Unknown well-known equation in Sakurai

  • Thread starter eXorikos
  • Start date
  • Tags
    Sakurai
In summary, the well-known equation in Sakurai (p391 in Modern Quantum Mechanics) is a normal identity involving distributions rather than functions. It involves the principal value and Dirac delta functions and can be found in the appendices of the book by Messiah. It is valid for analytic functions of an arbitrary self-adjoint operator and is provable by using a representation where it acts diagonally. The operator only needs to be bounded and self-adjoint to be considered physical in AQFT. However, strictly speaking, momentum is not an observable in AQFT, but this does not mean it is not observable in general.
  • #1
eXorikos
284
5
Unknown "well-known" equation in Sakurai

On p391 in Modern Quantum Mechanics is stated: Now we use the well-known equation:

[tex]\frac{1}{E - H_0 -i \epsilon} = Pr. \left( \frac{1}{E-H_0} \right) + i \pi \delta \left(E-H_0 \right)[/tex]

Frankly I've never heard of the equation or the notation Pr. My first thought is that the Pr. stands for Cachy principal value integral, but that's denoted [tex]P \int[/tex] or P.v. So I'm lost here...
 
Physics news on Phys.org
  • #2


Your guess is correct Pr does stand for the principal value.
 
  • #3


stone said:
Your guess is correct Pr does stand for the principal value.

and the equation must be understood as one involving distributions rather than functions. I.e., it becomes a normal identity only after multiplying it with a nice function and then integrating it. This provides the integral found missing by the OP.
 
Last edited:
  • #4


Here's one way to think about it. Start with [itex]1/(x-i\epsilon)[/itex], and multiply the numerator and denominator by [itex]x+i\epsilon[/itex] to get

[tex]{1\over x-i\epsilon}={x\over x^2+\epsilon^2} +i{\epsilon\over x^2+\epsilon^2}[/tex]

For small [itex]\epsilon[/itex], the imaginary part is a narrow spike at [itex]x=0[/itex], and the area under it is [itex]\pi[/itex] (do the integral over [itex]x[/itex] to get this). So, as [itex]\epsilon\to 0[/itex], the imaginary part is becoming [itex]i\pi\delta(x)[/itex].

It takes a little more work to see that the real part is becoming the principal value, which I leave as an exercise.
 
  • #5


This bothered me for quite a while too, as far too many authors claim this to be a well-known equation. Mattuck was the first to clear it up for me (in A Guide to Feynman Diagrams...) where he even sarcastically calls it "the so-called 'well-known theorem from complex function theory' ". Importantly, the written form

[tex]\frac{1}{x+i\epsilon} = P\frac{1}{x} - i\pi\delta(x)[/tex]

is only shorthand for

[tex]\int dx \frac{f(x)}{x+i\epsilon} = P\int dx \frac{f(x)}{x} - i\pi \int dx f(x) \delta(x)[/tex]

as pointed out by A. Neumaier.
 
  • #6


It should be noted that all these derivations should be questioned when applying the equation for operators instead of complex numbers :-) That doesn't mean that the equations necessarily become wrong; it only says that these simple derivations do not prove that they are right.
 
  • #7


t!m said:
This bothered me for quite a while too, as far too many authors claim this to be a well-known equation. Mattuck was the first to clear it up for me (in A Guide to Feynman Diagrams...) where he even sarcastically calls it "the so-called 'well-known theorem from complex function theory' ". Importantly, the written form

[tex]\frac{1}{x+i\epsilon} = P\frac{1}{x} - i\pi\delta(x)[/tex]

is only shorthand for

[tex]\int dx \frac{f(x)}{x+i\epsilon} = P\int dx \frac{f(x)}{x} - i\pi \int dx f(x) \delta(x)[/tex]

as pointed out by A. Neumaier.

It is well-known, though perhaps not to everyone. The formula can be found in the Appendices of the book by Messiah, but also in the wikipedia article
http://en.wikipedia.org/wiki/Dirac_delta_function under the name ''Sokhatsky–Weierstrass theorem'' (a name I didn't know before looking it up).
 
  • #8


tom.stoer said:
It should be noted that all these derivations should be questioned when applying the equation for operators instead of complex numbers :-) That doesn't mean that the equations necessarily become wrong; it only says that these simple derivations do not prove that they are right.

They remain correct when H is self-adjoint. The proof is by diagonalization and reduction to the scalar case.
 
  • #9


A. Neumaier said:
They remain correct when H is self-adjoint. The proof is by diagonalization and reduction to the scalar case.
I know. But usually the the one-to-one translation of matrix stuff to operators requires the operators to be compact - which is rarely the case in QM
 
  • #10


tom.stoer said:
I know. But usually the the one-to-one translation of matrix stuff to operators requires the operators to be compact - which is rarely the case in QM

No compactness is needed.

Cauchy's integral formula is valid for analytic functions of an arbitrary self-adjoint operator, and provable by using a representation where it acts diagonally (which exists according to the spectral theorem, in one of its formulations). One gets ''our'' formula from this.
 
  • #11


tom.stoer said:
I know. But usually the the one-to-one translation of matrix stuff to operators requires the operators to be compact - which is rarely the case in QM

Tongue-in-cheek: all physical operators are compact... otherwise they're not physical!

I've heard at least one very slick lecturer state: "and this integral obviously converges because it's physical" --- former head of theoretical physics at Bell...
 
  • #12


genneth said:
Tongue-in-cheek: all physical operators are compact... otherwise they're not physical!
Actually an operator need only be bounded and self-adjoint to be physical. Or in truth it need only have bounded projections and be self-adjoint.
 
  • #13


DarMM said:
Actually an operator need only be bounded and self-adjoint to be physical.

Do you want to rob momentum its character of being physical?
(This is why I don't like C^* algebras as foundation...)

DarMM said:
Or in truth it need only have bounded projections and be self-adjoint.

So the former was not in truth! But is the latter?

Self-adjoint alone is enough for being physical.

(Bounded projections is ambiguous. ''some projections are bounded'' is a nearly empty statement; ''all projections are bounded'' is far too strong, excluding momentum.)
 
  • #14


genneth said:
Tongue-in-cheek: all physical operators are compact... otherwise they're not physical!
No counter examples?

genneth said:
"and this integral obviously converges because it's physical" --- former head of theoretical physics at Bell...
I know a similar theorem that states that "numbers like 11, 13, 17 etc. are unphysical and do not exist in nature". There is one case where this theorem fails, namely the beta-function of QCD which reads

β(g) ~ g³ (11NC - 2NF) with NC=3

The factor "11" immediately proves that QCD is unphysical :-)
 
  • #15


A. Neumaier said:
Do you want to rob momentum its character of being physical?
(This is why I don't like C^* algebras as foundation...)



So the former was not in truth! But is the latter?

Self-adjoint alone is enough for being physical.

(Bounded projections is ambiguous. ''some projections are bounded'' is a nearly empty statement; ''all projections are bounded'' is far too strong, excluding momentum.)
Ah yes, good point! To explain more accurately, in Algebraic QFT using C*-algebras, one normally says that observables are self-adjoint bounded operators contained in a region of spacetime. So very strictly speaking AQFT would say that momentum is not an observable. However when I say very strictly, I mean very strictly. AQFT does not say that momentum is not observable, just that no complete measurement of it (enough to specify the point in the spectrum completely) is possible in a finite region. All you can observe in a finite region is operators whose eigenstates are ones whose support in momentum space is finite. The width of these regions is the resolution of the equipment and their number is determined by the highest and lowest momentum states the device can measure.

So momentum is not an observable in a very technical, strict manner. However more truthfully this is just a mathematical way of encoding that you cannot measure momentum "to a point", not that you cannot actually measure momentum.
 
  • #16


A. Neumaier said:
It is well-known, though perhaps not to everyone. The formula can be found in the Appendices of the book by Messiah, but also in the wikipedia article
http://en.wikipedia.org/wiki/Dirac_delta_function under the name ''Sokhatsky–Weierstrass theorem'' (a name I didn't know before looking it up).

Ah, thanks! Now I can at least refer to it by name instead of carrying on the tradition of calling it 'a well-known theorem.'

tom.stoer said:
I know a similar theorem that states that "numbers like 11, 13, 17 etc. are unphysical and do not exist in nature". There is one case where this theorem fails, namely the beta-function of QCD which reads

β(g) ~ g³ (11NC - 2NF) with NC=3

The factor "11" immediately proves that QCD is unphysical :-)

And what of the fractional quantum Hall effect? Those filling fractions are fantastically unphysical. I've never even heard of 7/15 ... :wink:
 
  • #17


DarMM said:
Ah yes, good point! To explain more accurately, in Algebraic QFT using C*-algebras, one normally says that observables are self-adjoint bounded operators contained in a region of spacetime. So very strictly speaking AQFT would say that momentum is not an observable. However when I say very strictly, I mean very strictly. AQFT does not say that momentum is not observable, just that no complete measurement of it (enough to specify the point in the spectrum completely) is possible in a finite region. All you can observe in a finite region is operators whose eigenstates are ones whose support in momentum space is finite. The width of these regions is the resolution of the equipment and their number is determined by the highest and lowest momentum states the device can measure.

So momentum is not an observable in a very technical, strict manner. However more truthfully this is just a mathematical way of encoding that you cannot measure momentum "to a point", not that you cannot actually measure momentum.

This happens to be true for momentum but has nothing to do with the problem of boundedness. One component of the electromagnetic field strength at a point x is local (and can in principle be measured arbitrarily well) but is not a bounded variable.

Whereas the projection of a momentum component to a bounded interval is bounded but cannot be measured exactly to the point (only arbitrarily well). But we'd discuss this in a new thread...
 
Last edited:
  • #18


t!m said:
And what of the fractional quantum Hall effect? Those filling fractions are fantastically unphysical. I've never even heard of 7/15 ... :wink:
Yeah, must be wrong, too :-)
 
  • #19


A. Neumaier said:
This happens to be true for momentum but has nothing to do with the problem of boundedness. The electromagnetic field strength at a point x is local (and can in principle be measured arbitrarily well) but is not a bounded variable.

Whereas the projection of a momentum component to a bounded interval is bounded but cannot be measured exactly to the point (only arbitrarily well).
Perhaps I'm missing something, please correct me if I am, but the electromagnetic field strength at a point x is not an observable, since it is not an operator, it is only an operator valued distribution.
 
  • #20


DarMM said:
Perhaps I'm missing something, please correct me if I am, but the electromagnetic field strength at a point x is not an observable, since it is not an operator, it is only an operator valued distribution.

I'll answer a new thread ''Boundedness of quantum observables'', since it is no longer related to the original question.
 

1. What is the Sakurai equation?

The Sakurai equation, also known as the Schrödinger–Sakurai equation, is a mathematical equation used in quantum mechanics to describe the time evolution of a quantum state. It is named after physicists Erwin Schrödinger and Jun Ishiwara Sakurai.

2. What is the significance of the Sakurai equation?

The Sakurai equation is one of the fundamental equations in quantum mechanics and is used to calculate the probability of finding a particle in a certain state at a given time. It is essential for understanding the behavior of particles at the atomic and subatomic levels.

3. How does the Sakurai equation differ from the Schrödinger equation?

The Sakurai equation is a modified version of the Schrödinger equation that takes into account the spin of particles. It includes a spin operator, which allows for the calculation of properties such as spin angular momentum and spin magnetic moment.

4. What is the role of the Sakurai equation in quantum field theory?

The Sakurai equation is a key component in quantum field theory, a theoretical framework that combines quantum mechanics with special relativity. It is used to describe the dynamics of quantum fields, which are fundamental to understanding the behavior of particles at high energies.

5. Are there any known limitations of the Sakurai equation?

The Sakurai equation is limited in its applicability to non-relativistic systems, meaning it cannot accurately describe the behavior of particles moving at speeds close to the speed of light. It also does not take into account certain effects, such as quantum tunneling, that are important in some systems.

Similar threads

  • Quantum Physics
Replies
3
Views
3K
  • Quantum Physics
Replies
4
Views
5K
Replies
3
Views
3K
Replies
2
Views
571
Replies
4
Views
2K
  • Quantum Physics
Replies
7
Views
2K
Replies
11
Views
1K
Replies
1
Views
724
Replies
2
Views
2K
Replies
9
Views
3K
Back
Top