Ballistic Transport in Periodic and Random Media

  • Chapter
  • First Online:
From Complex Analysis to Operator Theory: A Panorama

Part of the book series: Operator Theory: Advances and Applications ((OT,volume 291))

Abstract

We prove ballistic transport of all orders, that is, \(\Vert x^m\mathrm{e} ^{-\mathrm{i} tH}\psi \Vert \asymp t^m\), for the following models: the adjacency matrix on \(\mathbb {Z}^d\), the Laplace operator on \(\mathbb {R}^d\), periodic Schrödinger operators on \(\mathbb {R}^d\), and discrete periodic Schrödinger operators on periodic graphs. In all cases we give the exact expression of the limit of \(\Vert x^m\mathrm{e} ^{-\mathrm{i} tH}\psi \Vert /t^m\) as \(t\to +\infty \). We then move to universal covers of finite graphs (these are infinite trees) and prove ballistic transport in mean when the potential is lifted naturally, giving a periodic model, and when the tree is endowed with random i.i.d. potential, giving an Anderson model. The limiting distributions are then discussed, enriching the transport theory. Some general upper bounds are detailed in the appendix.

Dedicated to the memory of our friend Sergey Naboko

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

Access this chapter

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 139.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Hardcover Book
USD 179.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free ship** worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Similar content being viewed by others

Notes

  1. 1.

    It is known [33] that for \(f\in \mathcal {H}_{\mathrm {ac}}\), (1.2) strengthens to the fact that \(\lim _{t\to +\infty }\Vert \chi _K\mathrm{e} ^{-\mathrm{i} tH}f\Vert =0\). This implies that \(\lim _{t\to +\infty }\Vert \chi _{K^c}\mathrm{e} ^{-\mathrm{i} tH}f\Vert =\lim _{t\to +\infty }\Vert \mathrm{e} ^{-\mathrm{i} tH}f\Vert =\Vert f\Vert \). Taking \(K=\varLambda _r:=\lbrace \lvert x\rvert \leq r\rbrace \), we thus get \(\liminf _{t\to +\infty }\Vert x^m\mathrm{e} ^{-\mathrm{i} tH}f\Vert \geq \liminf _{t\to +\infty }r^m\Vert \chi _{\varLambda _r^c}\mathrm{e} ^{-\mathrm{i} tH}f\Vert = r^m\Vert f\Vert \). As r is arbitrary, this shows \(\lim _{t\to \infty } \Vert x^m\mathrm{e} ^{-\mathrm{i} tH}f\Vert =\infty \).

  2. 2.

    The fact that \(\nabla _{\theta }H(\theta )g=2(D+\theta )g\) for any \(g\in L^2(\mathbb {T}^d)\) is clear by definition of the derivative. Computing \(\nabla _{\theta }\mathrm{e} ^{-\mathrm{i} tH(\theta )}\) however is less clear. This is why we used the spectral decomposition of \(H(\theta )\) in Step 2 to estimate it.

  3. 3.

    Our operators U and \(H(\theta _{\mathfrak {b}})\) differ slightly from those of [21]. Namely, they consider \((\widetilde {U}\psi )_{\theta _{\mathfrak {b}}}(v_n)=\mathrm{e} ^{\mathrm{i} \theta _{\mathfrak {b}}\cdot v_n}(U\psi )_{\theta _{\mathfrak {b}}}(v_n)\), so they obtain instead the fiber operator \(\widetilde {H}(\theta _{\mathfrak {b}})=\mathrm{e} ^{\mathrm{i} \theta _{\mathfrak {b}}\cdot }H(\theta _{\mathfrak {b}})\mathrm{e} ^{-\mathrm{i} \theta _{\mathfrak {b}}\cdot }\), where \((\mathrm{e} ^{\pm \mathrm{i} \theta _{\mathfrak {b}}\cdot }f)(v_k)=\mathrm{e} ^{\pm \mathrm{i} \theta _{\mathfrak {b}}\cdot v_k}f(v_k)\). Since \(H(\theta _{\mathfrak {b}})\) and \(\widetilde {H}(\theta _{\mathfrak {b}})\) are unitarily equivalent, they share the same eigenvalues \(E_n(\theta _{\mathfrak {b}})\), moreover \(\widetilde {P}_n(\theta _{\mathfrak {b}})=\mathrm{e} ^{\mathrm{i} \theta _{\mathfrak {b}}\cdot }P_n(\theta _{\mathfrak {b}})\mathrm{e} ^{-\mathrm{i} \theta _{\mathfrak {b}}\cdot }\). We have avoided the introduction of “bridges” and quotient graphs and used fractional parts instead, which we think is more transparent for our purposes.

  4. 4.

    Some authors replace the last two terms by \(\psi (n-\mathfrak {e}_1+\mathfrak {e}_2) + \psi (n+\mathfrak {e}_1-\mathfrak {e}_2)\), this is just a different shearing convention and slightly changes the eigenvalue \(E_1(\theta )\).

References

  1. M. Aizenman, S. Warzel, Absolutely continuous spectrum implies ballistic transport for quantum particles in a random potential on tree graphs. J. Math. Phys. 53(9), 095205, 15 (2012)

    Google Scholar 

  2. N. Anantharaman, M. Ingremeau, M. Sabri, B. Winn, Absolutely continuous spectrum for quantum trees. Commun. Math. Phys. 383(1), 537–594 (2021)

    Article  MathSciNet  MATH  Google Scholar 

  3. N. Anantharaman, M. Sabri, Poisson kernel expansions for Schrödinger operators on trees. J. Spectr. Theory 9(1), 243–268 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  4. N. Anantharaman, M. Sabri, Recent results of quantum ergodicity on graphs and further investigation. Ann. Fac. Sci. Toulouse Math. (6) 28(3), 559–592 (2019)

    Google Scholar 

  5. J. Asch, Joachim, A. Knauf, Motion in periodic potentials. Nonlinearity 11(1), 175–200 (1998)

    Google Scholar 

  6. A.-M. Berthier, Spectral theory and wave operators for the Schrödinger equation, in Research Notes in Mathematics, vol. 71 (Pitman (Advanced Publishing Program), Boston, 1982)

    Google Scholar 

  7. P. Billingsley, Probability and measure, in Wiley Series in Probability and Mathematical Statistics, 3rd edn. (Wiley, New York, 1995)

    MATH  Google Scholar 

  8. H. Cartan, Calcul différentiel (Hermann, Paris, 1967) (French)

    MATH  Google Scholar 

  9. D. Damanik, M. Lukic, W. Yessen, Quantum dynamics of periodic and limit-periodic Jacobi and block Jacobi matrices with applications to some quantum many body problems. Commun. Math. Phys. 337(3), 1535–1561 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  10. M. Duerinckx, A. Gloria, C. Shirley, Approximate normal forms via Floquet-Bloch theory: Nehorošev stability for linear waves in quasiperiodic media. Commun. Math. Phys. 383(2), 633–683 (2021)

    Article  MATH  Google Scholar 

  11. J. Fillman, Ballistic transport for limit-periodic Jacobi matrices with applications to quantum many-body problems. Commun. Math. Phys. 350(3), 1275–1297 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  12. J. Fillman, Ballistic transport for periodic Jacobi operators on \(\mathbb {Z}^d\), in From Operator Theory to Orthogonal Polynomials, Combinatorics, and Number Theory. Operator Theory: Advances and Applications, vol. 285 (Birkhäuser, Cham, 2021), pp. 57–68

    Google Scholar 

  13. L. Ge, I. Kachkovskiy, Ballistic Transport for One-dimensional Quasiperiodic Schrödinger Operators (2020). https://arxiv.org/abs/2009.02896

  14. G.R. Grimmett, S. Janson, P.F. Scudo, Weak limits for quantum random walks. Phys. Rev. E 69(2), 026119 (2004)

    Google Scholar 

  15. B.C. Hall, Quantum theory for mathematicians, in Graduate Texts in Mathematics, vol. 267 (Springer, New York, 2013)

    Book  Google Scholar 

  16. R.P. Kanwal, Generalized Functions: Theory and Applications, 3rd edn. (Birkhäuser Boston Inc., Boston, MA, 2004)

    Book  MATH  Google Scholar 

  17. Y. Karpeshina, L. Parnovski, R. Shterenberg, Ballistic transport for Schrödinger operators with quasi-periodic potentials. J. Math. Phys. 62(5), Paper No. 053504, 12 (2021)

    Google Scholar 

  18. T. Kato, Perturbation theory for linear operators, in Classics in Mathematics (Springer, Berlin, 1995). Reprint of the 1980 edition

    Google Scholar 

  19. M. Keller, D. Lenz, S. Warzel, Absolutely continuous spectrum for random operators on trees of finite cone type. J. Anal. Math. 118(1), 363–396 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  20. A. Klein, Spreading of wave packets in the Anderson model on the Bethe lattice. Commun. Math. Phys. 177(3), 755–773 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  21. E. Korotyaev, N. Saburova, Schrödinger operators on periodic discrete graphs. J. Math. Anal. Appl. 420(1), 576–611 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  22. E. Korotyaev, N. Saburova, Spectral estimates for the Schrödinger operator on periodic discrete graphs. Algebra i Analiz 30(4), 61–106 (2018) (Russian), English transl., St. Petersburg Math. J. 30(4), 667–698 (2019)

    Google Scholar 

  23. H. Krüger, Periodic and Limit-periodic Discrete Schrödinger Operators (2011). https://arxiv.org/abs/1108.1584

  24. P. Kuchment, An overview of periodic elliptic operators. Bull. Am. Math. Soc. (N.S.) 53(3), 343–414 (2016)

    Google Scholar 

  25. Y. Last, Quantum dynamics and decompositions of singular continuous spectra. J. Funct. Anal. 142(2), 406–445 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  26. E. Le Masson, M. Sabri, \(L^p\) norms and support of eigenfunctions on graphs. Commun. Math. Phys. 374(1), 211–240 (2020)

    Google Scholar 

  27. C. Radin, B. Simon, Invariant domains for the time-dependent Schrödinger equation. J. Differ. Equ. 29(2), 289–296 (1978)

    Article  MATH  Google Scholar 

  28. M. Reed, B. Simon, Methods of Modern Mathematical Physics. I. Functional Analysis, 2nd edn. (Academic Press, Inc. [Harcourt Brace Jovanovich, Publishers], New York, 1980)

    Google Scholar 

  29. M. Reed, B. Simon, Methods of Modern Mathematical Physics. IV. Analysis of Operators (Academic Press [Harcourt Brace Jovanovich, Publishers], New York-London, 1978)

    Google Scholar 

  30. H. Saigo, H. Sako, Space-homogeneous quantum walks on \(\mathbb {Z}\) from the viewpoint of complex analysis. J. Math. Soc. Japan 72(4), 1201–1237 (2020)

    Google Scholar 

  31. H. Sako, Convergence theorems on multi-dimensional homogeneous quantum walks. Quantum Inf. Process. 20(3), Paper No. 94, 24 (2021)

    Google Scholar 

  32. B. Simon, Functional integration and quantum physics, in Pure and Applied Mathematics, vol. 86 (Academic Press, Inc. [Harcourt Brace Jovanovich, Publishers], New York-London, 1979)

    Google Scholar 

  33. G. Teschl, Mathematical methods in quantum mechanics. With applications to Schrödinger operators, in Graduate Studies in Mathematics, vol. 157, 2nd edn. (American Mathematical Society, Providence, RI, 2014)

    Google Scholar 

  34. C.H. Wilcox, Theory of Bloch waves. J. Analyse Math. 33, 146–167 (1978)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We thank Christopher Shirley and Jake Fillman for pointing out references, and for an interesting discussion reflected in Remark 5.10.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Anne BoutetdeMonvel .

Editor information

Editors and Affiliations

Appendix A: Upper Bounds and Derivatives

Appendix A: Upper Bounds and Derivatives

Here we first prove that the m-moments grow at most like \(t^m\). This already appeared in various forms: for continuous Schrödinger operators see [27] for \(m=1,2\); for discrete Schrödinger operators, upper bounds can be deduced from [1, Appendix B] for general moments but \(\psi =\delta _x\). Using the upper bounds, we then give rigorous proofs of the moment derivative formulas (A.11) and (A.22).

We start with some general remarks on lower bounds.

1.1 A.1 Lower Bounds

On \(L^2(X)\), if \(x_0\in X\) is fixed and \(\lvert x\rvert := d(x,x_0)\), then we have

$$\displaystyle \begin{aligned} {} \Vert\lvert x\rvert^j\phi\Vert^2=\langle\phi,\lvert x\rvert^{2j}\phi\rangle=\int_X\lvert x\rvert^{2j}|\phi|{}^2 &\leq\Bigl(\int_X\lvert x\rvert^{2m}\lvert\phi\rvert^2\Bigr)^{j/m}\Bigl(\int_X|\phi|{}^2\Bigr)^{(m-j)/m}\notag \\ &=\Vert\lvert x\rvert^m\phi\Vert^{2j/m}\Vert\phi\Vert^{2(m-j)/m}, \end{aligned} $$
(A.1)

where we used Hölder’s inequality with \(f=\lvert x\rvert ^{2j}|\phi |{ }^{2j/m}\) and \(g=|\phi |{ }^{(2m-2j)/m}\). It follows that \(\Vert \lvert x\rvert ^j\phi \Vert ^m\leq \Vert \lvert x\rvert ^m\phi \Vert ^j\Vert \phi \Vert ^{m-j}\). In particular,

$$\displaystyle \begin{aligned} {} \liminf_{t\to+\infty}\frac{\Vert\lvert x\rvert^m\mathrm{e}^{-\mathrm{i} tH}\psi\Vert^2}{t^{2m}} \geq\Big(\liminf_{t\to+\infty} \frac{\Vert\lvert x\rvert\mathrm{e}^{-\mathrm{i} tH}\psi\Vert^{2}}{t^2}\Big)^m\Vert\psi\Vert^{2-2m}. \end{aligned} $$
(A.2)

On \(\mathbb {R}^d\) we usually consider \(\Vert x^j\phi \Vert \) instead of \(\Vert \lvert x\rvert ^j\phi \Vert \), where \(x^j\phi :=(x_1^j\phi ,\dots ,x_d^j\phi )\). We have \(\Vert x_k^j\phi \Vert ^2\leq \Vert x_k^m\phi \Vert ^{2j/m}\Vert \phi \Vert ^{2(m-j)/m}\) by the same argument. Using the Plancherel identity gives the Gagliardo–Nirenberg inequality

$$\displaystyle \begin{aligned} \Vert D_k^j\phi\Vert\leq\Vert D_k^m\phi\Vert^{j/m}\Vert\phi\Vert^{(m-j)/m}. \end{aligned}$$

If H is a Schrödinger operator and

$$\displaystyle \begin{aligned} x^j_k(t):=\mathrm{e}^{\mathrm{i} tH}x^j_k\mathrm{e}^{-\mathrm{i} tH},\quad D^j_k(t):=\mathrm{e}^{\mathrm{i} tH}(-\mathrm{i}\partial_{x_k})^j\mathrm{e}^{-\mathrm{i} tH}, \end{aligned}$$

then this implies

$$\displaystyle \begin{aligned} {} \begin{aligned} \Vert x_k^j(t)\psi\Vert&\leq\Vert x_k^m(t)\psi\Vert^{j/m}\Vert\psi\Vert^{(m-j)/m},\\ \Vert D_k^j(t)\psi\Vert&\leq\Vert D_k^m(t)\psi\Vert^{j/m}\Vert\psi\Vert^{(m-j)/m}. \end{aligned} \end{aligned} $$
(A.3)

Lastly in this connection, recall the uncertainty principle \(\Vert \phi \Vert ^2\leq 2\Vert x_k\phi \Vert \Vert \partial _{x_k}\phi \Vert \). The above yields the generalization

$$\displaystyle \begin{aligned} \Vert\phi\Vert^2\leq 2\Vert x_k^m\phi\Vert^{1/m}\Vert\phi\Vert^{(m-1)/m}\Vert\partial_{x_k}^m\phi\Vert^{1/m}\Vert\phi\Vert^{(m-1)/m}, \end{aligned}$$

i.e.

$$\displaystyle \begin{aligned} {} \Vert\phi\Vert^2 \leq 2^m \Vert x_k^m\phi\Vert\cdot\Vert D_k^m\phi\Vert. \end{aligned} $$
(A.4)

Applying this to \(\phi =\mathrm{e} ^{-\mathrm{i} tH}\psi \), we thus get

$$\displaystyle \begin{aligned} {} \Vert x_k^j(t)\psi\Vert\cdot\Vert D_k^j(t)\psi\Vert\leq 2^{m-j}\Vert x_k^m(t)\psi\Vert\cdot\Vert D_k^m(t)\psi\Vert. \end{aligned} $$
(A.5)

More generally, for \(j,n\leq m\),

$$\displaystyle \begin{aligned} {} \Vert x_k^j(t)\psi\Vert\cdot\Vert D_k^n(t)\psi\Vert\leq 2^{\frac{2m-j-n}{2}}\Vert x_k^m(t)\psi\Vert^{\frac{2m+j-n}{2m}}\Vert D_k^m(t)\psi\Vert^{\frac{2m-j+n}{2m}}. \end{aligned} $$
(A.6)

The preceding estimates provide useful lower bounds for \(x^m(t)\psi \) and \(D^m(t)\psi \) in terms of lower moments.

