You are on page 1of 35

Statics & Stress and Strain

1 Introduction
Let’s start this first week of PhysicsWOOT off slowly. In fact, let’s start with zero velocity. And zero acceleration.
And zero angular velocity or angular acceleration. That is, let’s begin with statics.
Statics is the study of when nothing happens - points don’t accelerate, solids don’t move, fluids don’t flow, etc.
Although it’s rare to have an entire competition problem focused solely on statics, many problems have
components which are based on statics, often near the beginning, so it’s a useful place to start our study.
Statics shows up a lot because many systems have an equilibrium state in which they can sit at rest, such as a
mass on a spring with the spring at its rest length (and no external forces). Physicists often study a system by first
examining its equilibrium state with statics, then looking at what the system does when in motion with dynamics
(which we’ll get to the in other 15 classes).

2 Forces
The simplest rule of statics is:
If a system is in equilibrium, the sum of the forces on it is zero.
This follows from Newton’s second law, F~net = m~a. If the system is in equilibrium, meaning no parts are moving,
then ~a = 0, which means F~net = 0 as well.
Problem 2.1. A rock climber wears special shoes with a coefficient of friction against the rock of µ = 1.2. The
rock climber stands on a rock slope inclined at an angle θ to the horizontal. What is the greatest angle θ which
allows the rock climber to stand without slipping?

Solution:
The forces on the rock climber are a gravitational force downward, a normal force perpendicular to the rock
surface, and a friction force parallel to the rock surface.
Using a common trick for problems with inclined planes, we set up an x-y coordinate system with the x coordinate
parallel to the rock surface.

N
y

Ff

x
mg

The gravitational force can be broken down into x and y components:

mgx = mg sin θ (2.1)


mgy = mg cos θ (2.2)

© 2021 AoPS Incorporated 1


Statics & Stress and Strain
N
y

Ff
mg sin θ

x
mg
θ mg cos θ

The forces must sum to zero in each direction. We’ll begin with the y direction. The two forces are the normal force
and the component of gravitational force in the y direction. They must have equal magnitudes in order to balance
out, so
mg cos θ = N. (2.3)

In the x direction the forces must also sum to zero. In this direction, there is a friction force and the x component of
the gravitational force. These must be equal in magnitude, so

mg sin θ = Ff . (2.4)

If the slope is at its maximum, then the force of static friction is at its maximum, so

Ff = µN. (2.5)

Equation (2.5) contains two quantities we don’t want in our final answer: the normal force and the friction force. We
can use equation (2.3) to eliminate the normal force. We just substitute mg cos θ in for N in equation (2.5). Similarly,
equation (2.4) lets us eliminate the friction force in equation (2.5).
Making those substitutions, we have
g sin θmax = µg cos θmax (2.6)
which we can simplify to
µ = tan θmax (2.7)

The answer is
θmax = arctan(1.2) ≈ 50.2◦ . (2.8)

The equation µ = tan (θmax ) is very common in physics problems about inclined planes. We may see problems
equivalent to this one in the future, or sub-problems of longer problems equivalent to this one. When that happens,
we will simply assume that maximum angle is θmax = arctan(µ) and work from there.
In fact, the result is important enough that we will derive it another way. In the process, we’ll see a different
technique for solving problems where forces come to zero.

© 2021 AoPS Incorporated 2


Statics & Stress and Strain
The three force vectors on the person are a normal force, a gravitational force, and a friction force. Drawing just
these three vectors, we have

N
Ff

mg

If the person is in equilibrium, the sum of these forces must be the zero vector. This means that, added tip-to-tail,
the force vectors form a triangle.

Ff

mg

Because the normal force is perpendicular to the friction force, this is a right triangle. The climber is on the verge of
slipping, so Ff = µN. Also, the angle between the gravitational force and the normal force is θ. Adding these
details, we have

© 2021 AoPS Incorporated 3


Statics & Stress and Strain

µN

mg
θ
N

We can now read off from the picture


tan θ = µ.

Problem 2.2.
A rope with mass m is hung from the ceiling by both ends and a box with mass M is attached to the center of
the rope. The tangent to the rope at either end forms an angle α with the ceiling. What is the angle β between
the tangents to the rope at the box?

α α

(Kalda pr 12 )

Solution:
Because the rope is massive, there is gravitational force on it. We can’t assume tension in the rope is constant, and
furthermore, the rope bends, so different parts of it have different slopes. To deal with this, we begin by analyzing
some small, almost-straight piece of rope. The forces on it are a gravitational force downward and tension forces
to the left and to the right.

