News

My basil finally developed roots! I'm currently reading about quantum mechanics, ferrofluids, and language models.

Blog

[08/09/2024:] Motivating Ladder Operators II
[08/08/2024:] A Cool Way to Garden
[08/08/2024:] Life and Basil Limeade
[08/05/2024:] Motivating Ladder Operators
[08/04/2024:] Migrated to AstroJS
[07/31/2024:] The Classical in the Quantum

Notes

Working on notes on the quantum mechanics, derivatives, and uploading my previous course notes onto this blog!

Projects

Finally started a projects page! I've recently made some nice upgrades to my post component, so it looks pretty clean! ;)

🌊

I'm considering whether or not to continue this project using WebGL or Three.js.

I'm also researching methods for generating the 3D scenes I want for this project automatically.

In the meantime, I've decided to proceed with some preliminary prototypes of the other interactive parts of this project.

Orange Juice

I like orange juice. :)

Mlog


Motivating Ladder Operators

August 5, 2024
By Aathreya Kadambi

Recently I’ve been reading about spin and angular momentum operators in quantum mechanics, which both satisfy some very nice commutation relations:

[Ai,Aj]=iℏϡijkAk[A_i, A_j] = i\hbar\epsilon_{ijk} A_k
for A=LA = L or A=SA = S. Somehow, we are repeatedly able to use so called ladder operators to find that an operator can only take on a discrete set of eigenvalues. Apparently the two index cards in the picture above weren’t enough to figure it all out, so I’ve decided to utilize my new ability to make KaTeX+MDX blog posts.

Contents

  1. Jumbled Thoughts
  2. A Path to The Essence (Possibly)
  3. Random Scratch Work

Jumbled Thoughts

I’ll start out with a vague (potentially nonsensical) description of what I think is so interesting about the ladder operators technique, and then try to generalize it. We are able to construct so called β€œladder operators” which are such that when you act them on an eigenvector ff of AA with eigenvalue ΞΌ\mu, they have the property that they return a new eigenvector of AA with eigenvalue ΞΌΒ±c\mu \pm c. Namely,

A3(AΒ±f)=(Ξ»+c)(AΒ±f).A_3 (A_{\pm} f) = (\lambda + c)(A_{\pm} f).
Somehow, with this clever operator, we are able to actually discover that the set of eigenvalues of A3A_3 are discrete. Here the choice of 33 seems soemwhat arbitrary, it’s just to mimic the way it is done with LzL_z or SzS_z.

To me, it seems that the crux of the ladder argument is the above, that we can find an operator that β€œbumps up” an eigenvector to another eigenvector with increased eigenvalue. By doing this, we are actually breaking up R\R (the set of all possible eigenvalues) into a bunch of little ladders and offsets: cZ+rc\Z + r where r∈[0,c)r \in [0, c). Denote by EΞ»(Ξ©)E_\lambda(\Omega) the set of all eigenvectors of Ξ©\Omega corresponding to an eigenvalue Ξ»\lambda. Then AΒ±A_{\pm} give us maps:

AΒ±:EΞ»(A3)β†’Eλ±c(A3).A_{\pm} : E_{\lambda}(A_3) \rightarrow E_{\lambda \pm c}(A_3).

The other big trick is having an operator A2:=A12+A22+A32A^2 := A_1^2 + A_2^2 + A_3^2 that commutes with A3A_3:

[A3,A2]=[A3,A12]+[A3,A22]=iℏA1A2+iℏA2A1βˆ’iℏA2A1βˆ’iℏA1A2=0.[A_3, A^2] = [A_3, A_1^2] + [A_3, A_2^2] = i\hbar A_1A_2 + i\hbar A_2A_1 - i\hbar A_2A_1 - i\hbar A_1A_2 = 0.
This allows us to restrict our attention to just simultaneous eigenvectors of both A2A^2 and A3A_3, which allow us to fix an eigenvalue of A2A^2 and focus on eigenvalues of A3A_3. Somehow, with this fact, we are able to bound the eigenvalues of A3A_3, and combined with the previous ladder operators, we are able to actually discover that the eigenvalues of both A2A^2 and A3A_3 are discrete! A big step in this is the fact that:
Aβˆ“AΒ±=A2βˆ’A32Β±iℏA3,A_{\mp}A_{\pm} = A^2 - A_3^2 \pm i\hbar A_3,
which ends up following from the commutation relation and the construction of A2A^2.

