1 Introduction
The main result of the present paper is a complete computation of the open Gromov-Witten disk invariants of the Chiang Lagrangian [Reference Chiang8] in
$\mathbb {C}P^3.$
The Chiang Lagrangian is not fixed by any anti-symplectic involution [Reference Evans and Lekili10]. Multiple phenomena are observed that have not appeared in previous computations of open Gromov-Witten invariants for Lagrangian submanifolds fixed by anti-symplectic involutions. Our computation relies on direct geometric arguments combined with general structure theorems governing the invariants including the open WDVV equations of [Reference Solomon and Tukachinsky48] and results of the present paper.
1.1 Invariants
To formulate our results, we recall relevant background on open and closed Gromov-Witten invariants. Let
$(X,\omega )$
be a symplectic manifold of real dimension
$2n$
, let
$L \subset X$
be a Lagrangian submanifold and let J be an
$\omega $
-tame almost complex structure. For simplicity, assume L is a connected, spin Lagrangian submanifold with
$H^{*}(L;\mathbb {R}) \simeq H^{*}(S^n;\mathbb {R})$
and
$[L] = 0 \in H_n(X;\mathbb {R}).$
We refer the reader to [Reference Solomon and Tukachinsky46] for a fuller account of settings where open Gromov-Witten invariants can be defined. Write
$\widehat {H}^{*}(X,L;\mathbb {R}) = H^0(L;\mathbb {R})\oplus H^{>0}(X,L;\mathbb {R}).$
Let
$y : H^n(L;\mathbb {R}) \to \widehat {H}^{n+1}(X,L;\mathbb {R})$
denote the boundary map from the short exact sequence of the pair
$(X,L).$
Choose a map
$P_{\mathbb {R}} : \widehat {H}^{n+1}(X,L;\mathbb {R})\to H^n(L;\mathbb {R})$
left-inverse to
$y.$
For
$\beta \in H_2(X,L;\mathbb {Z})$
and
$k,l \geq 0,$
let

denote the open Gromov-Witten invariants of [Reference Solomon and Tukachinsky48]. The multilinear maps
$\overline {OGW}\!_{\beta ,k}$
are invariants of the symplectic manifold and the Lagrangian submanifold L up to Hamiltonian isotopy. When
$\beta $
belongs to the image of the natural map
$\varpi : H_2(X;\mathbb {Z}) \to H_2(X,L;\mathbb {Z}),$
the invariants
$\overline {OGW}\!_{\beta ,0}$
depend on the choice of
$P_{\mathbb {R}}.$
Explicit formulas for this dependence are given in Propositions 1.1 and 8.1 below.
The invariants
$\overline {OGW}\!$
can be described as follows. Let
$A_1,\ldots ,A_l \in \widehat {H}^{*}(X,L;\mathbb {R}).$
Assume first that either
$k> 0$
or
$\beta $
does not belong to the image of
$\varpi .$
The invariant
$\overline {OGW}\!_{\beta ,k}(A_1,\ldots ,A_l)$
counts J-holomorphic disks
$u: (D,\partial D) \to (X,L)$
representing
$\beta $
with boundary passing through k chosen points on L and interior passing through chosen cycles on X Poincaré dual to
$A_1,\ldots ,A_l,$
together with correction terms that compensate for disk bubbling. The correction terms come from bounding cochains in the Fukaya
$A_\infty $
algebra associated to
$L.$
The assumption
$H^{*}(L;\mathbb {R}) \simeq H^{*}(S^n;\mathbb {R})$
is used in [Reference Solomon and Tukachinsky46] to show vanishing of obstruction classes that arise in the construction of the bounding cochains following the method of Fukaya-Oh-Ohta-Ono [Reference Fukaya, Oh, Ohta and Ono14].
In the case
$k = 0$
and
$\beta \in \operatorname {\mathrm {Im}} \varpi ,$
it is necessary to add an additional correction term to the above-mentioned counts of J-holomorphic disks to obtain an invariant. The additional correction term counts J-holomorphic spheres in X passing through an
$n+1$
chain C in X with
$\partial C = L.$
This term compensates for the fact that the boundary of a J-holomorphic disk can collapse to a point in L forming a J-holomorphic sphere. The topological type of the chain C is characterized by the map
$P_{\mathbb {R}} : \widehat {H}^{n+1}(X,L;\mathbb {R})\to H^n(L;\mathbb {R})$
mentioned above, which is given by the integration of differential forms over
$C.$
See Remark 4.12 in [Reference Solomon and Tukachinsky48].
Suppose L is fixed by an anti-symplectic involution,
$\dim L = 2$
or
$3$
, and either
$k> 0$
or
$\beta \notin \operatorname {\mathrm {Im}} \varpi .$
Then, it is shown in [Reference Solomon and Tukachinsky46] that the invariants
$\overline {OGW}\!_{\beta ,k}(\cdots )$
coincide with invariants defined by straightforward counting of J-holomorphic disks [Reference Cho9, Reference Solomon45] or real J-holomorphic spheres [Reference Welschinger50, Reference Welschinger51]. When L is the Chiang Lagrangian, which is not fixed by any anti-symplectic involution, the invariants
$\overline {OGW}\!$
do not coincide with straightforward disk counts in general. See Section 1.2.2 below. Closely related invariants were defined in [Reference Welschinger52], and a comparison can be found in [Reference Chen5].
Although the definition of the invariants
$\overline {OGW}\!$
is quite abstract, they have a rich structure that makes them explicitly computable in many situations. This structure includes the open WDVV equations recalled in Theorem 2.2, the open axioms recalled in Proposition 2.2, and the wall-crossing formula recalled in Theorem 2.3. Moreover, the relative quantum cohomology ring of the pair
$(X,L)$
recalled in Section 2.8 encodes the invariants
$\overline {OGW}\!$
along with the genus zero closed Gromov-Witten invariants of X.
The genus zero closed Gromov-Witten invariants of X are given by multilinear maps

for
$\beta \in H_2(X;\mathbb {Z})$
and
$l \geq 0.$
For
$A_1,\ldots , A_l \in H^{*}(X;\mathbb {Q}),$
the invariant
${GW}_\beta (A_1,\ldots ,A_l)$
counts J-holomorphic maps
$S^2 \to X$
representing the class
$\beta $
and passing through chosen cycles Poincaré dual to
$A_1,\ldots ,A_l.$
We recall the basic properties of these invariants in Section 2.
1.2 Statement of results
Let
$L_\triangle $
denote the Chiang Lagrangian in
$\mathbb {C}P^3$
. In particular,
$L_\triangle $
is a rational homology sphere, so the open Gromov-Witten invariants
$\overline {OGW}\!_{\beta ,k}$
are defined. The definition of
$L_\triangle $
is recalled in Section 3. Write
$\Gamma _0 = [1] \in H^0(L_\triangle ;\mathbb {R}) = \widehat {H}^0(\mathbb {C}P^3,L_\triangle ;\mathbb {R}),$
and for
$j = 1,2,3,$
write

