Quantum Mechanics 1

Fact

Einstein found that for any photon we have the relation

E=hν

Superposition, Entanglement and Interaction Free Measurements

Definition

Spin is fundamental property of elementary particles dictating their internal(intrinsic) angular momentum

Example

linear superposition of spins

|Ψ=|,z+|,z
Example

tensor product states

(α1|u1+α2|u2)(β1|b1+β2|v2)=α1β1|u1|v1+α1β2|u1|v2+α2β1|u2|v1+α2β2|u2|v2

but nice we cant have

|u1|v1+|u2|v2

as that will imply

α1β1=α2β2=1 and α1β2=α2β1=0 which is a contradiction a s that would mean α1β1 and α2β2 is either 1 or 0. So we have an unfactorizable/entangled

Photoelectric effect

Fact

Key findings from photoelectric effect experiments:

  • critical frequency ν0 above which a current is measured but below nothing happens
  • light thought of as quantized bundles of energy(photon) with E=hνwhere h is the planck constant
Definition

For any surface work function(W) is the energy needed to release an electron to the vacuum around the surface

Definition

the compton wavelength of a particle of mass m is λc=hmc

Proposition

De Broglie(1924) stated that:
The wave partice duality is universal for all matter for each particle of momentum ρ we have an associated plane wave of wavelength λp=hρ known as the de Broglie wavelength

de Broglie Waves, Propagation and the Free Particle

we may rexpress de Broglie's relation as

p=hλ=h2π2πλ=h¯k

where h¯ is the reduced planck's constant and k is the wave number

something unusual arises.

Proposition

De Broglie wavelengths are not galliean invariant!

Proof.
First we recall in a galilean transformation we have

x=xvt  and  t=t

so to find the relations for velocity and momentum we take the derivative to see that

dxdt=dxdtv

so that we have v~=v~v for v~ observed in the S frame. With this we obtain

p=pmvλ=hp=hpmvλ

This is in stark contrast with ordinary waves. To see this we use the relation for phase kxwt where k=2πλ is the wave number. Then

ϕ=k(xωkt)=2πλ(xVt)=2πxλ2πVλt.

where we use the fact that ωk=2πf2π/λ=λf=V is the speed of the wave (not the same as the observer's relative velocity v). Since this quantity is Galilean invariant, observers at S and S' should see the same phase for the same point at the same time, meaning that

ϕ=2πλ(xvt)=2πλ(x+vtVt)=2πλx(Vv)t,

which further simplifies to

=2πλx2πλV(1vV)t,

should be equal to the phase ϕ=2πλx2πVλt seen by the moving observer. In particular, the wavenumber and
angular frequency satisfy k=2πλ and ω=2πλV(1vV), so that

ω=ω(1vV),k=kλ=λ
Question

So what is the implication of the wavelength of matter waves changing under gallilean transformations?

First assume the wavefuction form ei(kxωt) which you will learn later. Then you simply sub in the relations we found above. To do so for convenience we choose to write the phase in terms of the momentum and energy of the wave:

1hϕ=kxωt=pxp22mt

Where momentum undergoes the transformation

p=pmv

Thus, this must be equal to:

1hϕ=(p+mv)(x+vt)(p+mv)22mt=px+pσt+mvx+mv2tp22mtpσt12mv2t=px+mv(xvt)p22mt+12mv2t=px+mvxmv2tEt+12mv2t=pxEt+mvx12mv2t

The first two terms are the phase of the primed wavefunction. The last two terms gives rise to f(x,t) to be:

f(x,t)=exp(i(mvx12mv2t))
Proposition

Second de Broglie Relation:
Previously we had p=h¯k we find that we also have E=h¯ω

Phase and group velocities

Definition

The group velocity of a wave with phase kxωt is

vgroup=dωdk|k

evaluated at the wavenumber k that we're propagating.

In general the group velocity represents the velocity of a wave packet constructed under superpsotiions

So we then consider a superposition of plane waves ei(kxw(t)t) by

Ψ(x,t)=dkΦ(k)ei(kxω(k)t)

where Ψ(x,t) is the wave packet(plane waves collectively considered together as a packet) defined at position and time

ei(kxω(k)t) represents a wave of wavenumber k and angular frequency ω(k) and Φ(k) is the ampltidude of the wave and this wave number k. Now suppose that Φ(k) is peaked at k=k0.

now write ω as a taylor expansion

ω(k)=ω(k0)+(kk0)dωdk|k=k0+O((kk0)2),

because the value of k that matter must be very close to k0 given that Φ(k) peaks there

Ψ(x,t)=dκΦ(k)eikxeiω(k0)teikdωdk|k0teik0dωdk|k0te(negligible)=eiω(k0)teik0dωdk|k0tdkΦ(k)eikxeikdωdk|k0t.

taking the magnitude on both sides we get

|Ψ(x,t)|=|Ψ(xdωdk|k0t,0)|.

this shows that indeed the shape of the wave packet moves at the group velocity

In fact this taylor expansion justifies what is known as the principal of stationary phase which is based on the intuitition that say for an expression like f(x)sinx which is positive half the time and negative the half the time so it will contribute very little if f is slowly varying relative to sinx

Specifically in our case taking the phase ϕ(k)=kxω(k)t interpret slowly varying/stationary phase by

0=dϕ(k)dk|k0=xdω(k)dk|k0t=0,

where we get x=dωdk|k0t which also shows the group velocity is indeed the speed at which the wave packet is propagating.