© 2021 AoPS Incorporated 4


Statics & Stress and Strain

α α

T (x) β

dmg T (x + dx)

Because the little bit of rope we’re analyzing is not accelerating, the forces on it must sum to zero. This means that
the horizontal component of tension in the rope, which we will call Tx , is the same on both sides of the segment.
This in turn means it is the same everywhere in the rope - only the vertical component of tension varies throughout
the rope.
Consider the point where the rope meets the ceiling. The y-component of tension in the rope must support half the
total weight of the system (because by symmetry, the other point where the rope meets the ceiling supports the
other half). This gives us
1
Ty,end = (M + m)g. (2.9)
2
At the box, the y-component of tension only needs to support half the weight of the hanging box,
1
Ty,center = M g. (2.10)
2

The tension force in a rope always acts along the rope, meaning that the tangent of the angle that the rope makes
Ty
with the horizontal is equal to the ratio . For the point at the ceiling, that gives us
Tx
Ty,end
tan α = . (2.11)
Tx

β/2 is the complement to the angle that the rope makes with the horizontal at the center, so
 
β Ty,center
cot = . (2.12)
2 Tx

Equation (2.12) is the only thing we have to work with that includes β, so we’ll substitute things into that equation
until we’re left relating only β and α.
First, we can use (2.10) to eliminate Ty,center . Then we can use (2.11) and (2.9) together to eliminate Tx . The result is
1
 
β Mg
cot = 2 tan α. (2.13)
2 1
(M + m)g
2

© 2021 AoPS Incorporated 5


Statics & Stress and Strain
This simplifies to  
β M
cot = tan α. (2.14)
2 M +m
Solving for β gives us
 
M
β = 2arccot tan α . (2.15)
M +m

Exercises
2.1. A mass m is hung from the middle of a massless rope. The other ends of the rope hang over frictionless
pulleys, which connect to identical masses M. Find the angle θ that the rope makes with the horizontal.

M M

2.2. What is the minimum magnitude of force that could be used to dislodge a block of mass m resting on an
inclined plane of slope angle θ if the coefficient of friction is µ? Assume that when no force is applied, the block
does not slip.

F~

(Kalda pr 4 – link contains hints )


2.3. A rope rests on two slopes which form a V shape, as shown in the figure below. They are both inclined at an
angle θ to the horizontal. The rope has uniform mass density, and its coefficient of friction with the slopes is µ. The
system has left-right symmetry. What angle θ gives the largest possible fraction of the rope not touching the
slopes?

© 2021 AoPS Incorporated 6


Statics & Stress and Strain

θ θ

(IsaacPhysics – link contains hints)

3 Accelerated Frames
Although statics is generally applied to situations where nothing is accelerating, in situations of constant
acceleration, we can sometimes move into an accelerating frame and treat the situation as static.
Suppose a person is standing completely still. A train is rushing past them to the right, accelerating to the right at
acceleration ~a.

From the point of view of someone on the train, the person is accelerating to the left. Their acceleration is
−~a.

However, there is no force pushing the person to the left, so they shouldn’t be accelerating that way according to
Newton’s second law! This is why we usually do physics from an inertial reference frame, not an accelerating one.
In the accelerating frame, things with apparently no force on them can accelerate.
We can rescue the situation if we imagine there is a force −m~a on the person. (This is a force to the left.) This force
doesn’t have any physical origin, so we could call it a fictitious force. The reference frame of the train is a perfectly
valid one to use the analyze physics situations so long as we apply a fictitious force −m~a to every object of mass
m.
Similarly, we can use a reference frame that is rotating. In this frame, every stationary object experiences a
centrifugal force mω 2~r, where ~r is the vector from the axis of rotation to the object.
Moving objects in a rotating frame additionally experience a Coriolis force, but that is not needed for problems in
statics.

Problem 3.1. A rock climber wearing shoes with µ = 1.2 between the shows and train stands on top of a train.
The train accelerates to the right with acceleration ~a. The magnitude of ~a is |~a| = a. What is the maximum a
such that the rock climber does not slide off the train top?

© 2021 AoPS Incorporated 7


Statics & Stress and Strain
Solution: We go into a frame accelerating to the right at acceleration a. In this frame, neither the train nor the rock
climber is accelerating, so the situation is static.
The rock climber experiences a gravitational force mg downward and a fictitious force ma to the left. The vector
sum of these forces is identical to the situation as if gravitational acceleration were stronger and the train were
tilted by an angle
ma a
tan θ = = . (3.1)
mg g

ma

θ
mgeff mg

In this frame, the rock climber can think of themself as standing on a flat surface that is tilted at an angle θ. The
train is not tilted relative to Earth’s surface, but the rock climber only cares that the train feels tilted relative to the
effective gravitational acceleration they feel.
We saw in a previous problem that the rock climber can remain stationary on a surface tilted at angle θ to
gravitational acceleration so long as
tan θ ≤ µ. (3.2)
At the maximum acceleration, this inequality is saturated, so we can turn the ≤ sign into an = . We can also use
equation (3.1) to eliminate tan θ from equation (3.2). The result is
amax
=µ (3.3)
g

Solving for amax and putting in the value of µ, we get

amax = 1.2g . (3.4)

© 2021 AoPS Incorporated 8


Statics & Stress and Strain
Exercises
3.1. A block rests on an inclined surface with slope angle θ. The surface moves with a horizontal acceleration a
which lies in the same vertical plane as a normal vector to the surface. Determine the values of the coefficient of
friction µ that allow the block to remain still.

θ ~a

(Kalda pr 5 – link contains hints )

4 Torque
A point mass with zero net force on it is necessarily in equilibrium.
(We often think of “equilibrium” as implying no motion. A point mass with no force on it may not be stationary; it
could be moving at a constant velocity. Even in that case, we can change into a new inertial reference frame
moving along with the point mass, and in this frame the point mass is stationary.)
However, a real object, like a table or a rock or an apple, is not necessarily in equilibrium or stationary just because
there are no net forces on it. For example, imagine pushing forward on the left side of the object and pulling
backward with the same magnitude force on the right side of the object.

