You are on page 1of 36

1 Introduction to Scaling Laws

There are many different scaling laws. At one extreme, there are simple scaling laws that are easy to learn, easy to use, and very useful in everyday life. Scaling laws can be and should be introduced at the elementary-school level, and then reinforced and extended every year through middle school, high school, and beyond. Scaling laws are central to physics. This has been true since Day One of modern science. Galileo presented several important scaling results in 1638 (reference 1 orreference 2). This document is meant to be a tutorial, covering the simplest and most broadly useful scaling laws.
At the other extreme, there are more subtle scaling laws that are used to solve very deep and complicated problems at the frontiers of scientific research. The importance of scaling continues to the present day. There are dozens of references to scaling at the Nobel Prize site (reference 3). Dont let this scare you away. To repeat: this document is meant to be a tutorial, covering the simplest and most broadly useful scaling laws. You dont need to be a Nobel laureate to get a lot of value from scaling laws.

You may be familiar with a simple form of scaling in connection with scale models, such as in figure 1 and figure 2. (The two figures are the same, except that one has a larger scale than the other.) The same word shows up in connection with small-scale and large-scale maps. For more about the terminology, see section 2.

Figure 1: Small-Scale Model Train

Figure 2: Large-Scale Model Train

* Contents
1 Introduction to Scaling Laws 1.1 Area versus Length 1.2 Volume, Area, and Length in Three Dimensions 1.3 Non-Integer Scaling 1.4 Polygons, Polyhedra, etc. 1.5 The Pythagorean Theorem via Scaling 1.5.1 Proof 1.5.2 Discussion

1.6 Scaling One Thing but Not Another 2 Terminology 2.1 Examples of Good and Bad Terminology 2.2 ... k times greater than ... 2.3 Additional Terminology 3 Stiffness and Compliance 4 Breaking Strength 5 Finding the Scaling Law 5.1 Some Things Scale, and Some Dont 5.2 Keep Looking 6 Miscellaneous Examples of Scaling Laws 7 Non-Dimensional Scaling 7.1 Example: Apparent Horizon 7.2 Example: Mean Free Path 7.3 Example: Equilibrium and Activity 7.3.1 Introduction 7.3.2 Analysis 8 Fractals 9 References 1.1 Area versus Length

Perhaps the best known scaling law pertains to the relationship between length and area. In figure 3, when it comes to length, every length in the large square is twice as great as the corresponding length in the small square. When it comes to area, you can see that the area of the large square is not twice as great, but rather four times as great as the area of the small square.

Figure 3: Small and Large Squares

Now lets see what happens if we scale up the lengths by a factor of 3 rather than 2. In figure 4, when it comes to length, every length in the large square is three times as great as the corresponding length in the small square. When it comes to area, the area of the large square is nine times as great as the area of the small square.

Figure 4: Small and Larger Squares

The general rule here is the so-called square law:


( r a t i o o f a r e a s ) = ( r a t i o o f l e

n g t h s )
2

( 1 )

In words, the ratio of areas goes like the second power of the ratio of lengths. Equivalently, you can say that the ratio of areas goes like the square of the ratio of lengths. Continuing to examine the relationship between length and area, lets see what happens with triangles. In figure 5, when it comes to length, every length in the large triangle is twice as great as the corresponding length in the small triangle. So the question is, how does the area of the large triangle compare to the area of the small triangle?

Figure 5: Small and Large Triangles : Question

The right answer is that the area of the large triangle is four times the area of the small triangle, as you can see from figure 6.

Figure 6: Small and Large Triangles

So you can see that the square law in equation 1 is absolutely not limited to squares. In fact, squareness has got nothing to do with it. The key idea here is that areas depend on width times height. So if the width is scaled up by a factor of K and the height is scaled up by a factor of K, the area necessarily gets scaled up by a factor of K2.
1.2 Volume, Area, and Length in Three Dimensions

We can extend this idea into three dimensions, as shown in figure 7. When it comes to length, every length in the large cube is two times as great as the corresponding length in the small cube. When it comes to area, the surface area of the large cube is four times as great as the surface area of the small cube. When it comes to volume, the volume of the large cube is eight times as great as the volume of the small cube.

