Vacuum projection operator and normal ordering

In summary: W:##?In summary, the conversation discusses the derivation of an operator expression involving the vacuum projection operator and number operator in terms of creation and annihilation operators. The final expression is given in a normal-ordered form, and the question arises as to how one can freely commute an exponential operator past the vacuum projection operator. The answer is that since the vacuum projection operator is expressible in terms of the creation and annihilation operators, it can be freely moved within the normal ordering. Additionally, the book defines a quantity W as the vacuum projection operator, but writes it in a different form, which may cause confusion.
  • #1
"Don't panic!"
601
8
I've been reading this book, in which the author expresses the vacuum projection operator ##\vert 0\rangle\langle 0\vert## in terms of the number operator ##\hat{N}=\hat{a}^{\dagger}\hat{a}##, where ##\hat{a}^{\dagger}## and ##\hat{a}## are the usual creation and annihilation operators, respectively. I can follow most of the derivation, however, I don't quite understand the following step: $$:\text{exp}\bigg\lbrace\hat{a}^{\dagger}\frac{\mathrm{d}}{\mathrm{d}Z^{\ast}}\bigg\rbrace\,\vert 0\rangle\langle 0\vert\,\text{exp}\big\lbrace\hat{a}Z^{\ast}\big\rbrace :\bigg\vert_{Z^{\ast}=0} \ = \ :\text{exp}\big\lbrace\hat{a}^{\dagger}\hat{a}\big\rbrace\,\vert 0\rangle\langle 0\vert : \qquad (1)$$ How does one get from the left-hand side to the right-hand side of this equation? I'm assuming there are several steps that have been missed out, or am I missing something trivial?

Just in case the link is not viewable, let me elaborate on the details of the calculation a little further, in particular, how one arrives at eq. (1). Using the completeness relation for the basis of number-operator eigenstates, we have $$1\!\!1 \ = \ \sum_{n,m=0}^{\infty}\vert n\rangle\langle m\vert\,\delta_{n,m} \ = \ \sum_{n,m=0}^{\infty}\vert n\rangle\langle m\vert\,\frac{1}{\sqrt{n!m!}}\bigg(\frac{\mathrm{d}}{\mathrm{d}Z^{\ast}}\bigg)^{n}\big(Z^{\ast}\big)^{m}\bigg\vert_{Z^{\ast}=0}\qquad (2)$$ where we make use of the identity $$\frac{1}{\sqrt{n!m!}}\bigg(\frac{\mathrm{d}}{\mathrm{d}Z^{\ast}}\bigg)^{n}\big(Z^{\ast}\big)^{m}\bigg\vert_{Z^{\ast}=0} \ = \ \delta_{n,m}$$ Next, using that ##\vert n\rangle =\frac{(\hat{a}^{\dagger})^{n}}{\sqrt{n!}}\,\vert 0\rangle## we can rewrite (2) as $$\sum_{n,m=0}^{\infty}\frac{(\hat{a}^{\dagger})^{n}}{n!}\,\vert 0\rangle\langle 0\vert\,\frac{(\hat{a})^{m}}{m!}\bigg(\frac{\mathrm{d}}{\mathrm{d}Z^{\ast}}\bigg)^{n}\big(Z^{\ast}\big)^{m}\bigg\vert_{Z^{\ast}=0} \ = \ \sum_{n,m=0}^{\infty}\frac{(\hat{a}^{\dagger})^{n}\big(\frac{\mathrm{d}}{\mathrm{d}Z^{\ast}}\big)^{n}}{n!}\,\vert 0\rangle\langle 0\vert\,\frac{(\hat{a})^{m}\big(Z^{\ast}\big)^{m}}{m!}\bigg\vert_{Z^{\ast}=0} \\[1cm] \qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\,\, = \ \text{exp}\bigg\lbrace\hat{a}^{\dagger}\frac{\mathrm{d}}{\mathrm{d}Z^{\ast}}\bigg\rbrace\,\vert 0\rangle\langle 0\vert\,\text{exp}\big\lbrace\hat{a}Z^{\ast}\big\rbrace\bigg\vert_{Z^{\ast}=0}$$ This final expression is already normal-ordered as all creation operators are placed to the left of all annihilation operators. As such we can express this last line as given in eq. (1).

I understand this derivation up to the left-hand side of eq. (1), I just don't understand how the author then gets to the right-hand side of eq. (1).
 
Physics news on Phys.org
  • #2
So we would like to study the object
$$
\exp\left( \hat{a}^{\dagger} \frac{d}{dZ^{\ast}}\right) \exp\left( Z^{\ast} \hat{a} \right) \Bigg|_{Z^{\ast} = 0} = \sum_{k,k' = 0}^{\infty} \frac{1}{k! k'!} \left( \hat{a}^{\dagger} \frac{d}{dZ^{\ast}}\right)^k \left( Z^{\ast} \hat{a} \right)^{k'} \Bigg|_{Z^{\ast} = 0}
$$
Now the trick is to think about what terms in this sum could possibly survive setting ##Z^{\ast} = 0##. We have a set of terms ##(Z^{\ast})^{k'}##, and we have derivatives acting on them. If a derivative acts less than ##k'## times on a ##(Z^{\ast})^{k'}## term, it will vanish when you take##Z^{\ast} = 0##. But we also clearly have
$$
\frac{d^k }{dZ^{\ast k}}(Z^{\ast})^{k'} = 0, \quad k>k'.
$$
Therefore, the only terms which can survive the double sum are precisely the ##k=k'## ones. In this case we use
$$
\frac{d^k }{dZ^{\ast k}}(Z^{\ast})^{k} = k!
$$
and find
$$
\exp\left( \hat{a}^{\dagger} \frac{d}{dZ^{\ast}}\right) \exp\left( Z^{\ast} \hat{a} \right) \Bigg|_{Z^{\ast} = 0} = \sum_{k=0}^{\infty} \frac{1}{k!} \left( \hat{a}^{\dagger} \right)^k \left( \hat{a} \right)^{k}
$$
 
  • #3
king vitamin said:
So we would like to study the object
$$
\exp\left( \hat{a}^{\dagger} \frac{d}{dZ^{\ast}}\right) \exp\left( Z^{\ast} \hat{a} \right) \Bigg|_{Z^{\ast} = 0} = \sum_{k,k' = 0}^{\infty} \frac{1}{k! k'!} \left( \hat{a}^{\dagger} \frac{d}{dZ^{\ast}}\right)^k \left( Z^{\ast} \hat{a} \right)^{k'} \Bigg|_{Z^{\ast} = 0}
$$
Now the trick is to think about what terms in this sum could possibly survive setting ##Z^{\ast} = 0##. We have a set of terms ##(Z^{\ast})^{k'}##, and we have derivatives acting on them. If a derivative acts less than ##k'## times on a ##(Z^{\ast})^{k'}## term, it will vanish when you take##Z^{\ast} = 0##. But we also clearly have
$$
\frac{d^k }{dZ^{\ast k}}(Z^{\ast})^{k'} = 0, \quad k>k'.
$$
Therefore, the only terms which can survive the double sum are precisely the ##k=k'## ones. In this case we use
$$
\frac{d^k }{dZ^{\ast k}}(Z^{\ast})^{k} = k!
$$
and find
$$
\exp\left( \hat{a}^{\dagger} \frac{d}{dZ^{\ast}}\right) \exp\left( Z^{\ast} \hat{a} \right) \Bigg|_{Z^{\ast} = 0} = \sum_{k=0}^{\infty} \frac{1}{k!} \left( \hat{a}^{\dagger} \right)^k \left( \hat{a} \right)^{k}
$$
Thanks for your response. I understand this part though. What I don't understand is how one goes from: $$ : \text{exp}\bigg\lbrace\hat{a}^{\dagger}\frac{\mathrm{d}}{\mathrm{d}Z^{\ast}}\bigg\rbrace\,\vert 0\rangle\langle 0\vert\,\text{exp}\big\lbrace\hat{a}Z^{\ast}\big\rbrace : \bigg\vert_{Z^{\ast}=0}$$ to $$ : \text{exp}\big\lbrace\hat{a}^{\dagger}\hat{a}\big\rbrace\,\vert 0\rangle\langle 0\vert :$$ That is, why is one allowed to freely commute ##\text{exp}\big\lbrace\hat{a}Z^{\ast}\big\rbrace## past ##\vert 0\rangle\langle 0\vert##? Is it because the whole thing is normal ordered?
 