We now consider upper bounds.

1.2 A.2 Discrete Case

In the following, given a countable graph G, we fix some vertex \(o\in G\) regarded as an origin and denote \(\lvert x\rvert := d(x,o)\) and \(x^m\psi (x):=\lvert x\rvert ^m\psi (x)\).

Theorem A.1

Let \(H=\mathcal {A}+V\) be a Schrödinger operator on a countable graph G with maximal degree \(\leq \mathrm {D}\). We assume the potential V  is bounded. Then for any \(t\geq 0\) and \(m\in \mathbb {N}\), if \(\Vert x^m\psi \Vert <\infty \), then

$$\displaystyle \begin{aligned} {} \Vert x^m\mathrm{e}^{-\mathrm{i} tH}\psi\Vert\leq\sum_{r=0}^m p_{m-r}(t)\Vert x^r\psi\Vert, \end{aligned} $$
(A.7)

where \(p_k(t)\) is a polynomial in t of degree k with \(p_0(t)=1\), and the leading term of the top polynomial \(p_m(t)\) is \(\mathrm {D}^mt^m\). In particular,

$$\displaystyle \begin{aligned} {} \limsup_{t\to+\infty}\frac{\Vert x^m\mathrm{e}^{-\mathrm{i} tH}\psi\Vert}{t^m}\leq\mathrm{D}^m \Vert\psi\Vert. \end{aligned} $$
(A.8)

Note that each \(p_k(t)\) also depends on m, that is, for each fixed m there is a set of polynomials \(p_{0,m}(t),\dots ,p_{m,m}(t)\) with \(p_{k,m}\) of degree k such that (A.7) is satisfied with \(p_k\equiv p_{k,m}\). See Remark A.3 for a further comment.

Proof

By induction on m. The statement is trivial for \(m=0\) since \(\mathrm{e} ^{-\mathrm{i} tH}\) is unitary.

Let \(m=1\). For an operator O, recall we denote \(O(t):=\mathrm{e} ^{\mathrm{i} tH}O\mathrm{e} ^{-\mathrm{i} tH}\). In particular, \(x(t):=\mathrm{e} ^{\mathrm{i} tH}x\mathrm{e} ^{-\mathrm{i} tH}\) when O is the operator of multiplication by x. We formally have

$$\displaystyle \begin{aligned} {} \frac{\mathrm{d}}{\mathrm{d} t} x(t)\psi &=\mathrm{i} H\mathrm{e}^{\mathrm{i} tH}x\mathrm{e}^{-\mathrm{i} tH}\psi-\mathrm{i}\mathrm{e}^{\mathrm{i} tH}x H\mathrm{e}^{-\mathrm{i} tH}\psi\notag\\ &=\mathrm{i}\mathrm{e}^{\mathrm{i} tH}[H,x]\mathrm{e}^{-\mathrm{i} tH}\psi\notag\\ &=\mathrm{i}\mathrm{e}^{\mathrm{i} tH}[\mathcal{A},x]\mathrm{e}^{-\mathrm{i} tH}\psi\,. \end{aligned} $$
(A.9)

This calculation is formal because the first term \(\mathrm{i} H\mathrm{e} ^{\mathrm{i} tH}x\mathrm{e} ^{-\mathrm{i} tH}\psi \) in the derivative requires \(x\mathrm{e} ^{-\mathrm{i} tH}\psi \in \ell ^2(G)\), while the second term \(-\mathrm{i} \mathrm{e} ^{\mathrm{i} tH}x H\mathrm{e} ^{-\mathrm{i} tH}\psi \) requires \(\lim _{\delta \to 0}x\frac {\mathrm{e} ^{-\mathrm{i} (t+\delta )H}-\mathrm{e} ^{-\mathrm{i} tH}}{\delta }\psi =-\mathrm{i} xH\mathrm{e} ^{-\mathrm{i} tH}\psi \). See, e.g., [15, Lemma 10.17]. None of these facts is a priori clear (in fact the first point is partly what the theorem tries to prove, we only know that \(x\psi \in \ell ^2(G)\), a priori). Note that this formal calculation is justified however if instead of x we multiply by a bounded function.

So, similar to [27], given \(\epsilon >0\), we consider \(f_{\epsilon }(\lambda ):=\frac {\lambda }{1+\epsilon \lambda }\) for \(\lambda \geq 0\). Then multiplication by \(f_{\epsilon }(\lvert x\rvert )\) is a bounded operator and we have \(\frac {\mathrm{d} }{\mathrm{d} t} [f_{\epsilon }(\lvert x\rvert )](t)=\mathrm{i} \mathrm{e} ^{\mathrm{i} t H}[\mathcal {A},f_{\epsilon }(\lvert x\rvert )]\mathrm{e} ^{-\mathrm{i} tH}\psi \). But

$$\displaystyle \begin{aligned} [\mathcal{A},f_{\epsilon}(\lvert x\rvert)]\phi(x)=\sum_{y\sim x}[f_\epsilon(|y|) - f_\epsilon(\lvert x\rvert)]\phi(y) =:\sum_{y\sim x}\alpha_{x,y}\phi(y). \end{aligned}$$

Here \(|\alpha _{x,y}|=|f_{\epsilon }^{\prime }(\lambda )|\) for some \(\lambda \in [\lvert x\rvert -1,\lvert x\rvert +1]\). Hence, \(|\alpha _{x,y}|\leq \frac {1}{(1+\epsilon \lambda )^2}\leq 1\). We thus get

$$\displaystyle \begin{aligned} \Vert[\mathcal{A},f_{\epsilon}(\lvert x\rvert)]\phi\Vert^2\leq\mathrm{D}\sum_x\sum_{y\sim x}|\phi(y)|{}^2\leq\mathrm{D}^2\Vert\phi\Vert^2. \end{aligned}$$

Applying this to \(\phi =\mathrm{e} ^{-\mathrm{i} tH}\psi \), we get \(\Vert [\mathcal {A},f_{\epsilon }(\lvert x\rvert )]\mathrm{e} ^{-\mathrm{i} tH}\psi \Vert \leq \mathrm {D} \Vert \psi \Vert \). So using [8, Theorem 5.6.1],

$$\displaystyle \begin{aligned} \Vert(f_\epsilon(\lvert x\rvert))(t)\psi\Vert &= \Bigl\Vert(f_\epsilon(\lvert x\rvert))(0)\psi + \int_0^t\frac{\mathrm{d}}{\mathrm{d} s}(f_\epsilon(\lvert x\rvert))(s)\psi\,\mathrm{d} s\Bigr\Vert\\ &\leq \Vert f_\epsilon(\lvert x\rvert)\psi\Vert + \int_0^t \Vert[\mathcal{A},f_\epsilon(\lvert x\rvert)]\mathrm{e}^{-\mathrm{i} sH}\psi\Vert\,\mathrm{d} s\\ &\leq \Vert x\psi\Vert + t\mathrm{D}\Vert\psi\Vert\,. \end{aligned} $$

Since \(\epsilon >0\) is arbitrary, taking \(\epsilon \downarrow 0\) and using Fatou’s lemma, we get

$$\displaystyle \begin{aligned} {} \Vert x\mathrm{e}^{-\mathrm{i} tH}\psi\Vert^2=\sum_x\lvert x\rvert^2|(\mathrm{e}^{-\mathrm{i} tH}\psi)(x)|{}^2 &\leq\liminf_{\epsilon\downarrow 0}\sum_x\frac{\lvert x\rvert^2}{(1+\epsilon \lvert x\rvert)^2}|(\mathrm{e}^{-\mathrm{i} tH}\psi)(x)|{}^2\notag\\ &= \liminf_{\epsilon\downarrow 0} \Vert f_\epsilon(\lvert x\rvert)\mathrm{e}^{-\mathrm{i} tH}\psi\Vert^2\notag\\ &\leq (\Vert x\psi\Vert + t\mathrm{D}\Vert\psi\Vert)^2, \end{aligned} $$
(A.10)

where we used \(\Vert f_\epsilon (\lvert x\rvert )\mathrm{e} ^{-\mathrm{i} tH}\psi \Vert =\Vert (f_\epsilon (\lvert x\rvert ))(t)\psi \Vert \) since \(\mathrm{e} ^{\mathrm{i} tH}\) is unitary. This settles \(m=1\).

Now assume the statement holds for all \(k<m\). Let \(f_\epsilon (\lambda )=\frac {\lambda ^m}{1+\epsilon \lambda ^m}\). Here \(|f_\epsilon '(\lambda )|\leq m|\lambda |{ }^{m-1}\). Arguing as before, we get

$$\displaystyle \begin{aligned} \Vert[\mathcal{A},f_{\epsilon}(\lvert x\rvert)]\phi\Vert^2 &\leq \mathrm{D}m^2\sum_x(\lvert x\rvert+1)^{2(m-1)}\sum_{y\sim x}|\phi(y)|{}^2\\ &\leq \mathrm{D}^2m^2 \Vert(\lvert x\rvert+2)^{m-1}\phi\Vert^2\,. \end{aligned} $$

Hence,

$$\displaystyle \begin{aligned} \Vert(f_\epsilon(\lvert x\rvert))(t)\psi\Vert\leq \Vert f_\epsilon(\lvert x\rvert)\psi\Vert + \mathrm{D}m\sum_{q=0}^{m-1}\binom{m-1}{q}2^{m-1-q}\int_0^t \Vert x^q\mathrm{e}^{-\mathrm{i} sH}\psi\Vert\,\mathrm{d} s. \end{aligned}$$

Since \(f_\epsilon (\lvert x\rvert )\leq \lvert x\rvert ^m\), using the induction hypothesis we get

$$\displaystyle \begin{aligned} \Vert(f_\epsilon(\lvert x\rvert))(t)\psi\Vert\leq \Vert x^m\psi\Vert+\mathrm{D}m\sum_{q=0}^{m-1}\binom{m-1}{q}2^{m-1-q}\sum_{r=0}^q\tilde{p}_{q-r+1}(t)\Vert x^r\psi\Vert\,. \end{aligned}$$

As the RHS is independent of \(\epsilon \), taking \(\epsilon \downarrow 0\) and using Fatou’s lemma again yields \(\Vert x^m(t)\psi \Vert \leq \sum _{s=0}^m p_{m-s}(t)\Vert x^s\psi \Vert \). The above also shows the coefficient of \(\Vert x^m\psi \Vert \) is \(p_0(t)=1\). The top polynomial \(p_m(t)\) is found by taking \(q=m-1\) and \(r=0\) and equals \(\mathrm {D}m\tilde {p}_m(t)\), where \(\tilde {p}_m(t):=\int _0^t p_{m-1}(s)\,\mathrm{d} s\). As the leading term of \(p_{m-1}(s)\) is \(\mathrm {D}^{m-1}s^{m-1}\) by hypothesis, the leading term of \(\mathrm {D}m\int _0^t p_{m-1}(s)\,\mathrm{d} s\) is \(\mathrm {D}^m t^m\). □

A posteriori, the formal differentiation (A.9) is actually valid. Recall the notation \(x(t):=\mathrm{e} ^{\mathrm{i} tH}x\mathrm{e} ^{-\mathrm{i} tH}\).

Corollary A.2

Under the same assumptions, \(\lim \limits _{s\to t}x^m(s)\psi =x^m(t)\psi \) and

$$\displaystyle \begin{aligned} {} \frac{\mathrm{d}}{\mathrm{d} t}x^m(t)\psi=\mathrm{i}\mathrm{e}^{\mathrm{i} tH}[\mathcal{A},x^m]\mathrm{e}^{-\mathrm{i} tH}\psi\,. \end{aligned} $$
(A.11)

Proof

We have

$$\displaystyle \begin{aligned} {} &\Vert x^m(s)\psi - x^m(t)\psi\Vert = \Vert\mathrm{e}^{\mathrm{i} sH}x^m\mathrm{e}^{-\mathrm{i} sH}\psi - \mathrm{e}^{\mathrm{i} tH}x^m\mathrm{e}^{-\mathrm{i} tH}\psi\Vert\notag \\ &\quad \leq \Vert\mathrm{e}^{\mathrm{i} sH} x^m \mathrm{e}^{-\mathrm{i} sH}\psi - \mathrm{e}^{\mathrm{i} sH} x^m \mathrm{e}^{-\mathrm{i} tH}\psi\Vert + \Vert\mathrm{e}^{\mathrm{i} sH}x^m\mathrm{e}^{-\mathrm{i} tH}\psi - \mathrm{e}^{\mathrm{i} tH}x^m\mathrm{e}^{-\mathrm{i} tH}\psi\Vert\notag\\ &\quad = \Vert x^m \mathrm{e}^{-\mathrm{i} sH}\psi - x^m \mathrm{e}^{-\mathrm{i} tH}\psi\Vert + \Vert\mathrm{e}^{\mathrm{i} sH}x^m\mathrm{e}^{-\mathrm{i} tH}\psi - \mathrm{e}^{\mathrm{i} tH}x^m\mathrm{e}^{-\mathrm{i} tH}\psi\Vert\,. \end{aligned} $$
(A.12)

We know from Theorem A.1 that \(\phi =x^m\mathrm{e} ^{-\mathrm{i} tH}\psi \in D(H)=\ell ^2(G)\) for any \(t\geq 0\), so \(\lim _{s\to t}\mathrm{e} ^{\mathrm{i} sH}x^m\mathrm{e} ^{-\mathrm{i} tH}\psi =\mathrm{e} ^{\mathrm{i} tH}x^m\mathrm{e} ^{-\mathrm{i} tH}\psi \). This settles the second term in the RHS.

For the first term, we use Fatou’s lemma as in (A.10). Let \(f_\epsilon (\lambda )=\frac {\lambda ^m}{1+\epsilon \lambda ^m}\). We have \(f_{\epsilon }(\lvert x\rvert )\mathrm{e} ^{-\mathrm{i} tH}\psi =f_\epsilon (\lvert x\rvert )\mathrm{e} ^{-\mathrm{i} sH}\psi +\int _s^t\frac {\mathrm{d} }{\mathrm{d} \alpha }f_\epsilon (\lvert x\rvert )\mathrm{e} ^{-\mathrm{i} \alpha H}\psi \,\mathrm{d} \alpha \). Now

$$\displaystyle \begin{aligned} \Bigl\Vert\frac{\mathrm{d}}{\mathrm{d} \alpha}f_\epsilon(\lvert x\rvert)\mathrm{e}^{-\mathrm{i}\alpha H}\psi\Bigr\Vert&= \Vert f_\epsilon(\lvert x\rvert)H\mathrm{e}^{-\mathrm{i}\alpha H}\psi\Vert\\ &\leq \Vert x^m \mathrm{e}^{-\mathrm{i}\alpha H}H\psi\Vert \leq \sum_{r=0}^{m} p_{m-r}(\alpha)\Vert x^r H\psi\Vert \end{aligned} $$

for some polynomials \(p_k\), by Theorem A.1. These are uniformly bounded by some \(M(t)\) for all \(\alpha \in [t-1,t+1]\) and we get \(\Vert f_\epsilon (\lvert x\rvert ) \mathrm{e} ^{-\mathrm{i} sH}\psi - f_\epsilon (\lvert x\rvert ) \mathrm{e} ^{-\mathrm{i} tH}\psi \Vert \leq \left |t-s\right |M(t)\sum _{r=0}^{m}\Vert x^rH\psi \Vert \), with \(M(t)\) independent of \(\epsilon \).

By Fatou’s lemma, \(\Vert x^m \mathrm{e} ^{-\mathrm{i} sH}\psi - x^m \mathrm{e} ^{-\mathrm{i} tH}\psi \Vert ^2 \leq \liminf \limits _{\epsilon \to 0}\Vert f_\epsilon (\lvert x\rvert ) \mathrm{e} ^{-\mathrm{i} sH}\psi - f_\epsilon (\lvert x\rvert ) \mathrm{e} ^{-\mathrm{i} tH}\psi \Vert ^2\). Thus, \(\Vert x^m \mathrm{e} ^{-\mathrm{i} sH}\psi - x^m \mathrm{e} ^{-\mathrm{i} tH}\psi \Vert \leq \left |t-s\right |M(t)\sum _{r=0}^{m-1}\Vert x^rH\psi \Vert \to 0\) as \(s\to t\). Recalling (A.12), this completes the proof of the first claim.