Figure 7: Small and Large Cubes

The volume goes up by a factor of eight because the cube is twice as wide and twice as deep and twice as tall. Thats three factors of two. Meanwhile the surface area of the cube went up by a factor of four, not a factor of eight. Thats because on each face of the cube, the surface goes like width times height; the surface does not have any thickness. Each face of the cube is locally two dimensional, even though if you put all six faces together the total surface extends across three dimensions. Reconciling these facts requires a sophisticated notion of

dimensionality. We say that the surface is intrinsically two-dimensional but isembedded in three dimensions. The scaling laws for volume, area, and length can be expressed in terms of equations:
(ratio of areas) (ratio of volumes)

( 2 )

and if you want to get fancy, this can be expressed in one big equation
( r a t i o o f v o l u m e s )
( 1 / 3 )

( r a t i o o f a r e a s )
( 1 / 2 )

= ( r a t i o o f l e n g t h s )
1

( 3 )

In equation 3, if a quantity has extent in N spatial dimensions, we take the Nth root.
1.3 Non-Integer Scaling

It is straightforward to scale things by factors that are not integers, as illustrated in figure 8. When it comes to length, every length in the large square is 1.5 times as great as the corresponding length in the original square. When it comes to area, the area of the large square is 2.25 times as great as the area of the original square. You can confirm this graphically by counting squares on the far right of the diagram. Each of the tiny squares has 1/4th of the area of the original square. You can see that the large square has 9/4ths of the area of the original square.

Figure 8: Small and Largeish Squares

You can understand the factor of 2.25 in several ways. One way is to take the ratio of lengths (1.5) and multiply it by itself using long multiplication. You dont even need to do long multiplication if you remember that 15 squared is 225. Another way is to recognize 1.5 as 3/2, which you can square in your head to get 9/4, which reduces to 2.25. This corresponds to scaling up the lengths by a factor of 3 and then scaling down by a factor of 2. Of course the same thing works for triangles. When it comes to length, every length in the large triangle is 1.5 times as great as the corresponding length in the original triangle. When it comes to area, the area of the large triangle is 9/4ths as great as the area of the original triangle.

Figure 9: Small and Largeish Triangles 1.4 Polygons, Polyhedra, etc.

The scaling law for triangles is even more useful than the scaling law for squares, because it is easy to divide any polygon into triangles. By applying the scaling law to each of the triangles, you can easily prove that the scaling law must apply to any polygon. Going one step further down that road, note that the area and perimeter of any ordinary (non-fractal) plane figure can be approximated as closely as you want by a polygon. For example, a circle is (for most purposes) well approximated by a manysided regular polygon. In this way you can convince yourself that equation 1applies to a very wide class of figures. The same argument applies in three dimensions. First, convince yourself that equation 2 and equation 3 apply to any tetrahedron. Then note that any polyhedron can be divided into tetrahedra. Conclude that equation 2 and equation 3 apply to any polyhedron.
1.5 The Pythagorean Theorem via Scaling

1.5.1 Proof We start with the triangle ABC as shown in figure 10. We construct line CD perpendicular to AB.

Figure 10: Right Triangle

Now we have three triangles: the lower triangle ACD which has hypotenuse b, the upper triangle CBD which has hypotenuse a, and the whole triangle ABC which has

hypotenuse c. You can easily show1 that these are all similar to each other, so scaling arguments apply. The upper triangle is a scaled-down copy of the whole triangle, with each length scaled by a factor of a/c. Similarly the lower triangle has each length scaled by a factor of b/c. The whole triangle, naturally, is scaled relative to itself by a factor of c/c = 1. Area scales like length squared. You can also see that the area of the whole triangle is exactly covered by the two smaller triangles, which means:
( a

