Chapter 6: Constraints and Symmetry
The second part explores the usage of Lagrange mechanics in problems involving constraints and symmetries.
6.1: Contract Forces
6.11: Contact forces as Constraints
Contract forces arises by the virtue of physical contact between two objects. In origin, contact forces are electromagnetic. As two objects touch, the atoms push against one another through electromagnetic forces. Contact forces are not fundamental: they are the sum of microscopic interactions.
Consider a block on a horizontal ground. The net effect of the contact force is to constrain its vertical motion: the block can move sidewards and not up and down. More of such contract forces translate into constraints on the degree of freedom.

Consider a mechanical system parameterized by coordinates . However, due to some constraint forces, there are algebraic relations amongst these coordinates, given by:
where . This means that there are generalized coordinates or degrees of freedom.
The notation here can be confusing. represents constraints, and is the constraint in the system, representing equations that restrict the motion of the system. The constraints are usually expressed as functions of the generalized coordinates and time , as shown by the above equation.
For example, in the problem of a block on the horizontal ground, let the ground be at . Then, we can define , so that the algebraic relation implies . We started from three coordinates() and and specified the constraint (), where is the vertical position. Then, there are degrees of freedom. The Lagrange is given by:
6.12: A New Lagrangian
Let’s consider a new Lagrange defined as:
where is the original Lagrangian, The parameters labeled are called Lagrange multipliers, and are constraints equations. In the new Lagrangian, we introduced additional degrees of freedom labeled with .
We now assume that the constraint equations are not satisfied a priori. As a result, there are degrees of freedom and equations of motion. Using Euler’s equation, equations of motion for are (we apply Euler’s equation to the newly defined Lagrange , not the one we used in the example for the block. Note that there are two equations of motion for and , where is the generalized coordinates for the original Lagrange and is the generalized coordinates for the newly defined Lagrange.) :
Because is independent of and , we know that . The equation becomes:
The parameters labeled are called Lagrange multipliers. The equation for motion for is given by:
In terms of the original Lagrange , the equation for motion for becomes:
For example, with the block on a horizontal floor, is equal to:
Equations of motions are:
In this case is the normal force.
Meanings of constraints and Lagrange multiplers can be abstract. The term represents a Lagrange multiplier associated with the th constraint . The nature of Lagrange multipliers are scalar functions that adjust to enforce the constraints in the system. For example, in the problem of a block on a horizontal ground, is the constraint that (meaning that the block cannot move vertically and can only slide ways) and the multiplier is the normal force enforcing the constraint. Thus, physically, the multiplier can be interpreted as a force.
6.13: Constraints and Lagrange Multiplers
The addition term involving is defined as the generalized constraint force that enforce the constraints onto the dynamics.
We want to relate the generalized constraint force with the actual force. For every object located at position , denote the total constraint force acting on it by . We also know the relations that connect the position of every object to the generalized coordinates . The work done or energy associated with the constraint forces and be expressed as:
Differentiating , we get . Solving for , we get:
Then, work done is equal to
Rearranging the equation for , we get:
6.14: The Second Approach
In chapter four Lagrangian Mechanics , we define action as the integral of the Lagrangian with respect to time. In the process of extremizing the action, we would get (mathematics is not shown here. In order to derive the integral, integral by parts of multi-variable functions is required.):
for some functions of and . If the constraints on the generalized coordinates can be written in the form:
where and are arbitrary functions of and , we can write the following set of equations of motion:
Summary for Finding the Multiplier (Constraint Force)
- Define the constraints by writing down a constraint equation for each constraint.
- Define a set of generalized coordinates without applying the constraints yet.
- Write down the Lagrangian (again, with no constraints applied).
- Apply the modified Euler-Lagrange equations with constraints and Lagrange multipliers.
- You should now have the equations of motion for each coordinate with Lagrange multipliers.
- Solve for the Lagrange multipliers, which will give you the constraint forces.
Example 6.1: Rolling Down the Plane
Consider a hoop of radius R and mass M rolling down an inclined plane. The hoop rolls down an inclined plane without slipping. Find the Lagrangian and constraint force.

Solution
Steps
- Identifying constraints
- specifies position/velocity using Cartesian/polar/spherical/cylindrical coordinates
- find the Lagrangian
- Use the general form of solution to find coefficients
- Find the equations of motion for each generalized coordinate.
- After finding general equations of motion for each generalized coordinate, substitute identified constraints and coefficients we found form step 4 to find the Lagrange multiplier.
