced, these are related by g(π1, π2) = 0 whose solution
π1 = h(π2) leaves a function h(π2) of a single variable undetermined. And if
three dimensionless products, π1, π2, and π3, are produced, these are related
by a function j(π1, π2, π3) = 0 whose solution π1 = k(π2, π3) leaves a function
k(π1, π2) of two variables undetermined.
It is clear that in order to more completely determine a state or process,
we need to minimize the number NP of complete and independent
dimensionless products. According to the rule of thumb NP = NV − ND, we
do this by minimizing NV (the number of dimensional variables and
constants that describe the model) and, assuming we have such freedom, by
maximizing ND (the minimum number of dimensions in terms of which
these variables and constants can be expressed.) However, minimizing NV
and maximizing ND are not straightforward tasks. Both require skill and
judgment – the same kind of skill and judgment needed to construct a
model of a physical state or process.
1.8 Example: Pressure of an Ideal Gas
Many of these ideas are illustrated in the dimensional analysis of how the
pressure p of an ideal gas depends on quantities that describe its state. The
pressure a gas exerts on its container walls is the average rate per unit area
with which its molecules collide with and transfer momentum to the wall.
The ideal gas model treats these molecules as randomly and freely moving,
massive, point particles whose instantaneous collisions with other particles
and with the walls conserve their energy.
Therefore, we believe the ideal gas pressure p should depend on the
number density of the gas molecules N/V where N is the number of gas
molecules contained in volume V, the mass of each of the molecules m, and
their average or characteristic speed . These parameters should be
sufficient, since they are the elements out of which the momentum of the
gas particles, the rate at which they collide with the wall, and their energy
are composed. To include other variables or constants such as, for instance,
the acceleration of gravity g would be to introduce extraneous
dimensionless products and make our result not so much inaccurate as
uninformative. For convenient reference, we collect these symbols, their
descriptions, and their dimensional formulae in Table 1.1.
Table 1.1
Symbol Description Dimensional
Formula
p Pressure ML−1T−2
Symbol Description Dimensional
Formula
m Molecular mass M
(1.11)
M: 1 + β = 0,
(1.12a)
L: −1 − 3α + γ = 0,
(1.12b)
and
T: −2 − γ = 0.
(1.12c)
From equations (1.12) we find that α = − 1, β = − 1, and
(1.13)
Table 1.2
ω Frequency T−1
m Mass M
V Volume L3
(1.14)
and, therefore, that the exponents are constrained by
T: −1 − 2β − 2ε = 0,
(1.15a)
M: α + β + γ = 0,
(1.15b)
and
L: −3γ + 3δ + ε = 0.
(1.15c)
Equations (1.15) are solved by α = 1/2 − δ + 4ε/3,
β = − 1/2 − ε, and γ = δ − ε/3. Consequently,
(1.16)
(1.17)
(1.18)
(1.19)
(a) Show that the mass of the animal’s leg cannot enter into a dimensionless
product with υ, l, and g.
(b) Show that the one dimensionless product composed of these variables is
the walking Froude number υ2/gl.
(c) How does the walking speed υ depend on the leg length l?
answer
(d) Given that many humans prefer to walk at a speed of approximately
1.4 ⋅ m/s, estimate the walking Froude number of a typical human being.
(You will have to estimate the human leg length l.)
answer
(e) Estimate your own walking Froude number by measuring your normal
walking speed and your leg length.
1.5 Vibrating wire. A wire of length l and mass per unit length λ pulls with
a force (or tension) τ on two posts between which the wire is stretched tight
and to which each end is attached. Suppose the middle of the wire is
plucked. Use the Rayleigh algorithm to determine how the period Δt of the
wire’s oscillation depends upon l, λ, and τ.
answer
1.6 Falling through the center of the Earth. Suppose you drill a tunnel
through the Earth’s center as illustrated in Figure 1.3. Then you drop an
object of mass m from rest into the tunnel. What is the time required Δt for
the object’s passage from one side of the Earth to the other? This duration
should depend on the gravitational constant G, the density of the Earth ρ,
and the Earth’s radius R. It might also depend on the mass m of the object
dropped. These dimensional variables and constants, their descriptions, and
their dimensional formulae are collected in Table 1.3.
(a) Since there are 5 dimensional variables and constants expressed in terms
of 3 dimensions, there should be, according to the rule of thumb
NP = NV − ND, 2 dimensionless products. Use the Rayleigh algorithm to
find the two dimensionless products π1 and π2.
answer
(b) There is a good reason why the mass of the object m should be
eliminated from the list of variables and constants on which the duration Δt
depends. Again use the Rayleigh algorithm to find the single dimensionless
product implied by the variables and constants neglecting m. Derive an
expression for the duration Δt.
answer
R Radius of Earth L
m Mass of object M
Table 2.1
p Pressure ML−1T−2
w Weight MLT−2
A Area L2
(2.1)
M: 1 + α = 0,
(2.2a)
L: −1 + α + 2β = 0,
(2.2b)
and
T: −2 − 2α = 0.
(2.2c)
Since equations (2.2a) and (2.2c) are algebraically
equivalent, only one of them is needed. Thus, we find that α = − 1 and
β = 1. As expected, this solution produces the dimensionless product pA/w
or, equivalently, pA/mg and the result
(2.3)
Table 2.2
p Pressure (MLT−2)L−2 FL−2
w Weight (MLT−2) F
A Area L2 L2
Table 2.3
ω Frequency T−1
m Mass M
(2.4)
The constraints
T: −1 − 2β − 2ε = 0,
(2.5a)
M: α + β + γ = 0,
(2.5b)
V: −γ + δ = 0,
(2.5c)
and
L: ε = 0
(2.5d)
are solved by β = − 1/2, γ = 1/2 − α, δ = 1/2 − α, and
ε = 0. Therefore, ωmαkβργVδgε =(ωρ1/2V1/2/k1/2)(m/ρV)α where α is arbitrary,
and so the two dimensionless products are and ρV/m.
Because these two products are complete, we may form all others from
them. In particular, if we multiply the first by the inverse square root of the
second, we produce the convenient pair and ρV/m. These are
related by
(2.6)
where h(x) is an undetermined function. Given m = ρV, this result reduces
to
(2.7)
Table 2.4
Material Bulk moduli (in Pascals)
Concrete 30 ⋅ 109
Bone ∼9 ⋅ 109
Table 2.5
s Stretch L L
l Length L L
Table 2.6
s Stretch L
l Length L
(2.8)
f(x) ≈ C ⋅ xn
(2.9)
valid when either x < < 1 or x > > 1 and where both C
and n are undetermined and n may vanish.
Of course, dimensional analysis alone does not reveal the asymptotic
behavior of f(x). But our physics sense might. If x < < 1 or x > > 1, we
should ask ourselves, “How should f(x) behave as x → 0 or x → ∞? Should
f(x) approach a finite constant or should it become monotonically small or
large?” If our answer is “Yes” to either of these questions, the power law
limit f(x) → C ⋅ xn may be appropriate. And, if so, we may be able to
determine whether n = 0, n > 0, or n < 0.
To illustrate, we return to the stretched, hanging cable of Section 2.5.
