--- title: "Shor's algorithm" date: 2021-04-13 categories: - Quantum information - Cryptography - Algorithms layout: "concept" --- **Shor's algorithm** was the first truly useful quantum algorithm. It can solve important problems, most notably integer factorization, much more efficiently than any classical algorithm. It weakens widely-used cryptographic schemes, such as RSA and [Diffie-Hellman](/know/concept/diffie-hellman-key-exchange/). In essence, Shor's algorithm's revolutionary achievement is that it can efficiently find the periods $s_1, ..., s_A$ of a function $f(x_1, ..., x_A)$ on a discrete finite field, where: $$\begin{aligned} f(x_1, ..., x_A) = f(x_1 + s_1, ..., x_A + s_A) \end{aligned}$$ This is a so-called *hidden subgroup problem* for a *finite Abelian group*. With minimal modifications, Shor's algorithm can solve practically every such problem. ## Integer factorization Originally, Shor's algorithm was designed to factorize an integer $N$, in which case the goal is to find the period $s$ of the modular exponentiation function $f$ (for reasons explained later): $$\begin{aligned} f(x) = a^x \bmod N \end{aligned}$$ For a given $a$ and $N$. The period $s$ is the smallest integer satisfying $f(x) = f(x+s)$. To do this, the following $2q$-qubit quantum circuit is used, with $q$ chosen so that $N^2 \le 2^q < 2 N^2$: Here, $\mathrm{QFT}_q$ refers to the $q$-qubit [quantum Fourier transform](/know/concept/quantum-fourier-transform/), and the oracle $U_f$ calculates $f(x)$ for predetermined values of $a$ and $N$. It is an XOR oracle, working as follows: $$\begin{aligned} \Ket{x} \Ket{y} \quad \to \boxed{U_f(a, N)} \to \quad \Ket{x} \Ket{y \oplus f(x)} \end{aligned}$$ Execution starts by applying the [Hadamard gate](/know/concept/quantum-gate/) $H$ to the first $q$ qubits, yielding: $$\begin{aligned} \Ket{0}^{\otimes q} \Ket{0}^{\otimes q} \quad \to \boxed{H^{\otimes q}} \to \quad \Ket{+}^{\otimes q} \Ket{0}^{\otimes q} = \frac{1}{\sqrt{Q}} \sum_{x = 0}^{Q - 1} \Ket{x} \Ket{0}^{\otimes q} \end{aligned}$$ Where $Q = 2^q$, and $\Ket{x}$ is the computational basis state $\Ket{x_1} \cdots \Ket{x_q}$. Moving on to $U_f$: $$\begin{aligned} \frac{1}{\sqrt{Q}} \sum_{x = 0}^{Q - 1} \Ket{x} \Ket{0}^{\otimes q} \quad \to \boxed{U_f(a, N)} \to \quad \frac{1}{\sqrt{Q}} \sum_{x = 0}^{Q - 1} \Ket{x} \Ket{f(x)} \end{aligned}$$ Then we measure $f(x)$, causing it collapse as follows, for an unknown arbitrary value of $x_0$: $$\begin{aligned} f(x_0) = f(x_0 + s) = f(x_0 + 2s) = \cdots = f(x_0 + (L-1) s) \end{aligned}$$ Due to [entanglement](/know/concept/quantum-entanglement/), the unmeasured (top $q$) qubits change state into a superposition: $$\begin{aligned} \frac{1}{\sqrt{L}} \sum_{\ell = 0}^{L - 1} \Ket{x_0 + \ell s} \end{aligned}$$ Clearly, there is a periodic structure here, but we cannot measure it directly, because we do not know the value of $x_0$, which, to make matters worse, changes every time we run the algorithm. This is where the QFT comes in, which outputs the following state: $$\begin{aligned} \frac{1}{\sqrt{QL}} \sum_{k = 0}^{Q - 1} \bigg( \sum_{\ell = 0}^{L - 1} \omega_Q^{(x_0 + \ell s) k} \bigg) \Ket{k} \end{aligned}$$ Where $\omega_Q$ is a $Q$th root of unity. Measuring this state yields