Let $ \mathcal{O} $ be an order in an imaginary quadratic field $ K $, and Pic$ (\mathcal{O}) $ be the Picard group of $ \mathcal{O} $, i.e. the group of isomorphism classes of invertible fractional $ \mathcal{O} $-ideals under multiplication. The Hilbert class polynomial $ H_{\mathcal{O}}(x) $ attached to $ \mathcal{O} $ is defined to be
where $ [ \mathfrak{a}] $ denotes the isomorphism class of the invertible fractional $ \mathcal{O} $-ideal $ \mathfrak{a} $, and $ \mathfrak{j}( \mathbb{C}/\mathfrak{a}) $ stands for the $ \mathfrak{j} $-invariant of the complex elliptic curve $ \mathbb{C}/\mathfrak{a} $. It is well known that $ H_{\mathcal{O}}(x) $ has integral coefficients, and it is irreducible over $ \mathbb{Q} $ (see [3, §13] and [12, Chapter 10, App., p.144]).
Let $ p\in \mathbb{N} $ be a prime number, and $ \widetilde{H}_\mathcal{O}(x)\in \mathbb{F}_p[x] $ be the polynomial obtained by reducing $ H_\mathcal{O}(x)\in \mathbb{Z}[x] $ modulo $ p $. Suppose that $ p $ is non-split in $ K $ so that the roots of $ \widetilde{H}_\mathcal{O}(x) $ are supersingular $ \mathfrak{j} $-invariants, which are known to lie in $ \mathbb{F}_{p^2} $. It's natural to ask how many of them are actually in $ \mathbb{F}_p $. Castryck, Panny, and Vercauteren answered this question in [4, Theorem 26] for special cases when $ p\equiv 3 \!\! \pmod{4} $, $ K $ is of the form $ \mathbb{Q}(\sqrt{-l}) $ with $ l $ prime, $ l<(p+1)/4 $ and $ \mathcal{O} $ is an order containing $ \sqrt{-l} $. Their method as in [4, Section 5.2] counts the $ \mathbb{F}_p $-roots by constructing supersingular elliptic curves over $ \mathbb{F}_p $. We take a different approach here by reinterpreting the $ \mathbb{F}_p $-roots in terms of quaternion orders, which allows us to answer the question in more generality.
Our main result is as follows.
Theorem 1.1 Let $ K $ be an imaginary quadratic field and $ \mathcal{O} $ be an order in $ K $. Let $ p $ be a prime inert in $ K $ and strictly greater than $ | \operatorname{disc}({\mathcal{O}})| $, and $ \mathcal{H}_p $ be set of $ \mathbb{F}_p $-roots of $ \widetilde{H}_\mathcal{O}(x) $. If $ \mathcal{H}_p $ is nonempty, then it admits a regular (i.e. free and transitive) action by the $ 2 $-torsion subgroup Pic$ (\mathcal{O})[2]\subset \hbox{Pic}(\mathcal{O}) $. In particular, the number of $ \mathbb{F}_p $-roots of $ \widetilde{H}_\mathcal{O}(x) $ is either zero or $ |\hbox{Pic}(\mathcal{O})[2]| $.
Moreover, $ \mathcal{H}_p \neq \emptyset $ if and only if for every prime factor $ \ell $ of $ \operatorname{disc}( \mathcal{O}) $, either condition (i) or (ii) below holds for $ \ell $ depending on its parity:
${ \rm(i)} $ $ \ell \neq 2 $ and the Legendre symbol $ \frac{{-p}}{\ell}=1 $;
${\rm(ii)} $ $ \ell = 2 $ and one of the following conditions holds:
(a) $ p \equiv 7 \!\! \pmod{8} $;
(b) $ -p + \frac{ \operatorname{disc}(\mathcal{O})}{4} \equiv 0, \, 1 $ or $ 4 \!\! \pmod{8} $;
(c) $ -p + \operatorname{disc}(\mathcal{O})\equiv 1 \!\! \pmod{8} $.
The assumption that $ | \operatorname{disc}( \mathcal{O})| < p $ immediately implies that $ p $ does not divide the discriminant of $ H_\mathcal{O}(x) $ by an influential work of Gross and Zagier [8]. Therefore, $ \widetilde{H}_\mathcal{O}(x) $ has no repeated roots. We provide an alternative proof of this fact under the current assumptions in Corollary 2.7.
Remark 1.2 After the first of version of this manuscript appeared on the web, Jianing Li kindly informed us that a similar result to Theorem (1.1) has firstly been obtained in [25, Theorem 1.1] under the assumption that $ | \operatorname{disc}( \mathcal{O})|<4\sqrt{p/3} $. Moreover, Li et al. used a method similar to [25] and generalized it much further in a joint work [13]. Their result is as follows. Let $ \mathfrak{j}_0= \mathfrak{j}( \mathbb{C}/ \mathcal{O}) $, and put $ L:= \mathbb{Q}( \mathfrak{j}_0) $. If $ p $ coprime to the index $ [ \mathcal{O}_L: \mathbb{Z}[ \mathfrak{j}_0]) $ (e.g. if $ p\nmid \operatorname{disc}( \mathcal{O}) $), then they completely determined the factorization of $ \widetilde{H}_ \mathcal{O}(x) $ in $ \mathbb{F}_p[x] $. Partial results are also obtained without the co-primality condition. In particular, the results of Theorem (1.1) has been covered in [13, Theorem 4.1]. On the other hand, the current project was initiated in May 2021 during an online discussion between the authors. Unaware of the significant progress made by aforementioned works, we worked independently and obtained Theorem (1.1) by a completely different method: we count the $ \mathbb{F}_p $-roots by demonstrating a regular action using quaternion orders, whereas the aforementioned works count by studying the factorization of $ p $ in $ L $.
For the reader's convenience, we reproduce the celebrated formula of Gauss on the order of Pic$ (\mathcal{O}) $[2].
Theorem 1.3 [3, Proposition 3.11] Let $ r $ be the number of odd primes dividing $ \operatorname{disc}(\mathcal{O}) $. Define the number $ \mu $ as follows: if $ \operatorname{disc}(\mathcal{O})\equiv 1 \!\! \pmod{4} $, then $ \mu = r $, and if $ \operatorname{disc}(\mathcal{O})\equiv 0 \!\! \pmod{4} $, then $ \operatorname{disc}(\mathcal{O}) = -4n $, where $ n>0 $, and $ \mu $ is determined as follows:
Then $ |\hbox{Pic}( \mathcal{O})[2]|=2^{\mu-1} $.