Suppose we are most concerned with cables of, say, a few meters, more or
less, a mass density of less than one kilogram per meter, and common, solid
materials with a bulk modulus on the order of 109 ⋅ Pa or greater. In this
case the dimensionless ratio λg/lK ≤ 10−8. Since there is no reason f(λg/lK)
should oscillate or change exponentially or logarithmically as λg/lK
decreases, a power law asymptotic form makes sense. Therefore, we
approximate (2.8) with
(2.10)
Table 2.7
cS Sound speed LT−1
(2.11)
implies that when the exponents α, β, and γ satisfy
L: 1 − α − 3β + 2γ = 0,
(2.12a)
T: −1 − 2α − 2γ,
(2.12b)
and
M: α + β + γ = 0
(2.12c)
the form will be dimensionless. The
solution of equations (2.12) is α = − 1/2, β = 1/2, and γ = 0. Therefore,
(2.13)
Table 2.8
ω Frequency T−1
A Area of opening L2
V Resonator volume L3
(2.14)
T: −1 − 2δ = 0,
(2.15a)
L: 2α + 3β − 3γ − δ = 0,
(2.15b)
and
M: γ + δ = 0.
(2.15c)
Their solution, β = 1/3 − 2α/3, γ = 1/2, and δ = − 1/2,
reduces the dimensionless product to ωV1/3(ρ/p)1/2(A/V2/3)α where α is
arbitrary. Therefore,
(2.16)
(2.17)
(2.18)
2.9 Example: Two-Body Orbits
The moon orbits the earth, the earth orbits the sun, and the stars in a double
star system orbit each other. Each of these systems contains two bodies, one
of mass m1 and the other of mass m2, separated by a characteristic distance
r. Their gravitational attraction causes each body in a system to orbit their
mutual center of mass. This motion has a period Δt after which time the
massive bodies return to their initial positions and velocities. Thus, 4
dimensional variables, Δt, r, m1 and m2, and 1 dimensional constant, G,
completely characterize this two-body, gravitationally bound system. Their
symbols, descriptions, and dimensional formulae are collected in Table 2.9.
Table 2.9
Δt Period T
r Characteristic L
separation
m1 Mass 1 M
m2 Mass 2 M
(2.19)
and
T: 1 − 2δ = 0,
(2.20a)
L: α + 3δ = 0,
(2.20b)
and
M: β + γ − δ = 0
(2.20c)
renders dimensionless. The solution to these
constraints is α = − 3/2, γ = 1/2 − β, and δ = 1/2, and so
(2.21)
(2.22)
where the function f(x) is undetermined. This result expresses Kepler’s third
law, according to which Δt2 ∝ r3 where r is directly proportional to the
semi-major axis of a planet’s elliptical orbit around the sun. This is as far as
dimensional analysis takes us.
Symmetry considerations take us a little further. Replacing m1 with m2
and m2 with m1 should leave (2.22) unchanged. Therefore,
(2.23)
or, equivalently,
(2.24)
(2.25)
Essential Ideas
The dimensions in terms of which the dimensional formulae of the
dimensional variables and constants are expressed are not absolute, but
rather relative to the model adopted. These relative or imposed dimensions
encode idealizations and simplifications that define the model. The number
of effective dimensions is the minimum number of imposed dimensions or
combinations of imposed dimensions in terms of which the dimensional
variables and constants are expressed. The rule of thumb NP = NV − ND is
always observed when ND is the number of effective dimensions.
Problems
2.1 Gravitational constant. Use Newton’s law of gravitation
F = Gm1m2/r2 to determine the dimensional formula of the gravitational
constant G in terms of M, L, and T .
answer
2.2 Interior pressure. A non-rotating, self-gravitating, spherical body has a
mass density ρ and radius R. [In parts (a) and (b) below use the result of
Problem 2.1.]
(a) Use the Rayleigh algorithm to show that po = C ⋅ Gρ2R2 where po is the
pressure at the center of the sphere and C is an undetermined dimensionless
constant.
(b) Also determine an expression for the pressure p(r) at a distance r from
the center of the sphere.
answer
2.3 Gravitational instability. A self-gravitating, spherical body of mass
density ρ and radius R rotates with at an angular frequency ω. Find how the
critical frequency ω* at which the sphere begins to break up depends on ρ,
R, and G. [13]
answer
2.4 Escape velocity. Determine how the speed υ required for any object to
escape from a planet depends on the planet’s mass M and radius R. See
Figure 2.2.
answer
Figure 2.2. Escape velocity.
2.5 Transverse waveform. A long wire with mass per unit length ρ is
tightly strung between two posts as illustrated in Figure 2.3. Its tension τ is
essentially uniform throughout. The wire is plucked near one end and a
transverse waveform propagates as shown in Figure 2.3. Determine an
expression for the waveform’s speed of propagation υ in terms of τ and ρ.
answer
(a) How does the drag force FD that resists this motion depend on these
variables?
answer
(b) Suppose that the sphere is falling at terminal speed through the fluid.
Given that the drag force FD perfectly balances the sphere’s weight mg, find
how the terminal speed υ depends on mg, ρ, and r.
answer
2.7 Free vibrations. A body of mass m, size l, and bulk modulus K vibrates
with a natural (unforced) frequency ω. Find an expression for ω in terms of
m, l, and K.
answer
2.8 Soap bubble. The air pressure within a spherical soap bubble of radius
r and surface tension σ exceeds that outside the bubble by Δp. Since the
soap bubble is in an equilibrium state, force is an imposed dimension F
independent of M, L, and T.
(a) How is Δp related to r and the surface tension σ of the bubble? One
should find the, possibly, counterintuitive result that Δp ∝ r−1.
2.9 Blast wave. Energy E is quickly released at a point near the surface of
the Earth. The resulting blast wave propagates through air of density ρ and
creates a hemispherical shock of radius R at time t. How does R depend on
time t, ρ, and E? [14]
answer
Table 3.1
υ Wave speed LT−1
λ Wavelength L
a Wave amplitude L
d Fluid depth L
(3.1)
where f(x) is undetermined. However, (3.1) tells us only that the speed υ of
a small-amplitude water wave is proportional to the square root of the
acceleration of gravity g.
The limiting cases of deep-water waves for which d > > λ and for
shallow-water waves for which d < < λ yield to further analysis. In both
cases (3.1) reduces to an asymptotic form. Assuming this asymptotic form
is that of a power law, (3.1) becomes
(3.2)
where C and n are undetermined and may be different in the two cases
λ/d < < 1 and λ/d > > 1.
Because deep-water waves on the surface of the ocean seem to be
independent of ocean depth, it must be that in this limit n = 0 and therefore
the speed of a deep-water wave
(3.3)
(3.4)
Table 3.2
υ Wave speed LT−1
λ Wavelength L
(3.5)
M : β + δ = 0.
(3.6c)
The solution to equations (3.6) is β = − 1/4 − α/2,
γ = − 1/4 + α/2, and δ = 1/4 + α/2. Consequently, the dimensionless
products are υ[ρ/(σg)]1/4 and λ(gρ/σ)1/2. By multiplying the first by the
square root of the second, we produce υ(λρ/σ)1/2. This dimensionless
product and λ(gρ/σ)1/2 compose an independent pair. In this way we find
that the wave speed
(3.7)
Table 3.3
V Volume L3
Table 3.4
V Volume L3
σ Surface tension FL−1
(3.8)
(3.9)
(3.10)
(3.11)
Table 3.5
υ Terminal speed LT−1 LT−1
mg Weight MLT−2 F
r Radius L L
(3.12)
L : 1 − α + β + γ = 0,
(3.13a)
T : − 1 − α − 2β = 0,
(3.13b)
and
M:α+β=0
(3.13c)
in order to make υμα(mg)βrγ dimensionless. The solution
to these constraints is α = 1, β = − 1, and γ = 1. Therefore, the single
dimensionless product is υμr/mg, and so
(3.14)
FD = C′ ⋅ υrμ
(3.15)
where C′ = 1/C. Relation (3.15) with C′ = 6π is known as
Stokes’ law after George Stokes (1819–1903), the British mathematician,
scientist, and engineer who discovered it in 1851.