Although the net force would be zero and the center of mass of the object wouldn’t accelerate, the object would
begin to spin. This brings us to the second basic principle of statics.
When an extended body is in equilibrium, the sum of the torques on it must be zero.
Torque is calculated around some point, called an “origin” or “pivot.” We can choose any point to be the origin, but
some problems are easier to solve with a judicious choice of origin. In principle, though, the choice is arbitrary. If
the net torque on an object is zero about one origin, then it is zero about any origin, so long as the net force on the
object is also zero.

© 2021 AoPS Incorporated 9


Statics & Stress and Strain
Problem 4.1. Two students are carrying a thin tabletop together at constant velocity, one in front and one in
back. The student in front, being lazy, hoists their side up above their head to try to dump the weight of the table
on the person in back. Does this work? What if instead they were carrying a sofa?

Solution: First we’ll consider the thin table. The table has no acceleration and no angular acceleration, so the sum
of both the forces and the torques on it is zero.

There are three forces on the table - upward forces from each student and a downward force from gravity. The
upward forces must sum to the gravitational force,

F1 + F2 = mg. (4.1)

To prevent the table from rotating, the torques must sum to zero. We’ll choose the center of mass of the table as
the pivot point about which to calculate torques. This means we don’t have to consider the torque due to the
gravitational force. Gravitational forces can be considered to act at the center of mass, and therefore exerts no
torque about the center of mass.
F1 l1 = F2 l2 . (4.2)
Here, l1 and l2 are the moment arms of the forces. Because the students are equidistant from the center of the
mg
table, these moment arms are equal. This shows that F1 = F2 = .
2
Next, the lead student tilts the table.

The same equations still apply. The moment arms are still equal as well. Both have been reduced by a factor of
mg
cos θ, but they are equal to each other, so again F1 = F2 = . The lazy student’s plot
2
does not work for a table .

© 2021 AoPS Incorporated 10


Statics & Stress and Strain
Next we consider a couch, modeled as a thick rectangle. When the couch has not been rotated, the students
contribute equally to carrying its weight, as before. However, when the lead student lifts the couch up, the center of
mass of the couch moves backward towards the back student.

l1 l2

In this scenario, l2 < l1 . Revisiting the equation F1 l1 = F2 l2 , we see that F1 < F2 . The lazy student’s plot
does work with a couch .

In the previous problem, we assumed that the gravitational force on an object could be thought of as a single force
acting on the center of mass of the object. In fact, the gravitational force is many small forces, one on each piece
of the object. Let’s find the torque exerted due to gravity by more carefully summing over the entire mass
distribution of an object.
Suppose gravity points in the z direction. We will calculate the x component of torque about the origin. On some
little piece of mass dm and with y coordinate y, gravitational force exerts torque
dτx = ygdm.

The total torque in the x direction on an entire body is


Z
τx = ygdm.
entire body

Because g is a constant, we can write this as


Z
τx = g ydm.
entire body

The y component of the center of mass location is defined by


Z
1
ycenter of mass = ydm,
m entire body

where m is the total mass of the body.


This lets us write
τx = mgycenter of mass .

In other words, the torque in the x direction is the same as that cause by a force mg acting at the center of
mass.
The same applies to other components of torque, so we can think of gravity as being all applied directly at the
center of mass; the torque is the same as we would find if we integrated the gravitational torque over the entire
body.

© 2021 AoPS Incorporated 11


Statics & Stress and Strain
Problem 4.2. Consider the two possible climbing stances for a person wearing high-friction rock climbing-shoes.
(They do not have high-friction gloves.) Which stance will keep them more secure against slipping?

(Wylie, “Friction, fear, friends, and falling”, Quantum, July/Aug 1992)

Solution:
The weight of the climber must be supported by a combination of the feet and hands of the climber. As we saw in
the previous problem, when the center of mass is closer to being directly above the feet, more of the weight rests
on the feet and vice versa.

The climber has high friction on their feet, so they want their feet to support as much weight as possible to ensure
they are secure. So the climber should take the first stance , with their center of mass closer to being over their
feet.


Problem 4.3.
An ornery cat pushes a triangular wedge of cheese gradually over the edge of a countertop, with the thin edge
of the wedge hanging furthest off the countertop. The cat pushes the cheese very slowly. If the length of the
wedge is l, how far must the cat push the cheese off the countertop for the cheese to fall?

© 2021 AoPS Incorporated 12


Statics & Stress and Strain

Solution:
In order for the cheese to remain on the countertop, both the forces and the torques on it must sum to zero. The
forces are gravitational force and a normal force. We neglect the force from the cat because the cat pushes the
cheese very slowly. There is no obvious limit to how large the normal force on the cheese can be, so we examine
the torques.
Choosing the center of mass of the cheese as an axis, the gravitational force exerts no torque. Therefore the
normal force must also exert no torque about the center of mass.

mg

(In this picture, the normal force has been shifted slightly to the right to make it visibly-distinguishable from the
gravitational force. Ideally, they should both point through the center of mass of the wedge.)
The normal force from the counter can be distributed unevenly across the area of contact between the cheese and
the counter. However, the normal force must be applied somewhere where the cheese and counter are in
contact.
When the center of mass of the cheese is directly at the edge of the counter, the normal force could be applied
entirely at the very edge of the counter and exert no torque. However, if the cheese is pushed any further, the
normal force from the table will necessarily have a moment arm about the center of mass of the cheese and exert
a torque. The cheese therefore has no stable equilibrium on the table beyond this point.
The center of mass of the wedge lies at
l
h l
Z
1
xcm = (h − x)xdx = . (4.3)
1 0 l 3
hl
2

2
So the cat can push the wedge a distance l off the countertop before it falls.
3

So far, we have always calculated torques about the center of mass of the objects. That’s a good default, but an
even better idea is

© 2021 AoPS Incorporated 13


Statics & Stress and Strain
Choose the axis about which to calculate torque so that some torques are zero.
Specifically, if there are forces we don’t know and don’t want to calculate, choose the axis so that they exert no
torque.

