diff options
author | Prefetch | 2021-11-29 20:39:20 +0100 |
---|---|---|
committer | Prefetch | 2021-11-29 20:39:20 +0100 |
commit | c7d44d4d1c74152e49d4a21f41635aa21c5f6333 (patch) | |
tree | ad859c76b6663302d57a8e72e7db21ec024fcbcd | |
parent | 61271b92a793dd837d8326c7064cebd0a3fcdb39 (diff) |
Expand knowledge base
-rw-r--r-- | content/know/concept/berry-phase/index.pdc | 219 | ||||
-rw-r--r-- | content/know/concept/conditional-expectation/index.pdc | 1 | ||||
-rw-r--r-- | content/know/concept/hellmann-feynman-theorem/index.pdc | 97 | ||||
-rw-r--r-- | content/know/concept/self-energy/index.pdc | 55 | ||||
-rw-r--r-- | content/know/concept/wiener-process/index.pdc | 117 |
5 files changed, 463 insertions, 26 deletions
diff --git a/content/know/concept/berry-phase/index.pdc b/content/know/concept/berry-phase/index.pdc new file mode 100644 index 0000000..339599b --- /dev/null +++ b/content/know/concept/berry-phase/index.pdc @@ -0,0 +1,219 @@ +--- +title: "Berry phase" +firstLetter: "B" +publishDate: 2021-11-29 +categories: +- Physics +- Quantum mechanics + +date: 2021-11-25T20:42:45+01:00 +draft: false +markup: pandoc +--- + +# Berry phase + +Consider a Hamiltonian $\hat{H}$ that does not explicitly depend on time, +but does depend on a given parameter $\vb{R}$. +The Schrödinger equations then read: + +$$\begin{aligned} + i \hbar \dv{t} \ket{\Psi_n(t)} + &= \hat{H}(\vb{R}) \ket{\Psi_n(t)} + \\ + \hat{H}(\vb{R}) \ket{\psi_n(\vb{R})} + &= E_n(\vb{R}) \ket{\psi_n(\vb{R})} +\end{aligned}$$ + +The general full solution $\ket{\Psi_n}$ has the following form, +where we allow $\vb{R}$ to evolve in time, +and we have abbreviated the traditional phase of the "wiggle factor" as $L_n$: + +$$\begin{aligned} + \ket{\Psi_n(t)} + = \exp\!(i \gamma_n(t)) \exp\!(-i L_n(t) / \hbar) \: \ket{\psi_n(\vb{R}(t))} + \qquad + L_n(t) \equiv \int_0^t E_n(\vb{R}(t')) \dd{t'} +\end{aligned}$$ + +The **geometric phase** $\gamma_n(t)$ is more interesting. +It is not included in $\ket{\psi_n}$, +because it depends on the path $\vb{R}(t)$ +rather than only the present $\vb{R}$ and $t$. +Its dynamics can be found by inserting the above $\ket{\Psi_n}$ +into the time-dependent Schrödinger equation: + +$$\begin{aligned} + \dv{t} \ket{\Psi_n} + &= i \dv{\gamma_n}{t} \ket{\Psi_n} - \frac{i}{\hbar} \dv{L_n}{t} \ket{\Psi_n} + + \exp\!(i \gamma_n) \exp\!(-i L_n / \hbar) \dv{t} \ket{\psi_n} + \\ + &= i \dv{\gamma_n}{t} \ket{\Psi_n} + \frac{1}{i \hbar} E_n \ket{\Psi_n} + + \exp\!(i \gamma_n) \exp\!(-i L_n / \hbar) \: \ket{\nabla_\vb{R} \psi_n} \cdot \dv{\vb{R}}{t} + \\ + &= i \dv{\gamma_n}{t} \ket{\Psi_n} + \frac{1}{i \hbar} \hat{H} \ket{\Psi_n} + + \exp\!(i \gamma_n) \exp\!(-i L_n / \hbar) \: \ket{\nabla_\vb{R} \psi_n} \cdot \dv{\vb{R}}{t} +\end{aligned}$$ + +Here we recognize the Schrödinger equation, so those terms cancel. +We are then left with: + +$$\begin{aligned} + - i \dv{\gamma_n}{t} \ket{\Psi_n} + &= \exp\!(i \gamma_n) \exp\!