This paper is organized as follows. In section 2, we give a reinterpretation of $ \mathcal{H}_p $ in terms of quaternion orders. In section 3, we show that there is a regular action of Pic$ (\mathcal{O})[2] $ on $ \mathcal{H}_p $ whenever $ \mathcal{H}_p\neq \emptyset $, and provide a nonemptiness criterion for $ \mathcal{H}_p $. Throughout the paper, the prime $ p\in \mathbb{N} $ is assumed to be non-split in $ K $. The notation $ B_{p, \infty} $ is reserved for the unique quaternion $ \mathbb{Q} $-algebra ramified precisely at $ p $ and infinity. Given a set $ X $ and an equivalence relation on $ X $, the equivalence class of an element $ x\in X $ is denoted by $ [x] $.
Acknowledgements The first author would like to thank WIN5 research group members Sarah Arpin, Kristin Lauter, Renate Scheidler, Katherine Stange and Ha Tran for helpful comments on some initial ideas on this problem. The first author would also like to thank Kiran Kedlaya and Rachel Pries for helpful suggestions, and Nandagopal Ramachandran for helpful discussions. The second author thanks Chia-Fu Yu for helpful discussions.
As mentioned before, we are going to reinterpret the $ \mathbb{F}_p $-roots of $ \widetilde{H}_\mathcal{O}(x) $ in terms of quaternion orders. For this purpose, we first describe more concretely the reduction of singular moduli with complex multiplication by $ \mathcal{O} $. Assume that the prime $ p $ is non-split in $ K $. For the moment, we make no assumption on the discriminant of the order $ \mathcal{O}\subset K $.
Let $ \mathcal{E}\ell\ell( \mathcal{O}) $ be the set of isomorphism classes of elliptic curves over $ \overline{\mathbb{Q}} $ with complex multiplication by $ \mathcal{O} $. It is canonically identified with the singular $ \mathfrak{j} $-invariants with complex multiplication by $ \mathcal{O} $ (i.e. the roots of $ H_ \mathcal{O}(x)\in \mathbb{Z}[x] $). The Picard group Pic$ ( \mathcal{O}) $ acts regularly on $ \mathcal{E}\ell\ell( \mathcal{O}) $ via $ \mathfrak{a} $-transformation [19, §7] and [16, §1]:
More concretely, if we pick $ \mathfrak{a} $ to be an integral ideal of $ \mathcal{O} $ and write $ E[ \mathfrak{a}] $ for the finite group scheme $ \cap_{a\in \mathfrak{a}}E[a] $, then $ E^ \mathfrak{a}=E/E[ \mathfrak{a}] $ by [24, Corollary A.4]. Here $ E[a]=\ker(E\xrightarrow{a}E) $. See [16, Proposition 1.26] and [24, Appendix] for the functorial characterization of $ E^ \mathfrak{a} $. Alternatively, since $ \mathfrak{a} $ is an invertible $ \mathcal{O} $-ideal, $ E^ \mathfrak{a} $ can also be identified canonically with the Serre tensor construction $ \mathfrak{a}^{-1}\otimes_ \mathcal{O} E $ (see [1, §1] and [2, §1.7.4]). Fix a member $ E_0\in \mathcal{E}\ell\ell( \mathcal{O}) $. The regular action in (2.1) gives rise to a Pic$ ( \mathcal{O}) $-equivariant bijection $ \xi: \mathcal{E}\ell\ell( \mathcal{O})\to \hbox{Pic}( \mathcal{O}) $ that sends $ E_0 $ to the identity element $ [ \mathcal{O}]\in \hbox{Pic}( \mathcal{O}) $.
Similarly, let $ \mathcal{E}\ell\ell_{/\mathbb{\bar F}_{p}}^{ss} $ be the set of isomorphism classes of supersingular elliptic curves over $ \overline{\mathbb{F}}_p $, which is canonically identified with the set of supersingular $ \mathfrak{j} $-invariants in $ \mathbb{F}_{p^2} $. From [20, Theorem V.3.1], an elliptic curve $ E/ \overline{\mathbb{F}}_p $ is supersingular if and only if its endomorphism algebra $ \operatorname{End}^0(E):= \operatorname{End}(E)\otimes \mathbb{Q} $ is a quaternion $ \mathbb{Q} $-algebra. Assume that this is the case. Then $ \operatorname{End}^0(E) $ coincides with the unique quaternion $ \mathbb{Q} $-algebra $ B_{p, \infty} $ ramified precisely at $ p $ and infinity, and $ \operatorname{End}(E) $ is a maximal order in $ \operatorname{End}^0(E) $ by [24, Theorem 4.2]. For simplicity, put $ B:=B_{p, \infty} $ and let $ \operatorname{Typ}(B) $ be the type set of $ B $, that is, the set of isomorphism (i.e. $ B^\times $-conjugacy) classes of maximal orders in $ B $. We obtain the following canonical map, which is known to be surjective [23, Corollary 42.2.21]:
Let $ \mathcal{R} $ be a maximal order in $ B $, and Cl$ ( \mathcal{R}) $ be its left ideal class set, that is, the set of isomorphism (i.e. right $ B^\times $-equivalent) classes of fractional left ideals of $ \mathcal{R} $ in $ B $. Given a fractional left ideal $ I $ of $ \mathcal{R} $, we write $ \mathcal{R}_r(I) $ for the right order of $ I $, which is defined as follows:
Sending a fractional left $ \mathcal{R} $-ideal to its right order induces a surjective map
The Deuring correspondence [23, Corollary 42.3.7] establishes a bijection between Cl$ ( \mathcal{R}) $ and $ \mathcal{E} \ell \ell_{/ \mathbb{\bar F}_{p}}^{s s} $. One direction of this correspondence goes as follows. From the surjectivity of $ \rho $, we may always fix $ E_ \mathcal{R}\in \mathcal{E} \ell \ell_{/ \mathbb{\bar F}_{p}}^{s s} $ such that $ \operatorname{End}(E_ \mathcal{R})= \mathcal{R} $. Then the member of $ \mathcal{E} \ell \ell_{/ \mathbb{\bar F}_{p}}^{s s} $ corresponding to a left ideal class $ [I]\in $ Cl$ ( \mathcal{R}) $ is the $ I $-transform $ E_ \mathcal{R}^I $ of $ E_ \mathcal{R} $. If $ I $ is chosen to be an integral left ideal of $ \mathcal{R} $, then $ E_ \mathcal{R}^I $ can be identified with the quotient $ E_ \mathcal{R}/E_ \mathcal{R}[I] $ by [24, Corollary A.4] again. From [23, Corollary 42.3.7], we have
Let $ \mathfrak{P} $ be a place of $ \overline{\mathbb{Q}} $ lying above $ p $, and $ r_ \mathfrak{P}: \mathcal{E}\ell\ell( \mathcal{O})\to \mathcal{E} \ell \ell_{/ \mathbb{\bar F}_{p}}^{s s} $ be the reduction map modulo $ \mathfrak{P} $. For each $ E\in \mathcal{E}\ell\ell( \mathcal{O}) $, we write $ \widetilde{E} $ for the reduction of $ E $ modulo $ \mathfrak{P} $. From [12, §9.2], reducing $ E_0 $ modulo $ \mathfrak{P} $ gives rise to an embedding $ \iota: \mathcal{O}\hookrightarrow \mathcal{R}_0:= \operatorname{End}( \widetilde{E}_0) $. By an abuse of notation, we still write $ \iota $ for both of the following two induced maps:
For simplicity, we identify $ K $ with its image in $ B $ via $ \iota $ and write $ \mathcal{R}_0 \mathfrak{a} $ for $ \mathcal{R}_0 \iota( \mathfrak{a}) $.