© 2021 AoPS Incorporated 14


Statics & Stress and Strain
Problem 4.4.
A box is against a wall, held up by a rope as shown. There is no friction. Will the box remain in place against the
wall?

(Quantum, March/April 1992, pp 33)

Solution:
The problem does not provide the angle of the rope, the mass of the box, etc. making the problem appear difficult.
However, suppose we choose this point as our axis about which to calculate torque.

mg

Because this point is the intersection of the line along the gravitational force and the line along the tension from
the rope, neither of these forces exert any torque about this point. If the box is to remain stationary, the normal
force must also exert no torque about this point. But the normal force points directly outward from the wall and at
has a non-zero moment arm because it is below the chosen point, so it’s impossible for the normal force to exert
no torque. This means that no , the box will not remain stationary against the wall.


© 2021 AoPS Incorporated 15


Statics & Stress and Strain
Exercises
4.1. A uniform rectangular trap door of mass m is hinged along one of its sides. It is held open, at an angle θ to
the horizontal, with a force of magnitude F. This force acts on the open side, perpendicular to the trap door. Find
the magnitude of the force.

F~

hinge

(IsaacPhysics)
4.2. Find the angle that the force from the hinge on the door makes with the horizontal in the previous
problem.
4.3. Each of the following planar objects is placed, as shown in the figure, between two frictionless circles of
radius R. The mass density per unit area of each object is σ, and the radii to the points of contact make an angle θ
with the horizontal. For each case, find the horizontal force that must be applied to the circles to keep them
together. For what θ is this force maximum or minimum?
(a) An isosceles triangle with common side length L.
(b) A rectangle with height L.
(c) A circle.

© 2021 AoPS Incorporated 16


Statics & Stress and Strain

(Morin 2008, 2.7)

© 2021 AoPS Incorporated 17


Statics & Stress and Strain
5 Statics Via Energy
Suppose that an object is in equilibrium. There are no net forces on it. Now suppose that we move the object
through a small displacement d~x. How much work does this require? We can calculate it via the equation

dW = F~net · d~x. (5.1)

Because F~net = 0, we can simplify this to


dW = 0. (5.2)

In other words,
If an object is in equilibrium, then moving the object a small amount in any direction does not change its energy
(since no work needs to be done).
This statement is true only to first order, meaning that the derivative of the energy with respect to displacements of
the system is zero. For example, if a ball is at the bottom of a hill, it is in equilibrium there. Moving it a small
distance to the left or right does technically increase its energy some because it will climb at least a little way up
the hill. However, the hill is flat at the bottom, so the derivative of the energy with respect to moving the ball left or
right is zero. If we move the ball some distance dx to the right, its energy increases by some amount proportional
to dx2 . This is different from anywhere else on the hill, where the increase or decrease in energy would be
proportional to dx.
This means that any small displacement of the center of mass will not change the potential energy of the object, if
it’s in equilibrium. Likewise, equilibrium implies that small rotations do not change the potential energy when there
are no external forces.
The bottom of a hill is where the potential energy of the ball is at a minimum. Extending the example, we have the
general principle that in mechanics,
Equilibrium states for a system are where its potential energy is minimal, assuming no external forces.
(Maximum potential energy may also be an equilibrium, but it will be an unstable equilibrium, so that usually only
minima are physically relevant. More exotic example such as “saddle points” (a minimum in one direction and
maximum in another) are also possible, but uncommon in physics competitions. It is important to note that we are
discussing local minima and maxima, not necessarily global ones.)
We can find a static equilibrium either by requiring that the net force is zero or by requiring that the potential energy
doesn’t change when we make a small displacement.

Problem 5.1.
A device is built as shown. Each joint is flexible so that the device can change its shape, but the sides remain
vertical. When unloaded, the device is balanced at all angles.

© 2021 AoPS Incorporated 18


Statics & Stress and Strain

Identical masses m are placed on the arms of the device. The mass on the left is 2x from the edge and the mass
on the right is x from the edge.

How much additional mass must be added to the right hand side at distance 2x to maintain balance?

Solution:
It would appear that moving the mass further out creates more torque, and indeed it does create more torque.
However, we don’t know what the forces of tension are in the beams of the device, so we don’t know whether they
create torques.
The device has a single degree of freedom. It is called ”shear” and is similar to a rotation, but results in changing
the shape. If the device shears slightly, the mass on the left falls and the mass on the right rises. The amount that
the mass on the right rises does not depend on how far it is from the right hand side of the device.
If the masses are balanced, then the potential energy increase from the mass on the right rising cancels the
potential energy decrease from the mass on the left falling, and this is true regardless of how far the masses are
from the walls of the device.
Therefore, zero extra mass must be added to the device - it is already balanced.