For the derivative we first make some simplifications. Given \(\delta >0\),

$$\displaystyle \begin{aligned} {} &\Bigl\Vert\frac{x^m(t+\delta)\psi - x^m(t)\psi}{\delta} - \mathrm{i}\mathrm{e}^{\mathrm{i} tH}[H,x^m]\mathrm{e}^{-\mathrm{i} tH}\psi\Bigr\Vert \notag\\ &\quad =\Bigl\Vert\frac{\mathrm{e}^{\mathrm{i}\delta H} x^m\mathrm{e}^{-\mathrm{i}(t+\delta)H}\psi - x^m\mathrm{e}^{-\mathrm{i} tH}\psi}{\delta} - \mathrm{i} [H,x^m]\mathrm{e}^{-\mathrm{i} tH}\psi\Bigr\Vert \notag\\ &\quad \leq \Bigl\Vert\frac{\mathrm{e}^{\mathrm{i}\delta H} - I}{\delta}x^m\mathrm{e}^{-\mathrm{i}(t+\delta)H}\psi - \mathrm{i} H x^m\mathrm{e}^{-\mathrm{i}(t+\delta)H}\psi\Bigr\Vert \notag\\ &\qquad + \Vert Hx^m(\mathrm{e}^{-\mathrm{i} tH}\psi-\mathrm{e}^{-\mathrm{i}(t+\delta)H})\psi\Vert \notag\\ &\qquad + \Bigl\Vert x^m\Bigl(\frac{\mathrm{e}^{-\mathrm{i}(t+\delta)H} - \mathrm{e}^{-\mathrm{i} tH}}{\delta}\psi +\mathrm{i} H\mathrm{e}^{-\mathrm{i} tH}\psi\Bigr)\Bigr\Vert\notag\\ &\quad \leq \Bigl\Vert\Bigl(\frac{\mathrm{e}^{\mathrm{i}\delta H} - I}{\delta}-\mathrm{i} H\Bigr)x^m\mathrm{e}^{-\mathrm{i} tH}\psi\Bigr\Vert\notag\\ &\qquad +\Bigl\Vert\Bigl(\frac{\mathrm{e}^{\mathrm{i}\delta H} - I}{\delta}-\mathrm{i} H\Bigr)x^m(\mathrm{e}^{-\mathrm{i}(t+\delta)H}\psi - \mathrm{e}^{-\mathrm{i} tH}\psi)\Bigr\Vert\notag\\ &\qquad + \Vert Hx^m(\mathrm{e}^{-\mathrm{i} tH}\psi-\mathrm{e}^{-\mathrm{i}(t+\delta)H})\psi\Vert \notag\\ &\qquad + \Bigl\Vert x^m\Big(\frac{\mathrm{e}^{-\mathrm{i}(t+\delta)H} - \mathrm{e}^{-\mathrm{i} tH}}{\delta}\psi +\mathrm{i} H\mathrm{e}^{-\mathrm{i} tH}\psi\Big)\Bigr\Vert\,. \end{aligned} $$
(A.13)

For the first term, we know from Theorem A.1 that \(x^m\mathrm{e} ^{-\mathrm{i} tH}\psi \in D(H)=\ell ^2(G)\), so this term vanishes as \(\delta \to 0\). For the second term, we use the spectral theorem: \(\Vert \frac {\mathrm{e} ^{\mathrm{i} \delta H}-I}{\delta }\phi \Vert ^2=\int |\frac {\mathrm{e} ^{\mathrm{i} \delta \lambda }-1}{\delta }|{ }^2\,\mathrm{d} \mu _\phi (\lambda )\leq \int \lambda ^2\,\mathrm{d} \mu _\phi (\lambda )=\Vert H\phi \Vert ^2\). With this bound, we see the second and third terms vanish as \(\delta \to 0\) by the argument of (A.12) (note that H is bounded; see also Corollary A.5 for unbounded operators). So it remains to control the last term. For this we first use Fatou’s lemma to replace \(x^m\) by \(f_\epsilon (\lvert x\rvert )\) as follows.

We know that

$$\displaystyle \begin{aligned} \frac{\mathrm{d}}{\mathrm{d} s} f_\epsilon(\lvert x\rvert)\mathrm{e}^{-\mathrm{i} sH}\psi&=f_\epsilon(\lvert x\rvert)(-\mathrm{i} H\mathrm{e}^{-\mathrm{i} sH})\psi,\\ \frac{\mathrm{d}}{\mathrm{d} s} f_\epsilon(\lvert x\rvert)(-\mathrm{i} H\mathrm{e}^{-\mathrm{i} sH}\psi)&=-f_\epsilon(\lvert x\rvert) H^2\mathrm{e}^{-\mathrm{i} sH}\psi. \end{aligned} $$

It follows from [8, Theorem 5.6.2] that for small \(\delta \),

$$\displaystyle \begin{aligned} \Bigl\Vert f_\epsilon(\lvert x\rvert)\Big(\frac{\mathrm{e}^{-\mathrm{i}(t+\delta)H} - \mathrm{e}^{-\mathrm{i} tH}}{\delta}\psi +\mathrm{i} H\mathrm{e}^{-\mathrm{i} tH}\psi\Big)\Bigr\Vert \leq \frac{\lvert\delta\rvert}{2}\sup_{s\in [t-1,t+1]} \Vert x^mH^2\mathrm{e}^{-\mathrm{i} sH}\psi\Vert, \end{aligned}$$

where we used that \(f_\epsilon (\lvert x\rvert )\leq \lvert x\rvert ^m\). Using Theorem A.1 again, we may bound the RHS by \(\frac {\lvert \delta \rvert }{2}M(t)\sum _{r=0}^m\Vert x^r H^2\psi \Vert \), with \(M(t)\) independent of \(\epsilon \). Fatou’s lemma implies as before that the last term in (A.13) is now bounded by \(\frac {\lvert \delta \rvert }{2}M(t)\sum _{r=0}^m\Vert x^rH^2\psi \Vert \). Taking \(\delta \to 0\) finally completes the proof. □

Remark A.3

Theorem A.1 implies that

$$\displaystyle \begin{aligned} \limsup_{t\to+\infty}\frac{1}{t^{2m}}\sum_{x\in G}\lvert x\rvert^{2m}|\mathrm{e}^{-\mathrm{i} tH}\psi(x)|{}^2 \leq \mathrm{D}^{2m}\Vert\psi\Vert^2. \end{aligned}$$

This also implies a control for odd powers. Namely, if \(\Vert x^m\psi \Vert <\infty \), letting \(\psi _t:=\mathrm{e} ^{-\mathrm{i} tH}\psi \), we have by Cauchy–Schwarz that

$$\displaystyle \begin{aligned} \sum \lvert x\rvert^m |\psi_t(x)|{}^2 \leq\Bigl(\sum \lvert x\rvert^{2m}|\psi_t(x)|{}^2\Bigr)^{1/2}\Bigl(\sum|\psi_t(x)|{}^2\Bigr)^{1/2}. \end{aligned}$$

So \(\frac {1}{t^m}\sum \lvert x\rvert ^m |\psi _t(x)|{ }^2 \leq (\frac {1}{t^{2m}}\sum \lvert x\rvert ^{2m}|\psi _t(x)|{ }^2)^{1/2}\Vert \psi \Vert \) and thus

$$\displaystyle \begin{aligned} \limsup_{t\to+\infty}\frac{1}{t^{m}}\sum_{x\in G}\lvert x\rvert^{m}|\psi_t(x)|{}^2 \leq\mathrm{D}^{m}\Vert\psi\Vert^2. \end{aligned}$$

1.3 A.3 Continuous Case

Assume now that on \(\mathbb {R}^d\), we have a potential \(V\in C^{m-1}\) such that V  and its partial derivatives of order \(<m\) are bounded. Let \(H=H_0+V=-\Delta +V\). Then we claim that

$$\displaystyle \begin{aligned} {} \Vert D^m\phi\Vert\leq C_{m,V}\sum_{k=0}^m\Vert H^k\phi\Vert. \end{aligned} $$
(A.14)

Indeed, using \(X^m-Y^m=\sum _{p=0}^{m-1}X^p(X-Y)Y^{m-1-p}\), we have

$$\displaystyle \begin{aligned} \Vert D^m\phi\Vert^2=\langle\phi,D^{2m}\phi\rangle\leq\langle\phi,H_0^m\phi\rangle= \langle\phi,H^m\phi\rangle-\sum_{p=0}^{m-1}\langle\phi,H_0^pVH^{m-1-p}\phi\rangle. \end{aligned}$$

with the convention \(\sum _{p=0}^{-1}:= 0\).

Now (A.14) is clear for \(m=0\). If (A.14) holds for all \(p<m\), then using Cauchy–Schwarz, Leibniz formula, and our assumption on V , we get

$$\displaystyle \begin{aligned} |\langle\phi, H_0^p VH^{m-1-p}\phi\rangle| &\leq c_{d,p}\Vert D^p\phi\Vert\cdot\Vert D^p VH^{m-1-p}\phi\Vert\\ &\leq c_{V,m} \sum_{q=0}^p \Vert D^p\phi\Vert\cdot\Vert D^qH^{m-1-p}\phi\Vert. \end{aligned} $$

So by induction hypothesis,

$$\displaystyle \begin{aligned} \Vert D^m\phi\Vert^2 \leq \Vert\phi\Vert\cdot\Vert H^m\phi\Vert + c_{V,m}^{\prime}\sum_{p=0}^{m-1}\sum_{r=0}^p\sum_{q=0}^p\sum_{s=0}^q \Vert H^r\phi\Vert\cdot\Vert H^{m-1-p+s}\phi\Vert. \end{aligned}$$

Using \(ab\leq \frac {1}{2}(a^2+b^2)\), we thus get \(\Vert D^m\phi \Vert ^2 \leq c_{V,m}^{\prime \prime }\sum _{k=0}^m\Vert H^k\phi \Vert ^2\), implying (A.14). This implies that for \(D(t):=\mathrm{e} ^{\mathrm{i} tH}D\mathrm{e} ^{-\mathrm{i} tH}\) and \(\psi _t:=\mathrm{e} ^{-\mathrm{i} tH}\psi \), we have

$$\displaystyle \begin{aligned} {} \Vert D(t)^r\psi\Vert=\Vert D^r\psi_t\Vert\leq C_{r,V}\sum_{k=0}^r\Vert H^k\psi\Vert \leq C_{r,V}^{\prime}\Vert\psi\Vert_{H^{2r}}, \end{aligned} $$
(A.15)

independently of t, where we used that \(H^k\mathrm{e} ^{-\mathrm{i} tH}=\mathrm{e} ^{-\mathrm{i} tH}H^k\).

Theorem A.4

If \(V\in C^{m-1}(\mathbb {R}^d)\), if V  and its partial derivatives of order \(<m\) are bounded, and if \(\psi \in H^{2m}(\mathbb {R}^d)\), then

$$\displaystyle \begin{aligned} \Vert x^m\mathrm{e}^{-\mathrm{i} tH}\psi\Vert\leq 2^{m-1}\Bigl(\Vert x^m\psi\Vert+C_mt^m \sum_{k=0}^m\Vert H^k\psi\Vert\Bigr). \end{aligned}$$

A sketch of an earlier result can also be found in [27, Theorem 4.1]. We first give a formal proof, then indicate how to make it rigorous.

Proof (Formal)

Recall that \(x^m=(x_1^m,\dots ,x_d^m)\). In this proof we denote \(x^{2m}= x_1^{2m}+\dots +x_d^{2m}\) instead of \(\lvert x^m\rvert _2^2\) to avoid too cumbersome formulas.

Formally, \(\frac {\mathrm{d} }{\mathrm{d} t} x^{2m}(t)\psi =\mathrm{i} \mathrm{e} ^{\mathrm{i} tH}[-\Delta , x^{2m}]\mathrm{e} ^{-\mathrm{i} tH}\psi \) for \(\psi \in D(H)\). But, for F smooth on \(\mathbb {R}^d\) we have \([-\Delta ,F]\phi =-(\Delta F)\phi - 2\nabla F\cdot \nabla \phi =-\nabla \cdot [(\nabla F)\phi ]-\nabla F\cdot \nabla \phi \). In particular, for \(F(x)=x^{2m}\), since \(\nabla x^{2m}=2m(x_1^{2m-1},\dots ,x_d^{2m-1})\), we get

$$\displaystyle \begin{aligned} \frac{\mathrm{d}}{\mathrm{d} t}\Vert x^m(t)\psi\Vert^2&=\frac{\mathrm{d}}{\mathrm{d} t}\langle\psi,x^{2m}(t)\psi\rangle\\ &= \mathrm{i}\langle\psi_t,[-\Delta,x^{2m}] \psi_t\rangle \\ &= \langle\psi, D(t)\cdot[(\nabla x^{2m})(t)\psi] + (\nabla x^{2m})(t)\cdot D(t)\psi\rangle\\ &\leq 4m \sum_{j=1}^d\Vert x_j^m(t)\psi\Vert\cdot\Vert x_j^{m-1}(t)D_j(t)\psi\Vert, \end{aligned} $$

where \(D_j:=-\mathrm{i} \partial _{x_j}\) and \(D_j(t):=\mathrm{e} ^{\mathrm{i} tH}D_j\mathrm{e} ^{-\mathrm{i} tH}\). We have in general

$$\displaystyle \begin{aligned} \Vert x_j^n(t)D_j^k(t)\psi\Vert&=\langle\psi,D_j^k(t)x_j^{2n}(t)D_j^k(t)\psi\rangle^{1/2}\notag\\ {} &\leq\sum_{p=0}^k\binom{k}{p}(2n)\cdots(2n-p)\lvert\langle\psi,x_j^{2n-p}(t)D_j^{2k-p}(t)\psi\rangle\rvert^{1/2} \end{aligned} $$
(A.16)
$$\displaystyle \begin{aligned} {} &\leq\sum_{p=0}^k c_{p,k,n}\Vert x_j^m(t)\psi\Vert^{1/2}\Vert x_j^{2n-p-m}(t)D_j^{2k-p}(t)\psi\Vert^{1/2}. \end{aligned} $$
(A.17)

We may apply the same inequality to \(\Vert x_j^{2n-p-m}(t)D_j^{2k-p}(t)\psi \Vert \). Doing this \(\ell -1\) times we see that the term with highest power is

$$\displaystyle \begin{aligned} C_{k,n}\Vert x_j^m(t)\psi\Vert^{\frac{1}{2}+\frac{1}{4}+\dots+\frac{1}{2^{\ell-1}}}\Vert x_j^{2^{\ell-1}(n-m)+m}(t)D_j^{2^{\ell-1}k}(t)\psi\Vert^{\frac{1}{2^{\ell-1}}}. \end{aligned}$$

The Case\(m=2^\ell \) Let \(n=m-1=2^\ell -1\). Then by applying (A.17) \(\ell -1\) times, we get

$$\displaystyle \begin{aligned} {} \frac{\mathrm{d}}{\mathrm{d} t}\Vert x(t)^m\psi\Vert^2\leq C_m\sum_{j=1}^d\Vert x_j^m(t)\psi\Vert\cdot\Vert x_j^m(t)\psi\Vert^{1-\frac{1}{2^{\ell-1}}}\sum_{r=0}^{2^{\ell-1}}c_{r,m}\Vert x_j^r(t)D_j^r(t)\psi\Vert^{\frac{1}{2^{\ell-1}}}. \end{aligned} $$
(A.18)

In fact, the terms have the general form

$$\displaystyle \begin{aligned} x_j^{2^{\ell-1}(n-m)+m-2^{\ell-2}p_1-2^{\ell-3}p_2-\dots-p_{\ell-1}}(t)D_j^{2^{\ell-1}k-2^{\ell-2}p_1-\dots-p_{\ell-1}}(t)\psi. \end{aligned}$$

For \(m=2^\ell \), \(n=m-1\), \(k=1\), we see the powers of \(x_j(t)\) and \(D_j(t)\) match indeed.