Now we are ready to give a concrete description of $ r_ \mathfrak{P}:\mathcal{E}\ell\ell( \mathcal{O})\to \mathcal{E} \ell \ell_{/ \mathbb{\bar F}_{p}}^{s s} $.
Proposition 2.1 DDDIANDDThe reduction map $ r_ \mathfrak{P} $ fits into a commutative diagram asfollows:
Here $ \xi $ is the Pic$ ( \mathcal{O}) $-equivariant bijection that sendsthe fixed member $ E_0\in \mathcal{E}\ell\ell( \mathcal{O}) $ to$ [ \mathcal{O}]\in $ Pic$ ( \mathcal{O}) $, and $ \delta $ is the Deuring correspondence obtained by taking $ E_{ \mathcal{R}_0}= \widetilde{E}_0 $.
Proof According to [19, Proposition 15, §11], $ \mathfrak{a} $-transforms are preserved under good reductions1 This implies that for every $ [ \mathfrak{a}]\in $ Pic$ ( \mathcal{O}) $, we have
1A priori, the statement of [19,Proposition 15,§11] requires that $ \mathcal{O}= \mathcal{O}_K $, the maximal order of $ K $. Nevertheless, the result here holds for general $ \mathcal{O} $ here since $ \mathfrak{a} $ is an invertible $ \mathcal{O} $-ideal by our assumption.
so the left square commutes. The right triangle commutes because of (2.4).
Corollary 2.2 For any $ [ \mathfrak{a}]\in \rm{Pic}( \mathcal{O}) $, we have $ \operatorname{End}(\widetilde{E_0^ \mathfrak{a}})\simeq \mathfrak{a}^{-1} \mathcal{R}_0 \mathfrak{a} $.
Proof This follows directly from Proposition (2.1) since the right order of $ \mathcal{R}_0 \mathfrak{a} $ is precisely $ \mathfrak{a}^{-1} \mathcal{R}_0 \mathfrak{a} $.
Remark 2.3 Let $ \mathcal{O}_K $ be the ring of integers of $ K $, and $ f $ be the conductor of $ \mathcal{O} $ so that $ \mathcal{O}= \mathbb{Z}+f \mathcal{O}_K $. Write $ f=p^mf' $ with $ p\nmid f' $, and put $ \mathcal{O}':= \mathbb{Z}+f' \mathcal{O}_K $. According to [17, Lemma 3.1], $ \iota(K)\cap \mathcal{R}_0= \iota( \mathcal{O}') $. For any invertible fractional ideal $ \mathfrak{a} $ of $ \mathcal{O} $, we have $ \mathcal{R}_0 \mathfrak{a}=( \mathcal{R}_0 \mathcal{O}') \mathfrak{a}= \mathcal{R}_0( \mathcal{O}' \mathfrak{a}) $. It follows that the map $ \iota: \rm{Pic}( \mathcal{O})\to \rm{CI}( \mathcal{R}_0) $ factors through the following canonical homomorphism
From this, one easily deduces that $ \widetilde{H}_ \mathcal{O}(x)=( \widetilde{H}_{ \mathcal{O}'}(x))^{|\ker(\varpi)|} $.
Now assume that $ \mathcal{O} $ is maximal at $ p $ (i.e. $ p\nmid f $). From Remark 2.3, $ \iota: \mathcal{O}\to \mathcal{R}_0 $ is an {optimal embedding} of $ \mathcal{O} $ into $ \mathcal{R}_0 $, that is, $ \iota(K)\cap \mathcal{R}_0= \iota( \mathcal{O}) $. Given an arbitrary maximal order $ \mathcal{R} $ of $ B $, we write $ \rm{Emb}( \mathcal{O}, \mathcal{R}) $ for the set of optimal embeddings of $ \mathcal{O} $ into $ \mathcal{R} $. The unit group $ \mathcal{R}^\times $ acts on $ \rm{Emb}( \mathcal{O}, \mathcal{R}) $ by conjugation, and there are only finitely many orbits. Put $ m( \mathcal{O}, \mathcal{R}, \mathcal{R}^\times):=| \mathcal{R}^\times \backslash \rm{Emb}( \mathcal{O}, \mathcal{R})| $, the number of $ \mathcal{R}^\times $-conjugacy clasess of optimal embeddings from $ \mathcal{O} $ into $ \mathcal{R} $. We recall below a precise formula by Elkies, Ono and Yang for the cardinality of each fiber of the reduction map $ r_ \mathfrak{P}:\mathcal{E}\ell\ell( \mathcal{O})\to \mathcal{E} \ell \ell_{/ \mathbb{\bar F}_{p}}^{s s} $.
Lemma 2.4 [7, Lemma 3.3] Suppose that $ \mathcal{O} $ is maximal at $ p $. Then for any member $ E\in \mathcal{E} \ell \ell_{/ \mathbb{\bar F}_{p}}^{s s} $, we have
where $ \mathcal{R}= \operatorname{End}(E) $, and $ \varepsilon=1/2 $ or $ 1 $ according as $ p $ is inert or ramified in $ K $.
A priori, [7, Lemma 3.3] is only stated for the maximal order $ \mathcal{O}_K $. Nevertheless, the same proof there applies more generally to quadratic orders maximal at $ p $. Alternatively, using Proposition (2.1) and the Deuring lifting theorem2 [12, Theorem 14, §13.5] [8, Proposition 2.7], one easily sees that Lemma 2.4 is equivalent to the following purely arithmetic result, whose independent proof will be left for the interested reader.