Write
$\Delta _j=[\omega ^j] \in H^{*}(\mathbb {C}P^3;\mathbb {R})$
. Take
$P_{\mathbb {R}}$
to be the unique left-inverse of y such that
$\ker (P_{\mathbb {R}})=\mathrm {span}\{\Gamma _2\}$
. In light of Lemma 3.4 below, we identify
$H_2(\mathbb {C}P^3,L_\triangle ;\mathbb {Z}) \simeq \mathbb {Z}.$
Our main results are the following.
Theorem 1.1. The open WDVV equations imply the following relations for the invariants
$\overline {OGW}\!_{\beta , k}$
of
$(\mathbb {C}P^3, L_\triangle ).$
-
1. For
$k \geq 1, l \geq 1 $ , and
$I = \{ j_2,\ldots ,j_l\}$ ,
$$ \begin{align*} & \overline{OGW}\!_{\beta, k} (\Gamma_{j_1}, \Gamma_{j_2},\ldots,\Gamma_{j_l}) \\ &\quad = -\sum_{\substack{\varpi(\hat{\beta}) + \beta_1 = \beta \\ \beta_1 \not= \beta \\ I_1 \sqcup I_2 = I}} \sum_{i = 0}^{3} {GW}_{\hat{\beta}}(\Delta_{j_1 -1}, \Delta_1, \Delta_{I_1}, \Delta_i) \overline{OGW}\!_{\beta_1, k} (\Gamma_{3-i}, \Gamma_{I_2})\\ &\qquad +\sum_{\substack{\beta_1 +\beta_2 = \beta \\ k_1 + k_2 = k - 1 \\ I_1 \sqcup I_2 = I}}\binom{k-1}{k_1} \bigg( \overline{OGW}\!_{\beta_1, k_1} (\Gamma_{j_1 - 1}, \Gamma_1, \Gamma_{I_1}) \overline{OGW}\!_{\beta_2, k_2+2} (\Gamma_{I_2})\\ & \qquad -\overline{OGW}\!_{\beta_1, k_1 + 1}(\Gamma_{j_1 -1}, \Gamma_{I_1})\overline{OGW}\!_{\beta_2, k_2 + 1} (\Gamma_1, \Gamma_{I_2}) \bigg). \end{align*} $$
-
2. For
$k \geq 2, l \geq 0$ , and
$I = \{j_1,\ldots ,j_l\}$ ,
$$ \begin{align*} &\overline{OGW}\!_{\beta, k} (\Gamma_{j_1},\ldots,\Gamma_{j_l}) \overline{OGW}\!_{2, 0} (\Gamma_2, \Gamma_2) \\ & \quad = \sum_{\substack{ \varpi(\hat{\beta}) + \beta_1 = \beta+2 \\I_1 \sqcup I_2 = I}} \sum_{i = 0}^{3} {GW}_{\hat{\beta}}(\Delta_2, \Delta_2, \Delta_{I_1}, \Delta_i) \overline{OGW}\!_{\beta_1, k-1}(\Gamma_{3-i}, \Gamma_{I_2})\\ &\qquad + \sum_{\substack{\beta_1 + \beta_2 = \beta +2 \\ k_1 + k_2 = k-2 \\ I_1 \sqcup I_2 = I} } \binom{k-2}{k_1} \overline{OGW}\!_{\beta_1, k_1 + 1} (\Gamma_2, \Gamma_{I_1}) \overline{OGW}\!_{\beta_2, k_2 + 1} ( \Gamma_2, \Gamma_{I_2})\\ & \qquad -\sum_{\substack{\beta_1 + \beta_2 = \beta + 2 \\ k_1 + k_2 = k-2 \\ I_1 \sqcup I_2 = I \\ (\beta_1, k_1) \neq (\beta, k-2) }} \binom{k-2}{k_1} \overline{OGW}\!_{\beta_1, k_1 + 2}(\Gamma_{I_1}) \overline{OGW}\!_{\beta_2, k_2}(\Gamma_2, \Gamma_2, \Gamma_{I_2}). \end{align*} $$
-
3. For
$l \geq 2$ , and
$I = \{j_3,\ldots ,j_l\}$ ,
$$ \begin{align*} \overline{OGW}\!_{\beta, k} (\Gamma_{j_1},\ldots,\Gamma_{j_l})& =\overline{OGW}\!_{\beta, k} (\Gamma_{j_1-1},\Gamma_{j_2+1},\Gamma_{j_3},\ldots,\Gamma_{j_l})\\ &\quad + \sum_{\substack{ \varpi(\hat{\beta}) + \beta_1 = \beta \\\beta_1 \neq \beta \\I_1 \sqcup I_2 = I}} \sum_{i = 0}^{3} \bigg({GW}_{\hat{\beta}}(\Delta_1, \Delta_{j_2}, \Delta_{I_1}, \Delta_i) \overline{OGW}\!_{\beta_1, k}(\Gamma_{3-i}, \Gamma_{j_1-1}, \Gamma_{I_2})\\ &\quad -{GW}_{\hat{\beta}}(\Delta_1, \Delta_{j_1-1}, \Delta_{I_1}, \Delta_i) \overline{OGW}\!_{\beta_1, k}(\Gamma_{3-i}, \Gamma_{j_2}, \Gamma_{I_2})\bigg)\\ &\quad +\sum_{\substack{\beta_1 + \beta_2 = \beta \\ k_1 + k_2 = k \\ I_1 \sqcup I_2 = I }} \binom{k}{k_1} \bigg( \overline{OGW}\!_{\beta_1, k_1 }(\Gamma_1, \Gamma_{j_1-1}, \Gamma_{I_1}) \overline{OGW}\!_{\beta_2, k_2+1}(\Gamma_{j_2}, \Gamma_{I_2})\\ &\quad -\overline{OGW}\!_{\beta_1, k_1}(\Gamma_1, \Gamma_{j_2}, \Gamma_{I_1})\overline{OGW}\!_{\beta_2, k_2+1}(\Gamma_{j_1-1}, \Gamma_{I_2})\bigg). \end{align*} $$
Theorem 1.2. Consider the choice of spin structure and orientation on
$L_\triangle $
given in Section 3.2. Then

It follows from the open WDVV equations that
$\overline {OGW}\!_{2,0}(\Gamma _2, \Gamma _2)=\frac {35}{64}.$
We call the invariants
$\overline {OGW}\!_{1,1}, \overline {OGW}\!_{1,0}(\Gamma _2)$
and
$\overline {OGW}\!_{2,0}(\Gamma )$
basic invariants. Theorem 1.6 below shows that these invariants are given by straightforward counts of J-holomorphic disks without correction terms. Thus, we can compute the basic invariants using the theory of axial disks developed by Evans-Lekili [Reference Evans and Lekili10] and Smith [Reference Smith43, Reference Smith44] for computing the Floer cohomology of the Chiang Lagrangian and other homogeneous Lagrangian submanifolds. Since the Chiang Lagrangian is not fixed by an anti-symplectic involution, the techniques of real algebraic geometry used in other computations of open Gromov-Witten invariants [Reference Chen and Zinger7, Reference Horev and Solomon28, Reference Hugtenburg and Tukachinsky29, Reference Solomon and Tukachinsky48] are not available.
Corollary 1.3. The genus zero open Gromov-Witten invariants of
$(\mathbb {C}P^3, L_\triangle )$
are entirely determined by the open WDVV equations, the axioms of
$\overline {OGW}\!$
, the wall-crossing formula Theorem 2.3, the genus zero closed Gromov-Witten invariants of
$\mathbb {C}P^3$
, and the values computed in Theorem 1.2.
Samples values for the invariants
$\overline {OGW}\!$
computed using Theorems 1.1 and 1.2 are given in Tables 1 and 2.
Table 1 Values of invariants with boundary constraints only.

Table 2 Values of
$\overline {OGW}\!_{\beta ,0}(\Gamma _2^{\otimes l_2},\Gamma _3^{\otimes l_3})$
. The value of
$l_2$
is determined by
$\beta $
and
$l_3$
by the open degree axiom.

1.2.1 Monotonicity versus periodicity
In other computations [Reference Itenberg, Kharlamov and Shustin30, Reference Itenberg, Kharlamov and Shustin31, Reference Itenberg, Kharlamov and Shustin32, Reference Pandharipande, Solomon and Walcher40, Reference Solomon and Tukachinsky48, Reference Horev and Solomon28, Reference Farajzadeh Tehrani11, Reference Chen and Zinger7], the absolute values of
$\overline {OGW}\!_{\beta ,k}$
or invariants of a similar flavor are monotonically increasing in
$\beta .$
In the case of
$(\mathbb {C}P^3,L_\triangle )$
, we see that these values are not monotonically increasing. Interestingly, it appears that the only persistent violation of monotonicity occurs periodically:
$|\overline {OGW}\!_{\beta -1,\beta -1}|>|\overline {OGW}\!_{\beta ,\beta }|$
for
$\beta \in 8\mathbb {Z}$
at least up to
$\beta = 32.$
Another unusual periodicity is visible in the signs of the invariants presented in Table 1, which repeat with period
$16.$
Moreover, the sign is inverted when
$\beta $
is increased by
$8.$
1.2.2 Bounding cochain corrections and denominators
In [Reference Evans and Lekili10], Evans-Lekili proved that there are exactly two J-holomorphic disks representing the class
$\beta = 2 \in H_2(\mathbb {C}P^3,L_\triangle ;\mathbb {Z})$
passing through a specific choice of two points in
$L_\triangle $
. This does not coincide with our computation
$\overline {OGW}\!_{2,2}=-\frac {5}{4}$
. The discrepancy reflects correction terms arising from a bounding cochain in the Fukaya
$A_\infty $
algebra of
$L_\triangle .$
The correction terms need not be whole numbers. In fact, all invariants with only boundary constraints that are not whole numbers indicate the presence of nontrivial correction terms. Table 1 provides several examples. When interior constraints are present, open Gromov-Witten invariants may be non-integral even in the absence of correction terms [Reference Horev and Solomon28].
As far as we have calculated, for the Chiang Lagrangian
$L_\triangle $
, the denominators appearing in the invariants
$\overline {OGW}\!_{\beta ,k}$
with only boundary constraints are always powers of
$4.$
In the presence of interior constraints, powers of
$2$
appear as well. Examining the proof of the existence of bounding cochains in [Reference Solomon and Tukachinsky46], we can see where denominators of
$4$
arise. Namely, the proof uses a
$1$
-cochain h with real coefficients with coboundary equal to a
$2$
-cocycle g representing the generator of
$H^2(L_\triangle ;\mathbb {Z}) \simeq \mathbb {Z}/4\mathbb {Z}.$
Since
$[4g] = 0 \in H^2(L_\triangle ;\mathbb {Z}),$
there exists an integral
$1$
-cochain
$\tilde h$
with coboundary equal to
$4g$
. So, we can take
$h = \frac {\tilde h}4.$
This is the source of the powers of
$4$
in the denominators of the invariants
$\overline {OGW}\!_{\beta ,k}.$
In contrast, the invariants
$\overline {OGW}\!_{\beta ,k}$
with only boundary constraints for the Lagrangian submanifold
$\mathbb {R} P^{n} \subset \mathbb {C} P^{n}$
with n odd are shown in [Reference Solomon and Tukachinsky48] to be whole numbers. For
$n = 3,$
these invariants are shown in [Reference Solomon and Tukachinsky46] to be straightforward disk counts, but for
$n \geq 5$
, they presumably include correction terms arising from bounding cochains. Since
$H^{2i}(\mathbb {R} P^n;\mathbb {Z}) \simeq \mathbb {Z}/2$
for
$0 < i < n/2,$
by the reasoning of the preceding paragraph, one might expect to see powers of
$2$
in the denominators of the invariants
$\overline {OGW}\!_{\beta ,k}.$
However, since
$\mathbb {R} P^n$
is fixed by an anti-symplectic involution, (i.e., complex conjugation), holomorphic disks come in pairs. So, in the proof of the existence of bounding cochains in [Reference Solomon and Tukachinsky46], one needs only a
$(2i-1)$
-cochain h with coboundary equal to twice a
$2i$
-cocycle g representing the generator of
$H^{2i}(\mathbb {R} P^n;\mathbb {Z})$
. But
$[2g] = 0 \in H^{2i}(\mathbb {R} P^n;\mathbb {Z})$
, so h can be chosen to be integral. This explains the absence of denominators. A similar argument should apply for other Lagrangian submanifolds L fixed by an anti-symplectic involution. If all torsion in
$H^{2i}(L;\mathbb {Z})$
has order
$2$
, then the invariants
$\overline {OGW}\!_{\beta ,k}$
should be integral. More generally, if
$H^{2i}(L;\mathbb {Z})$
contains torsion of even order, the rate of growth of the power of
$2$
in the denominator of
$\overline {OGW}\!_{\beta ,k}$
as a function of
$\omega (\beta )$
should be slower when L is fixed by an anti-symplectic involution than otherwise. In particular, the behavior of denominators in the invariants
$\overline {OGW}\!_{\beta ,k}$
could potentially be used to prove that a Lagrangian submanifold is not fixed by an anti-symplectic involution.
1.2.3 Mixed disk and sphere invariants
As mentioned above, the invariants
$\overline {OGW}\!_{\beta ,k}$
for
$\beta \in \operatorname {\mathrm {Im}}\varpi $
and
$k = 0$
count J-holomorphic disks and J-holomorphic spheres together. In other computations [Reference Solomon and Tukachinsky48, Reference Hugtenburg and Tukachinsky29], these mixed disk and sphere invariants vanish for a natural choice of
$P_{\mathbb {R}}.$
Another family of invariants combining counts of J-holomorphic maps of different topology has also been shown to vanish in certain examples [Reference Farajzadeh Tehrani11].
We show that the mixed disk and sphere invariants of
$(\mathbb {C}P^3,L_\triangle )$
do not vanish. As shown in Lemma 3.4 below, the map
$\varpi : H_2(\mathbb {C}P^3;\mathbb {Z}) \to H_2(\mathbb {C}P^3,L_\triangle ;\mathbb {Z})\simeq \mathbb {Z}$
is given by multiplication by
$4.$
So, the nonvanishing of mixed disk and sphere invariants is visible in Table 2 for
$\beta = 4,8.$
In fact, Corollary 1.4 below asserts that this nonvanishing persists for any choice of
$P_{\mathbb {R}}.$
Let
$\rho : \widehat {H}^{4}(\mathbb {C}P^3,L_\triangle ;\mathbb {R})\rightarrow H^{4}(\mathbb {C}P^3;\mathbb {R})$
denote the natural map. Let
$P_{\mathbb {R}},P_{\mathbb {R}}^{\prime }$
be two choices of left inverse maps of y with associated invariants
$\overline {OGW}\!_{\beta ,k}$
and
$\overline {OGW}\!\,'\!\!_{\beta ,k}$
, respectively. The long exact sequence of the pair
$(\mathbb {C}P^3,L_\triangle )$
implies there exists a map