a $\Ket{k}$, with a probability $P(k)$: $$\begin{aligned} P(k) = \frac{1}{QL} \bigg| \sum_{\ell = 0}^{L - 1} \omega_Q^{(x_0 + \ell s) k} \bigg|^2 = \frac{1}{QL} \bigg| \omega_Q^{x_0 k} \sum_{\ell = 0}^{L - 1} \omega_Q^{\ell s k} \bigg|^2 = \frac{1}{QL} \bigg| \sum_{\ell = 0}^{L - 1} \omega_Q^{\ell s k} \bigg|^2 \end{aligned}$$ The last step holds because $|\omega_Q| = 1$. Surprisingly, this implies that we did not need to perform the measurement of $f(x)$ earlier! This makes sense: the period $s$ does not depend on $x_0$, so why would we need an implicit $x_0$ to determine $s$? So, what does the above probability $P(k)$ work out to? There are two cases: $$\begin{alignedat}{2} &\mathrm{if} \: \omega_Q^{sk} = 1: \qquad &&P(k) = \frac{1}{QL} |L|^2 = \frac{L}{Q} \\ &\mathrm{if} \: \omega_Q^{sk} \neq 1: \qquad &&P(k) = \frac{1}{QL} \Bigg| \frac{1 - \omega_Q^{sk L}}{1 - \omega_Q^{sk}} \Bigg|^2 \end{alignedat}$$ Where the latter case was evaluated as a geometric series. The condition $\omega_Q^{sk}\!=\!1$ is equivalent to asking if $sk$ is a multiple of $Q$, i.e. if $sk = cQ$, for an integer $c$. Recall that $L$ is the number of times that $s$ fits in $Q$, so $L\!=\!\lfloor Q / s \rfloor$. Assuming $Q/s$ is an integer, then $L\!=\!Q/s$ and $Q\!=\!s L$, which tells us that $\omega_Q^{sk}\!=\!\omega_{s L}^{s k}\!=\!\omega_L^k$. This implies that if $k$ is a multiple of $L$ (i.e. $k\!=\!c L$), then $\omega_L^k\!=\!1$, so $P(k) = L / Q$, which is exactly what we got earlier! In other words, the condition $\omega_Q^{sk}\!=\!1$ is equivalent to $Q/s$ being an integer. In that case, we have that $Q\!=\!sL$, which we substitute into $P(k)$ from earlier: $$\begin{aligned} \mathrm{if} \: (Q/s) \in \mathbb{N}: \qquad P(k) = \frac{L}{Q} = \frac{1}{s} \end{aligned}$$ And because $k$ is a multiple of $L$, and $L$ fits $s$ times in $Q$, there must be exactly $s$ values of $k$ that satisfy $P(k) = 1/s$. Therefore the probability of all other $k$-values is zero! This becomes clearer when you look at the sum used to calculate $P(k)$: if $Q\!=\!sL$, then it sums $\omega_L^{\ell k}$ over $\ell$, leading to perfect destructive interference for the "bad" $k$-values, leaving only the "good" ones. **So, to summarize: if** $Q/s$ **is an integer**, then measuring only yields $k$-values that are multiples of $L\!=\!Q/s$. Running Shor's algorithm several times then gives several $k$-values separated by $L$. That tells us what $L$ is, and we already know $Q$, so we *finally* find the period $s = Q/L$. That begs the question: what if $Q/s$ is not an integer? We cannot *check* this, since $s$ is unknown! Instead, we rewrite the probability $P(k)$ as follows: $$\begin{aligned} \mathrm{if} \: (Q/s) \not\in \mathbb{N}: \qquad P(k) = \frac{1}{QL} \Bigg| \frac{1 - \omega_Q^{sk L}}{1 - \omega_Q^{sk}} \Bigg|^2 = \frac{1}{QL} \Bigg| \frac{\sin(\pi s k L / Q)}{\sin(\pi s k / Q)} \Bigg|^2 \end{aligned}$$ This function peaks if $s k$ is close to a multiple of $Q$, i.e. $s k \approx c Q$, which we rearrange: $$\begin{aligned} \frac{k}{Q} \approx \frac{c}{s} \end{aligned}$$ We know the left-hand side, and, from the definition of $f(x)$, clearly $s \le N$. We chose $Q \sim N^2$, so $s$ is quite small, and consequently $c$ is too, since $k < Q$. In other words, $c/s$ is a "simple" fraction, so our goal is to find a "simple" fraction that is close to the "complicated" fraction $k/Q$. For example, if $k/Q\!