Skip to main content
Log in

Reflection Positive One-Parameter Groups and Dilations

  • Published:
Complex Analysis and Operator Theory Aims and scope Submit manuscript

Abstract

The concept of reflection positivity has its origins in the work of Osterwalder–Schrader on constructive quantum field theory. It is a fundamental tool to construct a relativistic quantum field theory as a unitary representation of the Poincaré group from a non-relativistic field theory as a representation of the euclidean motion group. This is the second article in a series on the mathematical foundations of reflection positivity. We develop the theory of reflection positive one-parameter groups and the dual theory of dilations of contractive hermitian semigroups. In particular, we connect reflection positivity with the outgoing realization of unitary one-parameter groups by Lax and Phillips. We further show that our results provide effective tools to construct reflection positive representations of general symmetric Lie groups, including the \(ax+b\)-group, the Heisenberg group, the euclidean motion group and the euclidean conformal group.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. This is an abstraction of the Markov condition for Osterwalder–Schrader positive processes that one finds in [15, 16].

  2. The Hilbert space \(L^2(X,\Sigma ,\mu )\) has a natural realization as a space of functions on \(\Sigma \) by the transform \(\widetilde{f}(E) := \langle f, \chi _E \rangle \). This is a realization as a reproducing kernel Hilbert space on \(\Sigma \) with the kernel \(K(E,F) = \mu (E \cap F)\). Here we only use that the characteristic functions of the elements of \(\Sigma \) form a total subset of \(L^2(X,\Sigma ,\mu )\). In this picture, we obtain for each smaller \(\sigma \)-algebra \(\Sigma _0 \subseteq \Sigma \) a natural restriction map which corresponds to the conditional expectation. This restriction map restricts to a unitary isomorphism of the closed subspace generated by the characteristic functions \(\chi _E, E \in \Sigma _0\), with \(L^2(X,\Sigma _0,\mu )\).

References

  1. Borchers, H.: The CPT-theorem in two-dimensional theories of local observables. Comm. Math. Phys. 143, 315–332 (1992)

    Article  MATH  MathSciNet  Google Scholar 

  2. van Dijk, G.: Introduction to Harmonic Analysis and Generalized Gelfand Pairs. Studies in Math, vol. 36. de Gruyter, Berlin (2009)

    Book  MATH  Google Scholar 

  3. Dixmier, J., Malliavin, P.: Factorisations de fonctions et de vecteurs indórdainment différentiables. Bull. Soc. Math. 2e Série 102, 305–330 (1978)

    MATH  MathSciNet  Google Scholar 

  4. Engel, K.-J., Nagel, R.: One-parameter Semigroups for Linear Evolution Equations. Graduate Texts in Mathematics, vol. 194. Springer-Verlag, New York (2000)

    MATH  Google Scholar 

  5. Gelfand, I.M., Ya Vilenkin, N.: Generalized Functions. Vol. 4: Applications of Harmonic Analysis. Translated by Amiel Feinstein. Academic Press, New York (1964)

    Google Scholar 

  6. Glimm, J., Jaffe, A.: Quantum Physics-A Functional Integral Point of View. Springer-Verlag, New York (1981)

    MATH  Google Scholar 

  7. Hilgert, J., Neeb, K.H.: Lie Semigroups and their Applications. Lecture Notes in Mathematics, vol. 1552. Springer Verlag, Berlin (1993)

    MATH  Google Scholar 

  8. Jaffe, A.: Quantum theory and relativity. In: Doran, R.S., Moore, C.C., Zimmer, R.J. (eds.) Group Representations, Ergodic Theory, and Mathematical Physics: A Tribute to George W. Mackey.Contemp. Math, vol. 449. Amer Math Soc, Providence (2008)

    Google Scholar 

  9. Jaffe, A., Ritter, G.: Quantum field theory on curved backgrounds. I. The euclidean functional integral. Comm. Math. Phys. 270, 545–572 (2007)

    Article  MATH  MathSciNet  Google Scholar 

  10. Jaffe, A., Ritter, G.: Quantum field theory on curved backgrounds. II. Spacetime symmetries. Comm. Math. Phys., arXiv:hepth/0704.0052v1

  11. Jaffe, A., Ritter, G.: Reflection positivity and monotonicity. J. Math. Phys. 49(5), 052301 (2008)

    Article  MathSciNet  Google Scholar 

  12. Jaffe, A., Pedrocchi, F.L.: Reflection Positivity for Majoranas, Preprint arXiv:mathph/305.1792

  13. Jorgensen, P.E.T., Ólafsson, G.: Unitary representations of Lie groups with reflection symmetry. J. Funct. Anal. 158, 26–88 (1998)

    Article  MathSciNet  Google Scholar 

  14. Jorgensen, P.E.T., Ólafsson, G.: Unitary representations and Osterwalder–Schrader duality. In: Doran, R.S., Varadarajan, V.S. (eds.) The Mathematical Legacy of Harish–Chandra. Proc. Symp. in Pure Math., vol. 68. Amer. Math. Soc., Providence (2000)

    Google Scholar 

  15. Klein, A.: Gaussian OS-positive processes. Z. Wahrscheinlichkeitstheorie und Verw. Gebiete 40(2), 115–124 (1977)

    Article  MATH  MathSciNet  Google Scholar 

  16. Klein, A.: The semigroup characterization of Osterwalder–Schrader path spaces and the construction of euclidean fields. J. Funct. Anal. 27, 277–291 (1978)

    Article  MATH  Google Scholar 

  17. Klein, A., Landau, L.: Periodic Gaussian Osterwalder–Schrader positive processes and the two-sided Markov property on the circle. Pac. J. Math. 94(2), 341–367 (1981)

    Article  MATH  MathSciNet  Google Scholar 

  18. Klein, A., Landau, L.: From the Euclidean group to the Poincaré group via Osterwalder–Schrader positivity. Comm. Math. Phys. 87, 469–484 (1982)

    Article  MathSciNet  Google Scholar 

  19. Loos, O.: Symmetric Spaces I: General Theory. Benjamin, New York (1969)

    MATH  Google Scholar 

  20. Lüscher, M., Mack, G.: Global conformal invariance and quantum field theory. Comm Math. Phys. 41, 203–234 (1975)

    Article  MathSciNet  Google Scholar 

  21. Lax, P.D., Phillips, R.S.: Scattering theory. Bull. Amer. Math. Soc. 70, 130–142 (1964)

    Article  MATH  MathSciNet  Google Scholar 

  22. Lax, P.D., Phillips, R.S.: Scattering Theory. Pure and Applied Mathematics, vol. 26. Academic Press, New York (1967)

    MATH  Google Scholar 

  23. Lax, P.D., Phillips, R.S.: The translation representation theorem. Integr. Equ. Oper. Theory 4, 416–421 (1981)

    Article  MATH  MathSciNet  Google Scholar 

  24. Magyar, M.: Continuous Linear Representations, North-Holland, Mathematical Studies vol. 168 (1992)

  25. Neeb, K.-H.: Operator valued positive definite kernels on tubes. Monatshefte für Math. 126, 125–160 (1998)

    Article  MATH  MathSciNet  Google Scholar 

  26. Neeb, K.-H.: Holomorphy and Convexity in Lie Theory. Expositions in Mathematics, vol. 28. de Gruyter Verlag, Berlin (2000)

    Book  Google Scholar 

  27. Neeb, K.-H.: On differentiable vectors for representations of infinite dimensional Lie groups. J. Funct. Anal. 259, 2814–2855 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  28. Neeb, K.-H., Ólafsson, G.: Reflection positivity and conformal symmetry, to appear in J. Funct. Anal. 266, 2174–2224 (2014). see arXiv:RT.1206.2039

  29. Neeb, K.-H., Merigon, S., Ólafsson, G.: Integrability of unitary representations on reproducing kernel spaces, arXiv:RT.1406.2681(submitted)

  30. Reed, S., Simon, B.: Methods of Mathematical Physics II: Fourier Analysis, Self-Adjointness. Academic Press, New York (1975)

    MATH  Google Scholar 

  31. Schwartz, L.: Théorie des distributions, 2nd edn. Hermann, Paris (1973)

    Google Scholar 

  32. Simon, B.: Functional Integration and Quantum Physics, 2nd edn. AMS Chelsea Publishing, Providence (2005)

    MATH  Google Scholar 

  33. Sinai Ya, G.: Dynamical systems with countable Lebesgue spectrum. I Izv. Akad. Nauk. USSR 25, 899–924 (1961)

    MATH  Google Scholar 

  34. Sz-Nagy, B., Foias, C., Bercovici, H., Kérchy, L.: Harmonic Analysis of Operators on Hilbert Space, 2nd edn. Springer, Universitext, Berlin (2010)

    Book  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Gestur Ólafsson.

Additional information

Communicated by Yuri Kondratiev.

K.-H. Neeb: Supported by DFG-grant NE 413/7-2, Schwerpunktprogramm “Darstellungstheorie”.

G. Ólafsson: The research of G. Ólafsson was supported by NSF grants DMS-0801010, DMS-1101337 and the Emerging Fields Project “Quantum Geometry” of the University of Erlangen.

Appendices

Appendix A: Rotation Invariant Reflection Positive Measures on \({\mathbb R}^d\)

In this section we take a closer look at rotation invariant tempered measures \(\nu \) on \({\mathbb R}^d\) which are reflection positive in the sense that their Fourier transform \(\widehat{\nu }\) is a reflection positive distribution w.r.t. (\({\mathbb R}^d, {\mathbb R}^d_+,\theta )\), where \(\theta (x_0,\mathbf {x}) = (-x_0,\mathbf {x})\). In Euclidean Quantum Field Theories, the Fourier transforms of these measures are the Schwinger distributions \(S_2(x,y) = \widehat{\nu }(x-y)\) describing the euclidean \(2\)-point “functions” (cf. [6, p. 91]).

The following lemma will be useful in our discussion of examples.

Lemma 7.1

For a non-zero positive Borel measure \(\rho \) on \([0,\infty [\), we consider the rotation invariant measure \(d\nu (x) := \Theta (x)dx\) on \({\mathbb R}^d\) with the density \(\Theta (x) := \int _0^\infty \frac{1}{y^2 + x^2}\, d\rho (y)\). Then the following are equivalent:

  1. (i)

    There exists a non-zero \(x \in {\mathbb R}^d\) with \(\Theta (x) < \infty \).

  2. (ii)

    \(\Theta (x) < \infty \) for every non-zero \(x \in {\mathbb R}^d\).

  3. (iii)

    \(L^2({\mathbb R}^d,\nu ) \not =\{0\}\).