such that
$\mathfrak {p}_{\mathbb {R}}\circ \rho =P_{\mathbb {R}}-P_{\mathbb {R}}^{\prime }.$
We use Poincaré duality to identify
$H^3(L_\triangle ;\mathbb {R})\simeq \mathbb {R}.$
Theorem 1.1. For
$\beta \in \operatorname {\mathrm {Im}} \varpi ,$
we have

Corollary 1.4. There is no map
$P_{\mathbb {R}}:\widehat {H}^{4}(\mathbb {C}P^3,L_\triangle ;\mathbb {R})\rightarrow H^3(L_\triangle ;\mathbb {R})$
such that
$\overline {OGW}\!_{\beta ,k}$
vanishes for every
$\beta \in \operatorname {\mathrm {Im}}\varpi $
and
$k=0.$
Proposition 1.1 is a special case of Proposition 8.1, which gives the analogous statement for X a general symplectic manifold and L a connected spin Lagrangian submanifold with
$H^{*}(L;\mathbb {R}) \simeq H^{*}(S^n;\mathbb {R})$
and
$[L] = 0 \in H_n(X;\mathbb {R}).$
In fact, Proposition 8.1 remains valid in all situations where the invariants
$\overline {OGW}\!$
are defined in [Reference Solomon and Tukachinsky48].
1.2.4 Relative quantum cohomology
It follows from Corollary 1.3 that we can compute the big relative quantum cohomology of
$(\mathbb {C}P^3,L_\triangle ).$
The resulting ring does not appear to have a tractable presentation by generators and relations. However, the small relative quantum cohomology of
$(\mathbb {C}P^3,L_\triangle )$
is more accessible.
Theorem 1.5. The small relative quantum cohomology of
$(\mathbb {C}P^3,L_\triangle )$
is given by

with

Computations of small relative quantum cohomology for a variety of pairs
$(X,L)$
are given in [Reference Hugtenburg and Tukachinsky29]. However, all Lagrangian submanifolds considered there are fixed by an anti-symplectic involution. The definitions of big and small quantum cohomology are recalled in Section 2.8.
1.2.5 Invariants without corrections
The proof of Theorem 1.2 depends on the following general result, which says when the invariants
$\overline {OGW}\!_{\beta ,k}$
can be computed without taking into account correction terms from bounding cochains. For simplicity, we continue as in Section 1.1 with L a connected spin Lagrangian with
$H^{*}(L;\mathbb {R}) \simeq H^{*}(S^n;\mathbb {R}).$
A similar result holds in other settings where the invariants
$\overline {OGW}\!$
can be defined. For
$\beta \in H_2(X,L;\mathbb {Z})$
, define

Let
$\sigma _k = (k-1)!$
for
$k \in \mathbb {Z}_{>0}$
and let
$\sigma _0 = 1.$
Theorem 1.6. Let
$k \in \mathbb {Z}_{\geq 0}$
and
$\beta \in H_2(X,L;\mathbb {Z})$
satisfy either
$k> 0$
or
$\beta \not \in \operatorname {\mathrm {Im}} \varpi $
. Let
$A_{j}\in \widehat {H}^{*}(X,L;\mathbb {R})$
. Let
$\bar {b}\in A^n(L;\mathbb {R})$
such that
$\text {PD}([\bar {b}])=pt$
, and let
$a_{j}$
be a representative of
$A_{j}$
. Suppose that for every

one of the following two conditions is satisfied:
-
1.
$1-\mu (\tilde {\beta }) +j(n-1) + \sum _{i\in I} (|A_{i}|-2)<0,$ or
-
2.
$1-\mu (\tilde {\beta }) +j(n-1) + \sum _{i\in I} (|A_{i}|-2)\ge n-1.$
Then,

Theorem 1.6 is a special case of Theorem 4.19. The basic invariants
$\overline {OGW}\!_{1,1}, \overline {OGW}\!_{1,0}(\Gamma _2)$
and
$\overline {OGW}\!_{2,0}(\Gamma _3)$
of Theorem 1.2 are exactly the invariants to which Theorem 1.6 applies for L the Chiang Lagrangian.
1.2.6 Orientation, spin structure and signs
We prove general formulas governing the dependence of the invariants
$\overline {OGW}\!_{\beta ,k}$
on the spin structure and orientation of L, which appear as the spin axiom 7 and the orientation axiom 8 of Proposition 2.2. See also Lemmas 4.5 and 4.6. We use the spin and orientation axioms to deduce vanishing results for certain open Gromov-Witten invariants, which are formulated in Corollaries 4.7 and 4.8.
A significant part of the present paper is devoted to computing the signs of the three basic invariants of Theorem 1.2. The spin and orientation axioms show that by changing the orientation and spin structure on
$L_\triangle $
, one can adjust the signs of any two of the basic invariants arbitrarily. Moreover, changing the orientation and spin structure only affects the signs of invariants but not their absolute values.
However, it can be seen from the proof of Lemma 6.1 that if the sign of one of the three basic invariants is changed while the two others are held fixed, then the absolute values of non-basic invariants change. Qualitatively, the denominators of invariants cease to be powers of two.
1.2.7 Arithmetic structure of the open WDVV equations
From the perspective of Theorem 1.1, it is surprising that only powers of
$2$
appear in the denominators of the invariants
$\overline {OGW}\!$
for the Chiang Lagrangian
$L_\triangle .$
Indeed, to use relation 2 for recursive computations, it is necessary to divide by
$\overline {OGW}\!_{2,0}(\Gamma _2,\Gamma _2) = \frac {35}{64}.$
A priori, this should introduce factors of
$5$
and
$7$
in denominators, but unexpected cancellations occur so that only factors of
$2$
remain.
To test the rigidity of these cancellations, we searched for rational numbers
$v \neq 1/4$
that upon substitution for the value of
$\overline {OGW}\!_{1,0}(\Gamma _2)$
in the recursive scheme of Corollary 1.3 yield
$\overline {OGW}\!_{k,k}$
with denominator a power of
$2$
for
$k = 1,\ldots ,M.$
As M increases, the magnitudes of the possible numerators and denominators of v increase. For
$M = 4,$
the v with smallest numerator and denominator that we found is