3.6 Example: Hydraulic Jump
Anyone with a kitchen sink has observed the hydraulic jump. As water falls
from the faucet and strikes the relatively flat bottom of the sink, it spreads
out in a smooth circular pattern that is surrounded by a stationary wave
form or jump. Beyond the jump, the water flows more deeply and roughly.
This jump develops in order to maintain a constant rate of flow given that
viscosity continually slows the water flowing along the sink bottom. Figure
3.2 illustrates the situation.
Table 3.6
R Position of jump L
μ Viscosity ML−1T−1
υ Speed LT−1
r Radius of column L
(3.16)
L : 1 − 3α − β + γ + δ = 0,
(3.17a)
M : α + β = 0,
(3.17b)
and
T : − β − γ = 0.
(3.17c)
The solution of (3.17) is β = − α, γ = α, and δ = − 1 + α. These determine
two dimensionless products, R/r and rρυ/μ, that are related by
(3.18)
(3.19)
I observe that a larger flow rate ρr2υ causes a larger jump position R.
Therefore, n > 0. I also observe that increasing r while preserving ρr2υ, say,
by raising a cookie sheet under a steady stream of water, weakly increases
R. Therefore, n < 1 and so 0 < n < 1. According to one model for which
there is experimental support, n = 1/3. [17] In this case (3.20) becomes
(3.21)
Given the above data, I find that, for my sink and faucet, C ≈ 0.25.
3.7 Example: Equilibrium Pipe Flow
Consider the equilibrium flow of an incompressible fluid at a volumetric
rate Q from left to right through a straight, horizontal pipe of length l and
uniform cross-sectional area A as illustrated in Figure 3.3. The quantity Q/A
is the average fluid speed υ over a cross-section. Thus, Q = υA. By
equilibrium flow I mean one for which the net force on each part of the
fluid, that is, on each fluid element, vanishes. Therefore, the velocity of
each fluid element is constant.
Given the interaction with the walls of the pipe and the retarding effect
of the fluid’s viscosity μ, a force must push the fluid from left to right in
order to keep it moving at constant speed. This force is maintained by a
pressure gradient ∇p pointing in the direction opposite to the flow, that is,
from right to left. Therefore, as the fluid moves from left to right over
length l, its pressure drops by an increment l∇p. How does the flow rate Q
depend on the cross-sectional area A, the pressure gradient ∇p, the fluid
viscosity μ, and the fluid mass density ρ?
Since this is an equilibrium problem, the dimension F is an appropriate
imposed dimension along with and independent of M, L, and T. As usual,
the dimension F enters only into those dimensional formulae that represent
counterbalancing forces. In particular, [Δp] = FL−3 and [μ] = FL−2T.
Note that imposing the dimension F, as we do here, is mathematically
distinct from using the relationship [F] = [m][a] and its consequence
F = MLT−2 to eliminate one of the dimensions M, L, or T in favor of F. The
former restricts the solution to one in which the net force on every fluid
element vanishes. In this case, forces are independent of mass times
acceleration. The latter applies Newton’s second law F = ma by adopting
the equivalence F = MLT−2.
The dimensional variables and constants, their descriptions, and their
dimensional formulae are collected in Table 3.7. Since there are 5
dimensional variables and constants (Q,∇p,μ,ρ, and A) and 4 imposed
dimensions (M, L, T, and F), there should be 1 dimensionless product. Table
3.7 reveals that, since the dimension M appears only once, the mass density
ρ cannot enter into this dimensionless product.
Table 3.7
Q Flow rate L3T−1
μ Viscosity FL−2T
A Cross-sectional area L2
Therefore, the dimensionless product assumes a form Q(∇P)αμβAγ
where the exponents, α, β, and γ, make
(3.22)
equal to 1. Thus,
L : 3 − 3α − 2β + 2δ = 0,
(3.23a)
T : − 1 + β = 0,
(3.23b)
and
F : α + β = 0.
(3.23c)
The solution to equations (3.23) is α = − 1, β = 1, and
γ = − 2. Therefore, the single dimensionless product is Qμ/(A2∇p) and so
(3.24)
Table 3.8
Q Flow rate L3T−1
μ Viscosity ML−1T−1
A Cross-sectional area L2
(3.25)
equals 1. Therefore,
3 − 2α − β − 3γ + 2δ = 0,
(3.26a)
−1 − 2α − β = 0,
(3.26b)
and
α + β + γ = 0.
(3.26c)
The solution to (3.26), β = − 1 − 2α, γ = 1 + α, and
δ = − 1/2 + 3α/2, produces two dimensionless products: Qρ/(μA1/2) and
ρA3/2∇p/μ2. Multiplying the first by the inverse of the second yields the
dimensionless product Qμ/(A2∇p) that, according to Section 3.7,
characterizes equilibrium flow in a straight pipe. The relation between this
dimensionless product and Qρ/(μA1/2),
(3.27)
(3.28)
where .
3.9 Scale Models
The result of a dimensional analysis is often that one dimensionless product
equals an unknown function of another. Relation (3.27) among the variables
describing pipe flow for arbitrary Reynolds number,
(3.29)
is one example.
Suppose our job is to design a large pipe of cross-sectional area A and
length l that transports motor oil. We want to test our design on a smaller-
scale version of that pipe. But pumping motor oil through a smaller pipe
with the same aspect ratio A1/2/l is not adequate. While proportionally
decreasing A1/2 and l preserves the geometry of the pipe, doing so does not
preserve the Reynolds number Qρ/(μA1/2) of the flow, and it is the Reynolds
number that determines whether the flow is laminar or turbulent. In fact,
preserving geometric similitude is not necessary at all. Rather, what needs
to be preserved, in this case, is dynamic similitude.
One preserves dynamic similitude by making the values of the two
dimensionless products Qμ/(A2∇p) and Qρ/(μA1/2) realized in the scale
model identical to those in the fully designed pipe. We do this by varying
the pressure gradient ∇p in the scale model and measuring the resulting
flow rate Q – the other quantities being constant – and so determining the
relationship between Qμ/(A2∇p) and in the scale model, that is,
by determining the function f(x). The pressure gradient ∇p in the scale
model is then chosen so that the two dimensionless products are the same in
the scale model and the fully designed pipe. Cars, ships, and planes, as well
as pipes, have been designed in this way.
However, scale models should not depart from the physics
incorporated in the full design. If, for instance, the scale model of a pipe is
too small, surface tension becomes significant and the scaling implied by
(3.29) breaks down.
Essential Ideas
Dimensional models of incompressible fluids introduce the dimensional
variables mass density ρ, flow speed υ, and pressure p as well as surface
tension σ and viscosity μ. The Reynolds number is a dimensionless product
that when relatively small produces laminar flow and when relatively high
produces turbulent flow.
Problems
3.1 Capillary waves. The context is Sections 3.2 and 3.3. Use the Rayleigh
algorithm to show that the speed υ of a capillary wave with wavelength λ in
an compressible fluid with mass density ρ and surface tension σ is given by
where C is an undetermined dimensionless number.