5. Wavefunction

To get a general form for a wavefunction we first choose from 4 potential periodic functions expressed in terms of phase kxωt

  1. sin(kxωt)
  2. cos(kxωt)
  3. ei(kx,ωt)=eikxeiωt
  4. ei(kxωt)=eikxeiωt

The wave function represents the probability of particle as position at a certain time. Hence at any time, the total the wave function cannot be all zero as that will mean the all the entire wave "vanished"

However for 1 and 2 clearly a certain periodic time intervals they will disappear. For example for (1)

Ψ(x,t)=sin(kxωt)+sin(kx+ωt)

as there is an equal probaility of the particle found moving in the +x and x directions but upon expanding this we get

Ψ(x,t)=2sin(kx)cos(ωt)

which is zero when ωt=(π2,3π2,5π2). In contrast for 3,4 take for example 3 we have again

Ψ=ei(kxωt)+ei(kxωt)=2coskxeiωt

which is again because there is an equal probability of moving +x and x but here it is never zero any time for all values of x. Using a similar process we obtain 2coskxeiωt

Question

Therefore because both 3 and 4 are valid solutions we superimpose in them to get the full solution?

Nope we cant

Ψ(x,t)=ei(kxωt)+ei(kxωt)=2cos(kxωt)

but this is the form of (2) which we have proven above is not valid! So we must choose between (3) and (4). The choice happens to be a matter of convnetion which is defined to be

Ψ(x,t)=ei(kxωt)

to be the free particle wave function where p=h¯k,E=h¯ω,E=p22m if you recall

We now want to find an operator that can extract information from the wave function, something like a derivative. More precisely notice that

h¯ixΨ(x,t)=h¯i(ik)Ψ(x,t)=h¯kΨ(x,t)=pΨ(x,t)

so it makes sense to identify this with what is known as the momentum operator p^

Definition

momentum operator is defined by

p^h¯ix

and so we have

p^Ψ=pΨ

this looks quite similar to linear algebra. Remember eigenvalues where we have Ax=λx we have a similar form here. With this inspiration the function Ψ(x,t) is called an eigenstate of the operator p^ with eigenvalue p.

Now that we have extracted the momentum it makes sense that we also want to extract information regarding the energy of the wave function. Now instead of taking the spatial derivative lets see what happens when we take the time derivative instead so now we obtain

ih^tΨ=(ih^)(iω)Ψ=h^ωΨ=EΨ

since Ψ=eikxeiωt. This results in yet another eigenvalue equation

It turns out we can extract more physics by combining this with our previous relation for momentum. In particular because E=p22m and pΨ=p^Ψ we have

EΨ=p2m(pΨ)=p2mh^ixΨ

where we have simply replace (pΨ) with the momentum operator. But because the p term in the p2mfraction outside is also a constant(relative to x) we can move it in like so to obtain

12mh^ix(pΨ)=12mh^ix(h^ixΨ)=h^22m2x2Ψ

this then motivates the following definition

Definition

energy operator:

E^=h¯22m2x2

so combining this with above we can see that

h¯22m2x2Ψ=ih¯tΨ

this is known as the free schrodinger equation. Looking at the energy operator E^=p^22m it looks like our usual "kinetic energy" in classical mechanics. The next natural step is to then add in a potential V(x,t) so that our total energy is E=p22m+V. So we then modify our energy operator expression using by defining

Definition

Hamiltonian operator

H^=p^22m+V(x,t)

this then allows us to write the general Schrodinger equation

Theorem

1-D Schrodinger Equation:
The the wave function Ψ in one dimension satisfies ih¯tΨ=H^Ψ

ih¯Ψt=(h¯22m2t2+V(x,t))Ψ
Definition

The position operator denoted by x^ multiplies functions by x

x^f(x)=xf(x)

Lets try to investigate the properties of these operators, in particular whether they commute.

Example

We determine whether x^ and p^ commuate by computing the difference between x^p^ϕ and p^x^ϕ for some function ϕ(x,t)

x^(p^ϕ)p^(x^ϕ)=x^(h¯ixϕ)p^(xϕ)=h¯ixxϕh¯ix(xϕ)

now using product rule on the second term we see that

=h¯ixxϕh¯iϕh¯ixxϕ=ih¯ϕ

therefore we realized that they don't commute and writing the operators on one side and the operand on the other we have

(x^p^p^x^)ϕ=ih¯ϕ
Definition

The commutator of two operators A^ and B^ is denoted by [A^,B^] refers to the operator A^B^B^A^

Proposition

We have the equality of operators

[x^,p^]=x^p^p^x^=ih¯

recall that we have our expression for the 1D schrodinger equation from earlier. Can we generalize this to 3D? That is