For this choice of
$v,$
the value for
$\overline {OGW}\!_{5,5}$
given by the recursive scheme has denominator
$7.$
Therefore, we reach the following conclusion. Suppose it were possible to formalize the heuristic argument of Section 1.2.2 to prove geometrically that the invariants
$\overline {OGW}\!_{\beta ,k}$
can only have powers of
$2$
for denominators. Then, it seems reasonable to expect that the basic invariant
$\overline {OGW}\!_{1,0}(\Gamma _2)$
could be deduced from the other two basic invariants, the open WDVV equations, the axioms of
$\overline {OGW}\!$
, and the closed Gromov-Witten invariants of
$\mathbb {C}P^3.$
If it were necessary to compute only two basic invariants geometrically, it would not be necessary to compute their signs. Indeed, these signs could be adjusted arbitrarily by changing the orientation and spin structure on
$L_\triangle $
as noted in Section 1.2.6. Thus, the computations of this paper could be simplified significantly. In fact, this strategy can be generalized for an arbitrary pair
$(X,L)$
as outlined below.
1.3 Directions for future research
Remarkably, Theorem 1.6 applies exactly for those invariants that are needed as initial values for the recursions given by Theorem 1.1. It is an interesting question to determine to what extent this phenomenon persists more generally.
We expect the techniques of the present paper to extend to the other Platonic Lagrangians studied by Smith [Reference Smith43] and more general homogeneous Lagrangians [Reference Bedulli and Gori3, Reference Gasparim, San Martin and Valencia23, Reference Konstantinov33]
Motivated by the discussion of Section 1.2.7, we formulate the following two hypotheses. These hypotheses should hold for a large class of Lagrangian submanifolds
$L \subset X$
, but some restrictions are necessary.
-
1. The invariants
$\overline {OGW}\!_{\beta ,k}$ are rational of the form
$\frac {m}{p_1^{a_1}\cdots p_N^{a_N}}$ where
$p_1,\ldots ,p_N$ are the primes that occur as orders of torsion elements in
$H^{*}(L;\mathbb {Z}).$
-
2. Let
$\mathcal {A}$ be a minimal set of the invariants
$\overline {OGW}\!_{\beta ,k}$ from which all other invariants can be deduced using the open WDVV equations, the axioms of
$\overline {OGW}\!$ , the wall-crossing formula, and the closed GW invariants of X. Let
$\mathcal {B} \subset \mathcal {A}$ be defined in the same way as
$\mathcal {A}$ except that we are also allowed to use hypothesis 1 to deduce the other invariants. Then
$\mathcal {B}$ is a proper subset of
$\mathcal {A}.$
Hypothesis 1 follows from the heuristic argument of Section 1.2.2 applied to L of general topology. It should admit a geometric proof. Hypothesis 2 is a statement about the arithmetic structure of the open WDVV equation. In the case of the Chiang Lagrangian, we can take
$\mathcal {A}$
to be the basic invariants
$\overline {OGW}\!_{1,1}, \overline {OGW}\!_{1,0}(\Gamma _2)$
and
$\overline {OGW}\!_{2,0}(\Gamma _3)$
. Then, the calculations presented in Section 1.2.7 indicate that the invariants
$\overline {OGW}\!_{1,1}$
and
$\overline {OGW}\!_{2,0}(\Gamma _3)$
suffice for
$\mathcal {B}.$
However, in the case
$(X,L) = (\mathbb {C} P^n, \mathbb {R} P^n)$
, one can see that hypothesis 2 does not hold as follows. It is shown in [Reference Solomon and Tukachinsky48, Corollary 1.9] that one can take
$\mathcal {A}$
to consist of a single invariant. It follows from either the spin axiom 7 or the orientation axiom 8 of Proposition 2.2 that by changing the spin structure or orientation on L, the sign of this single invariant can be changed without affecting the absolute value of the others – in particular, without violating hypothesis 1. Thus,
$\mathcal {B}$
cannot be smaller than
$\mathcal {A}.$
Another case where hypothesis 2 does not hold is when X is the quadric hypersurface in
$\mathbb {C} P^{n+1}$
given by
$\sum _{i = 0}^n z_i^2 - z_{n+1}^2 = 0$
and L is the real locus. This is explained in Remark 4.9 based on results of [Reference Hugtenburg and Tukachinsky29], which are related to the orientation axiom.
We observe the following common feature of the forgoing exceptions to hypothesis 2. In both cases, one can take
$\mathcal {A}$
to contain a single invariant, while there are one or more geometric degree of freedom affecting the values of the invariants: in the case
$(X,L) = (\mathbb {C} P^n,\mathbb {R} P^n)$
, there are choices of spin structure and orientation, and in the case of the quadric, there is the choice of orientation. The presence of more geometric degrees of freedom than the size of
$\mathcal {A}$
forces a measure of flexibility in the open WDVV equations. In contrast, for the Chiang Lagrangian,
$\mathcal {A}$
contains three invariants, while there are only two geometric degrees of freedom, the choices of spin structure and orientation. So, the equations can be sufficiently rigid for hypothesis 2 to hold.
1.4 Outline
In Section 2, we recall general definitions and results on closed and open Gromov-Witten invariants that will be useful in the other sections. This includes the closed and open axioms, the closed and the open WDVV equations, the wall-crossing formula, and the definitions of small and big relative quantum cohomology.
In Section 3, we recall from [Reference Smith43] a family of Lagrangian homology spheres in Fano threefolds called Platonic Lagrangians. We focus mainly on the Chiang Lagrangian
$L_\triangle $
, which is a special case of this construction, and recall some of its basic properties.
Section 4 contains results related to Fukaya
$A_\infty $
algebras, bounding cochains and the superpotential. We recall definitions and results on obstruction theory for bounding cochains from [Reference Solomon and Tukachinsky46]. These are used to prove Theorem 4.19, which gives a criterion for open Gromov-Witten invariants to coincide with straightforward counts of J-holomorphic disks. Theorem 1.6, which is needed for Section 5, is obtained as a special case. We also prove Lemmas 4.5 and 4.6, from which we deduce the spin axiom 7 and the orientation axiom 8 of Proposition 2.2.
The goal of Section 5 is to compute the basic invariants that serve as initial values for the recursive formulas. In Sections 5.1-5.3, we recall definitions and results about the anticanonical divisor, the Maslov class and axial disks. In Section 5.4, we give the definition of a spin Riemann-Hilbert pair, recall results on canonical orientations, and prove some new ones. Section 5.5 contains results concerning Riemann-Hilbert pairs arising from a certain type of axial disk. We use these results to compute the basic invariants geometrically in Sections 5.6-5.8.
Section 6 contains the proofs of Theorem 1.1 and Corollary 1.3, which give recursive formulas that determine all open Gromov-Witten invariants of
$(\mathbb {C}P^3,L_\triangle )$
from the three basic invariants.
In Section 7, we compute the small relative cohomology of
$(\mathbb {C}P^3, L_\triangle )$
.
In Section 8, we recall definitions and results concerning the map
$P_{\mathbb {R}}$
from [Reference Solomon and Tukachinsky48]. These results are used to prove Proposition 8.1, which shows how the invariants
$\overline {OGW}\!$
depend on the map
$P_{\mathbb {R}}.$
Proposition 1.1 is obtained as a special case. Finally, we prove Corollary 1.4.
2 Background
In this section, we recall definitions and properties of open and closed Gromov-Witten invariants and relative quantum cohomology.
2.1 Invariants
Let
$(X,\omega )$
be a closed symplectic manifold. For
$\beta \in H_2(X;\mathbb {Z})$
and
$l \geq 0,$
let

denote the standard closed genus zero Gromov-Witten invariant. This invariant counts J-holomorphic maps
$S^2 \to X$
representing the class
$\beta $
passing through cycles Poincaré dual to l cohomology classes on
$X.$
Let
$L \subset X$
be a closed Lagrangian submanifold. Genus zero open Gromov-Witten invariants are analogous invariants that count J-holomorphic maps
$(D,\partial D) \to (X,L)$
along with correction terms arising from bounding cochains. To simplify the exposition, we focus on the case that L is connected and spin, with
$H^{*}(L;\mathbb {R})\simeq H^{*}(S^n; \mathbb {R})$
and
$[L] = 0 \in H_n(X;\mathbb {R})$
. These conditions hold for the Chiang Lagrangian. Similar results are known [Reference Solomon and Tukachinsky46] in the case that L is only relatively spin or
$[L] \neq 0 \in H_n(X;\mathbb {R}).$
Consider the subcomplex of differential forms on X consisting of those with trivial integral on
$L,$

For an
$\mathbb {R}$
-algebra
$\Upsilon $
, write

Observe that

For
$\beta \in H_2(X,L;\mathbb {Z})$
and
$k,l \geq 0,$
let

denote the open Gromov-Witten invariants of [Reference Solomon and Tukachinsky46]. The definition is recalled in Section 4.2. For
$\beta \in H_2(X,L;\mathbb {Z})$
and
$k,l \geq 0,$
let

denote the enhanced invariants defined in [Reference Solomon and Tukachinsky48]. The definition is recalled in 8. These invariants combine counts of J-holomorphic disks and J-holomorphic spheres. With the exception of the case that
$\beta $
belongs to the image of the natural map

and
$k = 0,$
we have
$\overline {OGW}\! _{\beta ,k} = {OGW}_{\beta ,k}.$
In the exceptional case,
${OGW}_{\beta ,k} = 0,$
but
$\overline {OGW}\!_{\beta ,k}$
need not vanish. By the assumption
$[L] = 0 \in H_n(X;\mathbb {R}),$
there is a short exact sequence

For
$k=0$
and
$\beta \in \mathrm {Im \varpi }$
, the definition of the enhanced invariants
$\overline {OGW}\!_{\beta ,k}$
depends on the choice of a linear map

that is a left-inverse to the map y in the short exact sequence. A choice of
$P_{\mathbb {R}}$
is equivalent to a splitting of (2.1). We extend
$P_{\mathbb {R}}$
to a map
$P_{\mathbb {R}}: \widehat {H}^{*}(X,L;\mathbb {R}) \to H^n(L;\mathbb {R})$
by setting it to zero outside
$\widehat {H}^{n+1}(X,L;\mathbb {R})$
.
2.2 Axioms
The closed Gromov-Witten invariants satisfy the following properties [Reference Behrend4, Reference Fukaya and Ono22, Reference Kontsevich and Manin34, Reference Li and Tian36, Reference McDuff and Salamon38, Reference McDuff and Salamon37, Reference Ruan and Tian41, Reference Ruan and Tian42].
Theorem 2.1. Let
$\beta \in H_2(X,\mathbb {Z})$
and let
$A_1,\ldots ,A_k\in H^{*}(X;\mathbb {R})$
.
-
1. (Effective) If
$\omega (\beta )<0$ , then
${GW}_{\beta }(A_1,\ldots ,A_k)=0.$
-
2. (Symmetry) For each permutation
$\sigma \in S_k$ ,
$$\begin{align*}{GW}_{\beta}(A_{\sigma(1)},\ldots,A_{\sigma(k)})=(-1)^{s_\sigma(A)}{GW}_{\beta}(A_1,\ldots,A_k),\end{align*}$$
$s_\sigma (A):=\sum _{\substack {i<j\\ \sigma (i)<\sigma (j)}}|A_i|\cdot |A_j|.$
-
3. (Degree) If
$GW_{\beta }(A_1,\ldots ,A_k)\ne 0,$ then
$$ \begin{align*}\sum_{i=1}^k|A_i|-2k+6= 2n+ 2c_1(\beta).\end{align*} $$
-
4. (Fundamental Class) If
$(\beta ,k)\ne (0,3)$ and
$k\ge 1,$ then
$$ \begin{align*} GW_{\beta}(A_1,\ldots,A_{k-1},1)=0. \end{align*} $$
-
5. (Divisor) If
$(\beta ,k)\ne (0,3)$ ,
$|A_k|=2$ and
$k\ge 1,$ then
$$ \begin{align*} {GW}_{\beta}(A_1,\ldots,A_k)={GW}_{\beta}(A_1,\ldots,A_{k-1})\cdot \int_\beta A_k.\end{align*} $$
-
6. (Zero) If
$k \ne 3$ , then
$GW_{0}(A_1,\ldots ,A_k)=0$ . If
$k=3,$ then
$$ \begin{align*}GW_{0}(A_1,A_2,A_3)= \int_XA_1\smile A_2\smile A_3.\end{align*} $$
-
7. (Deformation invariance) The invariants
${GW}_\beta $ remain constant under deformations of the symplectic form
$\omega $ .
The enhanced open invariants
$\overline {OGW}\!$
satisfy the following properties.
Theorem 2.2. Let
$\beta \in H_2(X,L;\mathbb {Z})$
, let
$k \in \mathbb {Z}_{\geq 0}$
and let
$A_1,\ldots ,A_l\in \widehat {H}^{*}(X,L;\mathbb {R})$
.
-
1. (Effective) If
$\omega (\beta )<0,$ then
$\overline {OGW}\!_{\beta ,k}(A_1,\ldots ,A_k)=0.$
-
2. (Symmetry) For each permutation
$\sigma \in S_l$ ,
$$\begin{align*}\overline{OGW}\!_{\beta,k}(A_{\sigma(1)},\ldots,A_{\sigma(l)})=(-1)^{s_\sigma(A)}\overline{OGW}\!_{\beta}(A_1,\ldots,A_l),\end{align*}$$
$s_\sigma (A):=\sum _{\substack {i<j\\ \sigma (i)<\sigma (j)}}|A_i|\cdot |A_j|.$
-
3. (Degree) If
$\overline {OGW}\!_{\beta ,k}(A_1,\ldots ,A_l)\ne 0,$ then
$$\begin{align*}n-3+\mu(\beta)+k+2l=kn+\sum_{j=1}^l | A_j|.\end{align*}$$
-
4. (Unit/ Fundamental class)
$$\begin{align*}\overline{OGW}\!_{\beta,k}(1,A_1,\ldots,A_{l-1})= \left\{\begin{array}{ll} -1, & (\beta,k,l)=(\beta_0,1,1)\\ P_{\mathbb{R}}(A_1), & (\beta,k,l)=(\beta_0,0,2)\\ 0, & \mathrm{otherwise}. \end{array} \right. \end{align*}$$
-
5. (Zero)
$$\begin{align*}\overline{OGW}\!_{\beta_0,k}(A_1,\ldots,A_l)= \left\{\begin{array}{ll} -1, & (k,l)=(1,1)\ and\ A_1=1\\ P_{\mathbb{R}}(A_1\smile A_2), & (k,l)=(0,2)\\ 0, &\mathrm{otherwise}. \end{array} \right. \end{align*}$$
-
6. (Divisor) If
$| A_l|=2$ , then
$$\begin{align*}\overline{OGW}\!_{\beta,k}(A_1,\ldots,A_l)= \int_\beta A_l\cdot\overline{OGW}\!_{\beta,k}(A_1,\ldots,A_{l-1}).\end{align*}$$
-
7. (Spin) Changing the spin structure on L by the action of
$\alpha \in H^1(L;\mathbb {Z}/2\mathbb {Z})$ changes the sign of
$\overline {OGW}\!_{\beta ,k}(A_1,...,A_l)$ by the sign
$(-1)^{\alpha (\partial \beta )}.$
-
8. (Orientation) Changing the orientation of L changes the sign of
$\overline {OGW}\!_{\beta ,k}(A_1,...,A_l)$ by the sign
$(-1)^{k+1}.$
-
9. (Deformation invariance) The invariants
$\overline {OGW}\!_{\beta ,k}$ remain constant under deformations of the symplectic form
$\omega $ for which L remains Lagrangian.
Proof. Axioms 3–6 are given in Proposition 4.19 from [Reference Solomon and Tukachinsky48]. The analogs of Axioms 2 and 9 for the invariants
${OGW}$
are given in Theorem 4 of [Reference Solomon and Tukachinsky46]. The analogs of Axioms 7 and 8 for the invariants
${OGW}$
are given in Lemmas 4.5 and 4.6 below, respectively. The extension to the invariants
$\overline {OGW}\!$
is similar to the proof of Proposition 4.19 in [Reference Solomon and Tukachinsky48]. Axiom 1 follows from the fact that the Novikov ring
$\Lambda $
consists of power series in
$T^\beta $
for
$\omega (\beta ) \geq 0.$
Indeed, the enhanced superpotential
$\overline {\Omega }$
is defined in Section 1.3.3 in [Reference Solomon and Tukachinsky48] as an element of the ring
$R_W.$
2.3 Bases
Let
$\Delta _i \in H^{*}(X;\mathbb {R})$
for
$i = 0,\ldots ,N,$
be a basis and let
$\Gamma _i \in \ker (P_{\mathbb {R}}) \subset \widehat {H}^{*}(X,L;\mathbb {R})$
be the unique classes such that
$\rho (\Gamma _i) = \Delta _i$
. Let
$\Gamma _{N+1} = y(1).$
We also write
$\Gamma _\diamond = y(1).$
For convenience, set
$\Delta _{N+1} = 0.$
2.4 Novikov rings
Let
$\mu : H_2(X,L;\mathbb {Z}) \to \mathbb {Z}$
denote the Maslov index. Define Novikov coefficients rings

Gradings on
$\Lambda , \Lambda _c$
are defined by declaring
$|T^\beta | = \mu (\beta ).$
Let

Gradings on
$R_W,Q_W,Q_U,$
are defined by declaring
$|t_i| = 2-|\Gamma _i|,\, |s| = 1-n.$
2.5 Potentials
The closed Gromov-Witten potential is given by

The enhanced superpotential, which encodes the enhanced open Gromov-Witten invariants
$\overline {OGW}\!_{\beta ,k},$
is given by

2.6 WDVV equations
Let

and let
$g^{ij}$
denote the inverse matrix. The following WDVV equations hold for the closed Gromov-Witten potential [Reference Behrend4, Reference Fukaya and Ono22, Reference Kontsevich and Manin34, Reference Li and Tian36, Reference McDuff and Salamon38, Reference McDuff and Salamon37, Reference Ruan and Tian41, Reference Ruan and Tian42, Reference Witten53].
Theorem 2.1 (Closed WDVV equations).
For all quadruples of integers
$i,j,k,l\in \{0,\ldots , n\}$
, the function
$\Phi =\Phi ^{\mathbb {C}P^n}$
satisfies

The following is Corollary 1.6 from [Reference Solomon and Tukachinsky48].
Theorem 2.2 (Open WDVV equations).
For
$u, v, w = 0,\ldots , N+1,$
we have the open WDVV equations


2.7 Wall-crossing
Define

The following is Theorem 6 from [Reference Solomon and Tukachinsky48].
Theorem 2.3 (Wall crossing).
Suppose
$[L]=0$
. Then the invariants
$\overline {OGW}\!_{\beta ,k}$
satisfy

2.8 Relative quantum cohomology
Abbreviate

The underlying module of the big relative quantum cohomology is

The big relative quantum product

is given by

In Theorem 7 of [Reference Solomon and Tukachinsky48], it is shown that the product is associative and graded commutative. Though in [Reference Solomon and Tukachinsky48] the definition of
is based on a chain-level construction, the above explicit formula follows from Lemma 5.11 in [Reference Solomon and Tukachinsky48].
The underlying module of the small relative quantum cohomology is

The small relative quantum product

is given by specializing the big relative quantum product to
$t_0 = \cdots =t_{N+1} = 0.$
So, it too is associated and graded commutative. Explicitly,
is given by

3 The Chiang Lagrangian
In this section, we describe the construction of the Chiang Lagrangian along with an orientation and spin structure. Then, we prove some lemmas concerning its topology.
3.1 Geometric construction
The following is largely based on [Reference Smith43]. Consider the fundamental representation V of
$\text {SL}(2, \mathbb {C}).$
Projectifying, we obtain an action of
$\text {SL}(2, \mathbb {C})$
on
$\mathbb {C}P^1 = \mathbb {P}(V).$
Taking the d-fold symmetric product, we obtain an action of
$\text {SL}(2, \mathbb {C})$
on
$\text {Sym}^d \mathbb {C}P^1 = \mathbb {C}P^d.$
Equivalently, we can consider the d-fold symmetric power
$S^d V$
and projectify to obtain an action of
$\text {SL}(2, \mathbb {C})$
on
$\mathbb {P}(S^dV) = \mathbb {C}P^d.$
Fix
$d\ge 3$
, and a configuration C of d distinct points in
$\mathbb {C}P^1$
. Then, the
$\text {SL}(2, \mathbb {C})$
-orbit of C in
$\text {Sym}^d \mathbb {C}P^1 \cong \mathbb {P}S^dV$
is a three-dimensional complex submanifold, of which the
$\text {SU}(2)$
-orbit is a three-dimensional totally real submanifold. The stabilizer of C in
$\text {SL}(2, \mathbb {C})$
is a finite subgroup of
$\text {SU}(2)$
which we denote by
$\Gamma _C.$
In [Reference Aluffi and Faber1], Aluffi and Faber identify those configurations C for which the
$\text {SL}(2, \mathbb {C})$
-orbit has smooth closure
$X_C$
in
$\mathbb {P}S^dV$
. There are four cases: the vertices of an equilateral triangle on a great circle in
$\mathbb {C}P^1$
, which we denote by
$\triangle $
, and the vertices of a regular tetrahedron, octahedron and icosahedron in
$\mathbb {C}P^1$
, which we denote by
$T, O$
and I, respectively.
In each case, the restriction of the
$\text {SU}(2)$
-action to
$X_C$
with the Fubini- Study Kähler form is Hamiltonian with the moment map
$m: \mathbb {P}S^dV\rightarrow \mathfrak {su}(2)^{*}$
which is defined by