2Here the Deuring lifting theorem guarantees that the optimal embedding $ \iota: \mathcal{O}\to \mathcal{R}_0 $ is "non-special'', that is, every optimal embedding $ \varphi: \mathcal{O}\to \mathcal{R} $ is realizable as $ \operatorname{End}(E)\to \operatorname{End}( \widetilde{E}) $ for some $ E\in \mathcal{E}\ell\ell( \mathcal{O}) $.
Lemma 2.5 Keep $ \mathcal{O} $ and $ \varepsilon $ as in Lemma 2.4. Let $ \mathcal{R} $ be a maximal order in $ B $, and $ \varphi: \mathcal{O}\hookrightarrow \mathcal{R} $ be an optimal embedding. Denote the induced map $ \rm{Pic}( \mathcal{O})\to \rm{CI}( \mathcal{R}) $ by $ \varphi $ as well. Then for each $ [I]\in \rm{CI}( \mathcal{R}) $, we have
We immediately obtain the following corollaries from Lemma 2.4.
Corollary 2.6 Suppose that $ \mathcal{O} $ is maximal at $ p $. The $ \mathfrak{j} $-invariant of a supersingluar elliptic curve $ E/ \overline{\mathbb{F}}_p $ is a root of $ \widetilde{H}_\mathcal{O}(x) $ if and only if $ \mathcal{O} $ can be optimally embedded into $ \operatorname{End}(E) $.
This matches well with Corollary 2.2. Indeed, a classical result of Chevalley, Hasse and Noether [9, §4] says that any maximal order of $ B $ that contains a copy of $ \mathcal{O} $ optimally is isomorphic to $ \mathfrak{a}^{-1} \mathcal{R}_0 \mathfrak{a} $ for some $ [ \mathfrak{a}]\in $ Pic$ ( \mathcal{O}) $.
Corollary 2.7 If $ p>| \operatorname{disc}( \mathcal{O})| $, then the reduction map $ r_ \mathfrak{P}:\mathcal{E}\ell\ell( \mathcal{O})\to \mathcal{E} \ell \ell_{/ \mathbb{\bar F}_{p}}^{s s} $ is injective. In particular, $ \widetilde{H}_ \mathcal{O}(x) $ has no repeated roots.
We give a simple proof that is independent of the result of Gross and Zagier [8].
Proof Since $ p $ does not split in $ K $ and is strictly greater than $ |\hbox{disc}( \mathcal{O})| $, it is necessarily inert in $ K $. From Lemma 2.4, it suffices to show that $ |\hbox{Emb}( \mathcal{O}, \mathcal{R})|\leq 2 $ for any maximal order $ \mathcal{R} $ in $ B $. Since $ p>| \operatorname{disc}( \mathcal{O})| $, Kaneko's inequality [10, Theorem 2'] forces any two optimal embeddings $ \varphi, \varphi': \mathcal{O}\to \mathcal{R} $ to have the same image. On the other hand, $ \varphi $ and $ \varphi' $ share the same image if and only if $ \varphi'=\varphi $ or $ \bar\varphi $, the complex conjugate of $ \varphi $. The desired inequality $ |\hbox{Emb}( \mathcal{O}, \mathcal{R})|\leq 2 $ follows immediately.
Remark 2.8 In another direction, Elkies, Ono and Yang [7, Theorem 1.4] showed that there exists a bound $ N_p $ such that the reduction map $ r_ \mathfrak{P}:\mathcal{E}\ell\ell( \mathcal{O}_K)\to \mathcal{E} \ell \ell_{/ \mathbb{\bar F}_{p}}^{s s} $ is surjective whenever $ | \operatorname{disc}( \mathcal{O}_K)|>N_p $. This bound is first effectivized by Kane [11] conditionally upon the generalized Riemann hypothesis. Liu et al. further improved this bound in [14, Corollary 1.3].
Let us return to the task of interpreting $ \mathbb{F}_p $-roots of $ \widetilde{H}_ \mathcal{O}(x)\in \mathbb{F}_p[x] $ in terms of maximal orders in $ B $. For the rest of this section, we keep the additional assumption that $ p>| \operatorname{disc}( \mathcal{O})| $. We recall from [6, Proposition 2.4] a classical result on supersingular elliptic curves in characteristic $ p $.
Lemma 2.9 Let $ p>3 $ and let $ E $ be a supersingular elliptic curve over $ \overline{\mathbb{F}}_p $. Then $ \mathfrak{j}(E)\in \mathbb{F}_p $ if and only if there exists $ \psi\in \operatorname{End}(E) $ such that $ \psi^2 = -p $.
Recall that $ \mathcal{H}_p $ denotes the set of $ \mathbb{F}_p $-roots of $ \widetilde{H}_ \mathcal{O}(x) $, which can be identified canonically with a subset of $ \mathcal{E} \ell \ell_{/ \mathbb{\bar F}_{p}}^{s s} $.
Lemma 2.10 The map $ \rho: \mathcal{E} \ell \ell_{/ \mathbb{\bar F}_{p}}^{s s}\to \operatorname{Typ}(B) $ in (2.2) induces a bijection between $ \mathcal{H}_p $ and the following subset $ \mathcal{T}_p\subseteq \operatorname{Typ}(B) $:
Proof Combining Corollary 2.6 and Lemma 2.9, we see that $ \rho( \mathcal{H}_p)= \mathcal{T}_p $. Now it follows from [23, Lemma 42.4.1] that $ \rho: \mathcal{H}_p\to \mathcal{T}_p $ is injective, and hence bijective.
We give another characterization of $ \mathcal{T}_p $ by presenting the quaternion algebra $ B=B_{p, \infty} $ more concretely. Let $ d\in \mathbb{N} $ be the unique square-free positive integer such that $ K= \mathbb{Q}(\sqrt{-d}) $. The assumption that $ p $ is inert in $ K $ amounts to the equality $ \genfrac{(}{)}{}{}{{-d}}{{p}}=-1 $. Let $ \left( \frac{-d, -p}{ \mathbb{Q}} \right) $ be the quaternion $ \mathbb{Q} $-algebra with standard basis $ \{1, i, j, k\} $ such that
We identify $ K= \mathbb{Q}(\sqrt{-d}) $ with $ \mathbb{Q}(i) $, and $ \mathcal{O} $ with the corresponding order in $ \mathbb{Q}(i) $. Put $ \Lambda:= \mathcal{O}+j \mathcal{O} $, which is an order (of full rank) in the above quaternion algebra. Consider the following finite set of maximal orders:
Here the superscript "opt" stands for "$ \mathcal{O} $-optimal".