(-i L_n / \hbar) \: \ket{\nabla_\vb{R} \psi_n} \cdot \dv{\vb{R}}{t} +\end{aligned}$$ + +Front-multiplying by $i \bra{\Psi_n}$ gives us +the equation of motion of the geometric phase $\gamma_n$: + +$$\begin{aligned} + \boxed{ + \dv{\gamma_n}{t} + = - \vb{A}_n(\vb{R}) \cdot \dv{\vb{R}}{t} + } +\end{aligned}$$ + +Where we have defined the so-called **Berry connection** $\vb{A}_n$ as follows: + +$$\begin{aligned} + \boxed{ + \vb{A}_n(\vb{R}) + \equiv -i \braket{\psi_n(\vb{R})}{\nabla_\vb{R} \psi_n(\vb{R})} + } +\end{aligned}$$ + +Importantly, note that $\vb{A}_n$ is real, +provided that $\ket{\psi_n}$ is always normalized for all $\vb{R}$. +To prove this, we start from the fact that $\nabla_\vb{R} 1 = 0$: + +$$\begin{aligned} + 0 + &= \nabla_\vb{R} \braket{\psi_n}{\psi_n} + = \braket{\nabla_\vb{R} \psi_n}{\psi_n} + \braket{\psi_n}{\nabla_\vb{R} \psi_n} + \\ + &= \braket{\psi_n}{\nabla_\vb{R} \psi_n}^* + \braket{\psi_n}{\nabla_\vb{R} \psi_n} + = 2 \Re\{ - i \vb{A}_n \} + = 2 \Im\{ \vb{A}_n \} +\end{aligned}$$ + +Consequently, $\vb{A}_n = \Im \braket{\psi_n}{\nabla_\vb{R} \psi_n}$ is always real, +because $\braket{\psi_n}{\nabla_\vb{R} \psi_n}$ is imaginary. + +Suppose now that the parameter $\vb{R}(t)$ is changed adiabatically +(i.e. so slow that the system stays in the same eigenstate) +for $t \in [0, T]$, along a circuit $C$ with $\vb{R}(0) \!=\! \vb{R}(T)$. +Integrating the phase $\gamma_n(t)$ over this contour $C$ then yields +the **Berry phase** $\gamma_n(C)$: + +$$\begin{aligned} + \boxed{ + \gamma_n(C) + = - \oint_C \vb{A}_n(\vb{R}) \cdot \dd{\vb{R}} + } +\end{aligned}$$ + +But we have a problem: $\vb{A}_n$ is not unique! +Due to the Schrödinger equation's gauge invariance, +any function $f(\vb{R}(t))$ can be added to $\gamma_n(t)$ +without making an immediate physical difference to the state. +Consider the following general gauge transformation: + +$$\begin{aligned} + \ket*{\tilde{\psi}_n(\vb{R})} + \equiv \exp\!(i f(\vb{R})) \: \ket{\psi_n(\vb{R})} +\end{aligned}$$ + +To find $\vb{A}_n$ for a particular choice of $f$, +we need to evaluate the inner product +$\braket*{\tilde{\psi}_n}{\nabla_\vb{R} \tilde{\psi}_n}$: + +$$\begin{aligned} + \braket*{\tilde{\psi}_n}{\nabla_\vb{R} \tilde{\psi}_n} + &= \exp\!(i f) \Big( i \nabla_\vb{R} f \: \braket*{\tilde{\psi}_n}{\psi_n} + \braket*{\tilde{\psi}_n}{\nabla_\vb{R} \psi_n} \Big) + \\ + &= i \nabla_\vb{R} f \: \braket*{\psi_n}{\psi_n} + \braket*{\psi_n}{\nabla_\vb{R} \psi_n} + \\ + &= i \nabla_\vb{R} f + \braket*{\psi_n}{\nabla_\vb{R} \psi_n} +\end{aligned}$$ + +Unfortunately, $f$ does not vanish as we would have liked, +so $\vb{A}_n$ depends on our choice of $f$. + +However, the curl of a gradient is always zero, +so although $\vb{A}_n$ is not unique, +its curl $\nabla_\vb{R} \cross \vb{A}_n$ is guaranteed to be. +Conveniently, we can introduce a curl in the definition of $\gamma_n(C)$ +by applying Stokes' theorem, under the assumption +that $\vb{A}_n$ has no singularities in the area enclosed by $C$ +(fortunately, $\vb{A}_n$ can always be chosen to satisfy this): + +$$\begin{aligned} + \boxed{ + \gamma_n(C) + = - \iint_{S(C)} \vb{B}_n(\vb{R}) \cdot \dd{\vb{S}} + } +\end{aligned}$$ + +Where we defined $\vb{B}_n$ as the curl of $\vb{A}_n$. +Now $\gamma_n(C)$ is guaranteed to be unique. +Note that $\vb{B}_n$ is analogous to a magnetic field, +and $\vb{A}_n$ to a magnetic vector potential: + +$$\begin{aligned} + \vb{B}_n(\vb{R}) + \equiv \nabla_\vb{R} \cross \vb{A}_n(\vb{R}) + = \Im\Big\{ \nabla_\vb{R} \cross \braket{\psi_n(\vb{R})}{\nabla_\vb{R} \psi_n(\vb{R})} \Big\} +\end{aligned}$$ + +Unfortunately, $\nabla_\vb{R} \psi_n$ is difficult to evaluate explicitly, +so we would like to rewrite $\vb{B}_n$ such that it does not enter. +We do this as follows, inserting $1 = \sum_{m} \ket{\psi_m} \bra{\psi_m}$ along the way: + +$$\begin{aligned} + i \vb{B}_n + = \nabla_\vb{R} \cross \braket{\psi_n}{\nabla_\vb{R} \psi_n} + &= \braket{\psi_n}{\nabla_\vb{R} \cross \nabla_\vb{R} \psi_n} + \bra{\nabla_\vb{R} \psi_n} \cross \ket{\nabla_\vb{R} \psi_n} + \\ + &= \sum_{m} \braket{\nabla_\vb{R} \psi_n}{\psi_m} \cross \braket{\psi_m}{\nabla_\vb{R} \psi_n} +\end{aligned}$$ + +The fact that $\braket{\psi_n}{\nabla_\vb{R} \psi_n}$ is imaginary +means it is parallel to its complex conjugate, +and thus the cross product vanishes, so we exclude $n$ from the sum: + +$$\begin{aligned} + \vb{B}_n + &= \sum_{m \neq n} \braket{\nabla_\vb{R} \psi_n}{\psi_m} \cross \braket{\psi_m}{\nabla_\vb{R} \psi_n} +\end{aligned}$$ + +From the [Hellmann-Feynman theorem](/know/concept/hellmann-feynman-theorem/), +we know that the inner products can be rewritten: + +$$\begin{aligned} + \braket{\psi_m}{\nabla_\vb{R} \psi_n} + = \frac{\matrixel{\psi_n}{\nabla_\vb{R} \hat{H}}{\psi_m}}{E_n - E_m} +\end{aligned}$$ + +Where we have assumed that there is no degeneracy. +This leads to the following result: + +$$\begin{aligned} + \boxed{ + \vb{B}_n + = \Im \sum_{m \neq n} + \frac{\matrixel{\psi_n}{\nabla_\vb{R} \hat{H}}{\psi_m} \cross \matrixel{\psi_m}{\nabla_\vb{R} \hat{H}}{\psi_n}}{(E_n - E_m)^2} + } +\end{aligned}$$ + +Which only involves $\nabla_\vb{R} \hat{H}$, +and is therefore easier to evaluate than any $\ket{\nabla_\vb{R} \psi_n}$. + + + +## References +1. M.V. Berry, + [Quantal phase factors accompanying adiabatic changes](https://doi.org/10.1098/rspa.1984.0023), + 1984, Royal Society. +2. G. Grosso, G.P. Parravicini, + *Solid state physics*, + 2nd edition, Elsevier. diff --git a/content/know/concept/conditional-expectation/index.pdc b/content/know/concept/conditional-expectation/index.pdc index 50120e8..0384222 100644 --- a/content/know/concept/conditional-expectation/index.pdc +++ b/content/know/concept/conditional-expectation/index.pdc @@ -6,6 +6,7 @@ categories: - Mathematics - Statistics - Measure theory +- Stochastic analysis date: 2021-10-22T15:19:23+02:00 draft: false diff --git a/content/know/concept/hellmann-feynman-theorem/index.pdc b/content/know/concept/hellmann-feynman-theorem/index.pdc new file mode 100644 index 0000000..3b88cd8 --- /dev/null +++ b/content/know/concept/hellmann-feynman-theorem/index.pdc @@ -0,0 +1,97 @@ +--- +title: "Hellmann-Feynman theorem" +firstLetter: "H" +publishDate: 2021-11-29 +categories: +- Physics +- Quantum mechanics + +date: 2021-11-29T10:30:37+01:00 +draft: false +markup: pandoc +--- + +# Hellmann-Feynman theorem + +Consider the time-independent Schrödinger equation, +where the Hamiltonian $\hat{H}$ depends on a general parameter $\lambda$, +whose meaning or type we will not specify: + +$$\begin{aligned} + \hat{H}(\lambda) \ket{\psi_n(\lambda)} + = E_n(\lambda) \ket{\psi_n(\lambda)} +\end{aligned}$$ + +Assuming all eigenstates $\ket{\psi_n}$ are normalized, +this gives us the following basic relation: + +$$\begin{aligned} + \matrixel{\psi_m}{\hat{H}}{\psi_n} + = E_n \braket{\psi_m}{\psi_n} + = \delta_{mn} E_n +\end{aligned}$$ + +We differentiate this with respect to $\lambda$, +which could be a scalar or a vector. +This yields: + +$$\begin{aligned} + \delta_{mn} \nabla_\lambda E_n + &= \nabla_\lambda \matrixel{\psi_m}{\hat{H}}{\psi_n} + \\ + &= \matrixel{\nabla_\lambda \psi_m}{\hat{H}}{\psi_n} + + \matrixel{\psi_m}{\nabla_\lambda \hat{H}}{\psi_n} + + \matrixel{\psi_m}{\hat{H}}{\nabla_\lambda \psi_n} + \\ + &= E_m \braket{\psi_m}{\nabla_\lambda \psi_n} + E_n \braket{\nabla_\lambda \psi_m}{\psi_n} + \matrixel{\psi_m}{\nabla_\lambda \hat{H}}{\psi_n} +\end{aligned}$$ + +In order to simplify this, +we differentiate the orthogonality relation +$\braket{\psi_m}{\psi_n} = \delta_{mn}$, +which ends up telling us that +$\braket{\nabla_\lambda \psi_m}{\psi_n} = - \braket{\psi_m}{\nabla_\lambda \psi_n}$: + +$$\begin{aligned} + 0 + = \nabla_\lambda \delta_{mn} + = \nabla_\lambda \braket{\psi_m}{\psi_n} + = \braket{\nabla_\lambda \psi_m}{\psi_n} + \braket{\psi_m}{\nabla_\lambda \psi_n} +\end{aligned}$$ + +Using this result to replace $\braket{\nabla_\lambda \psi_m}{\psi_n}$ +in the previous equation leads to: + +$$\begin{aligned} + \delta_{mn} \nabla_\lambda E_n + &= (E_m - E_n) \braket{\psi_m}{\nabla_\lambda \psi_n} + \matrixel{\psi_m}{\nabla_\lambda \hat{H}}{\psi_n} +\end{aligned}$$ + +For $m = n$, we therefore arrive at the **Hellmann-Feynman theorem**, +which is useful when doing numerical calculations +to minimize energies with respect to $\lambda$: + +$$\begin{aligned} + \boxed{ + \nabla_\lambda E_n + = \matrixel{\psi_m}{\nabla_\lambda \hat{H}}{\psi_n} + } +\end{aligned}$$ + +While for $m \neq n$, we get the **Epstein generalization** +of the Hellmann-Feynman theorem, which is for example relevant for +the [Berry phase](/know/concept/berry-phase/): + +$$\begin{aligned} + \boxed{ + (E_n - E_m) \braket{\psi_m}{\nabla_\lambda \psi_n} + = \matrixel{\psi_m}{\nabla_\lambda \hat{H}}{\psi_n} + } +\end{aligned}$$ + + + +## References +1. G. Grosso, G.P. Parravicini, + *Solid state physics*, + 2nd edition, Elsevier. diff --git a/content/know/concept/self-energy/index.pdc b/content/know/concept/self-energy/index.pdc index d2908eb..c86f8c5 100644 --- a/content/know/concept/self-energy/index.pdc +++ b/content/know/concept/self-energy/index.pdc @@ -50,8 +50,6 @@ and $\hat{\Psi}_a \equiv \hat{\Psi}_{s_a}(\vb{r}_a, \tau_a)$: $$\begin{aligned} G_{ba} - %&= - \frac{\expval{\mathcal{T}\Big\{ \hat{K}(\hbar \beta, 0) \hat{\Psi}_b \hat{\Psi}_a^\dagger \Big\}}}{\expval{\hat{K}(\hbar \beta, 0)}} - %\\ &= - \frac{\displaystyle\sum_{n = 0}^\infty \frac{1}{n!} \Big( \!-\!\frac{1}{\hbar} \Big)^n \idotsint_0^{\hbar \beta} \expval{\mathcal{T}\Big\{ \hat{W}(\tau_1) \cdots \hat{W}(\tau_n) \hat{\Psi}_b \hat{\Psi}_a^\dagger \Big\}} \dd{\tau_1} \cdots \dd{\tau_n}} {\hbar \displaystyle\sum_{n = 0}^\infty \frac{1}{n!} \Big( \!-\!\frac{1}{\hbar} \Big)^n \idotsint_0^{\hbar \beta} @@ -141,12 +139,15 @@ $$\begin{aligned} These integrals over products of interactions and Green's functions are the perfect place to apply [Feynman diagrams](/know/concept/feynman-diagram/). -Conveniently, it even turns out that the factor $(-1)^p$ -is exactly equivalent to the rule that each diagram is multiplied by $(-1)^F$, +Conveniently, it turns out that the factor $(-1)^p$ +is equivalent to the rule that each diagram must be multiplied by $(-1)^F$, with $F$ the number of fermion loops. +Keep in mind that fermion lines absorb a factor $-\hbar$ each (see above), +and interactions $-1/\hbar$. -The denominator thus turns into a sum of all possible diagrams for each total order $n$ -(the order of a diagram is the number of interaction lines it contains). +The denominator turns into a sum of all possible diagrams +(including equivalent ones) for each total order $n$ +(the order is the number of interaction lines). The endpoints $a$ and $b$ do not appear here, so we conclude that all those diagrams only have internal vertices; we will therefore refer to them as **internal diagrams**. @@ -168,24 +169,25 @@ We thus find: $$\begin{aligned} G_{ba} - &= \frac{\displaystyle\sum_{n = 0}^\infty \frac{1}{n!} \Big( \!-\! \frac{\hbar}{2} \Big)^n - \bigg[ \sum_\mathrm{all\;ext}^{n} \binom{1 \: \mathrm{external}}{\mathrm{order} \; m \!\le\! n} - \binom{\mathrm{0\;or\;more\;internal}}{\mathrm{total\;order} \; (n \!-\! m)} \bigg]} - {\displaystyle\sum_{n = 0}^\infty \frac{1}{n!} \Big( \!-\! \frac{\hbar}{2} \Big)^n - \binom{\mathrm{0\;or\;more\;internal}}{\mathrm{total\;order} \; n}} + &= \frac{\displaystyle\sum_{n = 0}^\infty \frac{1}{2^n n!} + \bigg[ \sum_{m = 0}^{n} \frac{n!}{m! (n \!-\! m)!} \binom{1 \; \mathrm{external}}{\mathrm{order} \; m}_{\!\Sigma\mathrm{all}} + \binom{\mathrm{0\;or\;more\;internal}}{\mathrm{total\;order} \; (n \!-\! m)}_{\!\Sigma\mathrm{all}} \bigg]} + {\displaystyle\sum_{n = 0}^\infty \frac{1}{2^n n!} \binom{\mathrm{0\;or\;more\;internal}}{\mathrm{total\;order} \; n}_{\!\Sigma\mathrm{all}}} \end{aligned}$$ -Where the total order refers to the sum of the orders of all disconnected diagrams. -Note that the external diagram does not directly depend on $n$. -We can therefore reorganize: +Where the total order is the sum of the orders of all considered diagrams, +and the new factor is needed for all the possible choices +of vertices to put in the external part. +Note that the external diagram does not directly depend on $n$, +so we reorganize: $$\begin{aligned} G_{ba} - &= \frac{\displaystyle\sum_\mathrm{all\;ext}^{\infty} \binom{1 \: \mathrm{external}}{\mathrm{order} \; m} - \bigg[ \sum_{n = 0}^\infty \frac{1}{n!} \Big( \!-\! \frac{\hbar}{2} \Big)^n - \binom{\mathrm{0\;or\;more\;internal}}{\mathrm{total\;order} \; (n \!-\! m)} \bigg]} - {\displaystyle\sum_{n = 0}^\infty \frac{1}{n!} \Big( \!-\! \frac{\hbar}{2} \Big)^n - \binom{\mathrm{0\;or\;more\;internal}}{\mathrm{total\;order} \; n}} + &= \frac{\displaystyle\sum_{m = 0}^{\infty} \frac{1}{2^m m!