  4. (iv)

    \(\int _0^\infty \frac{d\rho (y)}{1+y^2} < \infty \). This implies that \(\rho \) is tempered.

If these conditions are satisfied, then the following assertions hold:

  1. (a)

    \(\lim _{\Vert x\Vert \rightarrow \infty } \Theta (x)x^2 = \rho ([0,\infty [)> 0.\)

  2. (b)

    \(L^2({\mathbb R}^d,\nu )\) contains a non-zero polynomial if and only if \(d = 1\) and \(\nu \) is finite, and then every polynomial in \(L^2({\mathbb R}^d,\nu )\) is constant. We have \(\nu ({\mathbb R}) = \pi \int _0^\infty \frac{d\rho (y)}{y}\).

Proof

The equivalence of (i)–(iv) is easy to see. Suppose that these conditions are satisfied.

(a) follows from

$$\begin{aligned} \lim _{x \rightarrow \infty } \Theta (x)x^2 = \lim _{x \rightarrow \infty } \int _0^\infty \frac{x^2}{y^2 + x^2}\, d\rho (y) = \int _0^\infty \, d\rho (y) > 0. \end{aligned}$$

(b) Let \(f = \sum _{j = 0}^N f_j : {\mathbb R}^d \rightarrow {\mathbb C}\) be a polynomial of degree \(N\), where the \(f_j\) are homogeneous of degree \(j\). For a suitably normalized measure \(\sigma \) on the sphere \({\mathbb S}^{d-1}\), we have the integral formula

$$\begin{aligned} \int _{{\mathbb R}^d} F(x)\, dx = \int _0^\infty \int _{{\mathbb S}^{d-1}} F(r x)\, d\sigma (x)\, r^{d-1}\, dr, \end{aligned}$$

so that

$$\begin{aligned} \int _{{\mathbb R}^d} F(x)\, d\nu (x) = \int _0^\infty \int _{{\mathbb S}^{d-1}} F(r x) \, d\sigma (x)\, \Theta (r)r^{d-1}\, dr. \end{aligned}$$

The function \(|f|^2\) is a polynomial of degree \(2N\), and this implies that \(h(r) := \int _{{\mathbb S}^{d-1}} |f(r x)|^2\, d\sigma (x)\) also is a polynomial of degree \(2N\). If

$$\begin{aligned} \int _{{\mathbb R}^d} |f(x)|^2\, d\nu (x) = \int _0^\infty h(r)\Theta (r)r^{d-1}\, dr \end{aligned}$$

is finite, (a) implies that \(\int _1^\infty h(r)r^{d-3}\, dr < \infty \), so that we obtain \(2N + d- 3 < -1\), i.e., \(2N + d < 2\). This leaves only the possibilities \(d = 1\) and \(N= 0\), so that \(f\) must be constant. But \(1 \in L^2({\mathbb R}^d,\nu )\) is equivalent to the finiteness of the measure \(\nu \). In view of

$$\begin{aligned} \nu ({\mathbb R}) = \int _0^\infty \int _{{\mathbb R}} \frac{dx}{y^2 + x^2}\, d\rho (y) = \int _0^\infty \frac{\pi }{y}\, d\rho (y), \end{aligned}$$

\(\nu \) is finite if and only if \(\int _0^\infty \frac{1}{y}\, d\rho (y) < \infty \).\(\square \)

Definition 7.2

We call a positive Borel measure \(\rho \) on \([0,\infty [\) tame if the conditions (i)–(iv) from the preceding lemma are satisfied.

Proposition 7.3

If \(\rho \) is tame, then \(\nu = \Theta \cdot dx\) with \(\Theta (x) := \int _0^\infty \frac{1}{y^2 + x^2}\, d\rho (y)\) is a tempered measure on \({\mathbb R}^d\) if and only if \(d > 2\) or

  1. (a)

    \(d = 1\) and \(\int _0^1 \frac{1}{y}\, d\rho (y) < \infty \).

  2. (b)

    \(d= 2\) and \(\int _0^1 \ln (y^{-1})\, d\rho (y) < \infty \).

In particular, \(\rho (\{0\}) = 0\) for \(d = 1,2\).

Proof

That \(\nu \) is a tempered measure means that, for some \(N \in {\mathbb N}\), the integral

$$\begin{aligned} \int _{{\mathbb R}^d} (1 + p^2)^{-N}\, d\nu (p) = \int _0^\infty \int _{{\mathbb R}^d} (1 + p^2)^{-N}\frac{1}{y^2 + p^2}\, dp\, d\rho (y) \end{aligned}$$

is finite. This is equivalent to the finiteness of the integral

$$\begin{aligned} \int _0^\infty \int _0^\infty (1 + r^2)^{-N}\frac{r^{d-1}}{y^2 + r^2}\, dr\, d\rho (y). \end{aligned}$$
(27)

If (27) is finite, we obtain in particular that

$$\begin{aligned} \int _0^\infty \int _1^2(1 + r^2)^{-N}\frac{r^{d-1}}{y^2 + r^2}\, dr\, d\rho (y) < \infty , \end{aligned}$$

and this is equivalent to the tameness of \(\rho \).

Now we split the double integral (27) according to \(r < 1\) and \(r > 1\) to find

$$\begin{aligned}&\int _0^\infty \int _1^\infty (1 + r^2)^{-N}\frac{r^{d-1}}{y^2 + r^2}\, dr\, d\rho (y) = \int _0^\infty I(y)\, d\rho (y) \end{aligned}$$

with

$$\begin{aligned} I(y) = \int _1^\infty (1 + r^2)^{-N}\frac{r^{d-1}}{y^2 + r^2}\, dr \le I(0). \end{aligned}$$

This integral is finite if and only if \(2N + 2 > d\). To see if \(I\) is \(\rho \)-integrable, we have to estimate the asymptotics of \(I(y)\) for \(y \rightarrow \infty \). The Monotone Convergence Theorem implies that \(\lim _{y \rightarrow \infty } I(y)= 0\). The asymptotics of \(I\) is the same as the asymptotics of the function

$$\begin{aligned} J(y) := \int _1^\infty \frac{r^{d-1-2N}}{y^2 + r^2}\, dr = \frac{1}{y^2} \int _1^\infty \frac{r^{d-1-2N}}{1+ \frac{r^2}{y^2}}\, dr \le \frac{1}{y^2} \int _1^\infty r^{d-1-2N}\, dr, \end{aligned}$$

provided \(2N > d\). Now the tameness of \(\rho \) implies that \(I\) is \(\rho \)-integrable for \(2N > d\).

Next we turn to the other part of the integral:

$$\begin{aligned} \int _0^\infty \int _0^1 (1 + r^2)^{-N}\frac{r^{d-1}}{y^2 + r^2}\, dr \, d\rho (y). \end{aligned}$$

Since \(1 \le 1 + r^2 \le 2\) for \(0 \le r \le 1\), it is finite if and only if the integral

$$\begin{aligned} \int _0^\infty \int _0^1 \frac{r^{d-1}}{y^2 + r^2}\, dr\, d\rho (y) = \int _0^\infty K(y)\, d\rho (y) \quad \text{ for } \quad K(y) := \int _0^1 \frac{r^{d-1}}{y^2 + r^2}\, dr, \end{aligned}$$

is finite. The integral \(K(y)\) is finite for \(y > 0\), and \(K(0) = \int _0^1 r^{d-3} \, dr\) is finite if and only if \(d > 2\). We thus have to estimate the asymptotics of \(K\) for \(y \rightarrow \infty \), and, for \(d \le 2\), also for \(y \rightarrow 0\). In view of

$$\begin{aligned} K(y) \le \int _0^1 \frac{dr}{y^2 + r^2} \le \frac{1}{y^2}, \end{aligned}$$

the finiteness of the integral \(\int _1^\infty K(y)\, d\rho (y)\) follows for every \(d\) from the tameness of \(\rho \). For \(d > 2\), the function \(K\) is continuous on \([0,\infty [\), hence \(\rho \)-integrable. This completes the proof for \(d > 2\).

For \(d = 1\) we have

$$\begin{aligned} K(y) = \int _0^1 \frac{1}{y^2 + r^2}\, dr = \frac{1}{y}\left( \arctan \frac{1}{y}- \arctan 0\right) = \frac{1}{y} \arctan \frac{1}{y} \sim \frac{\pi }{2y} \end{aligned}$$

for \(y \rightarrow 0\). We thus find the necessary and sufficient condition (a) for the finiteness of the \(\rho \)-integral of \(I\).

For \(d = 2\) we have

$$\begin{aligned} K(y) = \int _0^1 \frac{r}{y^2 + r^2}\, dr = \frac{1}{2} \big (\ln (y^2 + 1) - \ln (y^2)\big ) \sim \ln (y^{-1}) \end{aligned}$$

for \(y \rightarrow 0\). Therefore \(I\) is \(\rho \)-integrable over \([0,1]\) if and only if (b) is satisfied.\(\square \)

The following proposition can also be derived from [6, Prop. 6.2.5]. We include it for the sake of easier reference.

Proposition 7.4

If \(\rho \) is a tame measure on \([0,\infty [\) for which the measure \(\nu = \Theta \cdot dx\) is tempered, then the distribution \(\widehat{\nu } \in C^{-\infty }({\mathbb R}^d)\) is reflection positive for \(({\mathbb R}^d, {\mathbb R}^d_+, \theta )\) and \(\theta (x_0, \mathbf {x}) = (-x_0, \mathbf {x})\), i.e.,

Proof

Writing

$$\begin{aligned} \nu = \int _0^\infty \nu _y\, d\rho (y) \quad \text{ with } \quad d\nu _y(p)= \frac{dp}{y^2 + p^2}, \end{aligned}$$

we see that it suffices to show that the distributions \(\widehat{\nu _y}\) are reflection positive. First we observe that

$$\begin{aligned} \int _{{\mathbb R}^d} \theta \widehat{\psi } \cdot \overline{\widehat{\psi }}\, d\nu _y = \int _{{\mathbb R}^d} \overline{\widehat{\psi }(-p_0, \mathbf {p})} \frac{\widehat{\psi }(p_0, \mathbf {p})}{y^2 + p^2}\, dp \quad \text{ for } \quad p = (p_0, \mathbf {p})\in {\mathbb R}\times {\mathbb R}^{d-1}. \end{aligned}$$