- find the solution
STEP 1:
Before solving the problem, we first define the coordinate system. The coordinate system is shifted as shown in the figure. The hoop is described by three variables: the center of mass in two dimensions and a rotational angle . The position of center of mass is specified as:
From the coordinate we define, we observe two constraints: (1) The vertical position of the center of mass is fixed, so and . (2) The second constraint is that the hoop rolls without slipping, giving us the constraint of:
The two constraints suggest degree of freedom.
STEP 3:
Kinetic energy of the hoop involves linear and rotational kinetic energies, where rotational kinetic energy is equal to . Then, is equal to:
The potential energy of the hoop is equal to:
where is the angle tilted by the inclined plane. The Lagrangian is equal to:
Substituting constraints of and , the equation becomes:
Using Euler’s equation, the equation of motion is:
giving us:
STEP 4: Finding Coefficients
We are interested in knowing the frictional force, so we will apply the general form of solution we derived previously in the section. We have shown that the solution has a general form of:
Rearranging the constraint equation for , we get , matching the form of the general solution. Representing the constraint in the form and comparing the two equations, we get:
STEP 5-6:
We know that:
where is the generalized coordinate. In this problem, are identified as and .
STEP 7:
We know have a system of differential equations
giving us:
Then, is equal to:
The force is equal to:
Example 6.2: Stacking Barrels
Consider the problem of two cylindrical barrels, one on top of the other, as shown in Figure 6.3. The bottom barrel is fixed in position and orientation, but the top one, of mass m, is free to move. It starts rolling down from its initial position at the top, rolling without slipping due to friction between the barrels. The problem is to find the point along the lower barrel where the top barrel loses contact with it. That is, we need to find the moment when the normal force acting on the top barrel vanishes.

Solution
Steps
- Identifying constraints
- specifies position/velocity using Cartesian/polar/spherical/cylindrical coordinates
- Find the Lagrangian
- Use the general form of solution to find coefficients
- Find the equations of motion for each generalized coordinate.
- After finding general equations of motion for each generalized coordinate, substitute identified constraints and coefficients we found form step 4 to find the Lagrange multiplier.
- find the solution
STEP 1: Identifying Constraints
Before solving the problem, we need to identifies constraints first. We have three variables in the problem, and a rotational angle . Using polar coordinate, the first constraint we identify is . Another constraint is
Because radius and are fixed, we know that is fixed. Thus, there are three constraints in the total for this problem.
STEP 2: Position and Velocity
Using spherical coordinates, the center’s potion of the top barrel is specified using (the is also the barrel’s position of the center mass, assuming a uniform mass distribution). Differentiating , velocity of barrel is , where .
STEP 3: Find the Lagrangian
Because the lower barrel is stationary, only the upper barrel contributes to the system’s kinetic energy. Since the object rolls without slipping, we need to consider both translational and rotational kinetic energies, where . Then, the system’s kinetic energy of the system is equal to:
Potential energy of the system is:
The Lagrangian is equal to:
Substituting the constraint , we get:
Note that we do NOT substitute the constraint here when writing the Lagrangian. This is because is part of the generalized coordinates, which will be differentiated later to find equations of motion. If we substitute , we will get a wrong answer. This might sound confusing and abstract. An analogy to understand this is that we want to find the slope of at , where and . In order to find , we will substitute into instead of , giving us . Differentiating , we get , whereas if .
STEP 4: Find coefficients
From the Lagrangian, we can see that there are two generalized coordinates and . Thus, the general form of solution is equal to . Comparing the equation with the constraint , we get:
Note that we use the constraint instead of because both and , the generalized coordinates, are present, whereas the latter one includes unknown variable.
STEP 5: Find equations of motion
Using Euler’s equation, equations of motion for and are:
STEP 6: Finding the Lagrange multiplier
Because and , we get:
Substitute the constraint , is equal to:
where is normal force. Energy is:
Using the initial condition , we get . Then, energy . This implies that:
Normal force . T At the moment when barrels lose contact, , giving us:
Solving for , we get:
where is the critical angle at which the condition is satisfied and barrels loses contact.
Example 6.3:
A classic problem is that of a bob pendulum consisting of a point mass at the end of a rope of length l swinging in a plane (see Figure 6.4). We would like to determine the tension in the rope as a function of the angle θ. We start with two variables
Solution
Step 1: Identifying Constraint(s)
The first constraint we identified is that the string has a fixed length, reducing the degree of freedom of . Mathematically, the constraint can be expressed as:
For a pendulum, we usually use polar coordinates , where position is given by and velocity .