Theorem

3D Schrodinger Equation:

ih¯Ψt=(h¯22m2+V(x,t))Ψ

we will be dealing with it more in detail when we solve differential equations later but as a preview it is to be noted that we now have

[x^i,p^j]=ih¯δij

Probability Density, Current and Conservation

Now recall that the wave function is a probability function of a particle at (x,t) and returns a complex number. Therefore we must have

1=|Ψ(x,t)2|dx

First we assume some boundary conditions in particular

limx±Ψ(x,t)=0
Remark

While there are technically mathematical examples whereΨ(x,t)does not have a limit as x and the integrral of |Ψ|2 still converges those examples do not pop up in physical contexts and thus we are (mostly) safe to ignore them

in addition we require that the function Ψ does not oscillate too quickly in particular

Ψ(x,t)x is bounded as x±

Essentially we need some "regularity"

Definition

For a wave function Ψ(x,t) with |Ψ(x,t)|2dx equal to some finite number N we say that Ψ is normalizable and we let Ψ=ΨN be the associated normalized wavefunction

That is

|Φ(x,t)|2dx=|Ψ(x,t)|2Ndx=NN=1

So now want to check of the conversation of probability is satisfied that is

Ψ(x,t0)Ψ(x,t0)dx=1

Here we let Ψ(x,t0)Ψ(x,t0) be called the probability density and denote it by ρ(x,t). Then we define N(t)=ρ(x,t)dx and we then wish to show that N(t)=1 for all t>t0 given that N(t0)=1

We can show this by showing dNdt=0 so

dNdt=tρ(x,t)dx=(ΨtΨ+ΨΨt)dx

where we basically differentiated under integral since bounded. Then using the relation from Schrodinger equation

ih¯=H^Ψt=ih¯H^Ψ

But we also have a time deritive of the complex conjugate above so

ih¯(Ψt)=(H^Ψ)Ψt=ih¯(H^Ψ)

because the complex conjugate of the partial derivative is also the the partial derivative of the complex conjugate. Plugging these back we get

dNdt=ih¯((H^Ψ)ψΨ(H^Ψ))dx

and in order for this to be zero we essentially require

(H^Ψ)Ψdx=Ψ(H^Ψ)dx

to hold. This motivates the following definition

Definition

For any (linerar) opreator T the Hermitian conjugate of T denoted by T or T is the linear operator that satisfies

Ψ1TΨ2=(TΨ1)Ψ2

we say that T is Hermitian if T=T

We now plug in the from the definition of H^=h¯22m2x2+V(x,t) to get

dNdt=i((H^Ψ)\*ΨΨ\*(H^Ψ))dx=i((H^Ψ)\*ΨΨ\*(H^Ψ))dx=i(22m2Ψ\*x2Ψ+V(x,t)Ψ\*Ψ+22mΨ\*2Ψx2Ψ\*V(x,t)Ψ)dx,

We have used the fact that V(x,t) is always real(because it is some energy). Therefore these terms cancel out and we are left with

=i2m(2Ψ\*x2ΨΨ\*2Ψx2)dx.

A common way in physics to show an integral vanishes is by showing that is a total deriviatve

2Ψ\*x2ΨΨ\*2Ψx2=x(Ψ\*xΨΨ\*Ψx)

plugging this back in we get

dNdt=x[2im(Ψ\*ΨxΨΨ\*x)]dx=[h¯2im(ΨΨxΨΨx)]=0

because remember that Ψx and  Ψ x is bounded from our boundary conditions. Therefore our definition for the wave function is indeed normalizable

However we also notice inside the brackets we have the form zz=2iIm(z) so therefore we may rewrite this as

ρt=x[mIm(Ψ\*Ψx)].
Definition

The current density J(x,t) for the wavefunction Ψ(x,t) is given by

J(x,t)=h¯mIm(ΨΨx)

and we so may rewrite our above calculations in the simpler form

pt=Jxpt+Jx=0

which is a current conservation statement similar to what we might have seen in electromagnetism.

Fourier Transforms, Uncertainty and Time Evolution

We begin today by discussing wave packets and uncertainty.

Consider a wave function of the form

Ψ(x,0)=12πϕ(k)eikxdk

this is essentially a superposition of plane waves(eikx) of different wavelengths. Just see that the integral is simply a continuous sum

Theorem

Fourier Inversion:
Suppose Ψ(x,0)=12πϕ(k)eikxdk we can recover coefficients via fourier inversion

ϕ(k)=12πΨ(x,0)eikxdx

we briefly touched on this in our earlier discussions on group velocity but not we focus on how to understand uncertainties of position and momentum through this result

first consider Φ(k) symmetric and centered about k=k0 with width Δk

attachments/Quantum Mechanics 1 2025-06-12 16.37.34.excalidraw.png

Proposition

Ψ(x,0) is real if and only if Φ(k)=Φ(k) is real

Proof.
Begin by complex conjugating the expression (1.1) for Ψ(x,0):

Ψ\*(x,0)=12πΦ\*(k)eikxdk=12πΦ\*(k)eikxdk.

In the second step we let kk in the integral, which is allowable because we are integrating over

all k, and the two sign flips, one from the measure dk and one from switching the limits of integration,

cancel each other out. If Φ\*(k)=Φ(k) then

Ψ\*(x,0)=12πΦ(k)eikxdk=Ψ(x,0),

as we wanted to check. If, on the other hand we know that Ψ(x,0) is real then the equality of Ψ\* and

Ψ gives

12πΦ\*(k)eikxdk=12πΦ(k)eikxdk.

This is equivalent to

12π(Φ\*(k)Φ(k))eikxdk=0.

This equation actually means that the object over the brace must vanish. Indeed, the integral is

computing the Fourier transform of the object with the brace, and it tells us that it is zero. But a

function with zero Fourier transform must be zero itself (by the Fourier theorem). Therefore reality

but the condition that Ψ(k)=Ψ(k) is not true in our case unless k0=0(essentially we need Ψ(k) to be symmetric about zero for this condition to be true)

Momentum Space and Expectation Values

Recalling that we have the fourier decompositions

Ψ(x)=12πΦ(k)eikxdk,Φ(k)=12πΨ(x)eikxdk.

you might decide to combine both to get a single expression for the wave function like so

Ψ(x)=12π[12πΨ(x)eikxdx]eikxdk,

rearranging in this way we seemed to have isolated a perculiar "function"(spoiler it isn't a function)

Ψ(x)=Ψ(x)12πeik(xx)dkdx.
Definition

The Dirac delta function is a "function" given by

δ(xx)=12πeik(xx)dk.

we have essentially demonstrated the sifting property of the dirac delta function that is

Ψ(x)=Ψ(x)δ(xx)dx

where

δ(xx)={1x=x0xx

note that δ isn't even a function. The cases the above is just an informal way of describing the effect of integrating Ψ(x) with the dirac delta function δ(xx). You should refer to Differential Analysis to learn what exactly δ is and why it isn't a function.

Recall that we prove our normalization condition by showing that

Ψ(x)Ψ(x)dx=1

We now want to do the same for Φ(x).

First using different variables of integration k, k' for the two terms (but the same x):

Ψ\*(x)Ψ(x)dx=(12πΦ\*(k)eikxdk)(12πΦ(k)eikxdk)dx.

From here, we know that we can't do the integrals over k in general, since those are in the most general form possible, and thus it makes sense to try to do the x-integral. Rewriting the order of the integrals, we have

=Φ\*(k)Φ(k)(12πei(kk)xdx)dkdk,

and now the inner integral is indeed the delta function δ(kk) , just with different dummy variables, so this simplifies to

=Φ\*(k)Φ(k)δ(kk)dkdk=Φ\*(k)Φ(k)dk.

Essentially we just proved the following

Theorem

Parseval-Plancheral Identity:
For any function Ψ(x) with Fourier transform Φ(k), we have

|Ψ(x)|2dx=|Φ(k)|2dk.

Recalling that Ψ(x,t) has a probabilistic interpretation and that Ψ(x,t)Ψ(x,t)dx=1 plancheral identity hints that Φ(k) perhaps too as a probabilistic interpretation.

First recall from our discussions above regarding the momentum operator that we associate our plane waves(eikx) with eigenstates of momentum p^Ψ(x,t)=pΨ(x,t). Knowing that p=h¯k let us attempt to write plancheral relation in terms of momentum instead of k. In doing so we obtain

Ψ(x)=12πΦ~(p)eipx/dp,Φ~(p)=12πΨ(x)eipx/dx.

by performing change of variables p=h¯k and Φ~(p)=Φ(k). We write this in a more symmetric form by letting Φ~(p)Φ(p)h¯ to get

Ψ(x)=12πΦ(p)eipx/dp,Φ(p)=12πΨ(x)eipx/dx.dx|Ψ(x)|2=dp|Φ(p)|2.

with this we make the following proposition

Proposition

|Φ(p)2| is the probability to find the particle with momentum in the range (p,p+dp)

let us verify this by discussing the expectation values of operators which is one of the first steps towards a full interpretation of quantum mechanics.

Definition

Let Q be a random variable which can take values in a finite set {Q1,,Qn} with probabilities p1,,pn, respectively. Then Q has expectation or expected value

Q=i=1npiQi.

so since we took ΦΦ to be our probability density function(for particle to be found in [x,x+dx]) our expected value for x will then be

x^xΨ\*(x,t)Ψ(x,t)dx.

similarly for momentum we have the expectation of the momentum operator as

p^pΦ\*(p,t)Φ(p,t)dp.

but what form does this exactly take? lets see

p^=pΦ\*(p,t)Φ(p,t)dp=dppdx2πeipx/Ψ\*(x,t)dx2πeipx/Ψ(x,t)=12πdxΨ\*(x,t)dxΨ(x,t)dppeipx/eipx/=12πdxΨ\*(x,t)dxΨ(x,t)dp(ix)eipx/eipx/=dxΨ\*(x,t)dxΨ(x,t)(ix)12πdpeipx/eipx/.

letting p=h¯u in the final integral we have

12πdp eipx/eipx/=12πdu eiu(xx)=δ(xx).

as a result

p^=dxΨ\*(x,t)dxΨ(x,t)(ix)δ(xx)=dxΨ\*(x,t)(ix)dxΨ(x,t)δ(xx),

where we have changed the order of integration and so finally by sifting property

p^=dxΨ\*(x,t)(ix)Ψ(x,t),

so we have shown

p^=dxΨ\*(x,t)p^Ψ(x,t),p^=ix.

note the position of the p^ operator! It acts on Ψ(x)!
this motivates the definition of any operator Q

Definition

Expectation of Operator is defined by

Q^=dxΨ(x,t)Q^Ψ(x,t)

Hermitian Operators, Measurement and Uncertainty

Definition

Define the inner product which takes in two wavefunctions Ψ1,Ψ2 and outputs

(Ψ1,Ψ2)=Ψ1(x)Ψ2(x)dx
Proposition

For any Hermitian operator Q^ the expectation value Q^Ψ is real

Proof. Notice that

((Q^)ω)\*=(Ψ\*Q^Ψ)\*dx

(if we want to take the complex conjugate of an integral, we can just conjugate the integrand). The integrand is
the product of Ψ\* and Q^Ψ, which are each complex-valued (an important note here: we are not thinking of just
conjugating Q^ on its own, because it is acting on a wavefunction), and thus we can rewrite this as

=Ψ(Q^Ψ)\*dx=(Q^Ψ)\*Ψdx.

where we've used the fact that (Ψ\*)\*=Ψ. And now we use the Hermiticity of Q^ to move the Q^ from one term to
the other:

=Ψ\*(Q^Ψ)dx=Q^Ψ.

Proposition

The eigenvalues of a hermitian operator Q^ are always real

Proof. Let q1 be an eigenvalue of Q^ with associated eigenvector Ψ1. There are many ways to show this result, but we'll apply Proposition 70 to Q^ in the state Ψ1. Then we have the real number

Q^Ψ1=Ψ1\*Q^Ψ1dX=Ψ1\*q1Ψ1dx

(by the eigenvalue condition), and then taking q1 out of the integral gives us

=q1Ψ1\*Ψ1dx.

The integrand of this integral is always a positive real number, so the integral is real. Because our left-hand side Q^Ψ1 was also real, our eigenvalue q1 is indeed real.

Remark

the proves here are similar to how you prove a hermitian form has real eigenvalues

ΨTΨ=(TΨ)Ψ  and  XTAX=(AX)TX

where the proof follows by

qΨΨ=qΨΨ  and  λXTX=λ¯XTX

Proposition

For any Hermitian operator Q^, consider its eigenvalues and eigenfunctions given by

q^ψ1=q1ψ1,q^ψ2=q2ψ2,

(the set of which can be finite or infinite). then the eigenfunctions can be organized to satisfy the orthonormality relation

Ψi\*(x)Ψj(x)dx=δij

Proof . We'll consider the case where the eigenvalues of different eigenvectors are always different (meaning that qiqj if ij). In this case, we can always first normalize our eigenfunctions by rescaling. Then we have

Ψi\*Q^Ψjdx=Ψi\*qjΨjdx=(qjΨi\*Ψj)

but also by Hermiticity we have

Ψi\*Q^Ψjdx=(Q^Ψi)\*Ψjdx=(qiΨi)\*Ψj=(qjΨi\*Ψj)

where we know that qi\*=qi by the previous proposition. Equating the two boxed expressions, because qiqj, we must have ΨiΨj=0 as desired

Remark

Again this is analogous to the linear algebra version in that for any symmetric/hermitian form on vector space V there exists an orthogonal basis for V. Here we have also assumed that that eigenfunctions Ψi indeed span the space of Ψ. We will talk about this more in Quantum Mechanics 2

Proposition

The eigenfunctions of any Hermitian operator Q^ form a set of basis functions, meaning that any (reasonable) wavefunction Ψ can be written as a superposition$$\Psi = \alpha_1 \Psi_1 + \alpha_2 \Psi_2 + \cdots = \sum_i \alpha_i \Psi_i.$$

the previous proposition makes it easy to calculate these coefficients αi. Consider

(Ψi,Ψ)=Ψi\*jαjΨjdx

(we should make sure not to reuse indices here, which is why we sum over j), and now we can switch the sum and
integral to get

=jαjΨi\*Ψjdx.

But orthonormality tells us that the only term that survives is i=j:

=jαjδij=αi.

Thus, we compute the coefficient αi (and thus how much of the wavefunction Ψ is "along" the state Ψi) by integrating
Ψi\*Ψ. Additionally, we can compute through expansion that

1=|Ψ|2dx=Ψ\*Ψdx=(iαiΨi)\*jαjΨjdx,

and now moving the sum over i, j out, we get

ijαi\*αjΨi\*Ψjdx=ijαi\*αjδij=iαi\*αi=i|αi|2.

In particular we have just shown a pretty powerful result

Fact

If we have a state which is a superposition of orthonormal basis vectors then the sums of squares of cofficients gives the normalization condition

We can sum our discussions up here so far with what is known as the measurement postulate

Suppose we measure a Hermitian operator Q^ in a state Ψ. Then the possible results that we may obtain are the eigenvalues q1,q2,, and the probability of measuring qi is

pi=|αi|2=|(Ψi,Ψ)|2.

If the outcome of the measurement is qi, then the state of the system becomes Ψi (this is called the collapse of the wavefunction).

Definition

Let Q be a random variable which takes on values Q1,,Qn with probabilities p1,,pn, and suppose its expectation value is Q=Q=ipiQi (we use Q and Q interchangedby). Then the variance of Q is

(ΔQ)2=ipi(QiQ)2,

and the uncertainty or standard deviation of Q is ΔQ=(ΔQ)2.

also note that

(ΔQ)2=pi(QiQ)2=piQi22piQiQ+piQ2.

noting that Q¯ is a constant we have

=Q22QipiQi+Q2ipi=Q22QQ+Q2=Q2Q2.

this is basically the usual VarX=EX2(EX)2 from elementary probability theory

With this we now define

Definition

Let Q^ be a Hermitian operator. We define the uncertainty of Q^ in the state Ψ, denoted (ΔQ^)Ψ2, to be

(ΔQ^)Ψ2=Q2ΨQ^Ψ2.
Lemma

We may write the uncertainty of Q^ in the state Ψ as a single expectation

(ΔQ^)Ψ2=(Q^Q^)2Ψ,

and we can also write it as an integral

(ΔQ^)Ψ2=|(Q^Q^)Ψ|2dx.

Proof . For the first claim, we do a direct computation (remembering that Q^ is thought of as the operator which multiplies any state by the constant Q^). We expand the square, noticing that Q^ and Q^ commute (because the former is a number)

(Q^Q^)2=Q^22QQ^+Q^2.

(Notice that (A+B)2=A2+AB+BA+B2A2+2AB+B2 if A and B don't commute.) Now the expectation of a sum is the sum of the individual expectations, so we can further simplify to

=Q^22Q^Q^+Q^2

(where we've used the fact that we can take constants out of expectations: cQ^=cQ^). Combining the last two terms indeed gives us the definition of the uncertainty of Q^, as desired.
For the second claim, we start with the expression (Q^Q^)2 and plug in the definition of an expectation value:
we find that

(ΔQ^)2=Ψ\*(x)(Q^Q^)2Ψ(x)dx=Ψ\*(x)(Q^Q^)Ψ(x)dx.

We now think of the first (Q^Q) as an operator acting on the rest of the integrand – because Q^ is Hermitian and so is Q (multiplying by a real constant is equivalent no matter which part of the integrand it's done on), the whole
term (Q^Q) is Hermitian, and thus by the definition of Hermiticity we have

=((Q^Q^)Ψ(x))\*(Q^Q^)Ψ(x)dx=|(Q^Q^)Ψ(x)|2dx,

as desired.

Corollary

A state Ψ is an eigenstate of a Hermitian operator Q^ (meaning that Q^Ψ=λΨ) if and only if (ΔQ^)2=0. In addition, we have λ=Q^ (so that we can write Q^Ψ=Q^ΨΨ).

Proof. We show the second claim first: to do so, we integrate

Q^=Ψ\*Q^Ψdx=Ψ\*λΨdx=λ|Ψ|2dx=λ,

The second part simply follows directly from the lemma knowing now that λ=Q^ we subsitute it into the lemma to get

(ΔQ^)Ψ2=|(Q^λ)Ψ|2dx

and the 1st part clearly follows

Stationary States and the Particle on a Circle

stationary states

Consider the

iΨt=H^Ψ(x,t)=(22m2x2+V(x))Ψ(x,t),
Definition

a stationary state is a quantum state with all observables indpedent of time

Theorem

A distinguising featuer of stationary states are such that solutions of the wavefunction take a "separable" form

Ψ(x,t)=g(t)ψ(x)

we now demonstrate this upon rearranging the schrodinger equation we have and subbing in Ψ=g(t)ψ(x) we have

i1g(t)dg(t)dt=1ψ(x)H^ψ(x)=E

as we know from basic PDE theory both sides are in terms of different indepent variables so they must be equal to some constant which we define to be E as above. So on one hand we have

idgdt=Eg,g(t)=eiEt/

on the other hand we have what we call the time independent schrodinger equation

H^ψ(x)=Eψ(x)

which shows that E must be real because eigenvalues of Hermitian operators must be real as proven previously.

Now to show that such a separable solution admits a solution that is time independent consider that now we have(subbing our relation for g(t))

Ψ(x,t)=eiEt/ψ(x),ER,H^ψ=Eψ

First consider what happens if E was not real. We attempt to set the boundary normalization condition for Ψ as usual

1=dxΨ\*(x,t)Ψ(x,t)=dxeiE\*t/heiEt/hψ\*(x)ψ(x)

= ei(E\*E)t/hdxψ\*(x)ψ(x)=e2Im(E)t/hdxψ\*(x)ψ(x).
which implies that the normalization condition is time dependent which is not what we want. Instead if E were real we would get instead

dx ψ\*(x)ψ(x)=1.

which is time indepedent as desired.

Question

So why do we want the normalization condition to be time independent?

Because it implies expectation value of H^ is also time independent

H^Ψ=dx Ψ\*(x,t)H^Ψ(x,t)=dxΨ\*(x,t)EΨ(x,t)=EdxΨ\*(x,t)Ψ(x,t)=E,
Fact

In fact we can generalize our observations regarding stationary states

  • (1) The expectation value of any time-independent operator Q^ on a stationary state Ψ is time-independent:
QΨ(x,t)=dxΨ\*(x,t)Q^Ψ(x,t)=dxeiEt/ψ\*(x)Q^eiEt/ψ(x)=dxeiEt/eiEt/ψ\*(x)Q^ψ(x)=dxψ\*(x)Q^ψ(x)=Qψ(x),=dxeiEt/eiEt/ψ\*(x)Q^ψ(x)=dxψ\*(x)Q^ψ(x)=Qψ(x),
  • (2) In general superposition of stationary states with different energies not stationary. This is clear because a stationary state requires a factorized solution of the Schrödinger equation: if we add two factorized solutions with different energies they will have different time dependence and the total state cannot be factorized. We now show that that a time-independent observation values in such a state. Consider a superposition Ψ(x,t)=c1eiEt/ψ1(x)+c2eiEt/ψ2(x), where ψ1 and ψ2 are H^ eigenstates with energies E1 and E2, respectively. Consider a Hermitian operator Q^. With the system in state , its expectation value is
QΨ=dxΨ\*(x,t)Q^Ψ(x,t)=dx(c1\*eiEt1t/ψ1\*(x)+c2\*eiEt2t/ψ2\*(x))(c1eiEt1t/Q^ψ1(x)+c2eiEt2t/Q^ψ2(x))

Particle on Circle

returning to solving for H^ψ(x)=Eψ(x) which is our time independent schrodinger equation if you recall to fully solve for stationary states. Assume that the eigenstates and their energies can counted and that in forms a complete of orthonormal functions

ψi(x)ψj(x)=δij

subbing in the definition of H^(hamiltonian operator) we have

d2ψdx2=2m2(EV(x))ψ.

However to understand the general properties of ψ we need to restrict the types of potentials V(x), specifically we will consider the cases where V(x) is continuous and piecewise discontinuous(finite number discontinuities):

Next we claim that we require ψ to be continuous if not ψ will contain delta functions and ψ will be a derivative of delta functions which we do not wnt

Question

So let us consider, how does V(x) having finite discontinuous then why is then implied that ψ must be continuous?

Consider a discontinuity at a like say a single "step" function

Integrate both sides of the above wave equation from aϵ to a+ϵ, and then take ϵ0. We find

aϵa+ϵdxddx(dψdx)=2m2aϵa+ϵdx(EV(x))ψ(x).

The left-hand side integrand is a total derivative so we have

dψdx|a+ϵdψdx|aϵ=2m2aϵa+ϵdx(V(x)E)ψ(x).

By definition, the discontinuity in the derivative of ψ at x=a is the limit as ϵ0 of the left-hand side:

Δa(dψdx)=limϵ0(dψdx)a+ϵdψdx|aϵ.Δa(dψdx)=limϵ02m2aϵa+ϵdx(V(x)E)ψ(x).

The potential V is discontinuous but not infinite around x=a, nor is ψ infinite around x=a and, of course, E is assumed finite. As the integral range becomes vanishingly small about x=a the integrand remains finite and the integral goes to zero. We thus have

Δa(dψdx)=0.
Example

we now consider a particle confined to a circle of circumference L and assume V(x)=0. Then we will have

22md2ψdx2=Eψ(x).

Before we solve this, let us show that any solution must have E0. For this multiply the equation by ψ\*(x) and integrate over the circle x[0,L). Since ψ is normalized(recall above for stationary states normalization condition for wave function implies normalization of ψ(x)) we get

22m0Lψ\*(x)d2ψdx2dx=Eψ\*(x)ψ(x)dx=E.

The integrand on the left hand side can be rewritten as

22m0L[ddx(ψ\*dψdx)dψ\*dxdψdx]dx=E.

and the total derivative can be integrated where we have assumed ψ is well normalized

22m[(ψ\*dψdx)|x=L(ψ\*dψdx)|x=0]+22m0L|dψdx|2dx = E.

Since ψ(x) and its derivatives are periodic, the contributions from x=L and x=0 cancel out are left with

E = 22m0L|dψdx|2dx0,

with this we are now ready to solve. Applying the periodicity condition ψ(x+L)=ψ(x) we have

eik(x+L)=eikxeikL=1kL=2πn

for some integer n. We'll index the allowed values of k as kn=2πnL, and from this we find that the nth momenta and energy levels are

pn=kn=2πnL,En=2kn22m=22m4π2n2L2=2π22n2mL2.

To find our energy eigenstates, we now just need to find the normalization constants for ψn(x)eiknx(we want to find a set of orthonormal eigenfunctions so we can use parasevel theorem etc)

And here's where it's nice that our particle is on a circle rather than on a line: exponentsals like this have |eiknx|2=1, so we would not be able to normalize the integral over all real numbers, but we can normalize the integral over the range [0, L]. Explicitly, the way we do this is to write ψn(x)=Neiknx, and then require

1=0L|ψn(x)|2dx=0LN2dx=LN2N=1L.

(Note that we can choose N to be positive real – an overall phase doesn't change the state.) This gives us our final answer:

ψn(x)=1Leiknx=1Le2πinx/L.

Thus, we've found the energy eigenstates for the time-independent Schrödinger equation, and to obtain our stationary states, we must add in the time phase: for each n, we have

Ψn(x,t)=ψn(x)eiEnt/n.

but recall now that in general superposition of stationary states with different energies is not a stationary state. Now observe that En=En so they are in fact degenerate energy eigensates therefore our stationary state will take the form

Ψ=a1Ψna2Ψn

Infinite and finite square well

Classically any region where the potential exceeds the energy of the particle is forbidden. Not so in quantum mechanics.

But even in quantum mechanics a particle can’t be in a region of infinite
potential.

So now consider the case of the infinite square well where

V(x)={0,0<x<a,x0,x0

so

ψ(x)=0 for x<0 and for x>a.

we will give more intuition on why this is so later but for now let us assume these conditions. Now recall that because we require our wave function to be continuous we also have

ψ(x=0)=0,ψ(x=a)=0
Question

what about ψ(x) is it continuous?

it is clear that the ψ(x) vanishes outside the iterval afterall ψ(x) vanishes there as well. However if we were to also claim that it vanishes at x=0,a it turns out that is not possible because that will imply the solution to the schrodinger equation is indentically zero. if a solution should exists we must have that ψ is discontiuous at an infinite wall.

In the region x[0,a] the potential vanishes and the Schrödinger equation takes the form

d2ψdx2=2mE2ψ,

and as we did before, one can show that the energy E must be positive (do it!). This allows
define, as usual, a real quantity k such that

k22mE2E=2k22m.

The differential equation is then

d2ψdx2=k2ψ,

and the general solution can be written as

ψ(x)=c1coskx+c2sinkx,

bound states and shooting method

Definition

A bound state is a normalizable energy eigenstate over the real line (meaning that ψ0 as x±).

Proposition

Energy eigenstates of the time-independent Schrodinger equation, where V(x) is real, can be chosen to be real.

note eigenstates not eigenvalue which will always be real for hermitian operators

Proof. Consider the Schrodinger equation H^ψ=Eψ. If we take the complex conjugate of both sides, we obtain H^ψ\*=Eψ\* (because the potential V(x) is real, we can check that (H^ψ)\*=H^ψ\*). Thus, ψ\* and ψ are two energy eigenstates of the same energy. If they are linearly independent, then taking ψr=12(ψ\*+ψ) and ψi=12i(ψψ\*) gives us two new real, linearly independent energy eigenstates of the same energy (which we can work with instead). And if ψ\* and ψ start off linearly dependent, then we just have one unique energy eigenstate, and at least one of those linear combinations is nonzero. Thus we can always form enough new energy eigenstates that are real as long as V(x) is real

Proposition

If a one-dimensional potential V(x) is even (that is, V(x)=V(x)), then the energy eigenstates can be chosen to be either even or odd (under xx).

problem sets

PS3

  1. Exercises with commutators [10 points] Let A, B, and C be linear operators.

for (a) by direct definition of commutators above

[A,BC]=[A,B]C+B[A,C][A,BC]ϕ(x)=[A,B]Cϕ(x)+B[A,C]ϕ(x)ABCϕ(x)BCAϕ(x)=ABCϕ(x)BACϕ(x)+BACϕ(x)BCAϕ(x)

similarly for (b)

[AB,C]=A[B,C]+[A,C]B[AB,C]ϕ(x)=A[B,C]ϕ(x)+[A,C]Bϕ(x)ABCϕ(x)CABϕ(x)=ABCϕ(x)ACBϕ(x)+ACBϕ(x)CABϕ(x)

for (c) just expand to get

A(BCCB)(BCCB)A+B(CAAC)(CAAC)B+C(ABBA)(ABBA)C=ABCACBBCA+CBA+BCABACCAB+ACB+CABCBAABC+BAC

for (d)
First, note that:

x^nϕ(x)=xnϕ(x),p^nϕ(x)=(i)nnϕ(x)xn

Thus,

[x^n,p^]ϕ(x)=xniϕ(x)x+ix(xnϕ(x))=xniϕ(x)x+i(nxn1ϕ(x)+xnϕ(x)x)=inxn1ϕ(x)