For each \(\mathbf {p}\in {\mathbb R}^{d-1}\), the function \(h_\mathbf {p}(p_0) := \widehat{\psi } (-p_0, \mathbf {p})\) is a Schwartz function with \(\mathop {\mathrm{supp}}\nolimits (\widehat{h}_{\mathbf {p}}) \subseteq ]0, \infty [\), and

$$\begin{aligned} \int _{{\mathbb R}^d} \theta \widehat{\psi } \cdot \overline{\widehat{\psi }}\, d\nu _y = \int _{{\mathbb R}^{d-1}} \int _{\mathbb R}\frac{\overline{h_{\mathbf {p}}(p_0)} h_{\mathbf {p}}(-p_0)}{p_0^2 + y^2 + \mathbf {p}^2}\, dp_0 \, d\mathbf {p}. \end{aligned}$$

If \(y = 0\), then \(d > 2\), so that \(\{0\}\) is a Lebesgue zero set in \({\mathbb R}^{d-1}\). Therefore it suffices to show that, for \(f \in \mathcal {S}({\mathbb R})\) with \(\mathop {\mathrm{supp}}\nolimits (\widehat{f}) \subseteq {\mathbb R}_+\) and \(y > 0\), the measure \(d\nu _y^1(x) := \frac{dx}{x^2 + y^2}\) satisfies

$$\begin{aligned} \int _{\mathbb R}f(x) \overline{f(-x)}\, \frac{dx}{x^2 + y^2} = \int _{\mathbb R}f(x) \overline{f(-x)}\, \nu _y^1(x) \, dx \ge 0. \end{aligned}$$

In view of \(\widehat{\nu _y^1}(p)= \frac{\pi }{y}\ e^{-y|p|}\), this follows from

$$\begin{aligned}&\int _{\mathbb R}(\widehat{f} * \overline{\widehat{f}})(p) e^{-y |p|}\, dp \!=\! \int _{{\mathbb R}_+} (\widehat{f} * \overline{\widehat{f}})(p) e^{-y p} \, dp \!=\! \mathcal {L}(\widehat{f}* \overline{\widehat{f}})(y) = |\mathcal {L}(\widehat{f})(y)|^2 \!\ge \! 0. \end{aligned}$$

More conceptually, the preceding calculation means that \(\widehat{\nu _y^1}\) is a reflection positive function on \({\mathbb R}\), resp., its restriction to the semigroup \({\mathbb R}_+\) is also positive definite with respect to the trivial involution (cf. [28]).\(\square \)

Note that Proposition 7.8 below implies in particular that, for \(d = 1\), the preceding proposition does not cover all reflection positive measures. Example that are not covered arise for \(\widehat{\nu }(x) = |x|^{-s}\) (on \({\mathbb R}^\times \)) for \(s > 1\).

Lemma 7.5

Suppose that \(\rho \) is a positive Borel measure on \([0,\infty [\) whose Laplace transform \(\mathcal {L}(\rho )\) exists on \({\mathbb R}_+\). Then the function \(\mathcal {L}(\rho )\) on \({\mathbb R}_+\) extends to a symmetric distribution on \({\mathbb R}\) if and only if \(\rho \) is tempered, i.e., \(\int _1^\infty \frac{1}{y^k}\, d\rho (y) < \infty \) for some \(k \in {\mathbb N}\). More precisely, for \(\ell \in {\mathbb N}_0\),

$$\begin{aligned} \int _0^1 x^\ell \mathcal {L}(\rho )(x)\, dx < \infty \quad \Leftrightarrow \quad \int _1^\infty \frac{1}{y^{\ell + 1}}\, d\rho (y) < \infty . \end{aligned}$$

In particular, \(\mathcal {L}(\rho )\) is locally integrable on \([0, \infty [\) if and only if \(\int _1^\infty \frac{1}{y}\, d\rho (y) < \infty \).

Proof

Since the Laplace transform of the finite measure \(\rho \vert _{[0,1]}\) extends to a continuous function on \([0,\infty [\), we may w.l.o.g. assume that \(\rho ([0,1]) = 0\). We put \(D^\times (x) = \mathcal {L}(\rho )(|x|)\) for \(x \in {\mathbb R}^\times \). In view of [31, Thm. VIII, §VII.4], \(D^\times \) extends to a distribution on \({\mathbb R}\) if and only if there exists an \(\ell \in {\mathbb N}_0\) with

$$\begin{aligned} \int _0^1 x^\ell D^\times (x)\, dx < \infty . \end{aligned}$$
(28)

We rewrite this integral as follows

$$\begin{aligned} \int _0^1 x^\ell D^\times (x)\, dx&= \int _0^1 x^\ell \int _1^\infty e^{-y x}\, d\rho (y) dx = \int _1^\infty \int _0^1 x^\ell e^{-y x}\, dx\, d\rho (y) \\&= \int _1^\infty \int _0^y \frac{u^\ell }{y^\ell } e^{-u}\, \frac{du}{y}\, d\rho (y) = \int _1^\infty \int _0^y u^\ell e^{-u}\, du\, \frac{1}{y^{\ell +1}}\, d\rho (y). \end{aligned}$$

In view of \(0 < \int _0^\infty u^\ell e^{-u}\,du < \infty \), it follows that \(\int _0^1 x^\ell D^\times (x)\, dx\) is finite if and only if \(\int _1^\infty \frac{1}{y^{\ell +1}}\, d\rho (y)\) is finite.\(\square \)

Theorem 7.6

The reflection positive distributions on \({\mathbb R}\) which are represented by a locally integrable function are the Fourier transforms \(D = \widehat{\nu }\) of measures of the form

$$\begin{aligned} \nu = c \delta _0 + \Theta \cdot dx, \qquad \Theta (x) = \frac{1}{\pi } \int _0^\infty \frac{y}{y^2 + y^2}\, d\rho (y), \end{aligned}$$
(29)

where \(c \ge 0\) and \(\rho \) is a positive Radon measure on \({\mathbb R}_+\) satisfying

(30)

Proof

Let \(D \in C^{-\infty }({\mathbb R})\) be reflection positive. Then its restriction to \({\mathbb R}_+\) is positive definite w.r.t. the involution \(x^* = x\), so that [28, Thm. 4.13] implies \(D\vert _{{\mathbb R}_+}\) is represented by an analytic function which is the Laplace transform \(\mathcal {L}(\rho )\) of a positive Radon measure \(\rho \) on \([0,\infty [\). If \(D\) is represented by a locally integrable function, \(\mathcal {L}(\rho )\) is locally integrable, so that Lemma 7.5 leads to

$$\begin{aligned} \int _1^\infty \frac{1}{y}\, d\rho (y) < \infty . \end{aligned}$$

We now have almost everywhere on \({\mathbb R}\)

$$\begin{aligned} D(x) = \mathcal {L}(\rho )(|x|) = \int _0^\infty e^{-y|x|}\, d\rho (y) = \widehat{\nu }(x) \quad \text{ for } \quad \nu = \rho (\{0\})\delta _0 + \Theta \cdot dx, \end{aligned}$$

provided \(\nu \) is a tempered measure. Here we used that \(e^{-y|x|}\) is the Fourier transform of the measure \(\frac{1}{\pi }\frac{y \cdot dx}{y^2 + x^2}\). In view of Proposition 7.3, the temperedness of \(\nu \) is equivalent to (30).

Suppose, conversely, that (30) is satisfied. Then Proposition 7.3 implies that \(\nu \) is tempered, so that \(D := \widehat{\nu }\) is a positive definite distribution. That it is reflection positive follows from the reflection positivity of the functions \(\widehat{\nu }_y(x) = e^{-y|x|}\).\(\square \)

Remark 7.7

The integral representation from Theorem 7.6 does not extend to all reflection positive distributions on \({\mathbb R}\). Any distribution of the form \(E := P\left( {1\over i}{d\over dx}\right) \delta _0\), where \(P\) is a non-negative even polynomial, is reflection positive. In this case \(E \vert _{{\mathbb R}_+} = 0\).

We also conclude that, for any reflection positive distribution \(D\) on \({\mathbb R}\), the distribution \(D + E\) is another extension of \(D\vert _{{\mathbb R}_+}\), so that reflection positive extensions are not unique.

Proposition 7.8

For every tempered measure \(\rho \) on \([0,\infty [\), there exists a reflection positive distribution \(D\) on \({\mathbb R}\) with \(D\vert _{{\mathbb R}_+} = \mathcal {L}(\rho )\).

Proof

In view of Lemma 7.5, we may assume that \(\int _1^\infty \frac{1}{\lambda ^{4N}}\, d\rho (\lambda ) < \infty .\) Then the measure

$$\begin{aligned} d\widetilde{\rho }(\lambda ) := \frac{1}{1 + \lambda ^{4N}}\, d\rho (\lambda ) \end{aligned}$$

on \([0,\infty [\) is finite. Its Laplace transform therefore defines a continuous reflection positive function \(\widetilde{D}(x) := \mathcal {L}(\widetilde{\rho })(|x|)\) ([28, Cor. 3.3]). For the polynomial \(P(x) := 1 + x^{4N}\), we now obtain the positive definite distribution \(D := P({1\over i}{d\over dx}) \widetilde{D},\) and its restriction to \({\mathbb R}_+\) is given by

$$\begin{aligned} D(x) \!=\! P\Big ({1\over i}{d\over dx}\Big ) \mathcal {L}(\widetilde{\rho })(x) \!=\! \int _0^\infty (1 + \lambda ^4)e^{-\lambda x}\, d\widetilde{\rho }(\lambda ) = \int _0^\infty e^{-\lambda x}\, d\rho (\lambda ) = \mathcal {L}(\rho )x). \end{aligned}$$

Therefore \(D\) is a reflection positive extension of \(\mathcal {L}(\rho )\).\(\square \)

Appendix B: Lorentzian Invariant Tempered Measures on the Light Cone

For the following theorem, we recall the \(L^{\mathop {\uparrow }}\)-invariant measures \(\mu _m\) from Definition 6.10. It is our version of the Lehmann Spectral Formula from Quantum Field Theory ([6, Thm. 6.2.4]) describing the \(2\)-point functions of Poincaré invariant field theories. The proof given below provides many details skipped in [6].

Theorem 8.1

For every \(L^{\mathop {\uparrow }}\)-invariant Borel measure \(\mu \) on the closed forward light cone \(\overline{V_+}\), there exists a \(\sigma \)-finite Borel measure \(\rho \) on \([0,\infty [\) and a constant \(c \ge 0\) so that

$$\begin{aligned} \mu = c \delta _0 + \int _0^\infty \mu _m\, d\rho (m), \end{aligned}$$
(31)

where \(\rho (\{0\}) = 0\) for \(d = 1\).

  1. (a)

    The measure \(\mu \) is tempered if and only if the following conditions are satisfied:

    1. (C1)

      \(\rho \) is a tempered measure on \([0,\infty [\).