Step 2: The Lagrangian
The Lagrangian is equal to:
Equations of motion are:
Substituting the constraints of and , we get:
6.3: Cyclic Coordinates and Generalized Momenta
In chapter 4 The Variational Principle, we defined a cyclic coordinate as a generalized coordinate is missing but the corresponding velocity is presents in the Lagrangian. The generalized momentum is defined as:
and is conserved.
The question is: why is a particular coordinate cyclic, from a physical point of view?
In order to answer the question, consider the example of a moving object with mass under a uniform gravitational field, where gravity acts in direction. The Lagrangian is equal to:
where and are cyclic coordinates. The corresponding generalized momenta and are conserved:
The answer for the conservation is quit clear. The physical environment is invariant under displacements in and directions, but not under , the vertical direction. A change in means that gets closer or farther from the ground, but a change in and make no difference. We say that there is a symmetry under horizontal displacements, but not under vertical displacement.
If a generalized coordinate is missing from the Lagrangian, it means that there is a symmetry under changes in that coordinate: the physical environment, whether a potential energy or a constraint, is independent of that coordinate. If the environment possesses a symmetry, the corresponding generalized momentum will be conserved.
Note: The concept of “symmetry” here might sound abstract and confusing. It is better illustrated in example 6.5.
Example 6.5: A star orbiting s spheroidal galaxy
A particular galaxy consists of an enormous sphere of stars somewhat squashed along one axis, so it becomes spheroidal in shape, as shown in the figure. A star at the outer fringes of the galaxy experiences the general gravitational pull of the galaxy. Are there any conserved quantities in the motion of this star?

Solution
There is no given quantities, such as mass and radius, so we cannot solve the problem mathematically using Lagrange mechanics (The statement can be proved mathematically, but the derivation could be cumbersome. We will use symmetry to solve it in a smarter way. Check the last part for the mathematical derivation).
The problem can be solved easily by identifying symmetries. Recall that if an environment possess symmetry, the corresponding generalized momentum is conserved. Imaging moving translationally, we observe that any translational movement will make one nearer or farther from the galaxy, so no translational symmetries are presents. Thus, no linear momentum is conserved.
Consider rotating about the axis. The shape of the galaxy remains unchanged. Thus, the angular momentum is conserved, where is the azmuthal angle.
Example 6.6: A Charged Particle Outside a Charged Rod
An infinite straight dielectric rod is oriented in the z direction, and given a uniform electric charge per unit length. A point charge is free to move outside it, as shown in Figure 6.7. Are there any conserved quantities for the motion of the particle?
Solution
When finding symmetries, we consider translational and rotational movements. When moving alone and directions, and are not conserved. When rotating around the axis, a symmetry exists. Thus, and angular momentum are conserved.
6.4: A Less Straightforward Example
Consider a system consisted two masses and . Let the distances be and as shown in the figure.

The Lagrangian . The action is equal to:q
Consider the simple transformation
where is some arbitrary constant. This gives us:
so terms for kinetic energy remain unchanged after the transformation. In addition, we know that:
The action is equal to:
In the next section, we will dip deeper into infinitesimal transformations.
6.5: Infinitesimal Transformations
6.51: Direct Transformations
A direct transformation deforms the generalized coordinates of a system as in:
where is used to label a direct transformation. Note that can possibly be a function of time, where in section 6.4 we discussed the special case when is a constant.
6.52: Indirect Transformations
An indirect transformation affects the generalized coordinates indirectly, through the transformation of the time coordinate:
Note that again that the shift in time is assumed to be small, and the shift can be a function f of time. The small shifts in time bring small shifts in generalized coordinates and affect them indirectly, giving us:
where we use Taylor expansion in to linear order only. Rearranging the equation, we get
6.53: Combined Transformations
We want consider a transformation that includes both direct and indirect transformations. We write:
Thus, for an indirect transformation
where we need to provide a set of functions and to specify a particular transformation and determine .
The question seems to be complicated. In order to understand the formula, we consider physical meanings of the expression. The term represents the infinitesimal variation in the generalized coordinates . The second term accounts for how changes indirectly because time is changed. The last term represents the change in the generalized coordinate due to its time derivative , capturing the direct effect of the time shift on .