How does it all just work out so perfectly?

If I want to claim that it worked out so perfectly, I should identify what parts seem so coincidental. To me, it seems magical that we can discover that the collection of eigenvalues is discrete in nature from the initial commutation relationship. This leads me to wonder… is there a general use case for these ladder operators? It seems to me that there are a few crucial steps in this process:

  1. Bound the eigenvalues of the operator in question.
  2. Relate eigenvalues by the ladder operators AΒ±A_{\pm}.
  3. It must be that AΒ±kf=0A_{\pm}^k f = 0 for some kk.
  4. If kk is chosen minimally, it must mean that λ±kc=0\lambda \pm kc = 0, so that Ξ»=βˆ“kc\lambda = \mp kc with k∈Zk \in \Z.

One thing that particularly intruiges me is the possibility that the discretness comes from the very fact that we can bound the eigenvalues.

To examine this further, we can try to compare this to the discreization of energy for the particle in a box. In that case, the idea was to factor the hamiltonian operator: H∼A†AH \sim A^\dagger A. Here again, we try to factor L2∼A†AL^2 \sim A^\dagger A. In the case of the hamiltonian, we were left with something of the form:

H=A†A+12ℏωI=A†A+p(I)H = A^\dagger A + \frac{1}{2}\hbar\omega I = A^\dagger A + p(I)
where pp is a polynomial and notice that [H,I]=0[H, I] = 0. In the angular momentum case, we were left with
L2=A†A+L32βˆ“iℏL3=A†A+p(L3)L^2 = A^\dagger A + L_3^2 \mp i\hbar L_3 = A^\dagger A + p(L_3)
where pp is again a polynomial and notice that [L2,L3]=0[L^2, L_3] = 0. Very interesting indeed. But this doesn’t shed light on why they increment and decrement the eigenvalues of eigenvectors. Maybe we just find operators that both increment/decrement the eigenvalues and satisfy the above properties?

We can actually see the β€œladderness” of the ladder operators from the commutation relationship they have with the operator in question. In particular,

[H,AΒ±]=±ℏωAΒ±[H, A_{\pm}] = \pm \hbar\omega A_{\pm}
and for the angular momentum case,
[L2,AΒ±]=Β±0β‹…AΒ±[L^2, A_{\pm}] = \pm 0\cdot A_{\pm}
but we do have that:
[L3,AΒ±]=±ℏAΒ±.[L_3, A_{\pm}] = \pm \hbar A_{\pm}.
Mhm. Very peculiar indeed.

I think a very integral part of these ladder operators is that they satisfy: [A,DΒ±]=Β±cDΒ±[A, D_{\pm}] = \pm c D_{\pm} for some cc. This is important, because then:

A(D±f)=([A,D±]+D±A)f=(λ±c)(D±f).A(D_{\pm}f) = ([A, D_{\pm}] + D_{\pm}A)f = (\lambda \pm c)(D_{\pm}f).
where f∈Eλ(A)f \in E_{\lambda}(A). This encapsulates the idea that the ladder operator DD will increase of decrease the eigenvalue.

Now I’ll get into what I think may be the essence behind thesse ladder operator methods.

A Path to the Essence (Possibly)

To start, I’ll define ladder operator:

Definition. Consider any Hermitian operator AA. We say that an operator LL is a ladder operator for AA of step size cc if

[A,L]=cL.[A, L] = cL.

Theorem. Consider any Hermitian operator AA. If LL is a ladder operator for AA of step size cc, then L†L^\dagger is a ladder operator for AA of step size βˆ’cβˆ—-c^*.

Proof.