    2. (C2)

      \(\int _0^1 \frac{1}{m}\, d\rho (m) < \infty \) for \(d = 1\).

    3. (C3)

      \(\int _0^1 \ln (m^{-1})\, d\rho (m) < \infty \) for \(d = 2\). In particular, \(\rho (\{0\}) = 0\).

  2. (b)

    Let \(\mathop {\mathrm{pr}}\nolimits : {\mathbb R}^d \rightarrow {\mathbb R}^{d-1}, (p_0, \mathbf {p}) \mapsto \mathbf {p}\) be the projection map. Then the measure \(\widetilde{\mu }= \mathop {\mathrm{pr}}\nolimits _*\mu \) on \({\mathbb R}^{d-1}\) is tempered if and only if \(\mu \) is tempered and, in addition,

    $$\begin{aligned} \int _1^\infty \frac{1}{m}\, d\rho (m) < \infty . \end{aligned}$$

    If this condition is not satisfied, then every Borel subset \(E \subseteq {\mathbb R}^{d-1}\) satisfies \(\widetilde{\mu }(E) \in \{ 0, \infty \}\), so that \(L^2({\mathbb R}^{d-1}, \widetilde{\mu }) = \{0\}\).

Proof

In [30, Thm. IX.33], the decomposition (31) is stated only in the case \(d = 4\), but the proof works in the general case, where it leads to an integral representation of \(\mu \) in terms of a measure \(\rho \) on \([0,\infty [\), whose restriction to \(]0,\infty [\) is a Radon measure with possibly infinite mass for the interval \([0,1]\) (see also [2, Lemma 9.1.2/3] for a description of the ergodic \(\mathop {\mathrm{O}}\nolimits _{1,d-1}({\mathbb R})\)-invariant measures on \({\mathbb R}^d\)).

(a) Now the problem consists in characterizing the temperedness of \(\mu \) in terms of properties of the measure \(\rho \). To this end, we may w.l.o.g. assume that \(c = 0\). Temperedness of \(\mu \) is equivalent to the existence of an \(N \in {\mathbb N}\) such that the following integral is finite

$$\begin{aligned} \int _{{\mathbb R}^d} \frac{1}{(1 + p^2)^N}\, d\mu (p)&= \int _0^\infty \int _{{\mathbb R}^d} \frac{1}{(1 + p^2)^N}\, d\mu _m(p)\, d\rho (m) \\&= \int _0^\infty \int _{{\mathbb R}^{d-1}} \frac{1}{(1 + (m^2 + \mathbf {p}^2) + \mathbf {p}^2)^N} \frac{1}{\sqrt{m^2 + \mathbf {p}^2}}\, d\mathbf {p}\, d\rho (m). \end{aligned}$$

For \(d = 1\), the preceding formula simplifies to

$$\begin{aligned} \int _{{\mathbb R}} \frac{1}{(1 + p^2)^N}\, d\mu (p)&= \int _0^\infty \frac{1}{(1 + m^2)^N}\frac{1}{m}\, d\rho (m). \end{aligned}$$

There exists an \(N \in {\mathbb N}_0\) for which such an integral is finite if and only if (a) is satisfied and \(\rho \) is a tempered.

From now on we assume that \(d > 1\). Evaluating the integral over \({\mathbb R}^{d-1}\) in polar coordinates, we see that this is equivalent to the finiteness of the double integral

$$\begin{aligned} \int _0^\infty \int _0^\infty \frac{1}{(1 + m^2 + 2r^2)^N} \frac{1}{\sqrt{m^2 + r^2}} r^{d-2}\, dr\, d\rho (m). \end{aligned}$$
(32)

The finiteness of (32) implies in particular the finiteness of the integral

$$\begin{aligned} \int _0^\infty \int _1^2\frac{1}{(1 + m^2 + 2r^2)^N} \frac{1}{\sqrt{m^2 + r^2}} r^{d-2}\, dr\, d\rho (m), \end{aligned}$$

which is equivalent to the finiteness of the integral

$$\begin{aligned} \int _0^\infty \frac{1}{(1 + m^2)^N}\frac{1}{\sqrt{m^2 + 1}}\, d\rho (m) = \int _0^\infty \frac{1}{(1 + m^2)^{N+ \frac{1}{2}}}\, d\rho (m). \end{aligned}$$

This implies that \(\rho ([0,M]) < \infty \) for every \(M > 0\) and that \(\rho \) is tempered.

Suppose, conversely, that \(\rho \) is tempered. We discuss the integral (32) by splitting it into the two pieces corresponding to \(r < 1\) and \(r > 1\). For the \(r >1\)-part we obtain for \(N \ge N_0\) and \(N_0 > \frac{d-1}{2}\):

$$\begin{aligned}&\int _1^\infty \frac{1}{(1 + m^2 + 2r^2)^N} \frac{1}{\sqrt{m^2 + r^2}} r^{d-2}\, dr \\&\le \frac{1}{\sqrt{1 + m^2}} \int _1^\infty \frac{1}{(1 + m^2 + 2r^2)^{N_0}(1 + m^2 + 2r^2)^{N-N_0}} r^{d-2}\, dr \\&\le \frac{1}{\sqrt{1 + m^2}} \frac{1}{ (3 + m^2)^{N-N_0}} \int _1^\infty \frac{1}{(1 + 2r^2)^{N_0}} r^{d-2}\, dr. \end{aligned}$$

Our condition on \(N_0\) ensures that this integral is finite, and now the temperedness of \(\rho \) implies that

$$\begin{aligned} \int _0^\infty \int _1^\infty \frac{1}{(1 + m^2 + 2r^2)^N} \frac{1}{\sqrt{m^2 + r^2}} r^{d-2}\, dr\, d\rho (m) < \infty \end{aligned}$$

if \(N\) is sufficiently large.

Next we note that,

$$\begin{aligned}&\int _1^\infty \int _0^1 \frac{1}{(1 + m^2 + 2r^2)^N} \frac{1}{\sqrt{m^2 + r^2}} r^{d-2}\, dr\, d\rho (m) \\&\le \int _1^\infty \int _0^1 \frac{1}{(1 + m^2)^N} \frac{1}{m} \, dr\, d\rho (m) = \int _1^\infty \frac{1}{(1 + m^2)^N}\frac{1}{m}\, d\rho (m) < \infty . \end{aligned}$$

This already shows that \(\int _1^\infty \mu _m\, d\rho (m)\) is tempered if \(\rho \) is tempered.

It remains to consider the integral

$$\begin{aligned} \int _0^1 \int _0^1 \frac{1}{(1 + m^2 + 2r^2)^N} \frac{1}{\sqrt{m^2 + r^2}} r^{d-2}\, dr\, d\rho (m), \end{aligned}$$

which is finite if and only if

$$\begin{aligned} \int _0^1 \int _0^1 \frac{r^{d-2}}{\sqrt{m^2 + r^2}}\, dr\, d\rho (m) < \infty . \end{aligned}$$

Let

$$\begin{aligned} I(m) := \int _0^1 \frac{r^{d-2}}{\sqrt{m^2 + r^2}} \, dr, \end{aligned}$$

and observe that this defines a continuous function for \(m > 0\) with \(I(0)= \int _0^1 r^{d-3}\, dr < \infty \) if and only if \(d > 2\). Since \(\rho ([0,1]) < \infty \), we conclude that, for \(d\ge 3\), \(\mu \) is a tempered measure if and only if \(\rho \) has this property.

This leaves the case \(d = 2\). Then

$$\begin{aligned} I(m) = \int _0^1 \frac{1}{\sqrt{m^2 + r^2}}\, dr = \ln (1 + \sqrt{1 + m^2}) - \ln (m). \end{aligned}$$

We thus find condition (b), which is needed to ensure that \(\mu \) is a tempered measure.

(b) The measure \(\widetilde{\mu }\) is of the form \(\widetilde{\Theta }(\mathbf {p})\, d\mathbf {p}\) for

$$\begin{aligned} \widetilde{\Theta }(\mathbf {p}) = \int _0^\infty \frac{1}{\sqrt{m^2 + \mathbf {p}^2}}\, d\rho (m). \end{aligned}$$
(33)

For \(d = 1\), the measure \(\widetilde{\mu }\) is a point measure which is tempered if and only if it is finite. This is equivalent to

$$\begin{aligned} \widetilde{\Theta }(0) = \int _0^\infty \frac{1}{m}\, d\rho (m) < \infty . \end{aligned}$$

Suppose that \(d > 1\). If \(\widetilde{\mu }\) is finite on a set of positive Lebesgue measure, then there exists a non-zero \(\mathbf {p}\in {\mathbb R}^{d-1}\) with \(\widetilde{\Theta }(\mathbf {p}) < \infty \). This implies that compact subsets of \([0,\infty [\) have finite \(\rho \)-measure (\(\rho \) is a Radon measure), and that

$$\begin{aligned} \int _1^\infty \frac{1}{m}\, d\rho (m) < \infty . \end{aligned}$$
(34)

In particular, \(\rho \) is tempered. If (34) is not satisfied, then \(\widetilde{\Theta }(\mathbf {p}) = \infty \) implies that \(\widetilde{\mu }(E) = \infty \) for every Borel subset \(E \subseteq {\mathbb R}^{d-1}\) of positive Lebesgue measure and \(\widetilde{\mu }(E) = 0\) if \(E\) is a Lebesgue zero set.