3.2 Water drop oscillation. Use either the Rayleigh algorithm or a less
formal dimensional argument to determine how the oscillation frequency ω
of a small drop of incompressible fluid in free fall depends on the volume V
of the drop, its mass density ρ, and its surface tension σ.
answer
3.3 Capillary effect. When a small-diameter tube open at both ends is
inserted into water, as shown in Figure 3.4, the water is drawn a certain
distance h up the tube. This distance depends on the diameter d of the tube,
the surface tension of water σ, the mass density ρ of the water, and the
acceleration of gravity g. Find the 2 dimensionless products implied by
these dimensional quantities and cast them into an expression for h.
answer
[cp] = HΘ−1M−1
(4.1)
where cp is a heat capacity per unit mass at constant
pressure. We depend on context to distinguish between the traditional
symbols for temperature T and the dimension time T. Thus, [T] = Θ while,
if υ is a velocity, [υ] = LT−1.
When work is dissipated into or produced out of the internal energy of
a thermodynamic system, the first law of thermodynamics is necessarily
invoked. Then the dimension heat H loses its independence and should be
replaced with the dimension of energy ML2T−2. Yet when no work is done
on or by a system, heat H is an appropriate imposed dimension just as when
no part of a mechanical system is accelerated force F is an appropriate
imposed dimension.
Occasionally, one reads that the variable temperature does not need its
own dimension Θ because temperature can always be measured in energy
units with dimensional formula ML2T−2. The three examples in this chapter,
in Sections 4.4, 4.7, and 4.8, are counterexamples to this claim. Those in
Sections 5.7, 6.2, 6.3, and 6.6 illustrate states and processes in which
temperature can be measured in energy units.
4.3 Conduction and Convection
Heat conduction is accomplished when the energy of some atoms and
molecules is imparted to other nearby atoms and molecules. Solids, liquids,
gases, and plasmas all conduct heat. Thermal conductivity is a dimensional
constant that characterizes a particular material’s ability to conduct heat. In
general, thermal conductivity may vary from point to point and even change
in time as the material properties change.
The extent to which a material conducts heat at a point is proportional
to the temperature gradient at that point. A one-dimensional quantification
of this statement is
(4.2)
where q is the heat flux or rate at which heat flows past a point per unit
area, T is the temperature, and ∂T/∂x is the x component of the temperature
gradient. The negative sign in (4.2) ensures that the heat flux is always in a
direction opposed to the temperature gradient, that is, always from hot to
cold. The thermal conductivity k is the proportionality constant that turns
(4.2) into an equality. Therefore,
(4.3)
Table 4.1
Δt Cooking time T
m Turkey mass M
ΔT Temperature difference Θ
Since there are 6 dimensional variables and constants, Δt, m, ΔT, k, cp,
and ρ and 5 imposed dimensions, M, L, T, H, and Θ, our analysis should
produce 1 dimensionless product. This product will assume a form
with exponents, α, β, γ, δ, and ε, that render the form
dimensionless. Therefore,
(4.4)
where the 5 exponents α, β, γ, δ, and ε satisfy the 5
constraints
T: 1 − γ = 0,
(4.5a)
M: α − δ + ε = 0,
(4.5b)
Θ: β − γ − δ = 0,
(4.5c)
H: γ + δ = 0,
(4.5d)
and
L: −γ − 3ε = 0.
(4.5e)
Their solution, α = − 2/3, β = 0, γ = 1, δ = − 1, and ε = − 1/3, produces
(4.6)
(4.7)
(4.9)
where k is the thermal conductivity of the material through which the heat
conducts. However, (4.9) tells only half the story of heat conduction. For a
non-uniform heat flux q, that is a non-vanishing ∂q/∂x, inevitably leads to
regions into which more heat enters than leaves. Furthermore, the
temperature T of a region into which more heat enters than leaves increases
at a rate ∂T/∂t. Thus, we have
(4.10)
where the negative sign indicates that if more heat enters a region than
leaves, then ∂q/∂x < 0 and the temperature increases, that is, ∂T/∂t > 0.
Combining (4.9) and (4.10) we find, providing the thermal conductivity k is
uniform in space, that
(4.11a)
or, equivalently,
(4.11b)
where
(4.12)
(4.13)
where To, Δt, and l2 are dimensional constants that characterize the initial
temperature To and the time Δt required for a temperature T to penetrate a
distance l. Therefore, the dimensionless product DTΔt/l2, which appears in
the dimensionless equation (4.13), is equal to a dimensionless number C.
Thus, given Δt = C ⋅ l2/DT and DT = k/ρcp, we find that Δt = C ⋅ l2ρcp/k.
Also since l2 ∝ V2/3 and V = m/ρ we find that Δt = C′ ⋅ m2/3ρ1/3cp/k – which
reproduces (4.6).
Making the equations that govern a process or define a state
dimensionless and then extracting the dimensionless product or products
that parameterize these equations is a distinct method of dimensional
analysis. According to this method we should list all the relevant equations
and identify the dimensionless products entering into them. If there is only
one dimensionless product, that product is equal to a dimensionless number.
If there is more than one, the products are related to each other by an
unknown function.
The process of making the equations dimensionless takes the place of
the Rayleigh algorithm. Let’s call this procedure making the equations
dimensionless rather than the unattractive, if more common, non-
dimensionalization. Of course, making the equations dimensionless implies
that we know what those equations are. And sometimes we do not.
4.7 Example: Growth of Arctic Ice
For the purpose of being the first to reach the North Pole, the Norwegian
explorer Fridtjof Nansen (1861–1930) allowed his ship, the Fram (in
Norwegian “Forward”), to become frozen in the ice of the Laptev Sea (north of
Siberia). Nansen believed that the natural drift of polar ice would carry the
Fram across the North Pole. As it happened, Nansen made it only to 86 degrees
north latitude (a record), but he and his men survived. Nansen later wrote an
excellent account of his journey entitled Farthest North.
One of Nansen’s occupations that first winter of his expedition (1884–
1885) was to measure the thickness of the ice that surrounded the Fram. He
noted that,
From measurements that were constantly being made, it appeared that the
ice which was formed in the autumn in October or November continued to
increase in size during the whole of the winter and out into the spring, but
more slowly the thicker it became.
[18]
That the ice grew “more slowly the thicker it became” is confirmed by
dimensional analysis. Figure 4.2 illustrates the geometry.
Table 4.2
λ Ice thickness L
t Time T
L: 1 − γ − 3ε = 0,
(4.15a)
T: α − γ = 0,
(4.15b)
Θ: β − γ − ϕ = 0,
(4.15c)
H: γ + δ + ϕ = 0,
(4.15d)
and
M: −δ + ε − ϕ = 0.
(4.15e)
Constraints (4.15) are solved by α = γ = − 1/2, δ = − β,
ε = 1/2, and ϕ = β + 1/2. Therefore, the dimensionless products are
λ(hρ/tΔTk)1/2 and h/cpΔT, and the ice thickness λ is described by
(4.16)
h Height of chimney L
TH − TC Temperature increment Θ
TH Inside temperature Θ
From this list we may immediately eliminate the mass density ρ of the
air since its dimensional formula ML−3 cannot be combined with the others
in a dimensionless product. This leaves 6 dimensional variables and
constants, υ, g, h, cp, TH − TC, and TH, and 3 imposed dimensions, L, T, and
Θ. Therefore, the analysis should produce 3 dimensionless products. An
inspection of the table reveals these to be , cp(TH − TC)/gh, and
(TH − TC)/TH. Therefore, the speed υ of the airflow may be expressed as
(4.17)
(4.18)
(5.1a)
(5.1b)
(5.1c)
and
(5.1d)
(5.2)
where is the velocity of the charge. Equations (5.1) and (5.2) govern all
electrodynamic phenomena.a
Maxwell first composed his eponymous equations (5.1) in 1865 from
contributions made during the previous 80 years. He was especially
indebted to Michael Faraday (1791–1867) and urged those who would
understand his own work first to read Faraday’s 1,100-page Experimental
Researches in Electricity. Faraday pioneered the concept of field as opposed
to the action at a distance view that prevailed in his time. According to
action at a distance, charged particles exert forces directly on each other.