[A,L†]=ALβ€ βˆ’L†A=(LAβˆ’AL)†=(βˆ’[A,L])†=βˆ’cβˆ—L.[A, L^\dagger] = AL^\dagger - L^\dagger A = (LA - AL)^\dagger = (-[A, L])^\dagger = -c^*L.
β– \blacksquare

Theorem. Suppose AA is a Hermitian operator and LL is a ladder operator for AA of step size cc where c∈Rc \in \R. Then [A,LL†]=0[A, LL^\dagger] = 0.

Proof.

[A,LL†]=L[A,L†]+[A,L]L†=βˆ’cLL†+cLL†=0[A, LL^\dagger] = L[A,L^\dagger] + [A,L]L^\dagger = -cLL^\dagger + cLL^\dagger = 0
β– \blacksquare

Theorem. Suppose every eigenvalue of AA satisfies λ≀λmax⁑\lambda \le \lambda_{\max} for some eigenvalue Ξ»max⁑\lambda_{\max}. Then if LL is a ladder operator for AA of step size c∈R+c \in \R^+, and q(A)+L†L=Bq(A) + L^\dagger L = B for some operator BB which commutes with AA where pp and qq are polynomials, the only possible eigenvalues of AA are those in Ξ»maxβ‘βˆ’cN\lambda_{\max} - c\mathbb{N}.

Proof.

If λmax⁑\lambda_{\max} is an eigenvalue, consider a corresponding eigenvector vv which is also an eigenvector of BB with eigenvalue μ\mu (I think you might need that BB is self-adjoint for this, although I need to check again later). Then

A(Lv)=(LA+[A,L])v=(LA+cL)v=LAv+cLv=λmax⁑Lv+cLv=(λmax⁑+c)LvA(Lv) = (LA + [A, L])v = (LA + cL)v = LAv + cLv= \lambda_{\max} Lv + cLv = (\lambda_{\max} + c)Lv
but Ξ»max⁑+c\lambda_{\max} + c cannot be an eigenvalue. Thus, it must be that Lv=0Lv = 0, or equivalently, βˆ₯Lvβˆ₯=0\|Lv\| = 0. Now notice that:
βˆ₯Lvβˆ₯=⟨v,L†Lv⟩\|Lv\| = \langle v, L^\dagger L v\rangle
Now notice that
L†Lv=(q(A)βˆ’B)v=q(A)vβˆ’Bv=q(Ξ»max⁑)vβˆ’ΞΌv=(q(Ξ»max⁑)βˆ’ΞΌ)vL^\dagger Lv = (q(A) - B)v = q(A)v - B v = q(\lambda_{\max})v - \mu v = (q(\lambda_{\max}) - \mu) v
so that
0=βˆ₯Lvβˆ₯=q(Ξ»max⁑)βˆ’ΞΌ0 = \|Lv\| = q(\lambda_{\max}) - \mu
so q(λmax⁑)=μq(\lambda_{\max}) = \mu.

Hmm… there is something still missing here. The thing is, in the case of the energy eigenvalues, we can actually find BB as the identity operator so ΞΌ=1\mu = 1, and Ξ»min⁑\lambda_{\min} is some constant value. In the case of the angular momentum operator, we made AA the L2L^2 operator and BB the LzL_z operator. However, this isn’t quite the result I wanted. I was hoping for something more like β€œboundedness translates into discreteness,” but it looks like somehow, these polynomials end up messing things around.

It feels like beyond this step, the path diverges. Like in the case of the hamiltonian operator for the particle in a box, the fact that Lv=0Lv = 0 can be used to simplify the Shrodinger equation, and in the case of the angular momentum operator, we were able to use the identity of the polynomial in combination with the reverse statement with a Ξ»min⁑\lambda_{\min} to discritize the angular momentum. It seems that things may be more complicated than I thought. One similarity I do see, though, is that we have I=L†L+p(H)I = L^\dagger L + p(H) and L2=A†A+q(Lz)L^2 = A^\dagger A + q(L_z), and II and L2L^2 are similar in that they commute with everything else, LL and AA are the respective ladder operators of HH and LzL_z, and qq and pp are polynomials. I feel like there has to be something there.

I’ll have to think about it more, and I’ll make an update in a future post!