Let us assume that \(\rho \) is a Radon measure on \([0,\infty [\) satisfying (34). Temperedness of \(\widetilde{\mu }\) is equivalent to the existence of an \(N \in {\mathbb N}\) such that

$$\begin{aligned} \int _{{\mathbb R}^{d-1}} \frac{1}{(1 + \mathbf {p}^2)^N}\, d\widetilde{\mu }(p)&= \int _0^\infty \int _{{\mathbb R}^{d-1}} \frac{1}{(1 + \mathbf {p}^2)^N} \frac{1}{\sqrt{m^2 + \mathbf {p}^2}}\, d\mathbf {p}\, d\rho (m) < \infty . \end{aligned}$$

Evaluating the integral over \({\mathbb R}^{d-1}\) in polar coordinates, we see that this is equivalent to the finiteness of the double integral

$$\begin{aligned} \int _0^\infty \int _0^\infty \frac{1}{(1 + r^2)^N} \frac{1}{\sqrt{m^2 + r^2}} r^{d-2}\, dr\, d\rho (m). \end{aligned}$$
(35)

We discuss the integral (35) by splitting it into the two pieces corresponding to \(r < 1\) and \(r > 1\). For the \(r >1\)-part we obtain for \(N > \frac{d-1}{2}\):

$$\begin{aligned} \int _1^\infty \frac{1}{(1 + r^2)^N} \frac{1}{\sqrt{m^2 + r^2}} r^{d-2}\, dr \le \frac{1}{\sqrt{1 + m^2}} \int _1^\infty \frac{r^{d-2}}{(1 +r^2)^{N}}\, dr. \end{aligned}$$

Our condition on \(N\) ensures that this integral is finite, and now the finiteness of \(\int _0^\infty \, \frac{1}{\sqrt{1+m^2}}\, d\rho (m)\) implies the finiteness of the double integral

$$\begin{aligned} \int _0^\infty \int _1^\infty \frac{1}{(1 + r^2)^N} \frac{1}{\sqrt{m^2 + r^2}} r^{d-2}\, dr\, d\rho (m). \end{aligned}$$

Next we note that,

$$\begin{aligned} \int _1^\infty \int _0^1 \frac{1}{(1 + r^2)^N} \frac{ r^{d-2}}{\sqrt{m^2 + r^2}}\, dr\, d\rho (m) \le \int _1^\infty \frac{1}{m} \, d\rho (m) \cdot \int _0^1 \frac{r^{d-2}}{(1 + r^2)^N}\, dr < \infty . \end{aligned}$$

It remains to consider the integral

$$\begin{aligned} \int _0^1 \int _0^1 \frac{1}{(1 + r^2)^N} \frac{r^{d-2}}{\sqrt{m^2 + r^2}} \, dr\, d\rho (m), \end{aligned}$$

which is finite if and only if

$$\begin{aligned} \int _0^1 \int _0^1 \frac{r^{d-2}}{\sqrt{m^2 + r^2}}\, dr\, d\rho (m) < \infty . \end{aligned}$$

In (a) we have seen that this integral is finite if \(d > 2\), and that, for \(d = 2\), its finiteness is equivalent to (C3). This proves (b).\(\square \)

Remark 8.2

For the measure \(\nu = \Theta \cdot dx\) on \({\mathbb R}^d\), where \(\Theta (x) = \frac{1}{\pi } \int _0^\infty \frac{1}{m^2 + x^2}\, d\rho (m)\), the projection \(\widetilde{\nu }\) to \({\mathbb R}^{d-1}\) under \(\mathop {\mathrm{pr}}\nolimits (x_0,\mathbf {x}) := \mathbf {x}\), is of the form \(\widetilde{\Theta }(\mathbf {x})\, d\mathbf {x}\), where

$$\begin{aligned} \widetilde{\Theta }(\mathbf {x})&= \frac{1}{\pi } \int _0^\infty \int _{\mathbb R}\frac{1}{m^2 + \mathbf {x}^2 + x_0^2}\, dx_0\, d\rho (m)\\&= \int _0^\infty \frac{1}{\sqrt{m^2 + \mathbf {x}^2}} \frac{1}{\pi } \int _{\mathbb R}\frac{\sqrt{m^2 + \mathbf {x}^2}}{m^2 + \mathbf {x}^2 + x_0^2}\, dx_0\, d\rho (m) \\&= \int _0^\infty \frac{1}{\sqrt{m^2 + \mathbf {x}^2}}\, d\rho (m). \end{aligned}$$

This implies that \(\widetilde{\nu }= \widetilde{\mu }\) [cf. (33)].

Appendix C: Positive Definite Functions

In this appendix we collect some definitions and results concerning positive definite functions and kernels.

Definition 9.1

Let \(X\) be a set and \(\mathcal {F}\) be a complex Hilbert space.

  1. (a)

    A function \(K : X \times X \rightarrow B(\mathcal {F})\) is called a \(B(\mathcal {F})\) -valued kernel. A \(B(\mathcal {F})\)-valued kernel \(K\) on \(X\) is said to be positive definite if, for every finite sequence \((x_1, v_1), \ldots , (x_n,v_n)\) in \(X \times \mathcal {F}\),

    $$\begin{aligned} \sum _{j,k = 1}^n \langle K(x_j, x_k)v_k, v_j \rangle \ge 0. \end{aligned}$$
  2. (b)

    If \((S,*)\) is an involutive semigroup, then a function \(\varphi : S \rightarrow B(\mathcal {F})\) is called positive definite if the kernel \(K_\varphi (s,t) := \varphi (st^*)\) is positive definite.

Positive definite kernels can be characterized as those for which there exists a Hilbert space \(\mathcal {H}\) and a function \(\gamma : X \rightarrow B(\mathcal {H},\mathcal {F})\) such that

$$\begin{aligned} K(x,y) = \gamma (x)\gamma (y)^* \quad \text{ for } \quad x,y \in X \end{aligned}$$
(36)

(cf. [26, Thm. I.1.4]). Here one may assume that the vectors \(\gamma (x)^*v\), \(x \in X, v \in \mathcal {F}\), span a dense subspace of \(\mathcal {H}\). If this is the case, then the pair \((\gamma ,\mathcal {H})\) is called a realization of \(K\). The map \(\Phi : \mathcal {H}\rightarrow \mathcal {F}^X, \Phi (v)(x) := \gamma (x)v\), then realizes \(\mathcal {H}\) as a Hilbert subspace of \(\mathcal {F}^X\) with continuous point evaluations \(\mathop {\mathrm{ev}}\nolimits _x : \mathcal {H}\rightarrow \mathcal {F}, f \mapsto f(x)\). Then \(\Phi (\mathcal {H})\) is the unique Hilbert space in \(\mathcal {F}^X\) with continuous point evaluations \(\mathop {\mathrm{ev}}\nolimits _x\), for which \(K(x,y) = \mathop {\mathrm{ev}}\nolimits _x \mathop {\mathrm{ev}}\nolimits _y^*\) for \(x,y \in X\). We write \(\mathcal {H}_K \subseteq \mathcal {F}^X\) for this subspace and call it the reproducing kernel Hilbert space with kernel \(K\).

Example 9.2

(Vector-valued GNS construction) (cf. [26, Sect. 3.1]) Let \((\pi , \mathcal {H})\) be a representation of the unital involutive semigroup \((S,*)\), \(\mathcal {F}\subseteq \mathcal {H}\) be a closed subspace for which \(\pi (S)\mathcal {F}\) is total in \(\mathcal {H}\) and \(P : \mathcal {H}\rightarrow \mathcal {F}\) denote the orthogonal projection. Then \(\varphi (s) := P\pi (s)P^*\) is a \(B(\mathcal {F})\)-valued positive definite function on \(S\) with \(\varphi (\mathbf {1}) = \mathbf {1}_\mathcal {F}\) because \(\gamma (s) := P\pi (s) \in B(\mathcal {H},\mathcal {F})\) satisfies

$$\begin{aligned} \gamma (s)\gamma (t)^* = P\pi (st^*)P^* = \varphi (st^*). \end{aligned}$$

The map

$$\begin{aligned} \Phi : \mathcal {H}\rightarrow \mathcal {F}^S, \quad \Phi (v)(s) = \gamma (s)v = P \pi (s)v \end{aligned}$$

is an \(S\)-equivariant realization of \(\mathcal {H}\) as the reproducing kernel space \(\mathcal {H}_\varphi \subseteq \mathcal {F}^S\), on which \(S\) acts by right translation, i.e., \((\pi _\varphi (s)f)(t) = f(ts)\).

Conversely, let \(S\) be a unital involutive semigroup and \(\varphi : S \rightarrow B(\mathcal {F})\) be a positive definite function with \(\varphi (\mathbf {1}) = \mathbf {1}_\mathcal {F}\). Write \(\mathcal {H}_\varphi \subseteq \mathcal {F}^S\) for the corresponding reproducing kernel space and \(\mathcal {H}_\varphi ^0\) for the dense subspace spanned by \(\mathop {\mathrm{ev}}\nolimits _s^*v, s \in S, v \in \mathcal {F}\). Then \((\pi _\varphi (s)f)(t) := f(ts)\) defines a \(*\)-representation of \(S\) on \(\mathcal {H}_\varphi ^0\). We say that \(\varphi \) is exponentially bounded if all operators \(\pi _\varphi (s)\) are bounded, so that we actually obtain a representation of \(S\) by bounded operators on \(\mathcal {H}_\varphi \). As \(\mathbf {1}_\mathcal {F}= \varphi (\mathbf {1}) = \mathop {\mathrm{ev}}\nolimits _\mathbf {1} \mathop {\mathrm{ev}}\nolimits _\mathbf {1}^*\), the map \(\mathop {\mathrm{ev}}\nolimits _\mathbf {1}^* : \mathcal {F}\rightarrow \mathcal {H}\) is an isometric inclusion, so that we may identify \(\mathcal {F}\) with a subspace of \(\mathcal {H}\). Then \(\mathop {\mathrm{ev}}\nolimits _\mathbf {1} : \mathcal {H}\rightarrow \mathcal {F}\) corresponds to the orthogonal projection onto \(\mathcal {F}\) and \(\mathop {\mathrm{ev}}\nolimits _\mathbf {1} \circ \pi _\varphi (s) = \mathop {\mathrm{ev}}\nolimits _s\) leads to

$$\begin{aligned} \varphi (s) = \mathop {\mathrm{ev}}\nolimits _s \mathop {\mathrm{ev}}\nolimits _\mathbf {1}^* =\mathop {\mathrm{ev}}\nolimits _\mathbf {1} \pi _\varphi (s) \mathop {\mathrm{ev}}\nolimits _\mathbf {1}^*. \end{aligned}$$
(37)

If \(S = G\) is a group with \(s^* = s^{-1}\), then \(\varphi \) is always exponentially bounded and the representation \((\pi _\varphi , \mathcal {H}_\varphi )\) is unitary.

Lemma 9.3

Let \((S,*)\) be a unital involutive semigroup and \(\varphi : S \rightarrow B(\mathcal {F})\) be a positive definite function with \(\varphi (\mathbf {1}) = \mathbf {1}\). We write \((\pi _\varphi , \mathcal {H}_\varphi )\) for the representation on the corresponding reproducing kernel Hilbert space \(\mathcal {H}_\varphi \subseteq \mathcal {F}^S\) by \((\pi _\varphi (s)f)(t) := f(ts)\). Then the inclusion

$$\begin{aligned} \iota : \mathcal {F}\rightarrow \mathcal {H}_\varphi , \quad \iota (v)(s) := \varphi (s)v \end{aligned}$$

is surjective if and only if \(\varphi \) is multiplicative, i.e., a representation.