Faraday believed, rather, that fields are real objects that mediate the
interaction of charged particles. Maxwell’s effort to mathematize Faraday’s
pictorial understanding of electric and magnetic fields led him to formulate
equations (5.1). The second term on the right-hand side of the Ampere-
Maxwell law (5.1d) introduces Maxwell’s own contribution: the
displacement current term .
Maxwell showed that the Ampere-Maxwell law (5.1d) and Faraday’s
law (5.1b) together allow for self-supporting electromagnetic field
structures that detach from their sources and propagate through space at the
speed of light. In this way Maxwell predicted that light waves are part of
the spectrum of electromagnetic waves – a prediction Heinrich Hertz
(1857–1897) verified experimentally, if accidentally, in 1886.
Maxwell’s equations (5.1) require one new dimension, the electric
charge, whose SI unit is the coulomb and whose symbol is Q. Therefore, the
dimension of charge density ρ is QL−3 and of current density J is QT−1L−2.
From [ρ] = QL−3, [J] = QT−1L−2, and (5.2) we find that
[E] = MLT−2Q−1,
(5.3a)
and
[B] = MT−1Q−1.
(5.3b)
From Gauss’s law (5.1a) and from (5.3a) we find that
[εo] = M−1L−3T2Q2
(5.4a)
and from the Ampere-Maxwell law (5.1d) and from (5.3b)
we find that
[μo] = MLQ−2.
(5.4b)
Note that since [μoεo] = L−2T2, the dimensional constant
has the dimensional formula of speed LT−1. Furthermore, the
magnitude of , to three places 3.00 ⋅ 108 ⋅ m/s, is the speed of light
in vacuo. We will also make use of the potential difference or voltage V
defined by . Therefore, given (5.3a), [V] = ML2T−2Q−1 .
A useful approximation of Maxwell’s equations is one in which both
Maxwell’s displacement current term in (5.1d) and the right-hand side of
Faraday’s law (5.1b) are unimportant. In this approximation, electric and
magnetic fields decouple and each is generated separately from its own kind
of source, electric fields from charges and magnetic fields from currents.
This decoupling creates the separate realms of electrostatics and
magnetostatics. The vacuum permittivity εo enters into the first and the
vacuum permeability μo into the second.
5.2 Example: Oscillations of a Compass
Needle
By the end of the eighteenth century, small magnetized pieces of iron had
been used for six or seven centuries as compass needles to help guide
seafarers. Also by the end of the eighteenth century, the direction of
magnetic north relative to celestial north (magnetic declination) and the
deflection of a compass needle in the vertical direction (magnetic dip) had
been mapped over much of the Atlantic. Magnetic declination and dip
together describe the direction of what was eventually called the
geomagnetic field. But how can the magnitude of this field be measured?
In 1776, Jean Charles Borda (1733–1799), a French naval engineer
who would shortly be commanding French warships aiding the American
revolutionaries, had a clever idea. Just as the oscillation frequency ω of a
pendulum is directly proportional to the square root of the magnitude g of
the local acceleration of gravity so that , the oscillation frequency
ω of a compass needle should be proportional to the square root of the
magnitude B of the local geomagnetic field so that . By measuring
the oscillation frequency ω of a compass needle Borda hoped to measure B.
[22]
Because [ω] = T−1 and [B] = MT−1Q−1 the frequency ω must depend
on more than the magnitude of B. Certainly ω should also depend on how
strongly the compass needle responds to a magnetic field and on the
compass needle’s rotational inertia. The magnetic dipole moment μ of the
compass needle (whose symbol should not be confused with that of the
vacuum permeability μo) parameterizes the first and the needle’s moment of
inertia I the second. These symbols, their descriptions, and their
dimensional formulae are found in Table 5.1.
Table 5.1
ω Frequency T−1
(5.5)
equal to 1. Therefore,
T: −1 − α − β = 0,
(5.6a)
M: α + γ = 0,
(5.6b)
Q: −α + β = 0,
(5.6c)
and
L: 2β + 2γ = 0.
(5.6d)
Equations (5.6) are solved by α = − 1/2, β = − 1/2, and
γ = 1/2. In this way, we find that
(5.7)
Table 5.2
P Power radiated ML2T−3
q Charge Q
a Acceleration LT−2
(5.8)
equal to 1. Therefore,
M: 1 − γ + δ = 0,
(5.9a)
L: 2 + β − 3γ + δ = 0,
(5.9b)
T: −3 − 2β + 2γ = 0,
(5.9c)
and
Q: α + 2γ − 2δ = 0.
(5.9d)
The solution of (5.9) is: α = − 2, β = − 2, γ = − 1/2, and
δ = − 3/2. Therefore, is dimensionless and so
(5.10)
where C is an undetermined dimensionless number. Note that in (5.10) we
have replaced μo with c via . J. J. Larmor first derived (5.10)
with C = (6π)−1 in 1897.
Larmor’s result indicates that the planetary model of the hydrogen
atom, for a short time in vogue after Rutherford’s 1911 discovery of the
atomic nucleus, is untenable. For the electron, in this case, would
continually radiate energy and, therefore, continually descend into the
nucleus. Niels Bohr was aware of this problem when, in 1913, he devised
quantum conditions that not only saved the stability of the hydrogen atom
but also predicted its emission and absorption spectra.
5.4 Example: Child’s Law
The terms anode and cathode denote two parts of a vacuum tube, the simplest
geometry of which is illustrated in Figure 5.1. Two parallel, metallic plates
with surface area A are separated by a distance s . One of the metallic
plates is charged negative (the cathode) and the other positive (the anode) with
a potential difference ΔV between them. A flow of electrons of current density
J is emitted from the cathode and accelerated toward the anode.
Table 5.3
Jmax Maximum current QT−1L−2
density
s Plate separation L
e Electron charge Q
me Electron mass M
(5.11)
be equal to 1. Thus,
C: 1 − β + 2γ + δ,
(5.12a)
T: −1 − 2β − 2γ = 0,
(5.12b)
L: −2 + α + 2β − 3γ = 0,
(5.12c)
and
M: β − γ + ε = 0.
(5.12d)
These constraints are solved by β = − 7/2 + α, γ = − 3 + α,
δ = 3/2 − α, and ε = 1/2. The two dimensionless products produced,
and sΔVεo/e, are related by
(5.13)
(5.14)
Table 5.4
ωp Plasma frequency T−1
e Electron charge Q
me Electron mass M
(5.15)
Because this is not a heat transfer problem but rather one in which the
magnetic field compresses and heats the plasma, the dimension H is not
independent of M, L, and T. Boltzmann’s constant kB and the vacuum
permittivity μo should be included among the dimensional variables and
constants. Furthermore, since we are concerned only with the fully pinched
and balanced configuration, we need include neither a compression time nor
the inertia of the plasma particles. (But see Problem 5.5.) Table 5.5 collects
these variables and constants, their descriptions, and their dimensional
formulae.
Table 5.5
I Current QT−1
R Radius L
NL Density per unit length L−1
T Temperature Θ
(5.16)
implies that
Q: 1 − 2δ = 0,
(5.17a)
T: −1 − 2γ = 0,
(5.17b)
L: α − β + 2γ + δ = 0,
(5.17c)
and
M: γ + δ = 0.
(5.17d)
The solution of (5.17) is β = α − 1/2, γ = − 1/2, and
δ = 1/2. Note that only two of the three equations (5.17a), (5.17b), and
(5.17d) are linearly independent. Evidently, only 3 dimensions are effective.