( 4 )

1.5.2 Discussion Equation 4 says that the area of the upper triangle plus the area of the lower triangle is equal to the area of the whole thing. The first expresses this in dimensionless terms, using the area of the whole triangle as the unit of area. The second line expresses the same thing using conventional dimensions of area. The second line of equation 4 is the conventional way of stating the Pythagorean theorem. The first line is another way of expressing the same idea, and can also be interpreted as a thinly-disguised version of the trigonometric identity:
s i n
2

+ c o s

= 1

( 5 )

Here is another way to look at this proof: Whereas the traditional pictorial version of the theorem erects a square on the outside of each side of the triangle, we erect similar right triangles on each side of the original triangle. The scaling idea is the same no matter what shape we attach to the edges of the triangle, so we could use pentagons, as shown in figure 11: the area of the a-pentagon plus the area of the b-pentagon is equal to the area of the c-pentagon. We can generalize this to any shape, so long as the three attached shapes are all similar to each other.

Figure 11: Right Triangle with Pentagons

This can be considered a generalization relative to the usual statement of the theorem. This point was made by Euclid, 2300 years ago. See Proof #7 in reference 4, which is a collection of proofs of the Pythagorean theorem. The theorem can be applied to any shape, but the theorem is most easily proved using triangles, as infigure 10. I really like the proof shown in figure 10, because it is easy and well-nigh unforgettable, it yields additional insights beyond what was required, and the method is transferable to a host of other problems (although I rate the dot product proof slightly higher, based on the same criteria). Tangential remark: If youve never seen the dot product proof, here it is: For any nonzero vectors a and b in an Euclidean vector space, let c = a+b. Then
c

6 )

Two nonzero vectors are called perpendicular if and only if their dot product vanishes. This is the definition of perpendicular in this sort of vector space.
1.6 Scaling One Thing but Not Another

Consider the left half of figure 12. We see a circle inscribed in a square. The circle has area a2 while the square has area 4 a2. That means that compared to the square, the circle has /4ths of the area.

Figure 12: Scaling Y but Not X

Now let us turn our attention to the right half of figure 12. We see an ellipse inscribed in a rectangle, in the ordinary symmetric way. The right half of the figure is the same as the left half, except that the vertical lengths have been scaled by a factor of , such that b = a. (The horizontal lengths are unchanged.) The question is, what about the area? If we had scaled all of the lengths (horizontal and vertical) by a factor of , then the area would be scaled by a factor of 2 but that is not the case here. In fact we scaled one length and not the other, so we pick up only one factor of . Therefore we expect the rectangle to have area 4 a2 i.e. 4 a b. This result can be verified by applying the elementary formula for the area of a rectangle. As for the area of the ellipse, using the same scaling argument, we expect it to be scaled relative to the circle by one factor of . That is, we expect the area of the ellipse to be a2 i.e. a b. This result can be verified if you happen to remember the formula for the area of an ellipse ( times the semi-major axis times the semi-minor axis). Note that the dimensionless ratio /4 is unchanged by the scaling as we go from left to right in figure 12. That means that compared to the square, the circle has /4thsof the area ... and by the same token, compared to the rectangle, the ellipse has /4ths of the area.
a= = a rr e4 e a a o o ff c e

il rl c i lp e s e a ra e r a e a o fo ( f 7 s ) q r u e a c rt e a n g l e

If you want an absolutely rigorous proof of this result, you can express the area in terms of an integral, and then do a change of variable. The resulting factor of b/athen pops out in front of the integral, since an integral is a linear operator. That means you dont even need to evaluate the integral in order to obtain the scaling result. A more intuitive explanation for why this works has two parts: First, there is an areato-area scaling rule: We expect the area of one thing to scale like the first power of the area of another thing. So when we go from left to right in figure 12, we scale the area of the square by a factor of b/a, and also scale the area of the circle by a factor of b/a. As a corollary, since we have pairs of areas, the ratio of areas is a dimensionless number (/4ths for these particular pairs) and this dimensionless ratio scales like the zeroth power of the scale factor. That is, whenever a scale factor appears in a numerator in equation 7, it is canceled by an identical scale factor in the corresponding denominator. Tangential remark: Reference 5 uses figure 12 as the starting point for a discussion of the relationship between memorizing things and rederiving things.