We next apply (A.16) plus Cauchy–Schwarz to get

$$\displaystyle \begin{aligned} {} &\frac{\mathrm{d}}{\mathrm{d} t}\Vert x^m(t)\psi\Vert^2\notag\\ &\qquad \leq C_m\sum_{j=1}^d\Vert x_j^m(t)\psi\Vert^{2-\frac{1}{2^{\ell-1}}}\sum_{r=0}^{2^{\ell-1}} c_{r,m}\sum_{p=0}^rc_{p,r}\Vert x_j^{2r-p}(t)\psi\Vert^{\frac{1}{2^\ell}}\Vert D_j^{2r-p}(t)\psi\Vert^{\frac{1}{2^{\ell}}}\,. \end{aligned} $$
(A.19)

Recalling (A.5) and (A.15), we conclude that for \(m=2^\ell \),

$$\displaystyle \begin{aligned} {} \frac{\mathrm{d}}{\mathrm{d} t}\Vert x^m(t)\psi\Vert^2 &\leq C_{m,d}\sum_{j=1}^d\Vert x_j^m(t)\psi\Vert^{2-\frac{1}{m}}\Vert D_j^m(t)\psi\Vert^{1/m}\notag\\ &\leq C_{m,d,V}\Vert x^m(t)\psi\Vert^{2-\frac{1}{m}}\sum_{k=0}^m\Vert H^k\psi\Vert^{1/m}. \end{aligned} $$
(A.20)

Thus,

$$\displaystyle \begin{aligned} \frac{\mathrm{d}}{\mathrm{d} t}\langle\psi,x^{2m}(t)\psi\rangle^{\frac{1}{2m}}=\frac{1}{2m}\langle\psi,x^{2m}(t)\psi\rangle^{\frac{1}{2m}-1}\frac{\mathrm{d}}{\mathrm{d} t}\langle\psi,x^{2m}(t)\psi\rangle\leq c\sum_{k=0}^m\Vert H^k\psi\Vert^{1/m}. \end{aligned}$$

We thus have \(\Vert x^m(t)\psi \Vert ^{1/m}\leq \Vert x^m\psi \Vert ^{1/m}+ Ct\sum _{k=0}^m\Vert H^k\psi \Vert ^{1/m}\). The result follows in this case. □

The General Case For general m we let \(\ell \) such that \(2^{\ell } \leq m < 2^{\ell +1}\). Say \(m=2^\ell +q\) with \(0\leq q<2^\ell \). Then following the scheme, we apply (A.17) \(\ell -1\) times. Then \(\Vert x_j^r(t)D_j^r(t)\psi \Vert ^{\frac {1}{2^{\ell -1}}}\) in (A.18) is replaced by \(\Vert x_j^{r+q}(t)D_j^r(t)\psi \Vert ^{\frac {1}{2^{\ell -1}}}\). The proof must be slightly modified as now \(2r+2q\leq 2^\ell +2q=m+q\), i.e., the powers of \(x_j(t)\) in (A.19) can exceed m. So to the q highest terms \(r=2^{\ell -1}-q+1,\dots ,2^{\ell -1}\), we apply (A.17) once more to get \(\sum _{p=0}^r c_{p,q,r}\Vert x_j^m(t)\psi \Vert ^{\frac {1}{2^\ell }}\Vert x_j^{2r+2q-p-m}(t)D_j^{2r-p}(t)\psi \Vert ^{\frac {1}{2^\ell }}\). We can now apply (A.16) plus Cauchy–Schwarz to this and the lower terms as before. Then (A.19) is replaced by

$$\displaystyle \begin{aligned} {} \frac{\mathrm{d}}{\mathrm{d} t}\Vert x(t)^m\psi\Vert^2 &\leq C_m\sum_{j=1}^d\Vert x_j^m(t)\psi\Vert^{2-\frac{1}{2^{\ell-1}}}\Big(\sum_{r=2^{\ell-1}-q+1}^{2^{\ell-1}}\sum_{p=0}^r c_{p,q,r}\Vert x_j^m(t)\psi\Vert^{\frac{1}{2^\ell}}\notag\\ &\times\sum_{p'=0}^{2r-p}c_{p',r,p,q,m}\Vert x_j^{4r+4q-2p-2m-p'}(t)\psi\Vert^{\frac{1}{2^{\ell+1}}}\Vert D_j^{4r-2p-p'}(t)\psi\Vert^{\frac{1}{2^{\ell+1}}}\notag\\ &+\sum_{r=0}^{2^{\ell-1}-q} c_{r,m}\sum_{p=0}^rc_{p,r}\Vert x_j^{2r+2q-p}(t)\Vert^{\frac{1}{2^\ell}}\Vert D_j^{2r-p}(t)\psi\Vert^{\frac{1}{2^{\ell}}}\Big)\,. \end{aligned} $$
(A.21)

We may now apply (A.6) and (A.15) to get

$$\displaystyle \begin{aligned} &\Vert x_j^{4r+4q-2p-2m-p'}(t)\psi\Vert\cdot\Vert D_j^{4r-2p-p'}(t)\psi\Vert\\ &\quad \leq 2^{2m-r-2q}\Vert x_j^m(t)\psi\Vert^{\frac{2q}{m}}\Vert D_j^m(t)\psi\Vert^{\frac{2m-2q}{m}}. \end{aligned} $$

Recall \(q=m-2^{\ell }\), so \(\frac {2q}{2^{\ell +1}m}=\frac {1}{2^\ell }-\frac {1}{m}\). In the first sum of (A.21), \(\Vert x^m_j(t)\psi \Vert \) thus gets elevated to the power \(2-\frac {1}{2^{\ell -1}}+\frac {1}{2^\ell }+\frac {1}{2^\ell }-\frac {1}{m}=2-\frac {1}{m}\) as required. For the lower terms, the power is similarly \(2-\frac {1}{2^{\ell -1}}+\frac {2m+2q}{2^{\ell +1}m}=2-\frac {1}{m}\). This completes the formal proof. □

Proof (Completed)

To make the formal proof rigorous we consider the operator of multplication by

$$\displaystyle \begin{aligned} F_\epsilon(x):=\frac{x^{2m}}{1+\epsilon x^{2m}}=\frac{x_1^{2m}+\dots+x_d^{2m}}{1+\epsilon(x_1^{2m}+\dots+x_d^{2m})},\quad \epsilon>0. \end{aligned}$$

This is a bounded operator. For \(F_\epsilon (x)(t):=\mathrm{e} ^{\mathrm{i} tH}F_\epsilon (x)\mathrm{e} ^{-\mathrm{i} tH}\) we get

$$\displaystyle \begin{aligned} \frac{\mathrm{d}}{\mathrm{d} t}F_\epsilon(x)(t)\psi=\mathrm{i}\mathrm{e}^{\mathrm{i} tH}[-\Delta, F_\epsilon(x)]\mathrm{e}^{-\mathrm{i} tH}\psi\text{ for }\psi\in D(H). \end{aligned}$$

Again \([-\Delta ,F_\epsilon ]\phi =-(\Delta F_\epsilon )\phi - 2\nabla F_\epsilon \cdot \nabla \phi =-\nabla \cdot [(\nabla F_\epsilon )\phi ]-\nabla F_\epsilon \cdot \nabla \phi \). On the other hand \(\nabla F_\epsilon =2m(\frac {x_1^{2m-1}}{(1+\epsilon x^{2m})^2},\dots ,\frac {x_d^{2m-1}}{(1+\epsilon x^{2m})^2})\). So

$$\displaystyle \begin{aligned} &\frac{\mathrm{d}}{\mathrm{d} t}\langle\psi,F_\epsilon(x)(t)\psi\rangle=\langle\psi,D(t)\cdot [(\nabla F_\epsilon)(x)(t))\psi]+(\nabla F_\epsilon)(x)(t)\cdot D(t)\psi\rangle\\ &\qquad \quad =2m\sum_{j=1}^d\Bigl\langle\frac{x_j^{m-1}}{1+\epsilon x^{2m}}(t)D_j(t)\psi,\frac{x_j^{m}}{1+\epsilon x^{2m}}(t)\psi\Bigr\rangle\\ &\qquad \qquad \qquad \quad +\Bigl\langle\frac{x_j^{m}}{1+\epsilon x^{2m}}(t)\psi,\frac{x_j^{m-1}}{1+\epsilon x^{2m}}(t)D_j(t)\psi\Bigr\rangle\\ &\qquad \quad \leq 4m\sum_{j=0}^d\Bigl\Vert\frac{x_j^{m}}{1+\epsilon x^{2m}}(t)\psi\Bigr\Vert\,\Bigl\Vert\frac{x_j^{m-1}}{1+\epsilon x^{2m}}(t)D_j(t)\psi\Bigr\Vert. \end{aligned} $$