Proposition 2.11 Let $ \mathcal{R} $ be a maximal order in $ B $. We have $ [ \mathcal{R}]\in \mathcal{T}_p $ if and only if $ \mathcal{R}\simeq \mathscr{R} $ for some $ \mathscr{R}\in S^ \mathrm{opt} $. In particular, $ \mathcal{H}_p\neq \emptyset $ if and only if $ \left( \frac{-d, -p}{ \mathbb{Q}} \right)\simeq B $ and $ S^ \mathrm{opt}\neq \emptyset $.
Proof Clearly, if $ \mathcal{R}\simeq \mathscr{R} $ for some $ \mathscr{R}\in S^ \mathrm{opt} $, then $ [ \mathcal{R}]\in \mathcal{T}_p $. Conversely, suppose that $ [ \mathcal{R}]\in \mathcal{T}_p $, that is, $ \mathcal{R} $ contains a copy of $ \mathcal{O} $ optimally, and there exists $ \alpha\in \mathcal{R} $ with $ \alpha^2=-p $. Then $ \mathcal{R}\alpha $ is the unique two sided prime ideal of $ \mathcal{R} $ lying above $ p $. From [22, Exercise I.4.6], $ \mathcal{R} $ is normalized by $ \alpha $, which implies that $ \mathcal{O}_\alpha:=\alpha\mathcal{O}\alpha^{-1} $ is still a quadratic order optimally embedded in $ \mathcal{R} $. If $ \mathcal{O}_\alpha\neq \mathcal{O} $, then $ | \operatorname{disc}( \mathcal{O})|\geq p $ by Kaneko's inequality [10, Theorem 2'], contradicting to our assumption that $ | \operatorname{disc}( \mathcal{O})|<p $. Thus $ \mathcal{O}_\alpha= \mathcal{O} $, and conjugation by $ \alpha $ induces an automorphism $ \sigma\in \rm{Aut}( \mathcal{O}) $. If $ \sigma $ is the identity, then $ \alpha $ lies in the centralizer of $ \mathcal{O} $ in $ B $, which is just $ K $. This contradicts to the assumption $ | \operatorname{disc}( \mathcal{O})|<p $ again. It follows that $ \sigma $ is the unique nontrivial automorphism of $ \mathcal{O} $, i.e. the complex conjugation. We conclude that $ \Lambda_ \mathcal{R}:= \mathcal{O}+\alpha \mathcal{O}\subset \mathcal{R} $ is isomorphic to $ \Lambda $, and $ B=\Lambda_ \mathcal{R}\otimes_ \mathbb{Z} \mathbb{Q}\simeq \Lambda\otimes_ \mathbb{Z} \mathbb{Q}= \left( \frac{-d, -p}{ \mathbb{Q}} \right) $. Consequently, $ \mathcal{R} $ is isomorphic to some member of $ S^ \mathrm{opt} $. The last statement follows from the bijection $ \mathcal{H}_p\simeq \mathcal{T}_p $ in Lemma 2.10.
Lemma 2.12 The isomorphism $ \left( \frac{-d, -p}{ \mathbb{Q}} \right)\simeq B $ holds if and only if $ \genfrac{(}{)}{}{}{{-p}}{{\ell}}=1 $ for every odd prime factor $ \ell $ of $ d $.
Proof For the moment, let $ \ell $ be either a prime number or $ \infty $. Write $ (-d, -p)_\ell $ for the Hilbert symbol of $ -d $ and $ -p $ relative to $ \mathbb{Q}_\ell $ (where $ \mathbb{Q}_\infty= \mathbb{R} $). From [22, Corollaire II.1.2], $ \left( \frac{-d, -p}{ \mathbb{Q}} \right) $ is split at $ \ell $ if and only if $ (-d, -p)_\ell=1 $. Clearly, $ (-d, -p)_\infty=-1 $.
Now assume that $ \ell $ is an odd prime. By our assumption, $ p $ is an odd prime satisfying $ \genfrac{(}{)}{}{}{{-d}}{{p}}=-1 $. From [18, Theorem 1, § III.1], we easily compute that
Therefore, if $ \left( \frac{-d, -p}{ \mathbb{Q}} \right)\simeq B $, then necessarily $ \genfrac{(}{)}{}{}{{-p}}{{\ell}}=1 $ for every odd prime factor $ \ell $ of $ d $.
Conversely, if $ \genfrac{(}{)}{}{}{{-p}}{{\ell}}=1 $ for every odd prime factor $ \ell $ of $ d $, then $ (-d, -p)_2=1 $ by the product formula [18, Theorem 2, § III.2]. Hence this condition is also sufficient for the isomorphism $ \left( \frac{-d, -p}{ \mathbb{Q}} \right)\simeq B $.
Throughout this section, we assume that $ p $ is inert in $ K= \mathbb{Q}(\sqrt{-d}) $ and strictly greater than $ | \operatorname{disc}( \mathcal{O})| $. Assume further that the quaternion $ \mathbb{Q} $-algebra $ \left( \frac{-d, -p}{ \mathbb{Q}} \right) $ is ramified precisely at $ p $ and infinity, for otherwise $ \mathcal{H}_p=\emptyset $. Denote $ \left( \frac{-d, -p}{ \mathbb{Q}} \right) $ simply by $ B $ henceforth and let $ \{1, i, j, k\} $ be the standard basis of $ B $ as in (2.7). We identify $ K $ with the subfield $ \mathbb{Q}(i) $ of $ B $. Then conjugation by $ j $ stabilizes $ K $ and sends each $ x\in K $ to its complex conjugate $ \bar{x} $. Let $ \Lambda= \mathcal{O}+j \mathcal{O} $, and $ S^ \mathrm{opt} $ be the set of maximal orders in (2.8).
First, we assume that $ \mathcal{H}_p\neq \emptyset $ and exhibit a regular action of $ \rm{Pic}( \mathcal{O})[2] $ on $ \mathcal{H}_p $. Since the reduction map $ r_ \mathfrak{P}: \mathcal{E}\ell\ell( \mathcal{O})\to \mathcal{E} \ell \ell_{/ \mathbb{\bar F}_{p}}^{s s} $ is injective by Corollary 2.7, the regular action of $ \rm{Pic}( \mathcal{O}) $ on $ \mathcal{E}\ell\ell( \mathcal{O}) $ induces a regular action of $ \rm{Pic}( \mathcal{O}) $ on the image $ r_ \mathfrak{P}( \mathcal{E}\ell\ell( \mathcal{O})) $ (or equivalently, on the full set of roots of $ \widetilde{H}_ \mathcal{O}(x) $). We show that this action restricts to a regular $ \rm{Pic}( \mathcal{O})[2] $-action on $ \mathcal{H}_p $.