  • #4
"Don't panic!" said:
Thanks for your response. I understand this part though. What I don't understand is how one goes from: $$ : \text{exp}\bigg\lbrace\hat{a}^{\dagger}\frac{\mathrm{d}}{\mathrm{d}Z^{\ast}}\bigg\rbrace\,\vert 0\rangle\langle 0\vert\,\text{exp}\big\lbrace\hat{a}Z^{\ast}\big\rbrace : \bigg\vert_{Z^{\ast}=0}$$ to $$ : \text{exp}\big\lbrace\hat{a}^{\dagger}\hat{a}\big\rbrace\,\vert 0\rangle\langle 0\vert :$$ That is, why is one allowed to freely commute ##\text{exp}\big\lbrace\hat{a}Z^{\ast}\big\rbrace## past ##\vert 0\rangle\langle 0\vert##? Is it because the whole thing is normal ordered?

Yes, if you assume that ##|0\rangle \langle 0|## is expressible in terms of ##a## and ##a^\dagger##, then you can freely commute it past any ##a## or ##a^\dagger## within a normal ordering.

The line that I don't understand is how they go from:

##1 = : e^{\hat{a}^\dagger \hat{a}}: W ::##

to

##|0\rangle\langle 0|\ =\ :W:\ =\ : e^{- \hat{a}^\dagger \hat{a}}: ##

They know that ##: e^{\hat{a}^\dagger \hat{a}}: W :: = : e^{\hat{a}^\dagger \hat{a}}: e^{-\hat{a}^\dagger \hat{a}}::##. But I don't see how can cancel off the ##: e^{\hat{a}^\dagger \hat{a}}:##.

It's not true in general that ##:A:B::\ =\ :A:C::## implies that ##:B:\ =\ :C:##, is it?

And the other thing is that ##W## is defined to be##|0\rangle \langle 0|##, while the book is writing ##:W: = |0\rangle \langle 0|##.
 
  • #5
stevendaryl said:
Yes, if you assume that |0⟩⟨0||0⟩⟨0||0\rangle \langle 0| is expressible in terms of aaa and a†a†a^\dagger, then you can freely commute it past any aaa or a†a†a^\dagger within a normal ordering.

But I thought the point of this exercise to show precisely that ##\vert 0\rangle\langle 0\vert## is expressible in terms of ##a## and ##a^{\dagger}##? It is not a priori assumed, at least as far as I can tell.

stevendaryl said:
It's not true in general that :A:B:: = :A:C:::A:B:: = :A:C:::A:B::\ =\ :A:C:: implies that :B: = :C::B: = :C::B:\ =\ :C:, is it?

I think it possibly is. If one operates from the left by ##:A^{-1}:##, then we get ##:A^{-1}::A::B: =:A^{-1}A::B:=1\!\!1 :B:##. Similarly, ##:A^{-1}::A::C: =:A^{-1}A::C:=1\!\!1 :C:##, and so ##:B: = :C:##. That's legitimate, right? At least it is if ##:A::B:=:AB:## holds, which I'm not sure is true.

I find the last few steps of the authors derivation very confusing. It seems to be of some importance that the author introduces the ##Z^{\ast}## identity too (otherwise, one can essentially arrive at the same point without the need for introducing the ##Z^{\ast}## pieces.
 