A slightly different situation occurs when a system in equilibrium is capable of deforming and is under external
forces. (For example, a spring under compression is such a system.) In this case, the forces internal to the system

© 2021 AoPS Incorporated 19


Statics & Stress and Strain
must cancel the applied forces. This means that
If we make a small deformation of an object at equilibrium, the work done on it by external forces must equal the
increase in the object’s potential energy.

Problem 5.2.
Suppose you stretch a spring with spring constant k a length ∆x beyond its equilibrium. The potential energy
1
of such a spring is U = k(∆x)2 . Find the force F that the spring exerts on you.
2

Solution:
If we stretch the spring by a small additional distance dx, its potential energy increases by

dU = kx dx. (5.3)

The work we do in stretching the spring, assuming it was in equilibrium, is

Fapplied dx = dU. (5.4)

Setting equations (5.3) and (5.4) equal and canceling the dx, we find

Fapplied = kx (5.5)

By Newton’s third law, the spring exerts a force of the same magnitude and in the opposite direction on you, so

F = −kx (5.6)


In the previous example, we found the ordinary rule for the force exerted by a spring, which of course you already
knew as Hooke’s law. However, the method is useful in situations where we already know the potential energy and
want to find other quantities from it.

© 2021 AoPS Incorporated 20


Statics & Stress and Strain
Problem 5.3.
A soap bubble is a static, thin spherical shell of soap. The surface of the soap bubble has a surface tension
γ. This surface tension can be thought of as the energy needed per unit surface area to create the soap-air
interface. What is the difference in pressure ∆P between the inside and outside of the bubble in terms of the
bubble radius r and the surface tension?

P0

P0 + ∆P

Solution:
Some students may wish to begin this problem with dimensional analysis, especially if they are less familiar with
the physical quantities involved. The result from dimensional analysis is:
γ
∆P = c · . (5.7)
r

Now let us determine the missing constant. The potential energy of the surface is
U = γA = 2 · 4πr2 γ. (5.8)
(Recall that the bubble has both an inside and an outside surface, leading to a factor of 2 in the above
equation.)
If the bubble were to expand a small amount, the air inside it would do work
dW = ∆P dV = ∆P · 4πr2 dr (5.9)
We want the work done on the bubble by air to be equal to the increase in the surface energy of the bubble, so

dW = dU. (5.10)
Plugging equations (5.8) and (5.9) into (5.10), we get
d(8πr2 γ) = ∆P 4πr2 dr (5.11)
Solving for ∆P,

∆P = (5.12)
r

© 2021 AoPS Incorporated 21


Statics & Stress and Strain
Exercises
5.1. A light wire is bent into a right angle and a heavy ball is attached to the bend. The wire is placed onto supports
with height difference h and horizontal distance a. Find the position of the wire in its equilibrium. Express the
position as the angle between the bisector of the right angle and the vertical. Neglect any friction between the wire
and the supports; the supports have little grooves keeping all motion in the plane of the wire and the figure.

(Kalda pr 9 – link contains hints )

© 2021 AoPS Incorporated 22


Statics & Stress and Strain
5.2. Use an energy approach to find the mass m as a function of M and θ that results in equilibrium in the setup
below. The disk cannot slip on the wedge. The pulley is frictionless. The string is massless.

M
m
θ

(David Raymond, Virtual Work)


5.3. A mountain climber wishes to climb up a frictionless conical mountain. He wants to do this by throwing a
lasso (a massless rope with a loop) over the top and climbing up along the rope. Assume that the climber is of
negligible height, so that the rope lies along the mountain, as shown in the figure. The lasso is made of a segment
of rope tied to a loop of fixed length. When viewed from the side, the conical mountain has an angle α at its peak.
For what angles α can the climber climb up along the mountain?

(Morin 2008 2.16, adapted) note: this problem is significantly more challenging than most. It can be solved without
fancy math, but requires some thinking outside the box.

6 Fluid Surfaces
We deal with fluid statics in depth in the other year of PhysicsWOOT, but a simple rule follows from the above
discussion and solves some problems.
A liquid, such as a body of water, is in equilibrium when its potential energy is minimized. A condition for this
is
The surface of a liquid at equilibrium is perpendicular to the local gravitational acceleration.
Suppose that some part of the liquid surface is not perpendicular to the local gravitational acceleration. Then liquid
on the “uphill” side of that point could flow to the “downhill” side and release energy.

© 2021 AoPS Incorporated 23


Statics & Stress and Strain
Of course, this condition only works when the only source of potential energy of the water is gravity. For example,
later in the course we will study water as a dielectric (i.e. we will study how water interacts with electric fields). In
the case of an electric field interacting with the water, its equilibrium state might not have the surface
perpendicular to local gravitational acceleration any longer.

Problem 6.1.
A restaurateur builds a restaurant which slowly rotates at angular frequency ω = 0.03 s−1 . They place a cubical
150 L fish tank with a distance r = 10 m from the axis of rotation. Assuming that the floor is flat, how much
water can the fish tank hold?