Proposition 3.1 Let $ E_0\in \mathcal{E}\ell\ell( \mathcal{O}) $ be a member satisfying $ \mathfrak{j}( \widetilde{E}_0)\in \mathbb{F}_p $. Given $ [ \mathfrak{a}]\in \rm{Pic}( \mathcal{O}) $, we have $ \mathfrak{j}(\widetilde{E_0^ \mathfrak{a}})\in \mathbb{F}_p $ if and only if $ [ \mathfrak{a}] $ is a 2-torsion. In particular, $ \rm{Pic}( \mathcal{O})[2] $ acts regularly on $ \mathcal{H}_p $.
Proof Put $ \mathcal{R}_0: = \operatorname{End}( \widetilde{E}_0) $ and $ \mathcal{R}:= \mathfrak{a}^{-1} \mathcal{R}_0 \mathfrak{a} $ so that $ \operatorname{End}(\widetilde{E_0^ \mathfrak{a}})\simeq \mathcal{R} $ by Corollary 2.2. From Lemma 2.9, it is enough to show that there exists $ \alpha\in \mathcal{R} $ with $ \alpha^2=-p $ if and only if $ [ \mathfrak{a}]\in \rm{Pic}( \mathcal{O})[2] $. By Proposition 2.11, we may assume that $ \mathcal{R}_0\in S^ \mathrm{opt} $, that is, $ \mathcal{R}_0 $ is a maximal order in $ B $ satisfying $ \mathcal{R}_0\supseteq \mathcal{O}+j \mathcal{O} $ and $ \mathcal{R}_0\cap K= \mathcal{O} $. Then
First, suppose that $ [ \mathfrak{a}]\in \rm{Pic}( \mathcal{O})[2] $. Then $ \mathfrak{a}^{-1}\bar{\mathfrak{a}} = \mathcal{O} a $ for some $ a\in K^\times $. Moreover, $ N_{K / \mathbb{Q}}(\mathcal{O} a)=N_{K / \mathbb{Q}}\left(\mathfrak{a}^{-1} \overline{\mathfrak{a}}\right)=\mathbb{Z} $, so $ N_{K/ \mathbb{Q}}(a)=1 $. Therefore $ \alpha: =aj\in \mathcal{R} $ satisfies that $ \alpha^2 = a\bar{a}j^2 = -p $.
Conversely, suppose that $ \alpha \in \mathcal{R} $ is an element satisfying $ \alpha^2=-p $. From the proof of Proposition 2.11, we must have $ \alpha x= \bar{x}\alpha $ for every $ x\in \mathcal{O} $. Thus $ j^{-1}\alpha $ centralizes $ \mathcal{O} $, so there exists $ a\in K^\times $ such that $ \alpha=ja $. Moreover, $ N_{K/ \mathbb{Q}}(a)=1 $ since $ \alpha^2=j^2\bar{a}a $. Now we have
We claim that $\mathcal{R} \supset \mathfrak{a}^{-1} \overline{\mathfrak{a}}$. It suffices to show that $ \mathcal{R}_{\ell} \supset a \mathfrak{a}_{\ell}^{-1} \overline{\mathfrak{a}}_{\ell}$ for every prime $ \ell\in \mathbb{N} $, where the subscript $ _\ell $ indicates $ \ell $-adic completion at $ \ell $. If $ \ell\neq p $, then $ (-p)\in \mathcal{R}_\ell^\times $, so the containment follows directly from (3.2). If $ \ell=p $, then $ \mathcal{R}_p $ coincides with the unique maximal order of the division quaternion $ \mathbb{Q}_p $-algebra $ B_p $. More concretely, $ \mathcal{R}_p=\{z\in B_p\mid \operatorname{nrd}(z)\in \mathbb{Z}_p\} $, where $ \operatorname{nrd}(z) $ denotes the reduced norm of $ z\in B_p $. On the other hand, for any $ x_p\in \mathfrak{a}_p^{-1} $ and $ y_p\in \mathfrak{a}_p $, we have $ x_py_p\in \mathcal{O}_p $, and hence $ \operatorname{nrd}(a x_p\bar{y}_p)= \operatorname{nrd}(x_p) \operatorname{nrd}(\bar{y}_p)= \operatorname{nrd}(x_py_p)\in \mathbb{Z}_p $. Since $ \mathfrak{a}_{p}^{-1} \overline{\mathfrak{a}}_{p}$ is generated by elements of the form $ x_p\bar{y}_p $, it follows that $ \mathcal{R}_{\ell} \supset a \mathfrak{a}_{p}^{-1} \overline{\mathfrak{a}}_{p} $. The claim is verified. Now $a \mathfrak{a}^{-1} \overline{\mathfrak{a}} \subseteq \mathcal{R} \cap K=\mathcal{O}$, which implies that $ a \overline{\mathfrak{a}} \subseteq \mathfrak{a} $. Comparing discriminants on both sides, we get $\operatorname{disc}(a \overline{\mathfrak{a}})=N_{K / \mathbb{Q}}(a)^{2} \operatorname{disc}(\overline{\mathfrak{a}})=\operatorname{disc}(\mathfrak{a}) $. Therefore, $a \overline{\mathfrak{a}}=\mathfrak{a} $, so $ [ \mathfrak{a}]\in \rm{Pic}( \mathcal{O})[2] $.