2 Terminology
2.1 Examples of Good and Bad Terminology

Part of the skill in making good scaling arguments involves using good terminology and avoiding bad terminology.

Here is some bad terminology: In figure 6, for example, it is dangerous to say that the big triangle is twice as large, or twice as big, or has twice the size of the small square. Thats because depending on the context, largeness and bigness and size could refer to the lengths involved, or could equally well refer to areas or volumes. Example: In a very practical sense, a two-liter bottle of water is twice as big as a one-liter bottle, even though its height and diameter are only 1.26 times as big. This comes a shock to some people, and it is certainly inconvenient, but if you want to be careful you must avoid any unqualified statement about largeness, bigness, size, et cetera. If you mean length, say length. If you mean area, say area. And so forth. As an example, note that in section 2.2, we do not compare object A to object B; we compare the volume of A to the volume of B. Here is some good terminology: People who are adept at making scaling arguments say things like this routinely:

The area scales like the square of the length. The volume scales like the cube of the length. The volume scales like the three-halves power of the area.

The term scale factor is treacherous. Sometimes it refers to the ratio of lengths, but not always. On maps, it refers to the inverse ratio of lengths. That means that a largescale map has a small scale factor (such as 50 miles to the inch), while a small-scale map has a large scale factor (such as 500 miles to the inch). This is a source of endless confusion among non-experts, and even experts get it wrong sometimes. I try to avoid the term scale factor whenever possible. 2.2 ... k times greater than ... The phrase ... k times greater than ... is so problematic that it deserves detailed discussion. Consider the following scenario: VA denotes the volume of object A, while VB denotes the volume of object B. We are given that VA = 12 liters and VB = 36 liters. The first three of the following statements are good terminology, but then things go to pot: 1. [OK] VB is three times as great as VA. 2. 3. 4. 5. 6. [OK] B has 3 times the volume of A. [OK] VB is 300% of VA. [OK] VB is 200% greater than VA. [??] VB is greater than VA by a factor of 3. (Equivalently, VB is more than VA by a factor of 3.) [AVOID] VB is three times greater than VA.

7. 8.

[AVOID] VB is two times greater than VA. [AVOID] B is three times as big as A. Discussion: statement 1, statement 2, and statement 3 describe VB in absolute terms as a multiple of VA. In contrast, statement 4 describes VB in relative terms, as an increment relative to the amount of VA we started with; the increment, not the whole amount of VB, is given as a multiple of VA. In statement 5, the word greater (or more) is a comparative adjective, which might have hinted that we are describing VB by means of a relative increment, but the word factor overrides this hint and makes it clear that we are describing VB in absolute (not relative) terms. This may not be entirely logical from a grammatical point of view, but the meaning of this expression is reasonably well established. We deprecate both statement 6 and statement 7, because it is hard to know which of them is correct. The word times suggests that VB is being described in absolute terms, while the comparative adjective greater suggests that VB is being described by means of a relative increment. I cant say which interpretation is correct; some authors assume one interpretation, while others assume the other. I recommend avoiding both versions entirely, and using something like statement 1 or statement 4 instead. A different problem crops up in statement 8. Does the statement mean that every dimension of B (length, width, and height) is bigger by a factor of 3, or does it only mean that the volume is bigger by a factor of 3? When you are scaling some property, you need to be specific about which property you are scaling.
2.3 Additional Terminology