Proof

If \(\varphi \) is multiplicative, then \((\pi (s) \iota (v))(t) = \varphi (ts)v = \varphi (t)\varphi (s)v \in \iota (\mathcal {F})\). Therefore the \(S\)-cyclic subspace \(\iota (\mathcal {F})\) is invariant, which implies that \(\iota \) is surjective.

Suppose, conversely, that \(\iota \) is surjective. This means that each \(f \in \mathcal {H}_\varphi \) satisfies \(f(s) = \varphi (s)f(\mathbf {1})\). For \(v \in \mathcal {F}\) and \(t,s \in S\), this leads to

$$\begin{aligned} \varphi (st)v = \pi (t)(\iota (v))(s) = \varphi (s) (\pi (t)\iota (v))(\mathbf {1}) = \varphi (s) \iota (v)(t) = \varphi (s) \varphi (t)v. \end{aligned}$$

Therefore \(\varphi \) is multiplicative.\(\square \)

Remark 9.4

The preceding lemma can also be expressed without referring to positive definite functions and the corresponding reproducing kernel space. In this context it asserts the following. Let \(\pi : S \rightarrow B(\mathcal {H})\) be a \(*\)-representation of a unital involutive semigroup \((S,*)\), \(\mathcal {F}\subseteq \mathcal {H}\) be a closed cyclic subspace and \(P : \mathcal {H}\rightarrow \mathcal {F}\) the orthogonal projection. Then the function

$$\begin{aligned} \varphi : S \rightarrow B(\mathcal {F}), \quad \varphi (s) := P \pi (s)P^* \end{aligned}$$

is multiplicative if and only if \(\mathcal {F}= \mathcal {H}\).

The following lemma is an abstraction of [6, Thm. 6.2.2].

Lemma 9.5

Let \(V\) be a vector spaces over \({\mathbb K}\in \{{\mathbb R},{\mathbb C}\}\) and \(\beta : V \times V\rightarrow {\mathbb K}\) be a hermitian form on \(V\) (for \({\mathbb K}= {\mathbb R}\) this means that it is symmetric and bilinear). Then the kernel \(e^\beta \) is positive definite if and only if \(\beta \) is positive semidefinite.

Proof

If \(\beta \) is positive semidefinite, then \(\beta \) defines a positive definite kernel on \(V \times V\). Hence the kernels \(\beta ^n(x,y) := \beta (x,y)^n\) are also positive definite, see [26, Rem. I.17(b)], and therefore \(e^\beta = \sum _{n = 0}^\infty \frac{\beta ^n}{n!}\) is positive definite.

If, conversely, \(e^\beta \) is a positive definite kernel, let \(x_1,\ldots ,x_N\in V\) and \(c_1,\ldots ,c_N\in {\mathbb K}\). For \(t\in {\mathbb R}\) define

$$\begin{aligned} y_1=tx_1,\ldots , y_N=t x_N,\quad y_{N+1}=0,\ldots ,y_{2N}=0 \end{aligned}$$

and

$$\begin{aligned} d_1=c_1,\ldots , d_N=c_N,\quad d_{N+1}=-c_1,\ldots ,d_{2N}=-c_N. \end{aligned}$$

Then

$$\begin{aligned} 0 \le \sum _{i,j=1}^{2N}e^{\beta (y_i,y_j)}d_i\overline{d_j}= \sum _{i,j=1}^{N}(e^{t^2\beta (x_i,x_j)}-1)c_i\overline{c_j} \end{aligned}$$

and letting \(t\) tend to zero after dividing by \(t^2\) yields the claim.\(\square \)

Appendix D: Distribution Vectors and Tempered Measures

In this appendix we take a closer look at distribution vectors for representations of vector spaces by multiplication operators on \(L^2\)-spaces.

1.1 D.1 Representations of Cones

Let \(V\) be a finite-dimensional real vector space, \(\tau : V \rightarrow V\) be a linear involution and \(\Omega \subseteq V\) be an open convex cone invariant under the involution \(v \mapsto v^\sharp = -\tau (v)\). The cone

$$\begin{aligned} \widehat{\Omega } = \{ \alpha \in V_{\mathbb C}^* : (\forall x = x^\sharp \in \Omega )\ \alpha (x) \ge 0\} \end{aligned}$$

parametrizes the bounded characters of \(\Omega \) by assigning to \(\alpha \in \widehat{\Omega }\) the character \(e_\alpha (v) := e^{-\alpha (v)}\).

Any \(\sigma \)-finite measure \(\mu \) on \(\widehat{\Omega }\) defines a contraction representation of \(\Omega \) on \(\mathcal {H}:= L^2(\widehat{\Omega }, \mu )\) by

$$\begin{aligned} (\pi _\mu (v)f)(\alpha ) := e^{-\alpha (v)} f(\alpha ) \quad \text{ for } \quad v \in \Omega , \alpha \in \widehat{\Omega }. \end{aligned}$$
(38)

On the subspace \(V^c := i V^{-\tau } \oplus V^{\tau } \subseteq V_{\mathbb C}\) formula (38) defines a unitary representation \(\pi _\mu ^c\) because every \(\alpha \in \widehat{\Omega }\) is purely imaginary on \(V^c\).

The following lemma is a supplement to the generalized Bochner–Schwartz Theorem from [28] because it tells us which measures \(\mu \) on \(\widehat{\Omega }\) actually have Fourier–Laplace transforms defining distributions on \(\Omega \).

Lemma 10.1

Let \(V\) be a finite-dimensional vector space, \(\tau \in \mathop {\mathrm{GL}}\nolimits (V)\) an involution and \(\Omega \subseteq V\) be an open convex cone invariant under the involution \(v \mapsto v^\sharp = - \tau (v)\). Further, let \(\mu \) be a positive Borel measure on \(\widehat{\Omega }\) such that, for every \(\varphi \in C^\infty _c(\Omega )\) the Fourier–Laplace transform \(\widehat{\varphi }(\alpha ) := \int _{\Omega } \varphi (x) e^{-\alpha (x)}\, dx\) is \(\mu \)-integrable on \(\widehat{\Omega }\). Then \(\widehat{\mu }(\varphi ) := \mu (\overline{\widehat{\varphi }})\) defines a distribution on \(\Omega \). In the special case \(V = \Omega \) and \(\tau = \mathop {\mathrm{id}}\nolimits _V\), the measure \(\mu \) is tempered.

Proof

We consider the linear functional

$$\begin{aligned} E : C^\infty _c(\Omega ) \rightarrow {\mathbb C}, \quad E(\varphi ) := \mu (\widehat{\varphi }). \end{aligned}$$

Then, for every test function \(\psi \) and \(\psi ^\sharp := \psi ^* \circ \tau \), the regularized functional \(E_\psi (\varphi ) := E(\psi * \varphi * \psi ^\sharp ) = \mu (|\widehat{\psi }|^2 \widehat{\varphi })\) is continuous because it is the Fourier transform of the bounded measure \(|\widehat{\psi }|^2 \mu \).

Now we consider a sequence \(0 \le \psi _n \in C^\infty _c(\Omega )\) with \(\widehat{\psi }_n(0) = \int _{\Omega } \psi _n(x)\, d\mu _V(x) = 1\) and \(\mathop {\mathrm{supp}}\nolimits (\psi _n) \rightarrow \{0\}\) (a \(\delta \)-sequence). Then

$$\begin{aligned} \widehat{\psi }_n(\alpha ) = \int _{V} \psi _n(x) e^{-\alpha (x)}\, d\mu _V(x) \rightarrow e^{-\alpha (0)} = 1 \end{aligned}$$

holds pointwise. We also have \(\Vert \widehat{\psi }_n\Vert _\infty \le \Vert \psi _n\Vert _1 = 1\) for the sup-norm of \(\widehat{\Omega }\), so that Dominated Convergence implies that \(E_{\psi _n} \rightarrow E\) pointwise on \(C^\infty _c(V)\).

The Uniform Boundedness Theorem, applied to the restriction of \(E_{\psi _n}\) to the Fréchet spaces \(C^\infty _K(\Omega )\), \(K \subseteq \Omega \) compact, now implies that \(E\) is continuous on every subspace \(C^\infty _K(\Omega )\), hence continuous on \(C^\infty _c(\Omega )\). This means that \(E\) is a positive definite distribution.

For \(V = \Omega \) and \(\tau = \mathop {\mathrm{id}}\nolimits _V\), the Bochner–Schwartz Theorem ([31, Thm. XVIII, §VII.9]) further implies that \(\mu \) is tempered.\(\square \)

Remark 10.2

Suppose that \(D \in C^{-\infty }(\Omega )\) is a positive definite distribution, so that we can use the generalized Bochner–Schwartz Theorem ([28, Thm. 4.11]) to write it as the Fourier–Laplace transform of a measure \(\mu \) in the sense of

$$\begin{aligned} D(\varphi ) = \int _{\widehat{\Omega }} \overline{\widehat{\varphi }}\, d\mu \quad \text{ for } \quad \varphi \in C^\infty _c(\Omega ), \quad \text{ where } \quad \widehat{\varphi }(\alpha ) := \int _{\Omega } \varphi (x) e^{-\alpha (x)}\, dx \end{aligned}$$

is the Fourier–Laplace transform of \(\varphi \). Then we obtain an isomorphism

$$\begin{aligned} \Gamma : L^2(\widehat{\Omega }, \mu ) \rightarrow \mathcal {H}_D \subseteq C^{-\infty }(\Omega ), \quad \Gamma (f)(\varphi ) = \langle f, \widehat{\varphi } \rangle , \end{aligned}$$

under which \(D\) corresponds to the constant function \(1\). In particular, \(\widehat{\varphi } \in L^2(\widehat{\Omega }, \mu )\) for \(\varphi \in C^\infty _c(\Omega )\).

Remark 10.3

In the special case \(\tau = -\mathop {\mathrm{id}}\nolimits _V\) the integrability of the functions \(\widehat{\varphi }\) is equivalent to the existence of the Laplace transform

$$\begin{aligned} \widehat{\mathcal {L}}(\mu )(x) = \int _{\widehat{\Omega }} e^{-\alpha (x)}\, d\mu (\alpha ) \quad \text{ for } \quad x \in \Omega . \end{aligned}$$

This follows from

$$\begin{aligned} \int _{\widehat{\Omega }} \widehat{\varphi }\, d\mu = \int _\Omega \mathcal {L}(\mu )(x) \varphi (x)\, dx \end{aligned}$$

because \(\mathcal {L}(\mu )\) is continuous whenever it exists (by the Dominated Convergence Theorem).