Last edited:
  • #6
"Don't panic!" said:
But I thought the point of this exercise to show precisely that ##\vert 0\rangle\langle 0\vert## is expressible in terms of ##a## and ##a^{\dagger}##? It is not a priori assumed, at least as far as I can tell.

Well, there is a more direct argument to that effect.

##|0\rangle \langle 0|## is the operator ##\hat{A}## defined by its action on the basis states:

##\hat{A} |0\rangle = |0\rangle##
##\hat{A} |n\rangle = 0## if ##n \neq 0##

So we try to "implement" these rules in terms of creation and annihilation operators:

##\hat{A} = \sum_j c_j (\hat{a}^\dagger)^j (\hat{a})^j##

It's obvious that ##c_0 = 1##. Then we can get a recursive definition of ##c_j## for ##j > 0##. We want for ##n > 0##,

##\sum_{j=0}^{n} c_j (\hat{a}^\dagger)^j (\hat{a})^j |n\rangle = 0##

Using the actions of the creation and annihilation operators, this becomes:

##\sum_{j=0}^{n} c_j\ \frac{n!}{(n-j)!}= 0##

which we can write as:

##\sum_{j=0}^{n-1} c_j \frac{n!}{(n-j)!} + c_{n} n!##

So we have:

##c_{n} = - \sum_{j=0}^{n-1} \frac{c_j}{(n-j)!}##

Obviously, that let's us solve for each ##c_n## in terms of ##c_0, c_1, ..., c_{n-1}##.

So clearly, there is a way to write ##|0\rangle\langle 0|## in terms of creation and annihilation operators.
 
  • #7
stevendaryl said:
Well, there is a more direct argument to that effect.

##|0\rangle \langle 0|## is the operator ##\hat{A}## defined by its action on the basis states:

##\hat{A} |0\rangle = |0\rangle##
##\hat{A} |n\rangle = 0## if ##n \neq 0##

So we try to "implement" these rules in terms of creation and annihilation operators:

##\hat{A} = \sum_j c_j (\hat{a}^\dagger)^j (\hat{a})^j##

It's obvious that ##c_0 = 1##. Then we can get a recursive definition of ##c_j## for ##j > 0##. We want for ##n > 0##,

##\sum_{j=0}^{n} c_j (\hat{a}^\dagger)^j (\hat{a})^j |n\rangle = 0##

Using the actions of the creation and annihilation operators, this becomes:

##\sum_{j=0}^{n} c_j\ \frac{n!}{(n-j)!}= 0##

which we can write as:

##\sum_{j=0}^{n-1} c_j \frac{n!}{(n-j)!} + c_{n} n!##

So we have:

##c_{n} = - \sum_{j=0}^{n-1} \frac{c_j}{(n-j)!}##

Obviously, that let's us solve for each ##c_n## in terms of ##c_0, c_1, ..., c_{n-1}##.

So clearly, there is a way to write ##|0\rangle\langle 0|## in terms of creation and annihilation operators.

Okay, so do you think the author is assuming this then? What they are not assuming is that it can be written in the simple form of an exponential of the number operator.

What is the point in introducing the ##Z^{\ast}## identity then? One could equally well just use a kronecker delta to write it in terms of two exponentials, and then the last few steps will be the same. Maybe I'm missing something?

Also, is the choice of ##Z^{\ast}## significant? Is it implying something about the complex conjugate of some ##Z##? The identity holds regardless of my choice of variable.
 
  • #8
stevendaryl said:
It's not true in general that ##:A:B::\ =\ :A:C::## implies that ##:B:\ =\ :C:##, is it?
It depends how ##A## is quantified.
With ''##A## exists such that your formula holds'', your question has a positive answer, since you can take ##A=0##.
With ''for all ##A##, your formula holds'', your question has a negative answer, since you can take ##A=1##.
 
  • Like
Likes bhobba
  • #9
I don't understand the derivation, but the results I can convince myself of in a different way.