Random Scratch Work

Remark. Warning, the following stuff probably has errors, which is why it is down here. But the errors did help me realize and understand things better.

We use the well-ordering principle (or equivalently, we could frame this via induction). Let SβŠ†NS \subseteq \N be the set of all nβ‰₯0n \ge 0 such that the only eigenvalue of AA in Ξ»maxβ‘βˆ’c[n,n+1)\lambda_{\max} - c[n, n+1) is Ξ»maxβ‘βˆ’n\lambda_{\max} - n. Now let Sc=N\SS^c = \mathbb{N}\backslash S. Suppose for contradiction that ScS^c is nonempty. By the well-ordering principle, there exists a smallest value mm in ScS^c. Then there must be an eigenvalue in Ξ»maxβ‘βˆ’c[m,m+1)\lambda_{\max} - c[m, m+1) which is not Ξ»maxβ‘βˆ’cm\lambda_{\max} - cm. Denote it by Ξ»maxβ‘βˆ’c(m+r)\lambda_{\max} - c(m+r), where r∈[0,1)r \in [0,1). Denote the corresponding eigenvector by vv.

Now, if Lv≠0Lv \neq 0,

A(Lv)=(LA+[A,L])v=(LA+cL)v=LAv+cLvA(Lv) = (LA + [A, L])v = (LA + cL)v = LAv + cLv
=(Ξ»maxβ‘βˆ’c(m+r))Lv+cLv=(Ξ»maxβ‘βˆ’c(m+rβˆ’1))Lv= (\lambda_{\max} - c(m+r))Lv + cLv = (\lambda_{\max} - c(m+r-1))Lv
but Ξ»maxβ‘βˆ’c(mβˆ’1+r)<Ξ»maxβ‘βˆ’c(m+r)\lambda_{\max} - c(m-1+r) < \lambda_{\max} - c(m+r).

If m=0m = 0, then Ξ»maxβ‘βˆ’c(mβˆ’1+r)=Ξ»max⁑+c(1βˆ’r)>Ξ»max⁑\lambda_{\max} - c(m-1+r) = \lambda_{\max} + c(1-r) > \lambda_{\max}, which can’t possibly be an eigenvalue because Ξ»max⁑\lambda_{\max} was the largest eigenvalue, which is a contradiction. If mβ‰₯1m \ge 1, mβˆ’1m-1 is in ScS^c as well, which is a contradiction of the minimality of mm.

So since both cases lead to contradiction, it must be that Lv=0Lv = 0, or equivalently, βˆ₯Lvβˆ₯=0\|Lv\| = 0. Now notice that:

βˆ₯Lvβˆ₯=⟨v,L†Lv⟩\|Lv\| = \langle v, L^\dagger L v\rangle
Now notice that
L†Lv=(Aβˆ’p(B))v=Avβˆ’p(B)v=(Ξ»maxβ‘βˆ’c(m+r))vβˆ’p(ΞΌ)v=(Ξ»maxβ‘βˆ’c(m+r)βˆ’p(ΞΌ))vL^\dagger Lv = (A - p(B))v = Av - p(B) v = (\lambda_{\max} - c(m+r))v - p(\mu)v = (\lambda_{\max} - c(m+r) - p(\mu)) v
so that
0=βˆ₯Lvβˆ₯=Ξ»maxβ‘βˆ’c(m+r)βˆ’p(ΞΌ)0 = \|Lv\| = \lambda_{\max} - c(m+r) - p(\mu)

In either case, we have achieved a contradiction, so ScS^c is indeed empty, and so S=NS = \N. Thus, for all nn, the only eigenvalue of AA in Ξ»maxβ‘βˆ’c[n,n+1)\lambda_{\max} - c[n, n+1) is Ξ»maxβ‘βˆ’n\lambda_{\max} - n, so indeed, the only possible eigenvalues of AA are those in Ξ»maxβ‘βˆ’cN\lambda_{\max} - c\mathbb{N}

β– \blacksquare

By essentially the same proof but using L†L^\dagger in place of LL, we also have that:

Theorem. Suppose every eigenvalue of AA satisfies Ξ»β‰₯Ξ»min⁑\lambda \ge \lambda_{\min} for some Ξ»min⁑\lambda_{\min}. Then if LL is a ladder operator for AA of step size c∈R+c \in \R^+, the only possible eigenvalues of AA are those in Ξ»min⁑+cN\lambda_{\min} + c\mathbb{N}.