Example 10.4

The constant function \(1\) on \(\widehat{\Omega }\) need not be a distribution vector for the representation \(\pi _\mu ^c\) of \(V^c\), even if \(\widehat{\mu }\) defines a distribution on \(\Omega \). A simple example is provided for \(V = {\mathbb R}\), \(\tau = -\mathop {\mathrm{id}}\nolimits _V\) and \(\Omega = {\mathbb R}_+\) by the measure \(d\mu (t) := e^{\sqrt{t}} \, dt\) on \(\widehat{\Omega } = [0,\infty [\). Then the Laplace transform \(\mathcal {L}(\mu )\) is defined on \(\Omega \), so that each function \(\widehat{\varphi }\), \(\varphi \in C^\infty _c(\Omega )\), is integrable, but \(\mathcal {L}(\mu )\) does not extend to a distribution on \({\mathbb R}\) because \(\int _1^\infty x^{-\ell }\, d\mu (x) = \infty \) for every \(\ell \in {\mathbb N}\) (cf. Lemma 7.5).

Example 10.5

We now turn to the special case \(V = {\mathbb R}^d\), \(\Omega = {\mathbb R}^d_+\) and \(\tau (x) = (-x_0, \mathbf {x})\) for \(x = (x_0, \mathbf {x})\).

(a) Identifying \(\widehat{\Omega }\) with \([0,\infty [ \times {\mathbb R}^{d-1}\), the contraction representation of \(\Omega \) on \(L^2(\widehat{\Omega },\mu )\) is given by

$$\begin{aligned} (\pi _\mu (x)f)(p) = e^{-x_0 p_0} e^{-i \mathbf {x}\mathbf {p}} f(p) \quad \text{ for } \quad p = (p_0, \mathbf {p}). \end{aligned}$$

Here we use the embedding

$$\begin{aligned} \iota : [0,\infty [ \times {\mathbb R}^{d-1} \rightarrow \widehat{\Omega } \subseteq V_{\mathbb C}^* \cong {\mathbb C}^d, \quad \iota (p_0, \mathbf {p})(x) := p_0 x_0 + i \mathbf {x}\mathbf {p}. \end{aligned}$$

The corresponding unitary representation of \(V^c = i{\mathbb R}\oplus {\mathbb R}^{d-1} \cong {\mathbb R}^d\) is given by

$$\begin{aligned} (\pi _\mu ^c(x)f)(p) = e^{-i xp} f(p) = e^{-ix_0 p_0} e^{-i \mathbf {x}\mathbf {p}} f(p) \quad \text{ for } \quad x = (x_0, \mathbf {x}), p = (p_0,\mathbf {p}). \end{aligned}$$

In view of Corollary 10.8, the function \(1\) is a distribution vector for \(\pi ^c_\mu \) if and only if \(\mu \) is a tempered measure on \({\mathbb R}^d\).

(b) Note that \(\pi _\mu (x)1 \in \mathcal {H}\) for every \(x \in \Omega \) is equivalent to the existence of the Fourier–Laplace transform of \(\mu \) as a function on the open half space \(\Omega \). In view of \(|e^{-\alpha (x)}| = e^{-p_0 x_0}\) for \(\alpha = (p_0, i \mathbf {p})\), this in turn is equivalent to the square-integrability of all functions \(e_x(\alpha ) := e^{-\alpha (x)}\), \(x \in \Omega \), on \(\widehat{\Omega }\). This in turn is equivalent to the integrability of all functions \(e^{-t p_0}\), \(t > 0\). If \(\mu \) is supported by the forward light cone \(\overline{V_+} \subseteq {\mathbb R}^d\), then the temperedness of \(\mu \) implies the existence of its Laplace transform on \(\Omega \).

The following lemma makes the condition from Lemma 10.1 more explicit in terms of the product decomposition \(\Omega = {\mathbb R}_+ \times {\mathbb R}^{d-1}\).

Lemma 10.6

For \(V = {\mathbb R}^d\), \(\Omega = {\mathbb R}^d_+\) and \(\tau (x) = (-x_0, \mathbf {x})\), a positive Borel measure \(\mu \) on \(\widehat{\Omega }\) defines a distribution \(\widehat{\mu }\) on the open half space \(\Omega = {\mathbb R}^d_+\) if and only if all functions \(e_t(p_0,\mathbf {p}) := e^{-tp_0}\), \(t > 0\), on \(\widehat{\Omega }\) are distribution vectors for the representation \(\pi _\mu ^c\vert _{{\mathbb R}^{d-1}}\) of the subgroup \(\{0\} \times {\mathbb R}^{d-1}\) on \(L^2(\widehat{\Omega }, \mu )\). This in turn is equivalent to the temperedness of the measures \(\nu _t := \mathop {\mathrm{pr}}\nolimits _*(e_t \mu )\) on \({\mathbb R}^{d-1}\), where \(\mathop {\mathrm{pr}}\nolimits (x_0,\mathbf {x}) = \mathbf {x}\).

Proof

In view of Lemma 10.1, \(\widehat{\mu } \in C^{-\infty }(\Omega )\) is equivalent to the integrability of all function \(\widehat{\psi }\), \(\psi \in C^\infty _c(\Omega )\). Consider such a function of the form \(\psi (x_0,\mathbf {x}) = \psi _0(x) \psi _1(\mathbf {x})\) with \(\psi _0 \in C^\infty _c({\mathbb R})\) and \(\psi _1 \in C^\infty _c({\mathbb R}^{d-1})\). Then \(\widehat{\psi }(p_0,\mathbf {p}) = \widehat{\psi }_0(p_0) \widehat{\psi }_1(\mathbf {p})\) is \(\mu \)-integrable on \(\widehat{\Omega }\). If \(\psi _0 \ge 0\) is supported in the interval \([a,b] \subseteq {\mathbb R}_+\), then

$$\begin{aligned} \widehat{\psi }_0(p_0) = \int _0^\infty e^{-p_0 x_0}\psi _0(x_0)\, dx_0 \ge e^{-p_0 b} \int _0^\infty \psi _0(x_0)\, dx_0 \end{aligned}$$

implies that \(e^{-p_0 b} \widehat{\psi }_1(\mathbf {p}) \in L^2(\widehat{\Omega }, \mu )\) for every \(b > 0\) and \(\psi _1 \in C^\infty _c({\mathbb R}^{d-1})\). According to Corollary 10.8, this means that, for every \(b > 0\), the function \(e^{-b p_0}\) is a distribution vector for the representation of \({\mathbb R}^{d-1}\) on \(L^2(\widehat{\Omega }, \mu )\). In view of Lemma 10.7, this is equivalent to the temperedness of the measure \(\mathop {\mathrm{pr}}\nolimits _*(e^{-b p_0} \mu )\) on \({\mathbb R}^{d-1}\).

Suppose, conversely, that this condition is satisfied and let \(\varphi \in C^\infty _c(\Omega )\) and \(0 < a < b\) with

$$\begin{aligned} \mathop {\mathrm{supp}}\nolimits (\varphi ) \subseteq [a,b] \times {\mathbb R}^{d-1}. \end{aligned}$$

For \(\varphi _{x_0}(\mathbf {x}) := \varphi (x_0,\mathbf {x})\) and

$$\begin{aligned} \widehat{\varphi }(p_0,\mathbf {p}) = \int _0^\infty \int _{{\mathbb R}^{d-1}} \varphi (x_0, \mathbf {x}) e^{-p_0 x_0} e^{-i \mathbf {p}\mathbf {x}}\, d\mathbf {x}dx_0 = \int _0^\infty \widehat{\varphi }_{x_0}(\mathbf {p}) e^{-p_0 x_0} dx_0, \end{aligned}$$

we obtain the estimate \(|\widehat{\varphi }(p_0,\mathbf {p})| \le e^{-p_0 a} \int _a^b |\widehat{\varphi }_{x_0}(\mathbf {p})|\, dx_0.\) Therefore

$$\begin{aligned} \int _{\widehat{\Omega }} |\widehat{\varphi }(p_0,\mathbf {p})| \, d\mu (p)&\le \int _{\widehat{\Omega }} e^{-p_0 a} \int _a^b |\widehat{\varphi }_{x_0}(\mathbf {p})|\, dx_0 \, d\mu (p)\\&= \int _{{\mathbb R}^{d-1}} \int _a^b |\widehat{\varphi }_{x_0}(\mathbf {p})|\, dx_0 \, d p_*(e^{-p_0 a}\mu )(\mathbf {p}) \\&= \int _a^b \int _{{\mathbb R}^{d-1}} |\widehat{\varphi }_{x_0}(\mathbf {p})| \, d p_*(e^{-p_0 a}\mu )(\mathbf {p})\, dx_0. \end{aligned}$$

To see that this integral is finite, we first observe that the function \({\mathbb R}_+ \rightarrow C^\infty _c({\mathbb R}^{d-1}), x_0 \mapsto \varphi _{x_0}\) is continuous. Since the measure \(\nu := p_*(e^{-p_0 a}\mu )\) on \({\mathbb R}^{d-1}\) is tempered, the map \(\gamma : C^\infty _c({\mathbb R}^{d-1}) \rightarrow L^1({\mathbb R}^{d-1},\nu ), \psi \mapsto \widehat{\psi }\) is continuous, and hence

$$\begin{aligned} \int _{\widehat{\Omega }} |\widehat{\varphi }(p_0,\mathbf {p})| \, d\mu (p) \le \int _a^b \Vert \gamma (\varphi _{x_0})\Vert _1\, dx_0 < \infty . \end{aligned}$$

\(\square \)

1.2 D.2 Representations of Vector Groups

We have seen above that the temperedness of a measure on \({\mathbb R}^d\) is closely connected with distribution vectors. The following lemma provides a useful criterion to check this condition. Its main ingredient is the automatic continuity result from Lemma 10.1.

Lemma 10.7

Let \((X,{\mathfrak {S}},\mu )\) be a measure space. We write \(M(X,{\mathbb C})\) for the vector space of measurable functions \(X \rightarrow {\mathbb C}\). For \(H_j \in M(X,{\mathbb R})\), \(j = 1,\ldots , d\), we consider the continuous unitary representation of \({\mathbb R}^d\) on \(L^2(X,\mu )\), given by

Put \(R := \sqrt{\sum _{j = 1}^d H_j^2}.\) Then

$$\begin{aligned} \mathcal {H}^{-\infty } \cong \Big \{h \in M(X,{\mathbb C}) : (\exists n \in {\mathbb N}) \ \Vert (1+ R^{2})^{-n} f \Vert _2 < \infty \Big \}, \end{aligned}$$

where the pairing \(\mathcal {H}^{-\infty } \times \mathcal {H}^\infty \rightarrow {\mathbb C}\) is given by \((h,f) \mapsto \int _X h \overline{f} \, d\mu \).