Let ##n^{(j)}## be shorthand for ##n (n-1) (n-2) ... (n+1-j)##. Then look at the expression:

##S(n) = \sum_{j=0}^n (-1)^j n^{(j)}/j!##

For ##n=0##, only the first term survives, so ##S_0 = 1##. For ##n > 0##, this is the binomial expansion of ##(1-1)^n##, so it equals 0.

Now, let's look at a related expression: ##\hat{S} = \sum_{j=0}^\infty (-1)^j N^{(j)}/j!## where ##N## is the number operator. If we let it act on a basis state ##|n\rangle##, we get the following:

##\hat{S} |n\rangle = \sum_{j=0}^n (-1)^j n^{(j)} |n\rangle## (Because for ##j > n##, ##N^{(j)} |n\rangle = 0##)

So ##\hat{S} |n\rangle = S(n) |n\rangle##. Therefore, ##\hat{S}## is our operator that has eigenvalue 1 when acting on ##|0\rangle## and has eigenvalue ##0## when acting on any other basis state. So therefore it's equal to ##|0\rangle \langle 0|##.

So ##\hat{S}## is one expression for ##|0\rangle \langle 0|##.

Now, here's an amazing fact about normal ordering that you can prove by induction:

##: N^j :\ =\ N^{(j)}##

Just to show it in one case, with ##j=2##:

##: N^2 : ##
##\ = \ (\hat{a}^\dagger)^2 (\hat{a})^2 ##
##\ =\ \hat{a}^\dagger (\hat{a}^\dagger \hat{a}) \hat{a} ##
##\ =\ \hat{a}^\dagger (-1 + \hat{a} \hat{a}^\dagger) \hat{a} ##
##\ =\ - \hat{a}^\dagger \hat{a} + (\hat{a}^\dagger \hat{a})^2 ##
##\ =\ -N + N^2 ##
##\ =\ N (N-1) ##
##\ =\ N^{(2)}##

So
##\hat{S} = \sum_{j=0}^\infty (-1)^j N^{(j)}/j!##
##\ =\ : \sum_{j=0}^\infty (-1)^j N^{j}/j! :##
##\ = \ : e^{-N} : ##
 
  • Like
Likes A. Neumaier
  • #10
Thinking about it some more, I just don't think that the derivation is correct. What the author proves is:

##: e^{\hat{N}} |0\rangle \langle 0| : = 1##

Without further assumptions, I don't see how that implies that ##|0\rangle \langle 0| = : e^{- \hat{N}} :##
 
  • #11
stevendaryl said:
Thinking about it some more, I just don't think that the derivation is correct. What the author proves is:

##: e^{\hat{N}} |0\rangle \langle 0| : = 1##

Without further assumptions, I don't see how that implies that ##|0\rangle \langle 0| = : e^{- \hat{N}} :##

Ah, that's frustrating. I've seen the same argument as the one this author gives in an academic paper too. It would be helpful if they actually explained things in further detail. I can't find anywhere else that gives a detailed derivation.
 
  • #12
The simplest thing to do is to start with ##A = :\!\exp(-N){:}\ ## and then prove that ##A=|0\rangle\langle 0|##. What follows is essentially the same as stevendaryl's proof, but IMO is a little simpler.

First, Taylor expand:

##A=\sum_{j=0}^\infty {(-1)^j\over j!} {:}N^j{:}##

Then use

##{:}N^j{:} = (a^\dagger)^j a^j##

Then use

##a^j|n\rangle = \sqrt{n!/(n{-}j)!}\,|n{-}j\rangle##

##(a^\dagger)^j|n{-}j\rangle = \sqrt{n!/(n-j)!}\,|n\rangle##

Note that these hold even for ##j>n## if we take ##1/(n{-}j)!=0## for ##j>n##. We now have

##(a^\dagger)^j a^j|n\rangle = {n!\over(n{-}j)!}\,|n\rangle##.

Therefore

##A|n\rangle = \left[\sum_{j=0}^\infty {(-1)^j n!\over j!(n{-}j)!}\right]|n\rangle##

The factor in brackets is the binomial expansion of ##(1-1)^n##, and equals ##\delta_{n0}##, QED.
 
Last edited:
  • #13
Avodyne said:
The simplest thing to do is to start with ##A = :\!\exp(-N){:}\ ## and then prove that ##A=|0\rangle\langle 0|##. What follows is essentially the same as stevendaryl's proof, but IMO is a little simpler.

First, Taylor expand:

##A=\sum_{j=0}^\infty {(-1)^j\over j!} {:}N^j{:}##

Then use

##{:}N^j{:} = (a^\dagger)^j a^j##

Then use

##a^j|n\rangle = \sqrt{n!/(n{-}j)!}\,|n{-}j\rangle##

##(a^\dagger)^j|n{-}j\rangle = \sqrt{n!/(n-j)!}\,|n\rangle##

Note that these hold even for ##j>n## if we take ##1/(n{-}j)!=0## for ##n>j##. We now have

##(a^\dagger)^j a^j|n\rangle = {n!\over(n{-}j)!}\,|n\rangle##.

Therefore

##A|n\rangle = \left[\sum_{j=0}^\infty {(-1)^j n!\over j!(n{-}j)!}\right]|n\rangle##

The factor in brackets is the binomial expansion of ##(1-1)^n##, and equals ##\delta_{n0}##, QED.
Thanks, this derivation (and stevendaryl’s) makes a lot of sense. It is reasonable to take ##0^0=1## (there doesn’t seem to be a general consensus)?

Also, it would interesting to see why the author is able to write: $$: \text{exp}\Big\lbrace\hat{a}^{\dagger}\frac{d}{dZ^{\ast}}\Big\rbrace\vert 0\rangle\langle 0\vert\text{exp}\Big\lbrace\hat{a}Z^{\ast}\Big\rbrace\bigg\vert_{Z^{\ast}=0}: \ = \ : \text{exp}\big\lbrace\hat{a}^{\dagger}\hat{a}\big\rbrace\vert 0\rangle\langle 0\vert :$$ That is, why is it ok to commute the ##\hat{a}## past ##\vert 0\rangle\langle 0\vert##?
 
  • #14
"Don't panic!" said:
It is reasonable to take 00=100=10^0=1 (there doesn’t seem to be a general consensus)?
In a power series expansion, ##x^0## is always 1, by definition.
 
  • #15
A. Neumaier said:
In a power series expansion, ##x^0## is always 1, by definition.

Good point, fair enough.
 
  • #16
Fixed a typo in my post, "for ##n>j##" should be (and now is) "for ##j>n##".

Re the author's derivation: inside the normal-ordering colons, the factors can be written in any order, since, whatever order they are written in, they will be re-ordered by the normal-ordering. That said, I still can't follow every step. Perhaps some of the manipulations were explained earlier in the book.
 
Last edited:
  • #17
Avodyne said:
Fixed a typo in my post, "for ##n>j##" should be (and now is) "for ##j>n##".

Re the author's derivation: inside the normal-ordering colons, the factors can be written in any order, since, whatever order they are written in, they will be re-ordered by the normal-ordering. That said, I still can't follow every step. Perhaps some of the manipulations were explained earlier in the book.

I'm convinced that the argument in the reference proves that

1. ##:e^{\hat{a}^\dagger \hat{a}} |0\rangle \langle 0|: \ = \ 1##

What I don't understand is how that implies that

2. ##|0\rangle \langle 0| \ = \ :e^{-\hat{a}^\dagger \hat{a}}:##

I believe that 2. is true but I don't see how it is implied by 1. (I can see that 1. follows from 2., but not vice-versa)
 
  • #18
Avodyne said:
Perhaps some of the manipulations were explained earlier in the book.