Example 6.7: Translations
Consider a single particle in three dimensions, described by the three Cartesian coordinates , and . We also have the time coordinate . An infinitesimal constant spatial translation can be realized by
Solution
To specify a particular transformation, we need to provide the sets of function and to determine . Because we want an infinitesimal constant spatial translation, we know that the time coordinate remains unchanged, giving us:
By equation (7) for an indirect transformation, we know that \ is equal to:
Note that the notation here can be misleading. doe NOT represent exponentiation. A constant spatial translation in three dimensional by shifting in and directions. The shift is represented using , where and are three small constants for three shifts in and directions. A translation in space is then defined by:
Similarly, for a constant infinitesimal shift in time, the spatial coordinate remains unchanged, giving us:
Then,
The translation is defined by:
Note: It might be confusing at first to understand how we lead to the conclusions that and .
A constant translation in time means that every time coordinate is uniformly shifted by a constant small amount . Mathematically speaking, since , where is a constant and the symbol represents a change in time. This gives :
Similarly, for a shift in space coordinate, we know that at time , coordinates are:
At time , each coordinates is shifted by amounts of and . Then:
Thus, .
Example 6.8: Rotations
To describe constant rotations, we consider for simplicity a particle moving in two dimensions. We use the coordinates and .
Solution
Again, for a constant spatial transformation, the time coordinate remains unchaged, giving us:
where and represents two constants. A rotation in matrix form is given by:
The rotation in coordinates is given by:
Using small angle approximation, we know that and . The rotation is then equal to:
Example 6.9 Lorentz Transformation
6.6: Symmetry
We define a symmetry as a transformation that leaves the action unchanged in form, where action is defined as:
where is the Lagrangian. Using the general/combined transformation given by and ), it follows that:
where we use the Leibniz product’s rule , since is an infinitesimal change. The first term the is the change in the measure of the integrand:
The second term has two parts:
After some derivations (the note here is not complete. The complete derivation is in the textbook. Need to com back after studying mathematics from chapter two.), is equal to:
Now, given and , we can substitute these quantities into equation (8) to check is . If , there is a symmetry.
6.7: Noether’s Theorem
Noether’s theorem states:
For every symmetry, i.e., for every set of transformations that leave the action unchanged, there exists a quantity that is conserved under time evolution.
6.71: Proof of the Theorem
Consider a given symmetry . This means that the action is invariant under transformations. Mathematically, this can be expressed as . Using the equation for , we get:
Now suppose satisfies the Lagrange equation of motion, we know that:
When satisfies the Lagrange equations of motion, we can replace in the equation, allowing us to simplify the integrand. The expression for becomes:
More explanation
It might be confusing about why satisfies Euler’s equation. In order to understand this, recall that action is defined as the integral to the Lagrangian alone time , and symmetry is a transformation in whch . This gives us:
When the integral , we know that the integrand . Thus, satisfies the Lagrangian.
Connecting symmetries to conserved quantities:
- Noether's theorem states that every continuous symmetry of the action corresponds to a conserved quantity.
- The proof of Noether's theorem relies on the action being stationary. Hence, it specifically applies to trajectories that are solutions of the Euler-Lagrange equations.
The term can be rewritten using product rule, giving us:
Combining like terms, the integral is equal to:
Since the integration interval is arbitrary, we can conclude that:
where is is a conserved quantity called Noether charge, defined by:
6.72: Noether Charge
We have shown that the Neother charge is a conserved quantity. The proof gives a way to generalize the definition of symmetry, defined as .
where is some functions we can solve using equation (8). There are two cases for , the trivial and nontrivial cases. When is trivial, meaning that is a constant, is equal to:
The means that does not affect the variation of the action , and we revert to the standard form of Noether’s theorem where the conserved quantity is . When is nontrivial, meaning that is not a constant but depends on time, coordinates, velocity, e.t.c, . In this case, the integral is equal to:
Thus, there must be an additional term contributing to the conserved quantity, such that . We generalized the conserved quantity into , taking into account the non-trivial dependence of . Then,
Summary for finding Q
- Identifying and : The expression is shown below. Identify the generalized coordinate and find .
- computation: Use chain rule to calculate the integrand. If , there is a symmetry and a conserved quantity. If , there are no symmetry and conserved quantity.
- Q finding: Calculate the conserved quantity using equation (11). The expression is shown below.
Example 6.10
We start with the spatial translation transformation in a fixed ith direction, as in our previous example (6.87). Identity the conserved quantity if a symmetry exists.
Solution
STEP 1: Identity generalized coordinates and find
In the previous example, the transformation for a spatial translation is given by:
Consider a free non-relativistic particle whose Lagrangian is given by:
STEP 2: Use chain rule and equation (9) to find
The change in action is equal to:
Because is a constant, its derivative with respect to time is , giving us .