In Euclidean geometry, objects that are the same except for scale and except for isometries (such as translation, rotation, and reflection) are said to be geometrically similar. (Beware: This conflicts with the vernacular use of similar to mean merely approximately alike.) Objects that are

geometrically similar and have the same scale are said to be congruent. The word scale in this context has the same meaning as in terms like scale model, and largescale or small-scale maps. It is etymologically related to such things as the graduated scale on a burette. It is distantly related to the musical scale. It is unrelated to such things as the scales on a fish, or the scales used for weighing things. The simplest forms of scaling are also known as proportional reasoning but more generally, scaling laws go far beyond simple proportionality.

3 Stiffness and Compliance


Stiffness is the inverse of compliance. The meaning of stiffness and compliance can be seen in figure 13. The left end of the beam is held fixed, while the right end is free to move.

Figure 13: Bending a Beam

When we apply a force F, the beam is deflected a distance d relative to its resting position. Then we say the stiffness (or spring constant) is k, where
k F = d

8 )

and the compliance is 1/k. Now the interesting thing is that if the beam has length L, width w, and thickness t, the stiffness scales like w (which is unsurprising), and also scales like the cube oft/L.
t3 L k w ( 9 )

This cube law may comes as a surprise to some people. As discussed in section 4, breaking strength scales like cross-sectional area (in this case, width times thickness), and you might expect that stiffness would scale the same way, but it doesnt. A beam that is twice as thick is vastly stiffer than two beams in parallel, as shown in figure 14.

Figure 14: Bending Two Separate Beams

Look what happens if you bend two beams that are not connected, just placed in parallel. The stiffness scales like the number of beams, i.e. like the total thickness. But note that the free ends of the two beams dont line up, as you can see if you look closely at figure 14.

In contrast, if you have a double-thickness beam or a pair of beams glued together so that they cant slide relative to each other then the ends are forced to line up. That means when you bend it, the top half gets stretched and the bottom half gets compressed. This stretch and compression involves a lot of energy, and contributes greatly to the stiffness. An I-beam is designed to cultivate this effect, as shown in figure 15. If you try to wrap an I-beam around a cylinder of radius R, the inner flange will try to follow a circle of diameter 2R, while the outer flange will try to follow a circle of diameter 2(R+h), where h is the height of the web. Since the two flanges started out the same length, and the web will keep the two flanges lined up, you cant bend the I-beam without compressing the inner flange and stretching the outer flange. This makes an I-beam very much stiffer than a solid bar of comparable weight. See reference 6.

Figure 15: Bending an I-Beam

4 Breaking Strength
Breaking strength (unlike stiffness) scales simply like cross-sectional area. Following Galileo, lets see what this implies about the bones of animals. The weight of the animal scales like linear size cubed. If we just scaled up the bones in proportion, the breaking strength would scale like linear size squared, which is not enough to keep up with the weight. Therefore the bones of a large animal must be not just thicker, but disproportionately thicker.

The thickness must scale like linear size to the three-halves power. Galileo published this result in 1638 (reference 1, page 170 of the National Edition). Figure 16 is a copy of his original drawing, showing a scaled-up picture of a small animals bone, compared to an actual large animals bone. (He doesnt say exactly which animals he is comparing.)

Figure 16: Large Bone and Scaled-Up Small Bone

Galileo also says (ibid) I believe that a little dog might carry on his back two or three dogs of the same size, whereas I doubt if a horse could carry even one horse of his own size.

5 Finding the Scaling Law


5.1 Some Things Scale, and Some Dont

When we say something scales, or is scalable, we mean we can change the scale of the thing and it still makes sense. For example, a triangle is scalable, because if we change the size it is still a perfectly fine triangle. There exist plenty of things that are not scalable. As pointed out in section 4, an elephant is not just a scaled-up elephant-shrew. That means the animal as whole is not scalable. If you look at the skeleton of the shrew, you know what the natural size of the animal must be; if you tried to rescale it substantially, the result wouldnt make sense. There is a natural size-scale for this animal. Skeletal strength is not the only issue; the metabolism of any creature will generate heat in proportion to volume, while the ability to dissipate heat into the environment scales like surface area.