An inspection of the table reveals these to be QT−1, MT−2, and L or
alternatively and equivalently QT−1, MLT−2, and L. These latter identify the
total current QT−1 and the force MLT−2, that is, F, as effective dimensions.
With enough foresight we could we have adopted these as imposed
dimensions.
The dimensionless products assume the forms and
RNL. Therefore,
(5.18)
5.4 Plasma oscillation. The context is Section 5.6. Use the Rayleigh
algorithm to show that only one dimensionless product is
produced from the dimensional variables and constants ωp, me, εo, ne, and e.
5.5 Pinch effect compression time. The context is Section 5.7. Use the
Rayleigh algorithm to determine an expression for the time Δt required for a
current I to cause a “cold,” that is, a zero-pressure, plasma of initial radius
Ro composed of ions of mass m and density per unit length NL to become
fully compressed. (Include the vacuum permeability μo among the
dimensional variables and constants.)
answer
5.6 Pinch effect. Again, the context is Section 5.7 on the pinch effect.
Reformulate the table listing the symbols, names, and dimensional formula
of 5 dimensional variables and constants, I, R, NL, kBT, and μo. In doing so
express the dimensional formulae in terms of 3 effective dimensions: a
current QT−1, a pressure per unit length ML−2T−2, and a length L.
answer
Table 6.1
E/V Energy density ML−1T−2
M: 1 + α + β = 0,
(6.3a)
L: −1 + 2α + 2β = 0,
(6.3b)
and
T: −2 − α − 2β − γ = 0.
(6.3c)
Equations (6.3) are solved by α = 3, β = − 4, and γ = 3.
Therefore, the product (E/V)(hc)3(kBT)−4 is dimensionless and
(6.4)
Table 6.2
(dE/dν)V−1 Spectral energy ML−1T−1
density
ν Frequency T−1
(6.5)
implies
M: 1 + β + γ = 0,
(6.6a)
L: −1 + 2β + 2γ + δ = 0,
(6.6b)
and
T: −1 − α − 2β − γ − δ = 0.
(6.6c)
These are solved by β = − 3 − α, γ = 2 + α, Δ = 3. Therefore,
(dE/dν)V−1h2c3(kBT)−3 and hν/kBT are dimensionless and so
(6.7)
(6.8)
(6.10)
(6.11)
This result relates the function f(x) and the value of the number C, which
otherwise remain undetermined. A more detailed analysis finds that
f(x) = x3/(ex − 1) and so, given (6.11), C = 8π5/15.
6.4 Example: Bohr Model
Niels Bohr (1885–1962) used Planck’s constant h to forge a new
understanding of the hydrogen atom. According to the model Bohr
proposed in 1913, the hydrogen atom is composed of an electron, with mass
me and charge e, orbiting, in specially defined non-radiating orbits, a proton
with mass mp(=1836me). Since the electron and proton are bound together
electrostatically, the vacuum permittivity εo helps determine these orbits.
Because Bohr’s orbits are, by hypothesis, non-radiating and we assume the
electron speed in a hydrogen atom is non-relativistic, the speed of light c is
not relevant. Therefore, the dimensional constants me, e, εo, and h complete
the list of those that describe Bohr’s hydrogen atom.
Bohr’s model determines, among other quantities, a ground state
orbital radius r1 that characterizes the hydrogen atom’s spatial extent.
Dimensional analysis determines how r1 depends on me, e, εo, and h. Table
6.3 collects these symbols, their descriptions, and their dimensional
formulae.
Table 6.3
r1 Bohr radius L
me Electron mass M
e Electron charge Q
(6.12)
implies that
L: 1 − 3γ + 2δ = 0,
(6.13a)
M: α − γ + δ = 0,
(6.13b)
Q: β + 2γ = 0,
(6.13c)
and
T: 2γ − δ = 0.
(6.13d)
The solutions to equations (6.13) are α = 1, β = 2, γ = − 1,
and δ = − 2. Therefore, r1mee2/(εoh2) is dimensionless and
(6.14)
(6.15)
with C′ = 1. While this result suggests that we should use 4πεo in place of εo
and use ℏ in place of h, we have no reason to believe that this tactic will
always produce proportionality constants equal to 1.
6.5 Atomic Units
The dimensional constants me, e, εo, and h enter into all models in which
mechanics, electrostatics, and quantum physics play a role. Because atomic
physics draws on these subjects, εoh2/e2me (proportional to the Bohr radius)
characterizes lengths in atomic physics. Similarly, the time (see
Problem 6.2) and, of course, the mass me and charge e similarly
characterize atomic physics.
The rule of thumb NP = NV − ND suggests and the Rayleigh algorithm
confirms that only one length can be fashioned out of the 4 dimensional
constants, me, e, εo, and h, whose dimensional formulae are expressed in
terms of the 4 imposed dimensions, M, L, T, and Q. Similarly, only one time
follows from these same 4 dimensional constants. In all, the 4 dimensional
fundamental constants, me, e, εo, and h, produce 4 unique fundamental,
characteristic quantities: a length εoh2/e2me, a time , a mass me,
and a charge e. These compose what we call a characterizing system of
units or a scale – a scale of atomic-sized units. The elements of this scale,
their formulae, and their values to three places in SI units are listed in Table
6.4.
Table 6.4
atomic scale
Because atomic physicists exert much effort calculating with me, e, εo,
and h, they sometimes take the shortcut of setting each of these to unity,
that is, so that me = 1, e = 1, εo = 1, and h = 1. The effect of this shortcut is
that all masses produced in this way are in units of me. For instance,
mp = 1836 means mp = 1836me. Furthermore, all lengths are in units of
εoh2/e2me, all times are in units of , and all charges are in units of
e.
6.6 Example: Quantum Ideal Gas
A gas is ideal if its particles do not interact with one another at a distance or
through a field. We already know that the equation of state of a classical
ideal gas p = (N/V)kBT relates its pressure p, its particle density N/V, and its
absolute temperature T where kB is Boltzmann’s constant. How do quantum
effects modify this equation of state?
Certainly Planck’s constant h should enter into any quantum version of
p = (N/V)kBT. But the Rayleigh algorithm applied to the 4 dimensional
variables and constants p, N/V, kBT, and h that are denominated in terms of
the 3 imposed dimensions M, L, and T generates exactly one dimensionless
product pV/(NkBT).
Possibly we have ignored some dimensional variable or constant that
effectively brings Planck’s constant h into play. Since, by definition, no
force fields play a role, the dimensional constants G, εo, μo, and e are not
involved. Because we limit our model particles to non-relativistic speeds,
neither is the speed of light c. It seems that only the mass m of the ideal gas
particles remains to be considered. In retrospect, it may surprise us that m
does not enter directly into the classical equation of state p = (N/V)kBT. (But
review Section 1.7.)
Therefore, we allow the pressure p of a quantum ideal gas to depend
on N/V, kBT, h, and m. These 5 symbols, their descriptions, and their
dimensional formulae are collected in Table 6.5.
Table 6.5
p Pressure ML−1T−2
N/V Number density L−3
m Particle mass M
(6.16)
we have
M: 1 + β + γ + δ = 0,
(6.17a)
L: −1 − 3α + 2β + 2γ = 0,
(6.17b)
and
T: −2 − 2β − γ = 0.
(6.17c)
The solution to (6.17) is β = − 5/2 − 3α/2, γ = 3 + 3α, and
δ = − 3/2 − 3α/2. Therefore, ph3/[m3/2(kBT)5/2] and (N/V)h3/(mkBT)3/2 are
the 2 dimensionless products. Multiplying the first by the inverse of the
second produces pV/(NkBT). Therefore, we may write
(6.18)
(6.19)
where C and n are undetermined. Our physics sense takes us a little further.