The following assertions are equivalent:

  1. (i)

    The constant function \(1\) is a distribution vector.

  2. (ii)

    For the measurable map \(\eta := (H_1, \ldots , H_d) : X \rightarrow {\mathbb R}^d\), the measure \(\eta _*\mu \) on \({\mathbb R}^d\) is tempered.

  3. (iii)

    \(\widehat{\varphi } \circ \eta \in L^2(X,\mu )\) for every \(\varphi \in C^\infty _c({\mathbb R}^d)\).

If these conditions are satisfied, then the corresponding distribution on \({\mathbb R}^d\) is given by the Fourier transform of \(\eta _*\mu \).

Proof

Put \(\mathcal {H}:= L^2(X,\mu )\).

(a) For \(d = 1\) and the measurable function \(H : X \rightarrow {\mathbb R}\), the subspace of smooth vectors is

$$\begin{aligned} \mathcal {H}^\infty = \{ f \in \mathcal {H}: (\forall n \in {\mathbb N}) \ \Vert H^n f \Vert _2 < \infty \} \end{aligned}$$

and the natural Fréchet topology on this space is defined by the seminorms \(p_n(f) := \Vert (1 + H^2)^n f\Vert _2\). From

$$\begin{aligned} \Vert fh\Vert _1 = \Vert f(1+ H^{2})^n (1 + H^{2})^{-n}h\Vert _1 \le p_n(f) \Vert (1 + H^{2})^{-n}h\Vert _2 \end{aligned}$$

it follows that the space of distribution vectors can be identified with

$$\begin{aligned} \mathcal {H}^{-\infty } = \Big \{ h \in M(X,{\mathbb C}) : (\exists n \in {\mathbb N}) \ \Vert (1+ H^{2})^{-n} f \Vert _2 < \infty \Big \}. \end{aligned}$$

(b) In the general case it follows from (a) that the subspace of smooth vectors is

$$\begin{aligned} \mathcal {H}^\infty = \{ f \in \mathcal {H}: (\forall \mathbf {n}\in {\mathbb N}_0^d) \ \Vert H_1^{n_1} \cdots H_d^{n_d}f \Vert _2 < \infty \} \end{aligned}$$

and the natural Fréchet topology on this space is defined by the seminorms \(p_n(f) := \Vert (1 + R^{2})^n f\Vert _2.\) Therefore

$$\begin{aligned} \mathcal {H}^{-\infty } \cong \Big \{ h \in M(X,{\mathbb C}) : (\exists n \in {\mathbb N}) \ \Vert (1+ R^{2})^{-n} f \Vert _2 < \infty \Big \}. \end{aligned}$$

In particular, \(1\) is a distribution vector if and only if some function \((1 + R^{2})^{-n}\) is integrable, which is equivalent to the integrability of the function \((1 + \Vert x\Vert ^2)^{-n}\) with respect to the measure \(\eta _*\mu \) on \({\mathbb R}^d\), and this means that \(\eta _*\mu \) is tempered. This proves the equivalence of (i) and (ii).

(ii) \(\Rightarrow \) (iii) The integrated representation of \(L^1({\mathbb R}^d)\) on \(\mathcal {H}\) is given by \(U(\varphi ) f= (\widehat{\varphi } \circ (-\eta )) \cdot f\) and the integrated representation \(U^{-\infty }\) on \(\mathcal {H}^{-\infty }\) is given by the same formula. Hence (iii) follows from \(\widehat{\varphi } \circ (-\eta ) = U^{-\infty }(\varphi ) 1 \in L^2(X,\mu )\) for \(\varphi \in C^\infty _c({\mathbb R}^d)\).

(iii) First we use the Dixmier–Malliavin Theorem ([3, Thm. 3.1]) to see that every test function on \({\mathbb R}^d\) is a finite sum of convolution products \(\varphi * \psi \). Therefore (iii) implies that, for every \(\varphi \in C^\infty _c({\mathbb R}^d)\), the function \(\widehat{\varphi } \circ \eta \) is \(\mu \)-integrable, resp., \(\widehat{\varphi }\) is integrable w.r.t. \(\eta _*\mu \). Now Lemma 10.1, applied to \(\Omega = {\mathbb R}^d\) and \(\tau = \mathop {\mathrm{id}}\nolimits _{{\mathbb R}^d}\), implies that \(\eta _*\mu \) is tempered.

For the last assertion, we simply calculate:

$$\begin{aligned} D(\varphi ) := \langle 1, U^{-\infty }(\varphi ) 1 \rangle = \int _X \overline{\widetilde{\varphi }(\eta (x))}\, d\mu (x) = \int _{{\mathbb R}^d} \overline{\widetilde{\varphi }(x)}\, d(\eta _*\mu )(x) = (\eta _*\mu )^{\wedge }{}(\varphi ). \end{aligned}$$

\(\square \)

Applying Lemma 10.7 to \(X = V^* \cong {\mathbb R}^d\) and \(\eta (x) = -x\), we obtain:

Corollary 10.8

Let \(V\) be a finite-dimensional real vector space. For a positive Borel measure on \(V^*\), the following are equivalent:

  1. (i)

    \(\mu \) is tempered.

  2. (ii)

    \(1 \in L^2(V^*,\mu )^{-\infty }\) for the representation \((U_v f)(\alpha ) = e^{-i\alpha (v)} f(\alpha )\).

  3. (iii)

    \(\widehat{\varphi } \in L^2(V^*,\mu )\) for every \(\varphi \in C^\infty _c(V)\).

Remark 10.9

The first part of Lemma 10.7 remains true for vector-valued \(L^2\)-spaces \(L^2(X,Q,\mu ;\mathcal {H})\), where \(Q: X \rightarrow \mathop {\mathrm{Herm}}\nolimits _+(\mathcal {H})\) is an operator-valued density,

$$\begin{aligned} \langle f,g \rangle \!=\! \int _X \langle Q(x) f(x), g(x)\rangle \, d\mu (x) \quad \text{ and } \quad U_{\mathbf {t}}(f) \!:=\! e^{i \sum _{j = 1}^d t_j H_j} f, \, \mathbf {t}\!=\! (t_1,\ldots , t_d). \end{aligned}$$

Remark 10.10

(Projectable measures and “time-zero” subspaces)

(a) Let \(\mu \) be a measure on \({\mathbb R}^d\) and \(\mathop {\mathrm{pr}}\nolimits : {\mathbb R}^d \rightarrow {\mathbb R}^{d-1}, p = (p_0, \mathbf {p}) \mapsto \mathbf {p}\) be the projection. We call \(\mu \) projectable if \(\mathop {\mathrm{pr}}\nolimits _*\mu \) is tempered. In view of Corollary 10.8, this is equivalent to the integrability of all Fourier transforms \(\widehat{\varphi } \circ \mathop {\mathrm{pr}}\nolimits \), \(\varphi \in C^\infty _c({\mathbb R}^{d-1})\), which in turn means that \(1\) is a distribution vector for the canonical action of \({\mathbb R}^{d-1}\) on \(L^2({\mathbb R}^d,\mu )\). From Lemma 10.7 we now see that \(1\) is also a distribution vector for \({\mathbb R}^d\), which means that \(\mu \) is tempered.

In particular, we obtain an isometric embedding

$$\begin{aligned} \mathop {\mathrm{pr}}\nolimits ^* : L^2({\mathbb R}^{d-1}, \mathop {\mathrm{pr}}\nolimits _*\mu ) \hookrightarrow L^2({\mathbb R}^d,\mu ), \quad f \mapsto f \circ \mathop {\mathrm{pr}}\nolimits \end{aligned}$$

which maps onto the subspace of those functions not depending on the first argument \(p_0\). Accordingly their Fourier transform is supported in the \(x_0 = 0\) hyperplane. Therefore \(L^2({\mathbb R}^{d-1}, \mathop {\mathrm{pr}}\nolimits _*\mu )\) is also called the time-zero subspace of \(L^2({\mathbb R}^d,\mu )\).

(b) For the distribution \(D := \widehat{\mu }\), the projectability of \(\mu \) means that \(D\) can be restricted to the subspace \({\mathbb R}^{d-1}\) by

$$\begin{aligned} D_r(\varphi ) := \int _{{\mathbb R}^d} \overline{\widehat{\varphi }(\mathbf {p})}\, d\mu (p) = (\mathop {\mathrm{pr}}\nolimits _*\mu )^{\wedge }{}(\varphi ), \end{aligned}$$

respectively extended to the space \(C^\infty _c({\mathbb R}^{d-1})\). The scalar product \(\langle \varphi , \psi \rangle := D_r(\varphi ^* * \psi )\) on \(C^\infty _c({\mathbb R}^{d-1})\) leads to \(\mathcal {H}_{D_r} \cong L^2({\mathbb R}^{d-1}, \mathop {\mathrm{pr}}\nolimits _*\mu )\), which is naturally realized as a subspace of \(L^2({\mathbb R}^d,\mu )\). Accordingly, we see that, as a space of distributions on \({\mathbb R}^d\), we may consider \(\mathcal {H}_{D_r}\) as a subspace of \(\mathcal {H}_D\).

(c) The Lorentz invariant measures \(\mu _m\) are interesting examples of measures \(\mu \) on \({\mathbb R}^d\) for which the inclusion \(L^2({\mathbb R}^{d-1}, \mathop {\mathrm{pr}}\nolimits _*\mu ) \hookrightarrow L^2({\mathbb R}^{d}, \mu )\) is actually surjective.

(d) For \(d = 1\), a measure \(\mu \) is projectable if and only if it is finite, i.e., if its Fourier transform \(\widehat{\mu }\) is a continuous function. Then \(L^2({\mathbb R}^0, \mathop {\mathrm{pr}}\nolimits _*\mu )\) is the subspace of constant functions in \(L^2({\mathbb R},\mu )\).

(e) A similar picture prevails for operator-valued measures, distributions and \(L^2\)-spaces such as those discussed in Sect. 5. In this context the subspace \(\mathcal {H}\subseteq L^2({\mathbb R},Q;\mathcal {H})\) of constant functions plays the role of the time-zero subspace. This matches well with Proposition 5.2(iii), where \({\mathcal {E}}_0 \cong \mathcal {H}\) is identified as consisting of those functions whose Fourier transform is supported in \(\{0\}\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Neeb, KH., Ólafsson, G. Reflection Positive One-Parameter Groups and Dilations. Complex Anal. Oper. Theory 9, 653–721 (2015). https://doi.org/10.1007/s11785-014-0402-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11785-014-0402-2

Keywords

Navigation