Possibly, but unfortunately I don't have access to the rest of the book :(
stevendaryl said:
I'm convinced that the argument in the reference proves that

1. ##:e^{\hat{a}^\dagger \hat{a}} |0\rangle \langle 0|: \ = \ 1##

What I don't understand is how that implies that

2. ##|0\rangle \langle 0| \ = \ :e^{-\hat{a}^\dagger \hat{a}}:##

I believe that 2. is true but I don't see how it is implied by 1. (I can see that 1. follows from 2., but not vice-versa)

This is what leaves me confused too. It is discussed in this paper (right-hand side of page 3) and this paper (right-hand side of page 6) on the arXiv also (admittedly using a different approach), but without further explanation.
 
  • #19
The paper https://arxiv.org/pdf/0704.3116.pdf sort of explains it.

One representation of ##|0\rangle \langle 0|## is ##lim_{\lambda \rightarrow \infty} e^{-\lambda \hat{N}}##. Of course, that limit doesn't actually exist, but the meaning is this: for any state ##|\psi\rangle##,

##|0\rangle \langle 0|\psi\rangle = lim_{\lambda \rightarrow \infty} e^{-\lambda \hat{N}} |\psi\rangle##

The operator ##e^{-\lambda \hat{N}}## acting on ##|0\rangle## returns ##|0\rangle##, and when acting on any other basis state ##|n\rangle## returns something close to zero. So that's the same behavior as the operator ##|0\rangle \langle 0|##.

Now, we show an unexpected fact about the number operator and normal ordering:

##e^{-\lambda \hat{N}} = \ : e^{- (1 - e^{-\lambda}) \hat{N}}:## (equation 30, slightly rewritten, from that paper).

So taking the limit as ##\lambda \rightarrow \infty## gives:

##|0\rangle \langle 0| = \ :e^{- \hat{N}}:##

The way that the paper proves equation 30 makes a detour through "Bell numbers", but it seems to me that it can be proved more directly if you use the fact that I derived earlier:

##: N^n : \ = \ N(N-1)...(N+1-n)##

We can use the taylor expansion of ##(1+\alpha)^y##:

##(1+\alpha)^y = \sum_{j=0}^{\infty} \frac{\alpha^j}{j!} y (y-1) ... (y+1-j)##

So letting ##\alpha = e^{-\lambda} - 1##, we have:

##e^{-\lambda y} = \sum_{j=0}^{\infty} \frac{e^{(-\lambda} - 1)^j}{j!} y (y-1) ... (y+1-j)##

Pluggine ##N## in for ##y##:

##e^{-\lambda N} = \sum_{j=0}^{\infty} \frac{(e^{-\lambda} - 1)^j}{j!} N (N-1) ... (N+1-j)##
##= \sum_{j=0}^{\infty} \frac{(e^{-\lambda} - 1)^j}{j!}\ :N^j:##
##= : (\sum_{j=0}^{\infty} \frac{(e^{-\lambda} - 1)^j}{j!}\ N^j):##
##= :e^{(e^{-\lambda} - 1) N}:##
 
Last edited:
  • Like
Likes Avodyne
  • #20
stevendaryl said:
The paper https://arxiv.org/pdf/0704.3116.pdf sort of explains it.

One representation of ##|0\rangle \langle 0|## is ##lim_{\lambda \rightarrow \infty} e^{-\lambda \hat{N}}##. Of course, that limit doesn't actually exist, but the meaning is this: for any state ##|\psi\rangle##,

##|0\rangle \langle 0|\psi\rangle = lim_{\lambda \rightarrow \infty} e^{-\lambda \hat{N}} |\psi\rangle##

The operator ##e^{-\lambda \hat{N}}## acting on ##|0\rangle## returns ##|0\rangle##, and when acting on any other basis state ##|n\rangle## returns something close to zero. So that's the same behavior as the operator ##|0\rangle \langle 0|##.

Now, we show an unexpected fact about the number operator and normal ordering:

##e^{-\lambda \hat{N}} = \ : e^{- (1 - e^{-\lambda}) \hat{N}}:## (equation 30, slightly rewritten, from that paper).