Because we expect that as T → 0, the pressure p should not diverge, it must
be that n ≤ 2/3.
In the special case n = 2/3 (6.19) becomes
(6.20)
(6.21)
Table 6.6
P Power radiated ML2T−3
q Charge Q
a Acceleration LT−2
(6.22)
implies that
M: 1 − γ + ε = 0,
(6.23a)
L: 2 + β − 3γ + δ + 2ε = 0,
(6.23b)
T: −3 − 2β + 2γ − δ − ε = 0,
(6.23c)
and
Q: α + 2γ = 0.
(6.23d)
The solution of (6.23) is: α = − 2, β = − 2, γ = − α/2,
δ = 2 − α/2, and ε = − 1 − α/2. In this way we find the two dimensionless
products to be Pc2/a2h and . By multiplying the first by the inverse
square of the second, we eliminate h and produce, as expected, Pεoc3/(q2a2).
The square of the second independent, dimensionless product is q2/εoch.
Thus, we may write
(6.24)
6.3 Ideal gas. Show that, as claimed in Section 6.6, the Rayleigh algorithm
applied to the dimensional variables and constants P, N/V, kBT, and h
generates one dimensionless product PV/(NkBT).
7
Dimensional Cosmology
◈
7.1 The Fundamental Dimensional
Constants
Each subject we have explored thus far, mechanics and the mechanics of
fluids, thermal physics, electrodynamics, and quantum physics, has
introduced one or more fundamental dimensional constants: the
gravitational constant G, Boltzmann’s constant kB, the vacuum permittivity
εo, the speed of light c, the electron charge e, the electron mass me, and
Planck’s constant h. These subjects, considered expansively, encompass
most of what we know of the physical world, and these fundamental
constants quantify that knowledge. These seven constants, their descriptive
names and symbols, their numerical values to three places in SI units, and
their dimensional formulae are collected in Table 7.1.a
Table 7.1
Name Symbol Value in SI Dimensional
units Formula
(7.1)
(7.2)
or or . We may, for instance, substitute e2/
εo for in any term and preserve that term’s dimensional formula. Of
course, in doing so we shift its numerical value by a factor of 1042.
7.3 Example: Gravitational-Electrostatic
Mass
What quantity with the dimension of mass can be composed out of the three
fundamental constants e, εo, and G? The dimensional equivalence (7.2), that
is, , contains the answer. For if , then
, [e2/εo] = [G][me]2, and so . Therefore,
the quantity has the dimension of mass – the mass of a spherical
particle with charge e whose gravitational self-attraction is large enough to
balance its electrostatic self-repulsion. I call the gravitational-
electrostatic mass.
7.4 Example: Classical Electron Radius
What quantity having the dimension of length can be composed out of the
fundamental constants me, e, εo, and c? Since the Eddington-Dirac ratio
does not contain the speed of light c, this question cannot be answered by
using the dimensional equivalence (7.2) based on that dimensionless
product.
Instead, denote the length sought with l. Given that the 5 dimensional
quantities, l, me, e, εo, and c, are expressed in terms of 4 imposed
dimensions, M, L, T, and Q, as shown in Table 7.2, the Raleigh algorithm
applied to these quantities should produce 1 dimensionless product.
Table 7.2
l Characteristic length L
me Electron mass M
e Electron charge Q
Table 7.3
classical scale
Other classical masses, lengths, durations, and charges can be formed in similar fashion, that is,
by applying the Rayleigh algorithm to other sets of 4 dimensional constants taken from the 5 classical
ones: G, εo, c, e, and me. There are in all 5 ! /4 ! 1! or 5 such classical scales. These are collected in the
first five rows of Table 7.4 in Section 7.7.
Table 7.4
Constants Mass Length (meters) Time (seconds) Charge
(kilograms) (coulombs)
Constants Mass Length (meters) Time (seconds) Charge
(kilograms) (coulombs)
G , me , e , εo me ∼ 10−30 e ∼ 10−19
G , me , εo , c me ∼ 10−30
G , me , e , c me ∼ 10−30 e ∼ 10−19
G , e , εo , c e ∼ 10−19
me , e , εo , c me ∼ 10−30 e ∼ 10−19
classical
scale
G , me , e , h me ∼ 10−30 e ∼ 10−19
G , me , εo , h me ∼ 10−30
G , me , c , h me ∼ 10−30
G , e , εo , h e ∼ 10−19
me , e , εo , h me ∼ 10−30 e ∼ 10−19
atomic
scale
G,e,c,h e ∼ 10−19
me,e,c,h
Constants me ∼ 10−30
Mass Length (meters) Time (seconds) e ∼ 10−19
Charge
(kilograms) (coulombs)
G , εo , c , h
Planck scale
me , εo , c , h me ∼ 10−30
e , εo , c , h e ∼ 10−19
7.6 Fine Structure Constant
Restoring Planck’s constant h to the classical, dimensional constants (and
again excluding Boltzmann’s constant kB) composes a set of 6 fundamental,
dimensional constants: G, εo, c, e, me, and h. Since their dimensional
formulae are expressed in terms of 4 imposed dimensions, M, L, T, and Q,
they can be combined into 2 dimensionless products. One of these, we
know, is proportional to the Eddington-Dirac ratio . The
other, which we met for the first time in Section 6.7, is e2/(εohc). The
particular form e2/(2εohc) is called the fine structure constant.
Arnold Sommerfeld (1868–1951) observed that the ratio of the
velocity of an electron in the first Bohr orbit to the speed of light is the fine
structure constant. But the fine structure constant also carries with it another
meaning. Because it is proportional to the ratio of the potential energy of
two electrons separated by a distance d, that is, e2/(4πεod), to the energy of
a photon with wavelength d, that is, hc/d, the fine structure constant
regulates the interaction of electrons and photons. Indeed, e2/(2εohc)
characterizes a generalization of Maxwell’s equations, called quantum
electrodynamics or QED, in which electrons and photons interact.
The value of the fine structure constant was first thought to be exactly
1/137. Consequently, a number of physicists, including Wolfgang Pauli
(1900–1958), sought the meaning of this apparently special integer 137. As
it happened, Pauli was unsuccessful in this effort and balefully pointed out
the number of his hospital room, #137, to a visitor shortly before his death.
We now know that the fine structure constant is only approximately 1/137.
The current value of its inverse is, to 12 places, 137.035999139.
Nevertheless, since the equation
(7.5)
is accurate to three places, the number 137 is a handy way to remember the
value of the fine structure constant.
No one has been able to derive numerical values of either the
Eddington-Dirac ratio or the fine structure constant. Instead, these
dimensionless products or their constituent dimensional constants must be
measured. Yet there are physicists who believe the task of physics will not
be complete until the value of all the fundamental dimensional constants
can be derived from some very deep principle – now unknown to us.
The fine structure constant e2/(2εohc) provides a way of generating
dimensional quantities with the dimensions of M, L, T, and Q that contain
Planck’s constant h from quantities that do not. One simply uses
dimensional equivalences like
[e2/εo] = [hc]
(7.6)
or [e2] = [εohc] or [c] = [e2/hεo] to transform classical
quantities into quantum ones that contain Planck’s constant h. See, for
example, Section 7.9.
7.7 The 15 Scales
From the set of 6 fundamental dimensional constants, G, εo, c, e, me, and h,
4 can be taken, without regard to order, in 6 ! /4 ! 2! or 15 different ways.