This causes problems for large creatures (too much heat buildup) and for small creatures (too much heat loss). For details, see reference 7and reference 8. If something is scalable, the thing makes sense on many different size-scales; if something is not scalable, there is only one size-scale that makes sense for it. Often when a complex system is not scalable, it is because of a conflict between various subsystems. In the case of animals, the weight scales one way, and the strength of the bones scales another way. Either weight or strength is scalable separately, but when we put them together we get a conflict. Atoms are not scalable. The Bohr radius is a natural size-scale for atoms that is fixed by the fundamental physics. You cant make a scaled-up hydrogen atom unless you monkey with the fundamental physics, and nobody knows how to do that in practice. (The fundamental constants are called constants for a reason. That doesnt prove they are constants, but it means you should think twice before assuming they are easily variable.) When part of the system scales one way and part scales another way, we can often obtain useful scaling laws that work over some part of the domain, often a very large part. For example, as discussed in reference 9, we have a good scaling law for long-wavelength waves and another good scaling law for short-wavelength waves. Each of these laws is valid within its part of the domain; it is only in the crossover region that we see nonscalable behavior. Many scaling laws work exceedingly well over a wide range of practical situations, but any scaling law applied to anything made of atoms must break down eventually, if you push it far enough, because atoms are not scalable. Scaling laws are intimately connected to dimensional analysis; see reference 9 for an

introductory discussion of the capabilities and limitations of dimensional analysis.


5.2 Keep Looking

Always look for the scaling law. It might be obvious, or it might not. For example, if you cant find a scaling law for some quantity X, maybe there is a nifty scaling law for (1X). Keep looking. Note that many scaling laws take the form of a power law, but not all; see item 32 in section 6.
Always Look for the Scaling Law.

The rationale here is simple: If you find a scaling law, it greatly increases your understanding of what is going on. Scaling laws are easy to use, and are very powerful. Consider the relationship between scaling laws and detailed formal analysis. Neither is a substitute for the other; rather, they reinforce each other. The scaling law may sometimes tell you the answer you need, but even if it doesnt, it suggests how to do the analysis. The analysis, in turn, may reveal additional scaling laws. This is similar to the relationship between diagrams and formal analysis. Neither is a substitute for the other; rather, they reinforce each other.

6 Miscellaneous Examples of Scaling Laws


1. 2. 3. 4. Here are some interesting scaling laws. Some of them are self-evident, but some are not. The circumference scales like the diameter. The area scales like the square of the length. The volume scales like the cube of the length. The volume scales like the three-halves power of the area.

5.

The Pythagorean theorem can be elegantly proved using a scaling argument. Additional scaling arguments lead to a generalization of the theorem. See section 1.5. The period of a simple pendulum scales like the square root of its length. In free fall, starting from rest, the time for an object to fall a certain distance scales like the square root of the distance. Conversely, in free fall, starting from rest, the distance traveled scales like the square of the time. Suppose you are doing the special effects for a movie. You need to blow up the monsters castle. You cant afford to build a real castle, so you build a 1/Nth scale model. That is, all the linear dimensions are scaled down by a factor of N.

6. 7. 8. 9.

You need to scale the time by a factor of N. That is, you shoot the explosion at a higher frame rate, so that when it is played back everything is seen in slow motion. In accordance with item 7, this means that each falling object falls through given (scaled) length in the appropriate amount of (scaled) time. 10. The stopping distance for a car has two components. One component is related to reaction time, and scales linearly with the initial speed. The other component is related to braking, and scales like the square of the initial speed. This guarantees that the widely-touted three-second rule cannot possibly be accurate. Seereference 10. 11. 12. In a random walk, at any given time, the distance from the starting point scales like the square root of the elapsed time. Diffusion can be well described as a random walk, so we expect one thing to diffuse into another for a distance that scales like the square root of time. Note that diffusion cannot be represented by a speed (distance per time); it is distance per root time. Newtonian heating is closely analogous to diffusion, so we expect heat to penetrate a distance that scales like the square root of time. This means that cooking time for a roast should scale like the 2/3rds power of the mass. This is close but not identical to the familiar rule of thumb that predicts a time simply proportional to the mass (half an hour per pound). Scaling predicts that a 20-lb roast should take only 1.6 times as long as a 10-lb roast (not twice as long). The electric field of a point charge falls off like 1/r2. This can be understood (and visualized!) in terms of the continuity of electric field lines ... plus a basic scaling argument: The area of a Gaussian pillbox scales like r2. The mean free path of a particle in a gas of similar particles scales like the molar volume of the gas divided by the scattering cross-section of the particle.

13. 14.

15.

16.

This result cannot be reliably obtained via dimensional analysis alone. See section 7.2. 17. The distance to the apparent horizon scales like the square root of your height h above the surface (assuming the surface is reasonably well approximated as spherical, and assuming hR). That means that if you climb twice as high, you do not get to see twice as far (although you do get to see twice as much area). See section 7.1. In chemistry, it is easy to construct scenarios involving three quantities, all of which have the same dimensions, where one scales like system volume to the -1 power, one scales like system volume to the - power, and one scales like system volume to the 0 power. This is not what you would have predicted from dimensional analysis. Section 7.3 discusses an example. The voltage drop in a wire scales like the length and inversely like the crosssectional area at low frequencies but not at high frequencies, because of skindepth effects. The pressure drop in a pipe scales like the length and inversely like the square of the cross-sectional area, i.e. the fourth power of the diameter, assuming laminar flow and negligible compressibility. This fourth-power dependence is remarkably steep. This is a corollary of the Hagen-Poiseuille law. The energy in the pressurized gas inside a bubble scales like volume, i.e. like diameter cubed. The sign of the energy is such that it tends to make the bubble grow. The energy in the surface tension of a bubble scales like the surface area, i.e. like diameter squared. The sign of the energy is such that it tends to make the bubble shrink. Combining item 21 and item 22, we see that large bubbles tend to grow, while small bubbles tend to get crushed to nothingness. This is the fundamental basis of the much-discussed topic of nucleation. See reference 11. The speed of propagation of a long-wavelength wave in deep water (i.e. gravity wave) scales like the square root of the wavelength. See reference 9. The speed of propagation of a short-wavelength wave in deep water (i.e. capillary wave) scales inversely as the square root of the wavelength. See reference 9. The speed of propagation of a long-wavelength wave in shallow water scales like the square root of the depth. See reference 9. The hull speed of a boat scales like the square root of its length (assuming a displacement hull, as opposed to a hydrofoil).

18.

19.

20.

21.

22.

23.

24. 25.

26. 27.

28.

In thermodynamics, an intensive variable is one that scales like the zeroth power of the amount of stuff. An extensive variable is on that scales like the first power of the amount of stuff. (In this context, it is assumed that the stuff is homogeneous, so that it doesnt matter whether you measure the amount by mass, by volume, or by counting moles; all three methods have the same scaling properties.) Scaling laws are very important in fluid dynamics. For starters, two systems will exhibit similar amounts of turbulence if they have the same Reynolds number. The aerodynamic drag on a moving object scales like the area and like the square of the velocity, over a wide range of Reynolds numbers (assuming constant angle of attack). The terminal velocity of a falling object scales like the square root of the diameter (assuming constant density). The perceived loudness of a sound scales roughly like the logarithm of the amplitude of the sound-pressure wave. The size of the crater formed by an explosion scales like the cube root of the energy (within limits). The Secchi disk pattern has some interesting scaling properties, as discussed in reference 12. The energy per particle in an ideal gas scales like kT. If we add the assumption that external fields (such as gravity) can be ignored, then this is an example of dimensional scaling, since the energy kT is the only relevant energy in the problem. The energy per particle in a degenerate Fermi gas (such as the conduction electrons in a chunk of metal) scales like kT times (kT/EF), where EF is the Fermi energy. For ordinary metals, the Fermi temperature (EF/k) is thousands of degrees, so the heat capacity of the degenerate electron gas is smaller than it would be for a non-degenerate gas. However, at low temperatures, it is larger than the specific heat of the phonon gas, which scales like (T/D)3, where D is the Debye temperature. Therefore at low temperatures (below D) the electron specific heat is the dominant contribution to the overall specific heat of the material. This is an example of non-dimensional scaling, since there are three physically-relevant energies in the problem, kT, kD, and EF. The magnetization of a piece of iron (or almost any other ferromagnet) near the Curie point scales like the 0.37 power of the reduced temperature (reference 13). (Calculating this is not easy.) The force on an induced dipole in the field of a permanent dipole scales inversely like the 7th power of the distance. I doubt there is any way to obtain

29.

30.

31. 32. 33. 34. 35.

36.

37.

38.

this result by dimensional analysis. Its not even 100% obvious how to handle this in terms of non-dimensional scaling. Its easy to write F (r0/r)7 but its not obvious what to use as the reference distance r0.

7 Non-Di
(a) Identify a set of physically-relevant dimensionless groups, and (b) Determine the scaling exponent for each one.

In non-dimen

Dimensional a superficial un simplest prob There are som dimensional s


7.1 Example:

As mentioned (subject to mi This is a favo dimensional a There are two surface. The d of R (assumin distance to the Dimensional a Dimensional a dimensionless
7.2 Example:

As foreshadow the gas divide This example If you tried to There is a natu molar volume dimensional a You may have to form the di

whereupon w

where the exp about the expo So lets look a view facing o

You can see t layer. Call thi is

where P (with expect

which basical like D/. We also expec volume consta

That means w

as advertised.
7.3 Example:

7.3.1 Introdu Lets consider

and in particu the reaction co 100% reactan We choose co constant, and of Q increases Lets work ou brackets [] chemistry cou below.) Lets consider arrange the in temperature e column A in e

As a first step quickly that w expansion tak densities have equilibrium si As the final st unbound F ato ignore.) The final state was initially. more precisely There are at le equation (equa theoretically o the piston. Another thing

where the RH We see that in constant unde other conditio calling Kd the There are goo squared becau the recombina first-order pro In equation 17, nontrivial. We expansion/rea

I classica scale li

I column is much scales l like sys

There is no w This conflict b just shows the the dimension because it is m Here are the m

T analysi two ph you wh

S into acc situatio such a

In equation 17, the quantum-s However, it e

The bottom li not change Kd volume V. Th Note: Sometim nontrivial and If all you wan course. Note: The exa applied to rea with number d solvent-solute 7.3.2 Analys We can begin hydrogen, in w

where [np] is t neutral hydrog As usual, d

Comparing eq the inverse vo on immutable We can use th

where this K i any nicer or si The trick is th densities per u approximation

This (more-or

where {} de (See reference In equation 22, any of the fun the same dime power. When like V to the results using d scales the way We learned in always conve dimensions of 22 is in any w There is no es The key to un statistical volu not mere dime

There is, alas, quantum-stati instead, name quantity as dis throws away t awful.

Also beware t glance whethe for [], based Remark: The thermodynam entropy (as di case stat mech dimensionless need statistica that we have a

8 Fractal

You may have objects involv with nothin However, that rational numb not some perv The word fra Hausdorff dim According to given size it ta according to h using smalle For ordinary ( perimeter ... a In contrast, fo perimeter grow For more abou

You might also like