So indeed, the boundedness of the eigenvalues of an operator translates directly into discreteness! I should specify that by discrete, I mean that there is a minimum distance between two points in the set, or even more strongly, that the set is a subset of cZc\Z for some cc. Even more interestingly, this result seems to suggest that there cannot be two ladder operators of step sizes different in magnitude same operator. That does seem magical. (Just kidding, I had an error in this proof)

In other words, to identify whether an operator is discrete, you simply need to

  1. Find a ladder operator for it with step size cc.
  2. Find the minimum and/or maximum eigenvalues for the operator.

And boom. You have now shown that the eigenvalues of this operator are precisely those in the set:

{Ξ»min⁑+ck:k∈Z,0≀k≀λmaxβ‘βˆ’Ξ»min⁑c}\{\lambda_{\min} + ck : k\in \Z, 0 \le k \le \frac{\lambda_{\max} - \lambda_{\min}}{c}\}
This actually motivates setting up a β€œquantum number” for it. We might call it ll, and take l=Ξ»cl = \frac{\lambda}{c}. This way, the eigenvalues of the operator are those in the set:
{l∈Z:lmin⁑≀l≀lmax⁑}\{l \in \Z : l_{\min} \le l \le l_{\max}\}
where lmin⁑=λmin⁑cl_{\min} = \frac{\lambda_{\min}}{c} and lmax⁑=λmax⁑cl_{\max} = \frac{\lambda_{\max}}{c}..

If we then have bounds on the maximum and minimum possible eigenvalues for AiA_i (here AiA_i is self-adjoint so that all eigenvalues are real, which is an ordered set), we can actually see that there must be some maximal number of times that we can apply AΒ±A_{\pm} to an eigenvector ff of AA until we obtain zero. Else, we would have discovered a new eigenvector corresponding to a large enough or small enough eigenvalue, which is not possible. So there must be extreme eigenstates of AiA_i which can be obtained by repeated application of the ladder operators, such that AΒ±f=0A_{\pm} f = 0, or equivalently, βˆ₯AΒ±fβˆ₯=0\|A_{\pm} f\| = 0.

This is where the magical part happens. It just so happens that A+A_+ and Aβˆ’A_- are adjoints of each other, and so it is possible to write:

βˆ₯AΒ±fβˆ₯=⟨AΒ±f,AΒ±f⟩=⟨f,Aβˆ“AΒ±f⟩.\|A_\pm f\| = \langle A_\pm f, A_\pm f\rangle = \langle f, A_{\mp}A_{\pm} f\rangle.
It just so happens that ff is also an eigenvector of Aβˆ“AΒ±A_{\mp}A_{\pm} (whad’ya know?) so that now, βˆ₯AΒ±fβˆ₯\|A_{\pm} f\| is equal to the eigenvalue of ff with respect to Aβˆ“AΒ±A_{\mp}A_{\pm}. In the case of the angular momentum and spin operators, this gives us equations which the extreme states must satisfy. There is one last trick, which is that any two extreme states part of the same chain must differ by an element of cZc\Z, by the nature of the construction of the ladder operators. If we then factor out the right things, we obtain the fact that the set of eigenvalues of the operator AiA_i is discrete.

There is actually another trick that I glossed over, which was the trick of using some other operator that commutes with AiA_i; we use something like L2L^2 or S2S^2, which gives us a vehicle to do things, because somehow, we can express Aβˆ“AΒ±A_{\mp}A_{\pm} in terms of A2A^2 and AiA_i.



As a fun fact, it might seem like this website is flat because you're viewing it on a flat screen, but the curvature of this website actually isn't zero. ;-)

Copyright Β© 2024, Aathreya Kadambi

Made with Astrojs, React, and Tailwind.