where
$\varphi :\mathfrak {su}(2)\rightarrow \mathrm {Mat}_{(d+1)\times (d+1)}(\mathbb {C})$
is given by the representation
$S^dV.$
The
$\text {SU}(2)$
-orbits of the configurations
$\triangle , T, O$
and I all lie in the zero sets of the respective moment maps, so they are Lagrangian. We denote these orbits by
$L_C$
.
The Chiang Lagrangian is
$L_{\triangle } \subset X_{\triangle } = \mathbb {P}S^3V \cong \mathbb {C}P^3$
. It is the space of all the equilateral triangles on great circles in
$\mathbb {C}P^1.$
Topologically,
$L_{\triangle }$
is the quotient of
$\text {SU}(2)$
by the binary dihedral subgroup
$\Gamma _\triangle $
of order 12.
To work with the Chiang Lagrangian, we will need the following additional notations. Let
$W_\triangle $
denote the Zariski open
$\text {SL}(2, \mathbb {C})$
-orbit in
$\mathbb {C}P^3$
, and let
$Y_\triangle $
denote its complement.
$Y_\triangle $
consists of those
$3$
-point configurations in
$\mathbb {C}P^3$
where at least
$2$
of the points coincide. Let
$N_\triangle \subset Y_\triangle $
be the subvariety consisting of those configurations where all
$3$
points coincide. If
$[z_0:\ldots :z_3]$
are standard coordinates on
$\mathbb {P}S^3 V$
, then the roots of the polynomial

correspond (with multiplicity) to the
$3$
-tuple of points obtained by viewing
$[z]$
as a point of
$\text {Sym}^3 \mathbb {C}P^1$
. We count
$\infty $
as a root with multiplicity
$3 - \deg f$
. Consequently,
$Y_\triangle $
is defined by the vanishing of the discriminant
$\triangle (f)$
of f.
3.2 Orientation and spin structure
Let
$\alpha ,\beta ,\gamma $
be a basis of
$\mathfrak {su}(2)$
corresponding to infinitesimal right-handed rotations about a right-handed set of orthonormal axes. The infinitesimal group action gives rise to a frame for
$TL_\triangle ,$

where
$z\in L_\triangle .$
Let
$\mathfrak {o}_\triangle $
be the orientation on
$L_\triangle $
such that this frame is positively oriented. Let
$\mathfrak {s}_\triangle $
be the spin structure on
$L_\triangle $
such that this frame can be lifted to the associated double cover of the frame bundle.
3.3 Topological lemmas
Equip
$\mathbb {P}S^3V \cong \mathbb {C}P^3$
with the Fubini-Study Kähler form
$\omega $
such that
$\int _{\mathbb {C}P^1} \omega =1$
. The following lemma is given in Section 4.3 of [Reference Evans and Lekili10].
Lemma 3.1.
$H_1(L_\triangle ;\mathbb {Z})=\mathbb {Z} /4\mathbb {Z}.$
Lemma 3.2.
$H_2(L_\triangle ;\mathbb {Z})=0$
.
Proof. By Poincaré duality, it suffices to show
$H^1(L;\mathbb {Z})=0$
. By the universal coefficients theorem, we get the following short exact sequence

By Lemma 3.1
$H_1(L_\triangle ;\mathbb {Z})=\mathbb {Z}/4\mathbb {Z}$
. In addition,
$L_\triangle $
is connected, so
$H_0(L_\triangle ;\mathbb {Z})=\mathbb {Z}$
. Hence, we get an exact sequence
$ 0\rightarrow H^1(L_\triangle ;\mathbb {Z})\rightarrow 0, $
which yields
$H^1(L_\triangle ;\mathbb {Z})=0$
.
The following is Lemma 4.4.1 from [Reference Evans and Lekili10].
Lemma 3.3. The Chiang Lagrangian
$L_\triangle $
has minimal Maslov number equal to
$2.$
For
$\zeta \in H_2(\mathbb {C}P^3;\mathbb {Z})$
, we denote by
$c_1(\zeta )$
the evaluation of the first Chern class of
$\mathbb {C}P^3$
on
$\zeta $
. Let
$\xi = [\mathbb {C}P^1]\in H_2(\mathbb {C}P^3;\mathbb {Z})$
, so
$c_1(\xi )=4$
.
Lemma 3.4. The long exact sequence of the pair
$(\mathbb {C}P^3,L_\triangle )$
gives the short exact sequence

with
$H_2(\mathbb {C}P^3,L_\triangle ;\mathbb {Z}) \simeq \mathbb {Z} \simeq H_2(\mathbb {C}P^3;\mathbb {Z})$
, and the map
$\varpi $
is given by multiplication by 4.
Proof. Consider the long exact sequence

where

and Lemma 3.1 gives
$H_1(L_\triangle ;\mathbb {Z})\simeq \mathbb {Z}/4\mathbb {Z}.$
By Lemma 3.2, we have
$H_2(L_\triangle ;\mathbb {Z})=0$
, so we get the short exact sequence

We have two possibilities:
-
1. The short exact sequence (3.1) is split, so
$H_2(\mathbb {C}P^3,L_\triangle ;\mathbb {Z}) \simeq \mathbb {Z}\oplus \mathbb {Z}/4\mathbb {Z}$ . Thus, (3.1) is isomorphic to
$$\begin{align*}0\rightarrow \mathbb{Z} \overset{\varpi}{\rightarrow} \mathbb{Z}\oplus\mathbb{Z}/4\mathbb{Z} \rightarrow \mathbb{Z}/4\mathbb{Z}\rightarrow0, \end{align*}$$
$$\begin{align*}\varpi(x)= (x,0). \end{align*}$$
-
2. The short exact sequence (3.1) is not split, so
$H_2(\mathbb {C}P^3,L_\triangle ;\mathbb {Z})=\mathbb {Z}$ . Thus, (3.1) is isomorphic to
$$\begin{align*}0\rightarrow \mathbb{Z} \overset{\varpi}{\rightarrow} \mathbb{Z} \rightarrow \mathbb{Z}/4\mathbb{Z}\rightarrow0,\end{align*}$$
$$\begin{align*}\varpi(x)= 4x.\end{align*}$$
In case 1, the Maslov index is given by


Since
$\mu (\varpi (\xi ))=2c_1(\xi ) = 8$
, it follows that
$k=8$
is in contradiction to Lemma 3.3. Therefore, we must be in case 2.
In light of Lemma 3.4, let
$\tilde {\xi }\in H_2(\mathbb {C}P^3,L_\triangle ;\mathbb {Z})$
be the generator such that
$\varpi (\xi )=4\tilde {\xi }$
. Recall
$\Gamma _1=[\omega ]\in H^2(\mathbb {C}P^3,L_\triangle ;\mathbb {Z})$
.
Lemma 3.5. We have

Proof. We have

4 Bounding cochains and open Gromov-Witten invariants
In this section, we recall the definition of the Fukaya
$A_\infty $
algebra of a Lagrangian submanifold following [Reference Fukaya12, Reference Fukaya13, Reference Fukaya, Oh, Ohta and Ono17, Reference Fukaya, Oh, Ohta and Ono15, Reference Fukaya, Oh, Ohta and Ono16, Reference Fukaya, Oh, Ohta and Ono14, Reference Solomon and Tukachinsky47]. We recall the notion of bounding cochains from [Reference Fukaya, Oh, Ohta and Ono14] and the invariants
${OGW}_{\beta ,k}$
of [Reference Solomon and Tukachinsky46] mentioned in Section 2.1. Additionally, we recall results concerning bounding cochains from [Reference Solomon and Tukachinsky46]. We use these results to prove Theorem 4.19, from which Theorem 1.6 follows as a special case. These theorems give a class of open Gromov-Witten invariants which coincide with straightforward counts of J-holomorphic disks without corrections from bounding chains. Theorem 1.6 plays an essential role in the computation of the basic invariants of Theorem 1.2. We also prove Lemmas 4.5 and 4.6 concerning the affect of change of spin structure and orientation on the invariants
${OGW}.$
These lemmas are used to prove parts 7 and 8 of Proposition 2.2.
Throughout this section,
$(X, \omega )$
is a symplectic manifold of real dimension
$2n$
and
$L\subset X$
is a connected Lagrangian submanifold with relative spin structure
$\mathfrak {s}$
. The notion of a relative spin structure appeared in [Reference Fukaya, Oh, Ohta and Ono14]. See also [Reference Wehrheim and Woodward49]. In particular, a relative spin structure determines an orientation on L. Let J be an
$\omega $
-tame almost complex structure on X.
4.1 Moduli spaces
Denote by
$\mathcal {M}_{k+1,l}(\beta )$
the moduli space of J-holomorphic genus zero open stable maps
$u:(D, \partial D) \rightarrow (X,L)$
of degree
$\beta $
with one boundary component,
$k+1$
boundary marked points and l interior marked points. The boundary points are labeled according to their cyclic order. We denote elements in
$\mathcal {M}_{k+1,l}(\beta )$
by
$[u;z_0,\ldots ,z_k,w_1,\ldots ,w_l]$
, where
$z_i\in \partial D$
and
$w_i\in \operatorname {\mathrm {int}} D$
. Denote by