We have

$$\displaystyle \begin{aligned} \Bigl\Vert\frac{x_j^{m}}{1+\epsilon x^{2m}}(t)\psi\Bigr\Vert&=\Bigl\langle\psi_t,\frac{x_j^{2m}}{(1+\epsilon x^{2m})^2}\psi_t\Bigr\rangle^{1/2}\leq\langle\psi_t, F_\epsilon(x)\psi_t\rangle^{1/2}\\ &=\langle\psi,F_\epsilon(x)(t)\psi\rangle^{1/2} \end{aligned} $$

where \(\psi _t:=\mathrm{e} ^{-\mathrm{i} tH}\psi \). On the other hand,

$$\displaystyle \begin{aligned} \Bigl\Vert\frac{x_j^r}{1+\epsilon x^{2m}}(t)D_j^k(t)\psi\Bigr\Vert &=\Bigl\langle\psi,D_j^k(t)\frac{x_j^{2r}}{(1+\epsilon x^{2m})^2}(t)D_j^k(t)\psi\Bigr\rangle^{1/2}\\ &\leq\sum_{p=0}^k \binom{k}{p}\Big\lvert\Bigl\langle\psi_t,D_j^p\frac{x_j^{2r}}{(1+\epsilon x^{2m})^2}D_j^{2k-p}\psi_t\Bigr\rangle\Big\rvert^{1/2} \end{aligned} $$

and

$$\displaystyle \begin{aligned} \partial_{x_j}^p\frac{x_j^{2r}}{(1+\epsilon x^{2m})^2}=\sum_{\ell=0}^p\binom{p}{\ell}(2r)\cdots(2r-\ell+1) x_j^{2r-\ell}\partial_{x_j}^{p-\ell}\frac{1}{(1+\epsilon x^{2m})^2}. \end{aligned}$$

If \(f_\epsilon (u):=\frac {1}{(1+\epsilon u)^2}\) and \(g(x):= x^{2m}\) then by the Faà di Bruno formula,

$$\displaystyle \begin{aligned} &\partial_{x_j}^n\frac{1}{(1+\epsilon x^{2m})^2}=\partial_{x_j}^nf_\epsilon(g(x))=\sum_{\substack{(m_1,\dots,m_n)\\ \sum_{i=1}^nim_i=n}}c_{n,m_i}f_\epsilon^{(m_1+\dots+m_n)}(g(x))\\ &\quad \times\prod_{i=1}^n(\partial_{x_j}^ig(x))^{m_i}\,. \end{aligned} $$

But \(f_\epsilon ^{(q)}(u)=(-\epsilon )^q(q+1)!(1+\epsilon u)^{-2-q}\) and \(\partial _{x_j}^ig(x)=(2m)\cdots (2m-i+1)x_j^{(2m-i)}\). Thus,

$$\displaystyle \begin{aligned} \partial_{x_j}^n\frac{1}{(1+\epsilon x^{2m})^2} &=\sum_{\substack{(m_1,\dots,m_n)\\ \sum_{i=1}^nim_i=n}}\tilde{c}_{n,m_i} \frac{\epsilon^{m_1+\dots+m_n}}{(1+\epsilon x^{2m})^{2+m_1+\dots+m_n}}(2m)x_j^{(2m-1)m_1}\\ &\quad \times (2m)(2m-1)x_j^{(2m-2)m_2}\cdots(2m)\cdots(2m-n+1)x_j^{(2m-n)m_n}\\ &=\sum_{\substack{(m_1,\dots,m_n)\\\sum_{i=1}^nim_i=n}}C_{n,m_i} \frac{\epsilon^{m_1+\dots+m_n}}{(1+\epsilon x^{2m})^{2+m_1+\dots+m_n}}x_j^{2m(m_1+\dots+m_n)-n}\,, \end{aligned} $$

where we used \(\sum im_i=n\) in the last step. But

$$\displaystyle \begin{aligned} \epsilon ^{m_1+\dots+m_n}x_j^{2m(m_1+\dots+m_n)}\leq (1+\epsilon x^{2m})^{m_1+\dots+m_n}. \end{aligned}$$

Thus,

$$\displaystyle \begin{aligned} &\partial_{x_j}^p\frac{x_j^{2r}}{(1+\epsilon x^{2m})^2}\\ &\quad \leq\sum_{\ell=0}^p \binom{p}{\ell}(2r)\cdots(2r-\ell+1)x_j^{2r-\ell}\!\sum_{(m_1,\dots,m_{p-\ell})} C_{p-\ell,m_i}\frac{x_j^{\ell-p}}{(1+\epsilon x^{2m})^2}\\ &= \sum_{\ell=0}^p c_{p,\ell,r}\frac{x_j^{2r-p}}{(1+\epsilon x^{2m})^2}. \end{aligned} $$

It follows that

$$\displaystyle \begin{aligned} &\Bigl\Vert\frac{x_j^r}{1+\epsilon x^{2m}}(t)D_j^k(t)\psi\Bigr\Vert\\ &\quad \leq\sum_{p=0}^k c_{p,k,r}\Big\lvert\Bigl\langle\psi,\frac{x_j^{2r-p}}{(1+\epsilon x^{2m})^2}(t)D_j^{2k-p}(t)\psi\Bigr\rangle\Big\rvert^{1/2}\\ &\quad \leq\sum_{p=0}^k c_{p,k,r}\biggl\Vert\frac{x_j^m}{1+\epsilon x^{2m}}(t)\psi\biggr\Vert^{1/2}\,\biggl\Vert\frac{x_j^{2r-p-m}}{1+\epsilon x^{2m}}(t)D_j^{2k-p}(t)\psi\biggr\Vert^{1/2}. \end{aligned} $$

This proves the analog of (A.16)–(A.17). From here, the proof goes as before and we get

$$\displaystyle \begin{aligned} \Bigl\Vert\frac{x_j^m}{1+\epsilon x^{2m}}\psi\Bigr\Vert^{1/m}\leq\Vert x^m\psi\Vert^{1/m}+Ct\sum_{k=0}^m\Vert H^k\psi\Vert^{1/m}, \end{aligned}$$

independently of \(\epsilon \). The result follows by taking \(\epsilon \downarrow 0\), using Fatou’s lemma. □

Corollary A.5

Under the same assumptions on V , if \(\psi \in H^{2m+4}(\mathbb {R}^d)\) and \(x^m\psi \in L^2(\mathbb {R}^d)\), then

$$\displaystyle \begin{aligned} {} \frac{\mathrm{d}}{\mathrm{d} t}x^m(t)\psi=\mathrm{i}\mathrm{e}^{\mathrm{i} tH}[-\Delta,x^m]\mathrm{e}^{-\mathrm{i} tH}\psi. \end{aligned} $$
(A.22)

Proof

The proof is the same as that of Corollary A.2, using Theorem A.4. In more details, the fact that \(x^m(s)\psi \to x^m(t)\psi \) as \(s\to t\) for \(\psi \in H^{2m+2}(\mathbb {R}^d)\) is proved the same way by considering \(f_\epsilon (x):=\frac {x^m}{\sqrt {1+\epsilon x^{2m}}}\) instead. For the derivative, to deal with the second and third terms at the end of (A.13), we use that \(Hx^m=x^m H + [-\Delta ,x^m]\) instead. The term \(x^mH\) is dealt with as before. For the second term \([-\Delta ,x^m]\), let \(\phi _t^\delta :=\mathrm{e} ^{-\mathrm{i} tH}\psi -\mathrm{e} ^{-\mathrm{i} (t+\delta )H}\psi \). We have \(\Vert [-\Delta ,x^m]\phi _t^\delta \Vert \leq \sum _{i=1}^d [m(m-1)\Vert x_i^{m-2}\phi _t^\delta \Vert + 2m\Vert x_i^{m-1}\partial _{x_i}\phi _t^\delta \Vert ]\). The calculations (A.16)–(A.19) and their later generalization to all m imply that we may bound the second term by \(C\Vert x^m\phi _t^{\delta }\Vert ^{1-\frac {1}{m}}\Vert D^m\phi _t^{\delta }\Vert ^{\frac {1}{m}}\). The norm \(\Vert x^p\phi _t^{\delta }\Vert \to 0\) as \(\delta \to 0\) for \(p=m-2,m\) using the analog of (A.12), while \(\Vert D^m\phi _t^\delta \Vert \) is controlled using (A.15). Finally the last term in (A.13) is controlled using the same Fatou argument and we get (A.22). □

Rights and permissions

Reprints and permissions

Copyright information

© 2023 The Author(s), under exclusive license to Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

BoutetdeMonvel, A., Sabri, M. (2023). Ballistic Transport in Periodic and Random Media. In: Brown, M., et al. From Complex Analysis to Operator Theory: A Panorama. Operator Theory: Advances and Applications, vol 291. Birkhäuser, Cham. https://doi.org/10.1007/978-3-031-31139-0_10

Download citation

Publish with us

Policies and ethics

Navigation