Now we drop the assumption that $ \mathcal{H}_p\neq \emptyset $ and derive a non-emptiness criterion for $ \mathcal{H}_p $. From Proposition 2.11, $ \mathcal{H}_p\neq \emptyset $ if and only if $ S^ \mathrm{opt}\neq \emptyset $ (as we have already assumed that $ \left( \frac{-d, -p}{ \mathbb{Q}} \right)\simeq B_{p, \infty} $). For each prime $ \ell\in \mathbb{N} $, let us put
The local-global correspondence of lattices [5, Proposition 4.21] establishes a bijection between $ S^{\rm{opt}} $ and $ \prod_\ell S_\ell^{\rm{opt}} $, where the product runs over all prime $ \ell $. Since the reduced discriminant of $ B $ is $ p $ and the reduced discriminant of $ \Lambda $ is $ p \operatorname{disc}( \mathcal{O}) $ by [15, Lemmas 2.7 and 2.9], $ \Lambda $ is maximal at every prime $ \ell $ coprime to $ \operatorname{disc}( \mathcal{O}) $. Moreover, for each such $ \ell $, the maximal order $ \Lambda_\ell $ automatically satisfies the condition $ \Lambda_\ell\cap K_\ell = \mathcal{O}_\ell $ by its definition $ \Lambda_\ell= \mathcal{O}_\ell+j \mathcal{O}_\ell $. Hence for $ \ell \nmid \operatorname{disc}( \mathcal{O}) $, the set $ S_\ell^ \mathrm{opt} $ has a single element $ \Lambda_\ell $, and the bijection above simplifies as
Lemma 3.2 Let $ \ell $ be a prime factor of $ \operatorname{disc}( \mathcal{O}) $. Then $ S_\ell^ \mathrm{opt} \neq \emptyset $ if and only if $ -p\in N_{K/ \mathbb{Q}}( \mathcal{O}_\ell^\times) $. Moreover, if $ S_\ell^ \mathrm{opt}\neq \emptyset $, then there is a regular action of $ H^1(K/ \mathbb{Q}, \mathcal{O}_\ell^\times) $ on $ S_\ell^ \mathrm{opt} $, so any fixed member of $ S_\ell^ \mathrm{opt} $ gives rise to a bijection $ S_\ell^ \mathrm{opt}\simeq H^1(K/ \mathbb{Q}, \mathcal{O}_\ell^\times) $.
The Galois cohomological description of $ S_\ell^ \mathrm{opt} $ is nice to know but not used elsewhere in this paper.
Proof By our assumption, $ \operatorname{disc}( \mathcal{O}) $ is coprime to $ p $, so $ B $ splits at the prime $ \ell $. This allows us to identify $ B_\ell $ with the matrix algebra $ M_2( \mathbb{Q}_\ell) $. Let $ V_\ell= \mathbb{Q}_\ell^2 $ be the unique simple $ B_\ell $-module. Every maximal order $ \mathcal{R}_\ell $ in $ B_\ell $ is of the form $ \operatorname{End}_{ \mathbb{Z}_\ell}(L_\ell) $ for some $ \mathbb{Z}_\ell $-lattice $ L_\ell\subseteq V_\ell $, and $ L_\ell $ is uniquely determined by $ \mathcal{R}_\ell $ up to $ \mathbb{Q}_\ell^\times $-homothety. In other words, $ \operatorname{End}_{ \mathbb{Z}_\ell}(L_\ell) = \operatorname{End}_{ \mathbb{Z}_\ell}(L'_\ell) $ if and only if $ L_\ell =c L'_\ell $ for some $ c \in \mathbb{Q}_\ell^\times $. If $ \mathcal{R}_\ell \in S_\ell^ \mathrm{opt} $, then the inclusion $ \Lambda_\ell \subseteq \mathcal{R}_\ell $ puts a $ \Lambda_\ell $-module structure on $ L_\ell $. Moreover, the $ \Lambda_\ell $-lattice $ L_\ell $ is $ \mathcal{O}_{\ell \text {-optimal }}$ in the sense that $ \operatorname{End}_{ \mathbb{Z}_\ell}(L_\ell)\cap K_\ell = \mathcal{O}_\ell $. Conversely, if $ M_\ell $ is an $ \mathcal{O}_\ell $-optimal $ \Lambda_\ell $-lattice in $ V_\ell $, then $ \operatorname{End}_{ \mathbb{Z}_\ell}(M_\ell) $ is a member of $ S_\ell^ \mathrm{opt} $. We have established the following canonical bijection
Recall that $ \Lambda_\ell= \mathcal{O}_\ell+j \mathcal{O}_\ell $, where $ j^2 = -p $ and $ jx = \bar{x}j $ for any $ x\in \mathcal{O}_\ell $. If there exists $ a\in \mathcal{O}_\ell^\times $ satisfying $ a\bar{a}=-p $, then we can put a $ \Lambda_\ell $-module structure on $ \mathcal{O}_\ell $ as follows:
Since $ B_\ell=\Lambda_\ell\otimes_{ \mathbb{Z}_\ell} \mathbb{Q}_\ell $, this also puts a $ B_\ell $-module structure on $ K_\ell= \mathcal{O}_\ell\otimes_{ \mathbb{Z}_\ell} \mathbb{Q}_\ell $. Consequently, it identifies $ K_\ell $ with the unique simple $ B_\ell $-module $ V_\ell $, and in turn identifies $ \mathcal{O}_\ell $ with a $ \Lambda_\ell $-lattice $ L_\ell $ in $ V_\ell $. Necessarily, $ L_\ell $ is $ \mathcal{O}_\ell $-optimal since $ \operatorname{End}_{ \mathbb{Z}_\ell}(L_\ell)\cap K_\ell= \operatorname{End}_{ \mathcal{O}_\ell}(L_\ell)= \operatorname{End}_{ \mathcal{O}_\ell}( \mathcal{O}_\ell)= \mathcal{O}_\ell $. We have shown that if $ -p\in N_{K/ \mathbb{Q}}( \mathcal{O}_\ell^\times) $, then $ S_\ell^ \mathrm{opt}\neq \emptyset $.
Conversely, suppose that $ S_\ell^ \mathrm{opt}\neq \emptyset $ and let $ M_\ell $ be an $ \mathcal{O}_\ell $-optimal $ \Lambda_\ell $-lattice in $ V_\ell $. The inclusion $ \mathcal{O}_\ell\subset \Lambda_\ell $ equips $ M_\ell $ with an $ \mathcal{O}_\ell $-module structure satisfying $ \operatorname{End}_{ \mathcal{O}_\ell}(M_\ell)= \mathcal{O}_\ell $. Being a quadratic $ \mathbb{Z}_\ell $-order, $ \mathcal{O}_\ell $ is both Gorenstein and semi-local. It follows from [21, Characterization B 4.2] that $ M_\ell $ is a free $ \mathcal{O}_\ell $-module of rank one. Pick a basis $ e $ so that $ M_\ell = \mathcal{O}_\ell e $. Since $ M_\ell $ is at the same time a module over $ \Lambda_\ell $, we have $ je = ae $ for some $ a\in \mathcal{O}_\ell $. Necessarily, $ \bar{a}a=-p $ because
This also implies that $ a\in \mathcal{O}_\ell^\times $ since $ \ell\neq p $. Therefore, $ S_\ell^\rm{opt} \neq \emptyset $ if and only if $ -p\in N_{K/ \mathbb{Q}}( \mathcal{O}_\ell^\times) $.