} \binom{1 \; \mathrm{external}}{\mathrm{order} \; m}_{\!\Sigma\mathrm{all}} + \bigg[ \sum_{n = 0}^\infty \frac{1}{2^{n-m} (n \!-\! m)!} + \binom{\mathrm{0\;or\;more\;internal}}{\mathrm{total\;order} \; (n \!-\! m)}_{\!\Sigma\mathrm{all}} \bigg]} + {\displaystyle\sum_{n = 0}^\infty \frac{1}{2^n n!} + \binom{\mathrm{0\;or\;more\;internal}}{\mathrm{total\;order} \; n}_{\!\Sigma\mathrm{all}}} \end{aligned}$$ Since both $n$ and $m$ start at zero, @@ -194,18 +196,19 @@ we see that the second sum in the numerator does not actually depend on $m$: $$\begin{aligned} G_{ba} - &= \frac{\displaystyle\sum_\mathrm{all\;ext} \binom{1 \: \mathrm{external}}{\mathrm{order} \; m} - \bigg[ \sum_{n = 0}^\infty \frac{1}{n!} \Big( \!-\! \frac{\hbar}{2} \Big)^n - \binom{\mathrm{0\;or\;more\;internal}}{\mathrm{total\;order} \; n} \bigg]} - {\displaystyle\sum_{n = 0}^\infty \frac{1}{n!} \Big( \!-\! \frac{\hbar}{2} \Big)^n - \binom{\mathrm{0\;or\;more\;internal}}{\mathrm{total\;order} \; n}} + &= \frac{\displaystyle\sum_{m = 0}^{\infty} \frac{1}{2^m m!} \binom{1 \; \mathrm{external}}{\mathrm{order} \; m}_{\!\Sigma\mathrm{all}} + \bigg[ \sum_{n = 0}^\infty \frac{1}{2^n n!} \binom{\mathrm{0\;or\;more\;internal}}{\mathrm{total\;order} \; n}_{\!\Sigma\mathrm{all}} \bigg]} + {\displaystyle\sum_{n = 0}^\infty \frac{1}{2^n n!} \binom{\mathrm{0\;or\;more\;internal}}{\mathrm{total\;order} \; n}_{\!\Sigma\mathrm{all}}} \\ - &= \sum_\mathrm{all\;ext} \binom{1 \: \mathrm{external}}{\mathrm{order} \; m} + &= \sum_{m = 0}^{\infty} \frac{1}{2^m m!} \binom{1 \; \mathrm{external}}{\mathrm{order} \; m}_{\!\Sigma\mathrm{all}} \end{aligned}$$ In other words, all the disconnected diagrams simply cancel out, and we are left with a sum over all possible fully connected diagrams -that contain $a$ and $b$. Let $G(b,a) = G_{ba}$: +that contain $a$ and $b$. Furthermore, it can be shown using combinatorics +that exactly $2^m m!$ diagrams at each order are topologically equivalent, +so we are left with non-equivalent diagrams only. +Let $G(b,a) = G_{ba}$: <a href="fullgf.png"> <img src="fullgf.png" style="width:90%;display:block;margin:auto;"> diff --git a/content/know/concept/wiener-process/index.pdc b/content/know/concept/wiener-process/index.pdc index 11c7a6e..1919284 100644 --- a/content/know/concept/wiener-process/index.pdc +++ b/content/know/concept/wiener-process/index.pdc @@ -68,6 +68,123 @@ since each increment has mean zero (so it is a martingale), and all increments are independent (so it is a Markov process). +## Recurrence + +An important question about the Wiener process +is whether it is **recurrent** or **transient**: +given a hypersphere (interval in 1D, circle in 2D, sphere in 3D) +away from the origin, will $B_t$ visit it after a finite time $\tau\!<\!\infty$? +It is *recurrent* if yes, i.e. $P(\tau \!<\! \infty) = 1$, or *transient* otherwise. +The answer to this question turns out to depend on the number of dimenions. + +To demonstrate this, we model the $d$-dimensional Wiener process +as an [Itō diffusion](/know/concept/ito-calculus/) $X_t$, +which also allows us to shift the initial condition $X_0$ +(or resume a "paused" process): + +$$\begin{aligned} + X_t + = X_0 + \int_0^t \dd{B_s} +\end{aligned}$$ + +Consider two hyperspheres, the inner with radius $R_i$, +and the outer with $R_o > R_i$. +Let the initial condition $|X_0| \in \, ]R_i, R_o[$, +then we define the stopping times $\tau_i$, $\tau_o$ and $\tau$ like so: + +$$\begin{aligned} + \tau_i + \equiv \inf\{ t : |X_t| \le R_i \} + \qquad + \tau_o + \equiv \inf\{ t : |X_t| \ge R_o \} + \qquad + \tau + \equiv \min\{\tau_i, \tau_o\} +\end{aligned}$$ + +We stop when the inner or outer hypersphere is touched by $X_t$, +whichever happens first. + +[Dynkin's formula](/know/concept/dynkins-formula/) +is applicable to this situation, if we define $h(x)$ as follows, +where the *terminal reward* $\Gamma$ equals $1$ for $|X_\tau| = R_i$, +and $0$ for $|X_\tau| = R_o$, +such that $h(X_0)$ equals the probability +that we touch $R_i$ before $R_o$ for a given $X_0$: + +$$\begin{aligned} + h(X_0) + = \mathbf{E}\Big[ \Gamma(X_\tau) \Big| X_0 \Big] + = P\Big[|X_\tau| \!=\! R_i \:\Big|\: X_0\Big] +\end{aligned}$$ + +Dynkin's formula then tells us that $h(x)$ is given by the following equation, +with the boundary conditions $h(R_i) = 1$ and $h(R_o) = 0$: + +$$\begin{aligned} + 0 + = \hat{L}\{h(x)\} + = \frac{1}{2} \nabla^2 h(x) +\end{aligned}$$ + +Thanks to this problem's spherical symmetry, +$h$ only depends on the radial coodinate $r$, +so the Laplacian $\nabla^2$ can be written as follows +in $d$-dimensional [spherical coordinates](/know/concept/spherical-coordinates/): + +$$\begin{aligned} + 0 + = \nabla^2 h(r) + = \pdv[2]{h}{r} + \frac{d - 1}{r} \pdv{h}{r} +\end{aligned}$$ + +For $d = 1$, the solution $h_1(r)$ is as follows, +of which we take the limit for $R_o \to \infty$: + +$$\begin{aligned} + h_1(r) + = \frac{r - R_o}{R_i - R_o} + \quad\underset{R_o \to \infty}{\longrightarrow}\quad + 1 +\end{aligned}$$ + +The outer hypersphere becomes harder to reach for larger $R_o$, +and for $R_o \to \infty$ we are left with +the probability of hitting $R_i$ only. +This turns out to be $1$, so in 1D the Wiener process is recurrent: +it always comes close to the origin in finite time. + +For $d = 2$, the solution $h_2(r)$ is as follows, +whose limit turns out to be $1$, +so the Wiener process is also recurrent in 2D: + +$$\begin{aligned} + h_2(r) + = 1 - \frac{\log\!(r/R_i)}{\log\!(R_o/R_i)} + \quad\underset{R_o \to \infty}{\longrightarrow}\quad + 1 +\end{aligned}$$ + +However, for $d \ge 3$, the solution $h_d(r)$ +does not converge to $1$ for $R_o \to \infty$, +meaning the Wiener process is transient in 3D or higher: + +$$\begin{aligned} + h_d(r) + = \frac{R_o^{2 - d} - r^{2 - d}}{R_o^{2 - d} - R_i^{2 - d}} + \quad\underset{R_o \to \infty}{\longrightarrow}\quad + \frac{R_i^{d - 2}}{r^{d - 2}} + < 1 +\end{aligned}$$ + +This is a major qualitative difference. For example, consider a situation +where some substance is diffusing from a localized infinite source: +in 3D, the substance can escape and therefore a steady state can exist, +while in 2D, the substance never strays far from the source, +so no steady state is ever reached as long as the source continues to emit. + + ## References 1. U.H. Thygesen, |