Solution:
The fish tank rotates along with the restaurant, so its acceleration is

a = ω 2 r. (6.1)

r varies across the fish tank, but we assume that the fish tank is small compared to the size of the restaurant, so
this variation in r is not an important correction.
Like in the problem with the rock climber on the accelerating car, there is an effective gravitational acceleration that
points mostly down and slightly outward, away from the axis of the restaurant. The water surface will be
perpendicular to this effective gravitational acceleration, meaning it is tilted from the horizontal by an angle given
by
ω2 r
tan θ = . (6.2)
g
Thus, there is a portion of the fish tank that cannot be filled.

Its volume is
1 2
V = s · s tan θ. (6.3)
2
which we can rewrite as
1 3 ω2 r
V = s . (6.4)
2 g
Because the volume of the fish tank is s3 , we have that the volume that can’t be filled is

1 ω2 r
V = Vt (6.5)
2 g

where Vt is the volume of the tank.


The volume that can be filled is
1 ω2 r
 
Vf = Vt 1 − . (6.6)
2 g

© 2021 AoPS Incorporated 24


Statics & Stress and Strain
Plugging in the given numbers, we have
Vf = 149.93 L (6.7)
Evidently, the rotation of the restaurant has a minimal effect.


Exercises
6.1.
Suppose a deep cylindrical tank of water rotates at angular frequency ω. Find an equation for the height of the
surface of the water as a function of distance r from the center of the tank.
6.2.
A thin, right-angled U-tube contains water as shown in the figure. The U-tube is spinning about the left vertical part
of the tube. Find the minimum ω such that there is no water in the left vertical part of the tube. The original height
of the water on each side (when ω = 0) is h and the length of the horizontal connector at bottom is `.

7 Young’s Modulus
Often in physics, we treat all objects as rigid, meaning they cannot change shape. Real objects do change shape
though; even a diamond compresses a little bit when you push on it with your fingers.
In this section, we would like to be able to describe deformations of solid bodies. We’ll start with a quick review of
springs, then extend the results to solid bodies.
Problem 7.1.
Two springs, each of spring constant k, are attached end-to-end. What is the spring constant of the combined
spring?

k k

© 2021 AoPS Incorporated 25


Statics & Stress and Strain

Solution:
1
If we stretch the combined spring by ∆x, each constituent spring stretches by ∆x. Each constituent spring
2
experiences a tension of
1
T = k∆x (7.1)
2
This is the tension in the entire spring as well, so its effective spring constant is

1
keff = k. (7.2)
2


The above example shows that a spring constant is not a property of the material the spring is made of, like density
or thermal conductivity. Those are intensive properties, meaning that, for example, any object made of steel has
the same density, no matter its size. By contrast, not all objects made of steel have the same spring constant.
However, we can find an intensive property of steel that serves a similar role to the spring constant, in that it tells
us how stretchy or stiff the material is.

Problem 7.2.
A solid cylinder of radius r and length l can be thought of as a spring. Find its spring constant up to a constant,
undetermined factor. Find the dimensions of this factor.

Solution:
Let’s find how the spring constant depends on r and l.
From the previous problem, we know that putting two springs each with spring constant ks in series results in a
single spring of spring constant ks /2. Likewise, putting n springs in series results in a single spring with springs
constant ks /n.
We can think of the cylinder as a series of disks of constant height stacked on top each other. Whatever the spring
constant of a disk, the spring constant of the cylinder is inversely proportional to the number of disks. The number
of disks is a proxy for l, the length of the cylinder, so
1
k∝ . (7.3)
l

Next we want to think about how changing r affects the spring constant.
We can think of the cylinder as a stack of equal-height disks, but we can also think of it as a bundle of
equal-cross-section thin tubes. If each thin tube has a spring constant kt and there are n thin tubes, then the total
force they exert is n times as large as a single tube would. This means the effective spring constant is nkt . The
number of tubes is a proxy for the cross-sectional area of the cylinder, so

k ∝ A, (7.4)

where A is the cross sectional area.


Combining our two results,
πr2
k∝ . (7.5)
l

© 2021 AoPS Incorporated 26


Statics & Stress and Strain
That means we can write the formula for the spring constant as

πr2
k=Y (7.6)
l

for some constant Y.


This constant is called the Young’s modulus for the material the cylinder is made from.
Let’s examine the dimensions involved.

[k] = M T −2
r2
 
=L
l

Substituting these into the equation for the spring constant, we get

M T −2 = [Y ] · L (7.7)

or
[Y ] = M L−1 T −2 (7.8)


The Young’s modulus has the dimensions of pressure, and its value is often reported in gigapascals (GPa). A lot of
plastics have Young’s moduli of a few GPa and most metals are in the hundreds of GPa. Diamond is above
1000 GPa.
Many sources use E for the Young’s modulus, but we will usually use Y to avoid confusing it with energy.

Exercises
7.1.
A steel cable with an initial length of 20 m and a diameter of .05 m suspends an elevator car in its shaft. The
Young’s modulus of the steel is 200 GPa. Three people, having a total mass of 238 kg, enter the elevator car. Given
the elevator cable is made of steel, what is the amount of stretch that the cable experiences when the three
passengers enter the elevator car?
(Teacher Engineering)
7.2.
A rectangular prism with a square cross-section is used as a stiff spring. The prism is cut with two cuts into four
prisms of the same length and each one quarter the cross section. These prisms are put together to make a single
prism four times as long as the original. What is the spring constant of this long prism as a function of the spring
constant of the original prism?