So taking the limit as ##\lambda \rightarrow \infty## gives:

##|0\rangle \langle 0| = \ :e^{- \hat{N}}:##

The way that the paper proves equation 30 makes a detour through "Bell numbers", but it seems to me that it can be proved more directly if you use the fact that I derived earlier:

##: N^n : \ = \ N(N-1)...(N+1-n)##

We can use the taylor expansion of ##(1+\alpha)^y##:

##(1+\alpha)^y = \sum_{j=0}^{\infty} \frac{\alpha^j}{j!} y (y-1) ... (y+1-j)##

So letting ##\alpha = e^{-\lambda} - 1##, we have:

##e^{-\lambda y} = \sum_{j=0}^{\infty} \frac{e^{(-\lambda} - 1)^j}{j!} y (y-1) ... (y+1-j)##

Pluggine ##N## in for ##y##:

##e^{-\lambda N} = \sum_{j=0}^{\infty} \frac{(e^{-\lambda} - 1)^j}{j!} N (N-1) ... (N+1-j)##
##= \sum_{j=0}^{\infty} \frac{(e^{-\lambda} - 1)^j}{j!}\ :N^j:##
##= : (\sum_{j=0}^{\infty} \frac{(e^{-\lambda} - 1)^j}{j!}\ N^j):##
##= :e^{(e^{-\lambda} - 1) N}:##

Thanks for this.

Going back to the normal-ordered stuff, I managed to get hold of a copy of the book I was mentioning in my OP ("Nonequilibrium Quantum Transport Physics in Nanosystems: Foundation of Computational Nonequilibrium Physics in Nanoscience and Nanotechnology" by F. A. Buot). In it, the author states that normal-ordered products satisfy the following property: If ##A## is some operator, and ##:U\cdots V: \; =\; :W\cdots X:##, then $$:AU\cdots V: \; =\; :AW\cdots X:$$. Given this property, it is then clear that, for ##:AB: \; = 1\!\!1 =\; :1\!\!1:##, it follows that $$:A^{-1}AB: \; =\; :A^{-1}1\!\!1:\qquad \Rightarrow\qquad :B: \; =\; :A^{-1}: \;.$$
 

1. What is a vacuum projection operator?

A vacuum projection operator is a mathematical tool used in quantum field theory to project out the vacuum state from a larger Hilbert space. It is represented by a symbol, typically P0, and when applied to a state vector, it returns the component of that state that is in the vacuum state.

2. How is a vacuum projection operator related to normal ordering?

Normal ordering is a technique used in quantum field theory to rearrange operators in a way that isolates the vacuum state. The vacuum projection operator is often used in this process, as it can be used to project out the vacuum state from a larger operator expression.

3. What is the significance of normal ordering in quantum field theory?

Normal ordering is important in quantum field theory because it allows us to isolate the vacuum state and calculate the expectation value of operators in that state. This is necessary for understanding the behavior of quantum fields and making predictions about physical phenomena.

4. How is the vacuum projection operator used in particle physics?

In particle physics, the vacuum projection operator is used to calculate the vacuum expectation value of operators, which is an important aspect of understanding the behavior of quantum fields and predicting the behavior of particles. It is also used in the process of renormalization, which is a technique for dealing with divergences in quantum field theory calculations.

5. Can the vacuum projection operator be applied to any state vector?

No, the vacuum projection operator can only be applied to states that are in the vacuum state. Applying it to a state that is not in the vacuum state will result in a zero vector, as the projection operator will return the component of the state that is in the vacuum state, which in this case is nothing.

Similar threads

  • Quantum Physics
Replies
9
Views
1K
  • Quantum Physics
Replies
6
Views
805
Replies
2
Views
500
Replies
21
Views
2K
Replies
10
Views
1K
  • Quantum Physics
Replies
26
Views
1K
Replies
1
Views
579
  • Quantum Physics
Replies
4
Views
2K
  • Quantum Physics
Replies
8
Views
1K
  • Quantum Physics
Replies
4
Views
766
Back
Top