the evaluation maps associated to boundary marked points and to the interior marked points respectively.
Denote by
$\mathcal {M}^S_{k,l}(\beta )$
the moduli space of genus zero J-holomorphic open stable maps
$u:(D, \partial D) \rightarrow (X,L)$
of degree
$\beta $
with one boundary component, k unordered boundary points, and l interior marked points. It comes with evaluation maps as in the case of
$\mathcal {M}_{k,l}(\beta )$
.
We assume that all J-holomorphic genus zero open stable maps with one boundary component are regular and the moduli spaces
$\mathcal {M}_{k+1,l}(\beta ),\mathcal {M}^S_{k,l}(\beta )$
are smooth orbifolds with corners. In addition, we assume that the evaluation maps
$evb^\beta _0$
are proper submersions. These assumptions should hold for
$(X,L) = (\mathbb {C}P^3,L_\triangle )$
and J the standard complex structure on
$\mathbb {C}P^3$
by Remark 1.6 of [Reference Solomon and Tukachinsky47] in light of the transitive action of
$\text {SL}(4,\mathbb {C})$
on
$\mathbb {C} P^3$
and the transitive action of the subgroup
$\text {SU}(2) \subset \text {SL}(4,\mathbb {C})$
on
$L_\triangle $
. The arguments of the present section extend to arbitrary targets
$(X,\omega ,L)$
and arbitrary
$\omega $
-tame almost complex structures J given the virtual fundamental class techniques of [Reference Fukaya12, Reference Fukaya13, Reference Fukaya, Oh, Ohta and Ono14, Reference Fukaya, Oh, Ohta and Ono15, Reference Fukaya, Oh, Ohta and Ono16, Reference Fukaya, Oh, Ohta and Ono17, Reference Fukaya, Oh, Ohta and Ono19, Reference Fukaya, Oh, Ohta and Ono18, Reference Fukaya, Oh, Ohta and Ono20, Reference Fukaya, Oh, Ohta and Ono21, Reference Fukaya and Ono22]. Alternatively, it should be possible to use the polyfold theory of [Reference Hofer, Wysocki and Zehnder24, Reference Hofer, Wysocki and Zehnder27, Reference Hofer, Wysocki and Zehnder26, Reference Hofer, Wysocki and Zehnder25, Reference Li and Wehrheim35]. The relative spin structure
$\mathfrak {s}$
determines an orientation on
$\mathcal {M}_{k+1,l}(\beta ),\mathcal {M}^S_{k,l}(\beta )$
as in Chapter 8 of [Reference Fukaya, Oh, Ohta and Ono14].
4.2 The Fukaya
$A_\infty $
algebra
Denote by
$A^{*}(X;\mathbb {R})$
the ring of differential forms on X with coefficients in
$\mathbb {R}$
. For
$m>0$
, denote by
$\widetilde {A}^m(X,L;\mathbb {R})$
differential m-forms on X that vanish on L, and denote by
$\widetilde {A}^0(X,L;\mathbb {R})$
functions on X that are constant on L. Define

By Lemma 5.14 from[Reference Solomon and Tukachinsky46], if
$H^{*}(L;\mathbb {R})\simeq H^{*}(S^n;\mathbb {R})$
, then
$\widehat {H}^{*}(X,L;\mathbb {R})\simeq \widetilde {H}^{*}(X,L;\mathbb {R})$
. Denote by
$\beta _0$
the zero element of
$H_2(X,L;\mathbb {Z})$
.
Recall the definition of the Novikov ring
$\Lambda $
from Section 2.4. Let
$s,t_0,\ldots ,t_M$
, be formal variables with degrees in
$\mathbb {Z}$
. Set

thought of as differential graded algebras with trivial differential. Define a valuation

by

Whenever a tensor product (resp. direct sum) of modules with valuation is written, we mean the completed tensor product (resp. direct sum). Set

In particular, the gradings on C and D take into account the degrees of
$T^\beta , s,t_j$
, and the degrees of the differential forms. The valuation
$\nu $
induces valuations on
$\Lambda ,Q,C,D$
and their tensor products, which we also denote by
$\nu $
. Let

Let

Let
$\gamma \in \mathcal {I}_QD$
be a closed form with
$|\gamma |=2$
. For example, given closed differential forms
$\gamma _j\in \widetilde {A}^{*}(X,L;\mathbb {R})$
for
$j=0,\ldots ,M,$
take
$t_j$
of degree
$2-|\gamma _j|$
and
$\gamma :=\sum _{j=0}^Mt_j\gamma _j$
. Define maps

by

and for
$k\ge 0$
when
$(k,\beta )\ne (1,\beta _0)$
, by

The push-forward
$evb^\beta _{0_*}$
is given by integration over the fiber, which is well-defined because
$evb^\beta _0$
is a proper submersion. Define also

by

Furthermore, define

The following is Proposition 2.6 from [Reference Solomon and Tukachinsky47].
Theorem 4.1. The operations
$\{\mathfrak {m}_k^\gamma \}_{k\ge 0}$
define an
$A_\infty $
structure on C. That is, for any
$\alpha _1,\ldots ,\alpha _{k+1}\in C$
and
$k\in \mathbb {Z}_{\ge 0},$

The following is Proposition 3.1 from [Reference Solomon and Tukachinsky47].
Theorem 4.2. The operations
$\mathfrak {m}_k$
are R-multilinear. Namely, for any
$\alpha _1,\ldots ,\alpha _{k+1}\in C$
and
$a\in R$
,

Theorem 4.3. For
$k\ge 0$
and
$\omega (\beta )=0$
,

Proof. If
$\beta \ne \beta _0$
and
$\omega (\beta )=0$
, then
$\mathcal {M}_{k+1,l}=\emptyset $
. If
$\beta =\beta _0$
, see Proposition 3.8 in [Reference Solomon and Tukachinsky47].
Let
$\langle \ , \ \rangle $
denote the Poincaré pairing

The following is Proposition 3.1 from [Reference Solomon and Tukachinsky47].
Theorem 4.4. Let
$a\in R$
and
$\alpha _1,\ldots ,\alpha _{k+1}\in C$
. The Poincaré pairing satisfies the following:

The following is Proposition 3.3 from [Reference Solomon and Tukachinsky47].
Theorem 4.5. For
$\alpha _1,\ldots ,\alpha _{k+1}\in C$
,

4.3 Bounding pairs, the superpotential and invariants
Definition 4.1. A bounding pair with respect to J is a pair
$(\gamma ,b)$
where
$\gamma \in \mathcal {I}_Q D$
is closed with
$|\gamma |=2$
and
$b\in \mathcal {I}_R C$
with
$ |b|=1, $
such that

In this situation, b is called a bounding cochain for
$\mathfrak {m}^\gamma $
.
The definition of bounding pair appeared in [Reference Solomon and Tukachinsky46], and the definition of bounding cochain appeared in [Reference Fukaya, Oh, Ohta and Ono14].
Let
$\gamma \in \mathcal {I}_Q D $
,
$b\in \mathcal {I}_R C$
. The standard superpotential [Reference Fukaya13] is given by

Intuitively,
$\widehat {\Omega }$
counts J-holomorphic disks with constraints
$\gamma $
in the interior and b on the boundary. Modification is necessary in order to avoid J-holomorphic disks the boundary of which can degenerate to a point, forming a J-holomorphic sphere. We say that a monomial element of R is of type
$\mathcal {D}$
if it has the form
$a\, T^{\beta }s^0t_0^{j_0}\cdots t_N^{j_N}$
with
$a\in \mathbb {R}$
and
$\beta \in \operatorname {\mathrm {Im}}(\varpi )$
. Following [Reference Solomon and Tukachinsky46], in the present paper, the superpotential is defined by

Definition 3.12 from [Reference Solomon and Tukachinsky46] gives a notion of gauge equivalence between a bounding pair
$(\gamma ,b)$
with respect to J and a bounding pair
$(\gamma ',b')$
with respect to another almost complex structure
$J'.$
Let
$\sim $
denote the resulting equivalence relation. For a graded module M, we denote by
$(M)_j$
the degree j part of the module.
Define a map

by

By Lemma 3.16 in [Reference Solomon and Tukachinsky46],
$\varrho $
is well defined. The following is Theorem 1 from [Reference Solomon and Tukachinsky46].
Theorem 4.2. If
$(\gamma ,b)\sim (\gamma ',b')$
, then
$\Omega _J(\gamma ,b)=\Omega _{J'}(\gamma ',b')$
.
The following is Theorem 2 from [Reference Solomon and Tukachinsky46].
Theorem 4.3. Assume
$H^{*}(L;\mathbb {R})=H^{*}(S^n;\mathbb {R})$
. Then
$\varrho $
is bijective.
Fix
$\Gamma _0,\ldots ,\Gamma _M$
a basis of
$\widetilde {H}^{*}(X,L;\mathbb {R})$
, set
$|t_j|=2-|\Gamma _j|$
, and take

By Theorem 4.3, choose a bounding pair
$(\gamma ,b)$
such that

By Theorem 4.2, the superpotential
$\Omega =\Omega (\gamma ,b)$
is independent of the choice of
$(\gamma ,b)$
.
Definition 4.4. The open Gromov-Witten invariants of
$(X,L)$
,

are defined by setting

and extending linearly to general input.
The following is Proposition 5.2 from [Reference Solomon and Tukachinsky46].
Theorem 4.6. The invariants
$\mathrm {OGW}_{\beta ,k}$
are independent of the choice of a basis.
4.4 Changes of spin structure and orientation
The following Lemmas describe the dependence of the invariants
${OGW}_{\beta ,k}$
on the spin structure and the orientation of
$L.$
Lemma 4.5. Changing the spin structure on L by the action of
$\alpha \in H^1(L;\mathbb {Z}/2\mathbb {Z})$
changes
${OGW}_{\beta ,k}(\cdots )$
by multiplication by
$(-1)^{\alpha (\partial \beta )}.$
Proof. In the following proof, we add a superscript
$\mathfrak {s}$
to the
$A_\infty $
operations, superpotential and open Gromov-Witten invariants to indicate the spin structure used to construct them. By Lemma 2.10 of [Reference Solomon45], the orientations of
$\mathcal {M}_{k+1,l}(\beta )$
induced by
$\mathfrak {s}$
and
$\alpha \cdot \mathfrak {s}$
differ by a sign of
$(-1)^{\alpha (\partial \beta )}$
. So,