=\!0.332$, then probably $c/s\!=\!1/3$. This can be done rigorously using the **continued fractions algorithm**: write $k/Q$ as a continued fraction, until the non-integer part of the denominator becomes small enough. This part is then neglected, and we calculate whatever is left, to get an estimate of $c/s$. Of course, $P(k)$ is a probability distribution, so even though the odds are in our favour, we might occasionally measure a misleading $k$-value. Running Shor's algorithm several times "fixes" this. **So, to summarize: if** $Q/s$ **is not an integer**, the measured $k$-values are generally close to $c Q / s$ for an integer $c$. By approximating $k/Q$ using the continued fraction algorithm, we estimate $c/s$. Repeating this procedure gives several values of $c/s$, such that $s$ is easy to deduce by taking the least common multiple of the denominators. In any case, once we think we have $s$, we can easily verify that $f(x)\!=\!f(x\!+\!s)$. Whether $s$ is the *smallest* such integer depends on how lucky we are, but fortunately, for most applications of this algorithm, that does not actually matter, and usually we find the smallest $s$ anyway. You typically need to repeat the algorithm $\mathcal{O}(\log{q})$ times, and the QFT is $\mathcal{O}(q^2)$. The bottleneck is modular exponentiation $f$, which is $\mathcal{O}(q^2 (\log{q}) \log{\log{q}})$ and therefore worse than the QFT, yielding a total complexity of $\mathcal{O}(q^2 (\log{q})^2 \log{\log{q}})$. OK, but what does $s$ have to do with factorizing integers? Well, recall that $f$ is given by: $$\begin{aligned} f(x) = a^x \bmod N \end{aligned}$$ $N$ is the number to factorize, and $a$ is a random integer *coprime* to $N$, meaning $\gcd(a, N) = 1$. The fact that $s$ is the period of $f$ for a certain $a$-value, implies that: $$\begin{aligned} a^x = a^{x + s} \bmod N \quad \implies \quad 1 = a^s \bmod N \end{aligned}$$ Suppose that $s$ is even. In that case, we can rewrite the above equation as follows: $$\begin{aligned} (a^{s/2})^2 - 1 = 0 \bmod N \end{aligned}$$ In other words, $(a^{s/2})^2 \!-\! 1$ is a multiple of $N$. We then use that $(a\!-\!b) (a\!+\!b) = a^2\!-\!b^2$: $$\begin{aligned} \big( a^{s/2} - 1 \big) \big( a^{s/2} + 1 \big) = 0 \bmod N \end{aligned}$$ Because $s$ is even by assumption, the two factors on the left are integers, and as just mentioned, their product is a multiple of $N$. Then we only need to calculate: $$\begin{aligned} \gcd\!\big( a^{s/2}\!-\!1, N \big) > 1 \quad\:\: \mathrm{and} \quad\:\: \gcd\!\big( a^{s/2}\!+\!1, N \big) > 1 \end{aligned}$$ And there we have the factors of $N$! The $\gcd$ can be calculated efficiently in $\mathcal{O}(q^2)$ time. But what if $s$ is odd? No problem, then we just choose a new $a$ coprime to $N$, and keep repeating Shor's algorithm until we do find an even $s$. We do the same if $a^{s/2}\!\pm\!1$ is itself a multiple of $N$. ## References 1. J.S. Neergaard-Nielsen, *Quantum information: lectures notes*, 2021, unpublished. 2. S. Aaronson, *Introduction to quantum information science: lecture notes*, 2018, unpublished.