8 Stress and Strain


The Young’s modulus can tell us the spring constant of a cylinder via the equation
A
k=Y . (8.1)
l

© 2021 AoPS Incorporated 27


Statics & Stress and Strain
This is helpful if finding a spring constant is our goal, but it’s an equation specific to some particular cylinder. Let’s
try to find an equation that holds true for any object made of the same material.
By the definition of a spring constant, we could rewrite the spring constant equation as
F A
=Y . (8.2)
∆l l
We notice that both F and A appear in the equation. Recall that in fluids, a force per unit area is a basic object of
study - the pressure. So we might want to look at the equation with force per unit area on one side.

F ∆l
=Y . (8.3)
A l

We don’t quite want to call the left hand side “pressure”. Although it is a force per unit area, like pressure, it is not
the same in all directions inside the cylinder. We are only pressing on the top and bottom of the cylinder. Although
internally there may be some radial forces between different pieces of the cylinder, we cannot say that there is the
same force per unit area on any surface inside the cylinder, the same way we can when a fluid is under
pressure.
Because of these differences, the force per unit area in solids is called the stress. It is measured in pascals in the
SI system, where
1N 1J
1 Pa = 2 = 3 = 1 kg · m−1 · s−2 . (8.4)
m m
A common symbol for stress is σ.
∆l
The quantity is called the strain. It is dimensionless. A common symbol for strain is e.
l
We can rewrite our equation as
σ =Ye (8.5)
This is Hooke’s law for solid materials. It says that no matter the size or shape of an object, the stress is
proportional to the strain.
This law only applies to stretching or compressing objects in one dimension. It doesn’t apply to compression from
all directions at once or to twisting stresses. Those are covered in the optional section below.

9 Other Elastic Moduli (optional)


The Young’s modulus is the most common example of a stress-strain equation seen in physics competitions,
probably because it is the only one mentioned by name in the IPhO syllabus.
The full theory of stress and strain is quite complicated; here we present a basic introduction to the other types of
deformation besides stretching. This section is optional reading. USAPhO could ask questions on this material, but
any questions testing it would need to introduce the definitions and concepts before asking about them.
Imagine you are compressing an object from all sides. Then the object will shrink. The relevant elastic modulus in
this situation is called the bulk modulus.
The bulk modulus obeys the equation
∆V
∆P = −B . (9.1)
V
Here, ∆P is a pressure. The equation says that if you increase the pressure by ∆P and measure the change in the
volume V of the object, the change ∆V will satisfy that equation.

© 2021 AoPS Incorporated 28


Statics & Stress and Strain
The equation is quite similar to the equation for the Young’s modulus, but adapted to compression in all directions.
∆l
That’s why we use pressure - the force is the same in all directions. We have also replaced the linear strain, with
l
∆V
the volumetric strain, .
V
The equation often works only for small displacements, so we could write it as

dV
dP = −B . (9.2)
V

Problem 9.1.
Find the bulk modulus of an ideal gas for both an isothermal process and for a reversible adiabatic process.

Solution:
For an ideal gas,
P V = N kB T. (9.3)

If we have an isothermal process, dT = 0 and

P dV + V dP = 0. (9.4)

Dividing by dV, we get


dP
P +V = 0. (9.5)
dV

If we look back at equation (9.2) for the bulk modulus and solve for B, we get

dP
B = −V . (9.6)
dV
dP
Substituting −B for V in (9.5),
dV
P −B =0 (9.7)
or
B=P (isothermal). (9.8)

The bulk modulus simply is the pressure for an isothermal gas; the bulk modulus can be thought of as the extent to
which an object resists compression.
For an adiabatic process, we have
P V γ = constant. (9.9)
Taking a differential, we have
γP V γ−1 dV + V γ dP = 0. (9.10)
Dividing by V γ−1 dV,
V dP
γP + = 0. (9.11)
dV

The right hand term is again the negative of the bulk modulus, so

B = γP (adiabatic). (9.12)

© 2021 AoPS Incorporated 29


Statics & Stress and Strain
For an adiabatic process, the bulk modulus is higher. This is because if we adiabatically compress some gas, its
temperature increases. This increases the pressure and so increases the extent to which the gas resists being
compressed.

Finally, material can be deformed via shearing.

In this image, a shearing force is applied to the top of a cube of material. This is a force that is tangent to the face
on which is is applied. This is different from the tension and pressure forces we looked at before, which are
perpendicular to the surface where they are applied.
When a shear stress is applied to the top of a cube, the top slides sideways, and other parts of the cube slide
sideways lesser amounts. The amount of sliding is proportional to the distance from the bottom of the cube. If the
top of the cube slides a distance dx, the strain, called the “shear strain” or simply “shear”, is

dx
e= (9.13)
s

dx

Like the other types of strain we’ve studied, this is dimensionless.


The shear stress is defined by
Fshear
τ= , (9.14)
A
where A is the area of the surface to which the shear stress is applied. The shear stress has dimensions of
pressure and is measured in pascals.

© 2021 AoPS Incorporated 30


Statics & Stress and Strain
dx
Fluids in equilibrium do not support any shear, but solids do. In solids, the shear strain , shear stress τ , and
s
shear modulus G are related by the equation
dx
τ =G . (9.15)
s
This is the same stress-strain equation we saw for the bulk and Young’s moduli, but applied to the deformations
related to shear.
Problem 9.2.
A heavy sheet of metal of mass M is supported by four small, light cubes of rubber, one at each corner. The
shear modulus of the rubber is G and the length of the sides of each cube is s. Find the frequency of horizontal
oscillations of the sheet.