STEP 3: Find the conserved quantity
Thus, is a constant and we have a symmetry. By Noether’s theorem, there exists a conserved quantity. To find the associated conserved quantity, we apply the definition of Noether charge and find that:
Because is a constant, , must be conserved.
Thus, momentum is the Noether charge (the conserved quantity) with a symmetry. We then try to find conditions for the symmetry, and hence the conservation law, to be violated. For instance, for a simply harmonic oscillators, the Lagrangian is equal to:
When considering a spatial translation, the time coordinate remains unchanged, so . Using equation , is equal to:
where is the generalized coordinate. Because , the equation for becomes:
Using chain rule to multivariable functions, is equal to:
Note that because is a constant (time independent). Rearranging the integral, is equal to:
Example 6.11: Time Translation and the Hamiltonian
let us consider time translational invariance. Due to its particular usefulness, we want to treat this example with greater generality. We focus on a system with an arbitrary number of degrees of freedom labeled by , with a general Lagrangian . We propose the transformation
Identity the conserved quantity if a symmetry exists.
Solution
STEP 1: Identifying and :
Before solving the problem to find the symmetry, recall that is equal to:
where is the generalized coordinate and is the Lagrangian. In order to find , we need to find . From section 6.5, we have derived the formula for a combined transformation, given by:
Given that , we know that , where given by the problem. Then, is equal to:
STEP 2: Compute
Using chain rule, is equal to:
where . Using chain rule for multi-variable functions, we know that is equal to:
Substituting the expression into , terms cancel out, and the integral is equal to:
where is a constant.
STEP 3: Finding the conserved quantity :
If there is a symmetry, by definition, . This implies that a symmetry exists if and only if:
Doesn’t the equation look familiar? It is in the same form of equation (10), where and the conserved quantity is equal to (equation (11)):
Substituting into equation (11), the conserved quantity is equal to:
Because is a constant, the quantity inside that parenthesis must be conserved, denoted by . Factoring out the term , we get get:
which we recognize as the Hamiltonian of the system, introduced in chapter 4 Lagrangian Mechanics . Thus, the conserved quantity is .
Example 6.12: Rotations and Angular Momentum
Consider the problem of a nonrelativistic particle of mass m moving in two dimensions, in a plane labeled by and . We add to the problem a central force with a Lagrangian of the form
Solution
STEP 1: Identifying and :
Before solving the problem, recall that is equal to
In the given Lagrangian, we identify the generalized coordinate as , so . Because time is not mention, the infinitesimal variation , such that .
STEP 2: Compute :
Substitute into (9), is equal to:
Let . Using chain rule for multivariable functions, partial derivatives are simplified into:
As derived in example 6.8, constant rotations can be expressed by
where and . ( and are indices representing dimensions. In this problem, the motion is in a plane, so dimensions are and , giving us and ). Then, the first term of the integrand is equal to:
Differentiating with respect to time, we get:
Then, the second term of the integrand is equal to:
is equal to:
A symmetry exists if . We have computed the first term of the integrand in terms of indices and . Representing and using and , the expression is equal to:
Similarly, the second term is equal to:
Thus, and a symmetry exists.
STEP 3: Find the conserved quantity :
Using equation (11), we know that when , the conserved quantity us equal t0:
Because , is equal to:
Dropping the constant , the conserved quantity is equal to:
Example 6.13: Lorentz and Galilean Boosts
consider instead the limit of small speeds, i.e., a nonrelativistic system with Galilean symmetry. We take a single free particle in one dimension with Lagrangian
Identity the conserved quantity if a symmetry exists.
Solution:
Again, before solving the problem, recall that a symmetry exists if and is equal to:
In the given Lagrangian, we identify the generalized as (Because and do not appear, the motion is one dimensional, where and . Since time is not given, the infinitesimal variation in time is also .) By Galiean transformation, we know that:
where is velocity. Then, and . In addition, because is cyclic in the Lagrangian, . The integral for is equal to:
If a symmetry exists, , so the integrand and must be a constant. As discussed in section 6.7, there are two cases for : the trivial and non-trivial case. When is trivial, meaning that the function is independent of time, coordinate, or velocity, the conserved quantity is given by equation (11). When is non-trivial, the conserved quantity is equal to , where is the integrand for and is found using (11).
In this problem,
is a function dependent of velocity, so it is non-trivial and the conserved quantity is . Using equation (11), is equal to:The conserved quantity is equal to:
Dropping the constant term , the conserved quantity is equal to: