Skip to main content

Complex Vector Spaces and Quantum Mechanics

  • Chapter
  • First Online:
Book cover Quantum Mechanics for Pedestrians 1: Fundamentals

Part of the book series: Undergraduate Lecture Notes in Physics ((ULNP))

  • 3700 Accesses

Abstract

In our complex vector space, we can define a scalar product. The properties of orthogonality and completeness lead to the important concept of a complete orthonormal system. The measurment process can be formulated by means of suitable projection operators.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Institutional subscriptions

Notes

  1. 1.

    Of course, we treat these technical aspects not as an end in themselves, but because they are of fundamental importance for the physical description of natural phenomena in the context of quantum mechanics.

  2. 2.

    For the notation \(\cong \), see Chap. 2.

  3. 3.

    Later on, we will meet vector spaces where this is no longer the case; keyword ‘identical particles’ or ‘superselection rules’.

  4. 4.

    We note that the superposition principle contains three pieces of information: (1) The multiplication of a state by a scalar is meaningful. (2) The addition of two states is meaningful. (3) Every linear combination of two states is again an element in the vector space.

  5. 5.

    As we know, only the zero vector has length zero.

  6. 6.

    We recall that \(^{*}\) means complex conjugation.

  7. 7.

    In the bra-ket notation, one cannot identify the dimension of the corresponding vector space (the same holds true for the familiar vector notations v or \(\overrightarrow{v}\), by the way). If necessary, this information must be given separately.

  8. 8.

    In the two-dimensional vector space that we are currently addressing, such a system consists of course of two vectors; as stated above, the zero vector is excluded a priori from consideration.

  9. 9.

    We repeat the remark that for products of numbers and vectors, it holds that \(c\cdot \left| z\right\rangle =\left| z\right\rangle \cdot c\). Because \(\left\langle h\right| \left. z\right\rangle \) is a number, we can therefore write \(\left\langle h\right| \left. z\right\rangle \left| h\right\rangle \) as \(\left| h\right\rangle \left\langle h\right| \left. z\right\rangle \).

  10. 10.

    In equations such as (4.17), the \(1\) on the right side is not necessarily the number \(1\), but is generally something that works like a multiplication by \(1\), i.e. a unit operator. For instance, this is the unit matrix when working with vectors. The notation \(1\) for the unit operator (which implies writing simply \(1\) instead of \(\left( \begin{array}{cc} 1 &{} 0 \\ 0 &{} 1 \end{array} \right) \), for instance) is of course quite lax. On the other hand, as said before, the effect of multiplication by the unit operator and by \(1\) is identical, so that the small inaccuracy is generally accepted in view of the economy of notation. If necessary, ‘one knows’ that \(1\) means the unit operator. But there are also special notations for it, such as \(\mathbb {E}\), \(I_{n}\) (where \(n\) indicates the dimension) and others. An analogous remark applies to the zero operator. By the way, we recall that in the case of vectors we write quite naturally \(\overrightarrow{a}=0\) and not \(\overrightarrow{a}=\overrightarrow{0}\).

  11. 11.

    For the summation we use almost exclusively the abbreviation \( \sum \nolimits _{n}\) (instead of \(\sum \nolimits _{n=1}^{\infty }\) or \(\sum \nolimits _{n=1}^{N}\) etc.). In the shorthand notation, the range of values of \(n\) must follow from the context of the problem at hand, if necessary.

  12. 12.

    For the connection between inner product and projection, see Appendix F, Vol. 1.

  13. 13.

    As we shall see in Chap. 13, a projection operator in quantum mechanics must meet a further condition (self-adjointness).

  14. 14.

    In other words, due to the process of measuring, a superposition such as \( \left| z\right\rangle =a\left| h\right\rangle +b\left| v\right\rangle \) ‘collapses’ e.g. into the state \(\left| h\right\rangle \).

  15. 15.

    In order to make it clear once more: If, for example, we measure an arbitrarily-polarized state \(\left| z\right\rangle =a\left| h\right\rangle +b\left| v\right\rangle \) with \(\left| a\right| ^{2}+\left| b\right| ^{2}=1\) and \(ab\ne 0\), we find with probability \(\left| a\right| ^{2}\) a horizontal linearly-polarized photon. This does not permit the conclusion that the photon was in that state before the measurement. It simply makes no sense in this case to speak of a of a well-defined value of the linear polarization (\(+1\) or \(-1\)) before the measurement.

  16. 16.

    Essentially, this operator is the \(x\) component of the orbital angular momentum operator for the angular momentum \(1\); see Chap. 16, Vol. 2.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jochen Pade .

Rights and permissions

Reprints and permissions

Copyright information

© 2014 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Pade, J. (2014). Complex Vector Spaces and Quantum Mechanics. In: Quantum Mechanics for Pedestrians 1: Fundamentals. Undergraduate Lecture Notes in Physics. Springer, Cham. https://doi.org/10.1007/978-3-319-00798-4_4

Download citation

Publish with us

Policies and ethics