Solution:
Let’s suppose we displace the sheet a small distance dx to the right. Then each cube experiences a shear strain
dx
e= . (9.16)
s
If the heavy sheet where held in place with the displacement dx, the force needed to hold it would be the sum of
the shear stresses on the tops of each cube times the area over which those stresses are applied.
dx
F = 4G · A = 4Gsdx. (9.17)
s

When the sheet is oscillating, the cubes exert this same force for the same displacement, so we can think of the
sheet as experiencing an effective spring constant
F
keff = = 4Gs. (9.18)
dx

The oscillation frequency of the sheet is


r r
1 keff 1 4Gs
f= = . (9.19)
2π M 2π M
We can simplify this to
r
1 Gs
f= . (9.20)
π M

© 2021 AoPS Incorporated 31


Statics & Stress and Strain

These three elastic moduli - the bulk modulus, the shear modulus, and the Young’s modulus - completely
characterize the elastic behavior of any solid that obeys Hooke’s law. In fact, only two of the three are needed. For
example, the Young’s modulus Y can be computed from the shear modulus G and bulk modulus B via the equation
9BG
Y = (9.21)
3B + G
Although deriving this result is beyond the scope of our course, competition problems can sometimes quote these
sorts of results and expect you to use them without understanding their origin.

Exercises
9.1.
Fluids have zero shear modulus. What does this imply about their Young’s modulus? What does it imply about their
bulk modulus?
9.2.
In some solids, the bulk modulus is much higher than the shear modulus. In such a solid, what is the approximate
Young’s modulus as a function of the shear modulus alone?
9.3.
The center of the Earth has significantly higher density than the crust of the Earth. This is partially because denser
elements, such as iron, tend to migrate towards the center of the Earth. However, another significant factor is that
the pressure at the center of the Earth is so high it compresses the material there, increasing its density. This
exercise begins to explore the phenomenon.
Suppose a certain planet is made from rocks with a bulk modulus B1 . A second planet is made from rocks with a
bulk modulus B2 . (Both planets are homogenous, but made from different materials.) When uncompressed, the
rocks from the two planets have the same density. The density of rock at the center of the two planets is the same.
What is the ratio of radii of the two planets?

10 Further Reading and Problems


The sources in this section are optional and contain more complete explanations of the material here, additional
practice problems, or both.

10.1 Statics
Halliday, Resnick, Krane Physics, 5th ed. 2002.
The recommended introductory text, also useful as refresher material. It has many good problems. Material on
statics is spread between various chapters in volume 1, including chapters 3, 4, 5, and 9.
MIT OpenCourseWare, 8.01 Static Equilibrium.
Morin, Problems and Solutions in Introductory Mechanics, 2014. Chapter 9.
Morin, David. Introduction to classical mechanics: with problems and solutions. Cambridge University Press, 2008.
Chapter 2.
Kalda, Jaan. Problems on Mechanics. Section 3
IsaacPhysics statics problems

© 2021 AoPS Incorporated 32


Statics & Stress and Strain
10.2 Accelerated Frames
Halliday, Resnick, Krane Physics, 5th ed. 2002. Section 5.6
Morin, David. Introduction to classical mechanics: with problems and solutions. Cambridge University Press, 2008.
Chapter 10.
Morin, Problems and Solutions in Introductory Mechanics, 2014. Chapter 12.

10.3 Statics and Energy


Feynman, Richard. Conservation of Energy
Chulalongorn University. Virtual Work

10.4 Young’s modulus, Stress and strain


Lewin, Walter. Elasticity and Young’s Modulus (video, 50 minutes)
Roylance, David. Introduction to Elastic Response
Taylor, John R. Classical mechanics. University Science Books, 2005. Chapter 16 (advanced)

11 Exercise answers
 m 
2.1: θ = arcsin .
2M
mg(µ cos θ − sin θ)
2.2: Fmin = mg sin(arctan(µ) − θ). An equivalent equation is F = p .
1 + µ2
arctan(µ)
2.3: θ = .
2
|g sin θ − a cos θ|
3.1: µ ≥ , assuming g + a tan θ > 0.
g cos θ + a sin θ
mg cos θ
4.1: F = .
2
2 − cos2 θ
4.2: tan φ = .
sin θ cos θ
gσL2 cos2 θ
4.3a: F = .
2
σgRL(1 − cos θ) cos θ
4.3b: F = .
sin θ
σgπR2 (1 − cos θ)2
4.3c: F = .
sin(2θ)
h
5.1: tan(2α) = .
a
1
5.2: m = M sin θ.
2
5.3: α = 60◦ .

© 2021 AoPS Incorporated 33


Statics & Stress and Strain
ω2 r2
6.1: h = h0 + .
2g
2√
6.2: ωmin = hg.
`
7.1: 1.2 × 10−4 m.
1
7.2: k 0 = k.
16
9.1: Y = 0. No inferences about B are possible.
9.2: Y ≈ 3G.
r
r2 B2
9.3: = .
r1 B1

© 2021 AoPS Incorporated 34


Art of Problem Solving is an ACS WASC Accredited School.
Thanks to our sponsors:

You might also like