Employing the Rayleigh algorithm on each of these 15 sets of 4 constants,
we derive quantities with the dimensions of M, L, T, and Q.
Of course, the Rayleigh algorithm produces a particular quantity only
when that quantity exists. And occasionally it does not. For instance,
because the dimensional formulae of the 4 constants G, me, c, and h do not
contain the dimension Q, the Rayleigh algorithm cannot generate, from
these constants, a quantity with the dimension Q. Therefore, this particular
set of 4 possible dimensional quantities remains incomplete.
Each of the 15 sets of dimensional quantities derived from 4
fundamental dimensional constants composes a scale. The 15 scales,
classical and quantum, complete and incomplete, are listed in Table 7.4.
Three of them have been given special names: the classical scale, based on
me, e, εo, and c, the atomic scale based on me, e, εo, and h, and the Planck
scale based on G, εo, c, and h. Attending each dimensional combination is
its value in SI units to the nearest order of magnitude.d,e,f
7.8 Planck Scale
Max Planck (1858–1947) was quite taken with the “naturalness” of what
later were called Planck scale units. [25]g Unlike SI or Gaussian units or
other units in widespread use, Planck units do not depend on the properties
of arbitrarily chosen, if universally recognized, measures like the kilogram,
nor do they depend on the properties of any one particle. The dimensional
formulae and the values of quantities that make up the Planck scale are
listed in Table 7.5. The Planck scale is the only one of the 15 possible ones
in which Planck’s constant h appears in all four characteristic quantities.
Table 7.5
Planck scale
One is drawn to especially large and small magnitudes. For instance, the
7.2 Fine structure constant. Show that the fine structure constant
e2/(2ε0hc) is dimensionless by using the dimensional formulae in Table 7.1
of Section 7.1.
7.3 Planck scale. Derive the Planck length, time, and charge from the
classical electron length h/mec, time h/mec2, and charge e by substituting
dimensional equivalences built into the Eddington-Dirac number
and the fine structure constant e2/(2hc) as illustrated in
Section 7.9.
7.4 Electron scales? Speed of light scale? According to Section 7.8, “The
Planck scale is the only one of the 15 in which Planck’s constant h appears
in all four characteristic quantities.” But there other scales in which a
particular fundamental constant appears in all four characteristic quantities.
(a) Identify two scales that could be called an “electron mass scale” because
me appears in all four characteristic quantities.
answer
(b) Identify two scales that could be called an “electron charge scale.”
answer
(c) Identify one scale that could be called a “speed of light scale.”
answer
cOne could always add to the classical or any other scale consisting of a
mass, length, and time, a characteristic temperature defined by a quantity
with the dimensional formula of energy ML2T−2 divided by Boltzmann’s
constant kB.
1.2 f = C ⋅ υ/r.
1.3 a = C ⋅ υ2/r.
1.5 .
2.3 .
2.4 .
2.5 .
2.9 R = C ⋅ E1/5t2/5/ρ1/5.
3.2 .
3.4 t = C ⋅ μr2/mg.
3.6 m = C ⋅ ρυ6/g3.
5.1 .
5.2 (a) The dimensionless products are Ed2εo/q and υ/c. Therefore,
E = (q/d2εo) ⋅ f(υ/c) where f(x) is undetermined. (b) The dimensionless
products are Bd2εoc/q and υ/c. Therefore, B = (q/(d2εoc)) ⋅ g(υ/c) where
g(x) is undetermined.
5.3 and
.
5.5 (a) The dimensionless products are mΔt2/(μoI2) and NLR. (b)
. In typical pinches NLR > > 1 and
.
5.6 (a)
I Current (QT−1)
R Radius L
6.1 .
7.4 (a) “Electron mass scales” are based on G, me, εo, and c, and on G,
me, εo, and h. (b) “Electron charge scales” are based on G, e, εo, and c,
and on G, e, εo, and h. (c) A “speed of light scale” is based on G, εo, c,
and h. The “speed of light scale” is identical to the Planck scale.
References
10. The concept, although not the name, of imposed dimension originates
with Percy Bridgman’s highly recommended text Dimensional Analysis
(New Haven, CT: Yale University Press, 1922). See, in particular, pp. 9–11,
63–66, 67–69, and 77–78. Others, including H. E. Huntley, Dimensional
Analysis, (Mineola, NY: Dover, 1967), use the concept of imposed
dimensions.
16. See Aristotle, The Basic Works of Aristotle editor Richard McKeon,
(New York, NY: Random House, 1966) Physics, Book IV, chapter 8, p.
216a, lines 14–17. See also David C. Lindberg, The Beginnings of Western
Science (Chicago, IL: University of Chicago Press, 1992), pp. 59–60.
17. R. P. Godwin, The Hydraulic Jump (‘Shocks’ and Viscous Flow in the
Kitchen Sink), American Journal of Physics 61 (9), (1993) 829–832.
18. Fridtjof Nansen, Farthest North (Edinburgh, UK: Birlinn, 2002), p. 186.
19. C. Vuik, Some Historical Notes on the Stefan Problem, Nieuw Archief
voor Wiskunde, 4e serie 11 (2), (1993) 157–167.
22. Lloyd W. Taylor, Physics: The Pioneer Science (Mineola, NY: Dover,
1941), pp. 592–593.
24. L. Schiff, Quantum Mechanics 3rd Edition (New York, NY: McGraw
Hill, 1968) pp. 397 ff.
25. Max Planck, Theory of Heat Radiation (Mineola, NY: Dover, 2011), pp.
205–206. Number 164.
Index
Ampere-Maxwell law, 62
Analytical Theory of Heat, The, 50
arctic ice, growth of, 55–57
Aristotle (384-322 BCE), 39
asymptotic expansion, 24–25, 59
atomic scale, 79–80, 92
atomic time, 84
calorimetry, 49
capillary waves, 36–37, 47
Carnot, Sadi (1796-1832), 49
centripetal acceleration, 15
charge density, 61–62
Child, C. D. (1868–1933), 66
Child’s law, 66–68, 73
classical electron radius, 87–88
classical scale(s), 88–89, 91
classical time, 86
Compton wavelength, 93–94
Compton, A. H. (1892–1962), 74
convection, 51, 60
cooking a turkey, 51–53
current density, 61–62, 66
deep-water wave, 35
degeneracy pressure, 82
diffusion equation, 54
dimensional constants
fundamental, 85
table of, 85
dimensional cosmology, 85
dimensional formulae, 4–5, 8, 17, 19–20
dimensional homogeneity, 1–3, 5
dimensional model, 11
dimensionless product, 3–4, 6, 48
complete, 8
independent, 8
dimensionless products
number of, 9
dimensions
absolute vs. relative, 22
effective, 17–20
imposed, 21–22
number of, 8–9
Dirac, Paul (1902–1984), 86
displacement current, 62
Galileo (1584–1642), 1
geomagnetic field, 63
gravitational constant, 15, 31, 85
gravitational instability, 31
gravitational-electrostatic charge, 87
gravitational-electrostatic mass, 87
gravity wave, 37
ideal fluid, 32
ideal gas, 9–11, 19, 27, 59, 80, 84
quantized, 80–82
ideal gas constant, 59
imposed dimension, 20–21, 30
ionosphere, 68
laminar flow, 42
latent heat, 49
light waves, 62
Lorentz force law, 61
quantum electrodynamics, 89
quantum physics, 74
USDA data, 53
vacuum permeability, 61
vacuum permittivity, 61
vacuum tube, 66
variable of interest, 7, 10, 24, 26
virtual cathode, 68
viscosity, 39, 48
z-pinch, 70