Had we picked a different basis $ e' $ for $ M_\ell $, then $ e'=ue $ for some $ u\in \mathcal{O}^\times_\ell $. It follows that
Correspondingly, $ a $ is changed to $ u^{-1}\bar{u}a $. Therefore, we have defined the following map:
where $ a\sim a' $ if and only if there exists some $ u\in \mathcal{O}_\ell^\times $ such that $ a'=a(\bar{u}/u) $. We have already seen that $ \Phi $ is surjective. Suppose that $ \Phi([M_1])=\Phi([M_2]) $ for $ [M_r]\in \mathcal{M} $ with $ r=1, 2 $. By the above discussion, we can choose suitable $ \mathcal{O}_\ell $-base $ e_r $ for $ M_r $ such that they give rise to the same $ a\in \mathcal{O}_\ell^\times $. Then the $ \mathcal{O}_\ell $-linear map sending $ e_1 $ to $ e_2 $ defines a $ \Lambda_\ell $-isomorphism between $ M_1 $ and $ M_2 $. Since $ \rm{Aut}_{B_\ell}(V_\ell)= \mathbb{Q}_\ell^\times $, it follows that $ M_1 $ and $ M_2 $ are $ \mathbb{Q}_\ell^\times $-homothetic, so $ \Phi $ is injective as well.
Lastly, if the right hand side of (3.5) is nonempty, then it admits a regular action by $ H^1(K/ \mathbb{Q}, \mathcal{O}_\ell^\times)=\{b\in \mathcal{O}_\ell^\times \mid \bar{b}b=1\}/{\sim} $ via multiplication. The second part of the lemma follows by combining the bijections (3.4) and (3.5) with the above action.
Lemma 3.3 Let $ \ell $ be a prime factor of $ \operatorname{disc}( \mathcal{O}) $. Then $ -p\in N_{K/ \mathbb{Q}}( \mathcal{O}_\ell^\times) $ if and only if either condition (i) or (ii) below holds for $ \ell $ depending on its parity:
${ \rm(i)} $ $ \ell \neq 2 $ and $ \left(\frac{-p}{\ell}\right) = 1 $;
$ {\rm(ii)}$ $ \ell = 2 $ and one of the following conditions holds:
(b) $ -p + \frac{ \operatorname{disc}( \mathcal{O})}{4} \equiv 0, \, 1 $ or $ 4 \!\! \pmod{8} $;
(c) $ -p + \operatorname{disc}( \mathcal{O})\equiv 1 \!\! \pmod{8} $.
Proof For simplicity, put $ D:= \operatorname{disc}( \mathcal{O}) $ and $ \delta=\frac{1}{2}\sqrt{D} $. We claim that $ \mathcal{O}_\ell= \mathbb{Z}_\ell+ \mathbb{Z}_\ell\delta $. It is well known that $ \mathcal{O}= \mathbb{Z}+ \mathbb{Z}(D+\sqrt{D})/2 $. The claim is obviously true if $ 4|D $. If $ 4\nmid D $, then $ \ell\neq 2 $, so the claim is true in this case as well. Given an element $ a+b\delta\in \mathcal{O}_\ell $ with $ a, b\in \mathbb{Z}_\ell $, we have $ N_{K/ \mathbb{Q}}(a+b\delta)=a^2-b^2D/4 $. Therefore, $ -p \in N_{K/ \mathbb{Q}}(\mathcal{O}_\ell^\times) $ if and only if the equation
has a solution in $ \mathbb{Z}_\ell^2 $.
First, suppose that $ \ell $ is odd. Then equation (3.6) is solvable in $ \mathbb{Z}_\ell^2 $ if and only if $ \genfrac{(}{)}{}{}{{-p}}{{\ell}} = 1 $. Indeed, suppose $ \genfrac{(}{)}{}{}{{-p}}{{\ell}} = 1 $ so that $ -p $ is a square in $ \mathbb{F}_\ell $. By Hensel's lemma [23, Lemma 12.2.17], the equation $ x^2 = -p $ has a solution $ x_0\in \mathbb{Z}_\ell $. Hence $ (x_0, 0) $ is a solution of (3.6) in $ \mathbb{Z}_\ell^2 $. Conversely, suppose (3.6) has a solution $ (x_0, y_0)\in \mathbb{Z}_\ell^2 $. Reducing (3.6) modulo $ \ell $ shows that $ x_0 \!\! \pmod{\ell} $ is a square root of $ -p $ in $ \mathbb{F}_\ell $, i.e. $ \genfrac{(}{)}{}{}{{-p}}{{\ell}} = 1 $.
For the rest of the proof we assume that $ \ell=2 $, which implies that $ 4|D $. First, suppose that $ (x, y)\in \mathbb{Z}_2^2 $ is a solution of (3.6). Since $ x^2, \, y^2 \equiv 0, \, 1 \mbox{ or } 4 \!\! \pmod{8} $ and at least one of $ x, \, y $ lies in $ \mathbb{Z}_2^\times $ because $ p $ is odd, we see that the pair $ (x^2, y^2) $ takes on five possibilities modulo $ 8 $:
Each possibility puts the following respective constraint on $ p $ and $ D $:
We have proved the necessity part of the lemma for the case $ \ell=2 $.
Conversely, let us show that the above congruence conditions are also sufficient. From the discussion above, each of these conditions guarantees the existence of a solution $ (\tilde{x}, \tilde{y}) $ of equation (3.6) in $ ( \mathbb{Z}/8 \mathbb{Z})^2 $ such that either $ \tilde{x} $ or $ \tilde{y}D/4 $ lies in $ ( \mathbb{Z}/8 \mathbb{Z})^\times $. Now from a multivariate version of Hensel's lemma [23, Lemma 12.2.8], the pair $ (\tilde{x}, \tilde{y}) $ lifts to a solution of (3.6) in $ \mathbb{Z}_2^2 $. The sufficiency is proved.
Therefore, $ -p \in N_{K/ \mathbb{Q}}( \mathcal{O}_2^\times) $ if and only if one of the following conditions holds:
$ {\rm(a)} $ $ p \equiv 7 \!\! \pmod{8} $;
$ \rm(b) $ $ -p + \frac{ \operatorname{disc}( \mathcal{O})}{4} \equiv 0, \, 1 $ or $ 4 \!\! \pmod{8} $
$ \rm(c) $ $ -p + \operatorname{disc}( \mathcal{O})\equiv 1 \!\! \pmod{8} $.
Proof of Theorem 1.1 If $ \mathcal{H}_p\neq \emptyset $, then there is a regular action of $ \rm{Pic}( \mathcal{O})[2] $ on $ \mathcal{H}_p $ by Proposition 3.1. The criterion for the nonemptiness of $ \mathcal{H}_p $ follows from combining Proposition 2.11 with equation (3.3) and Lemmas 2.12, 3.2 and 3.3.