Let
$\mathfrak {f}_R : R \to R$
be the
$\mathbb {R}$
algebra automorphism that sends
$T^\beta $
to
$(-1)^{\alpha (\partial \beta )}T^\beta $
and acts trivially on the other formal variables. Then, the R module automorphism

is a strict
$A_\infty $
homomorphism between the
$A_\infty $
structures
$\mathfrak {m}_k^{\mathfrak {s}}$
and
$\mathfrak {m}_k^{\alpha \cdot \mathfrak {s}}$
and preserves the pairing
$\langle \cdot ,\cdot \rangle $
. So, if b is a bounding cochain for
$\mathfrak {m}_k^{\mathfrak {s}}$
with
$\int _L b = s$
, then
$\mathfrak {f}(b)$
is a bounding cochain for
$\mathfrak {m}_k^{\alpha \cdot \mathfrak {s}}$
with
$\int _L \mathfrak {f}(b) = s$
and
$\mathfrak {f}(\Omega ^{\mathfrak {s}}(\gamma ,b)) = \Omega ^{\alpha \cdot \mathfrak {s}}(\gamma ,\mathfrak {f}(b)).$
By Definition 4.4, we obtain

Lemma 4.6. Changing the orientation of L changes
${OGW}_{\beta ,k}(\Gamma _{i_1},\ldots ,\Gamma _{i_l})$
by multiplication by
$(-1)^{k+1}.$
Proof. In the following proof, we add a superscript
$\mathfrak {o}$
to the
$A_\infty $
operations, superpotential, Poincaré pairing,
$\int _L$
and open Gromov-Witten invariants to indicate the orientation used to construct them. By Lemma 2.9 from [Reference Solomon45], changing the orientation of L reverses the orientation of
$\mathcal {M}_{k+1,l}(\beta )$
. Since the relative orientation is not changed, it follows that

Let
$\mathfrak {f}_R : R \to R$
be the
$\mathbb {R}$
algebra automorphism that sends s to
$-s$
and acts trivially on the other formal variables. Then, the R module automorphism

is a strict
$A_\infty $
homomorphism between the
$A_\infty $
structures
$\mathfrak {m}_k^{\mathfrak {o}}$
and
$\mathfrak {m}_k^{-\mathfrak {o}}$
that satisfies

Similarly,
$\int _L^{\mathfrak {o}} = - \int _L^{-\mathfrak {o}}.$
So, if b is a bounding cochain for
$\mathfrak {m}_k^{\mathfrak {o}}$
with
$\int _L^{\mathfrak {o}} b=s$
, then
$\mathfrak {f}(b)$
is a bounding cochain for
$\mathfrak {m}_k^{-\mathfrak {o}}$
with
$\int _L^{-\mathfrak {o}} \mathfrak {f}(b)=s$
and
$\mathfrak {f}(\Omega ^{\mathfrak {o}}(\gamma ,b))=-\Omega ^{-\mathfrak {o}}(\gamma ,\mathfrak {f}(b))$
. By Definition 4.4, we obtain

We apply the preceding lemmas to prove vanishing results for the invariants
$\overline {OGW}\!$
of certain Lagrangian submanifolds
$L \subset X.$
The case
$n = 3$
and k even of the following corollary has been proved previously in [Reference Chen and Zinger6, Proposition 1.3] by a different method, which goes back to an observation of Mikhalkin [Reference Welschinger51, Remark 2.4].
Corollary 4.7. Consider
$(X,L) = (\mathbb {C} P^n,\mathbb {R} P^n)$
with n odd. Let
$\alpha \in H^1(\mathbb {R} P^n;\mathbb {Z})$
be the generator. The invariants
$\overline {OGW}\!_{\beta ,k}(\cdots )$
vanish when
$k + \alpha (\partial \beta )$
is even.
Proof. We identify
$H_2(\mathbb {C} P^n,\mathbb {R} P^n;\mathbb {Z})$
with
$\mathbb {Z}$
by the isomorphism taking the generator with positive symplectic area to
$1 \in \mathbb {Z}.$
It is shown in [Reference Solomon and Tukachinsky48, Corollary 1.9] that all the invariants
$\overline {OGW}\!_{\beta ,k}(\cdots )$
are determined by the open WDVV equations, the axioms of
$\overline {OGW}\!$
, the wall-crossing formula Theorem 2.3, the closed Gromov-Witten invariants of
$\mathbb {C} P^n,$
and the value of
$\overline {OGW}\!_{1,2}.$
If we simultaneously change the spin structure on
$\mathbb {R} P^n$
by
$\alpha $
and reverse the orientation, then Lemma 4.5 and Lemma 4.6 imply that
$\overline {OGW}\!_{\beta ,k}(\cdots )$
changes by multiplication by
$(-1)^{k + 1 + \alpha (\partial \beta )}.$
In particular
$\overline {OGW}\!_{1,2}$
remains unchanged. But since all the invariants
$\overline {OGW}\!_{\beta ,k}(\cdots )$
are determined from
$\overline {OGW}\!_{1,2}$
in a way that does not depend on the spin structure or the orientation of
$\mathbb {R} P^n,$
they must also remain unchanged. So, when
$k + 1+\alpha (\partial \beta )$
is odd,
$\overline {OGW}\!_{\beta ,k}(\cdots )$
must vanish.
The following was obtained by a different method in [Reference Hugtenburg and Tukachinsky29, Lemma 6.8].
Corollary 4.8. Take X to be the quadric hypersurface in
$\mathbb {C} P^{n+1}$
given by
$\sum _{i = 0}^n z_i^2 - z_{n+1}^2 = 0$
and L to be the real locus. Then
$\overline {OGW}\!_{\beta ,k}(\cdots )$
vanishes when k is even.
Proof. We identify
$H_2(X,L;\mathbb {Z})$
with
$\mathbb {Z}$
by the isomorphism taking the generator with positive symplectic area to
$1 \in \mathbb {Z}.$
It is shown in [Reference Hugtenburg and Tukachinsky29, Theorem 7] that all the invariants
$\overline {OGW}\!_{\beta ,k}(\cdots )$
are determined by the open WDVV equations, the axioms of
$\overline {OGW}\!$
, the wall-crossing formula Theorem 2.3, the closed Gromov-Witten invariants of
$\mathbb {C} P^n,$
and the value of
$\overline {OGW}\!_{1,3}.$
If we reverse the orientation of L, then Lemma 4.6 asserts that
$\overline {OGW}\!_{\beta ,k}(\cdots )$
changes by multiplication by
$(-1)^{k + 1}.$
In particular,
$\overline {OGW}\!_{1,3}$
remains unchanged. But since all the invariants
$\overline {OGW}\!_{\beta ,k}(\cdots )$
are determined from
$\overline {OGW}\!_{1,3}$
in a way that does not depend on the orientation of
$L,$
they must also remain unchanged. So, when
$k + 1$
is odd,
$\overline {OGW}\!_{\beta ,k}(\cdots )$
must vanish.
Remark 4.9. Let
$(X,L)$
be as in Corollary 4.8. It is shown in [Reference Hugtenburg and Tukachinsky29, Prop. 6.9] that the vanishing result of Corollary 4.8 implies that changing the sign of the invariant
$\overline {OGW}\!_{1,3},$
from which all the other invariants
$\overline {OGW}\!$
are determined by a recursion, does not affect the absolute value of other invariants. Consequently, hypothesis 2 of Section 1.3 cannot hold for this choice of
$(X,L).$
The proof of Corollary 4.8 is based on the tension between the fact that all invariants are determined recursively from a single invariant, and the freedom in the choice of the orientation of L, which affects the values of invariants through Lemma 4.6. Thus, it seems natural to restrict hypothesis 2 to
$(X,L)$
where such tension is not present. That is, the number of geometric degrees of freedom in the definition of invariants should not exceed the number of invariants from which all others are determined by recursions that do not depend on these degrees of freedom.
4.5 Obstruction theory
The goal of the present section is the proof of Lemma 4.18, which refines the existence result for bounding cochains given in Proposition 3.4 of [Reference Solomon and Tukachinsky46]. For this purpose, we recall relevant parts of the obstruction theoretic construction of bounding cochains given there. The basic idea goes back to [Reference Fukaya, Oh, Ohta and Ono14].
Let

denote the completed tensor algebra. For
$x\in \mathcal {I}_RC$
, abbreviate

Moreover, define

by

Any element
$\alpha \in C$
can be written as

Define a filtration
$F^E$
on R, C by

Definition 4.10. A multiplicative submonoid
$G\subset R$
is sababa if it can be written as a list

such that
$i<j\Rightarrow \nu (\lambda _i)\le \nu (\lambda _j).$
For
$j=1,\ldots ,m,$
and elements
$\alpha _j=\sum _i\lambda _{ij}\alpha _{ij}\in C$
decomposed as in (4.6), denote by
$G(\alpha _1,\ldots ,\alpha _m)$
the multiplicative monoid generated by
$\{\pm T^\beta \,|\,\beta \in \Pi _\infty \}$
,
$\{t_j\}_{j=0}^N$
, and
$\{\lambda _{ij}\}_{i,j}$
. The following is Lemma 3.3 from [Reference Solomon and Tukachinsky46].
Lemma 4.11. For
$\alpha _1,\ldots ,\alpha _m\in \mathcal {I}_R$
, the monoid
$G(\alpha _1,\ldots ,\alpha _m)$
is sababa.
For
$\alpha _1,\ldots ,\alpha _m\in \mathcal {I}_R$
, write the image of
$G=G(\alpha _1,\ldots ,\alpha _m)$
under
$\nu $
as the sequence
$\{E_0^G=0, E^G_1, E^G_2,\ldots \}$
with
$E_i^G<E^G_{i+1}$
. Let
$\kappa ^G_i\in \mathbb {Z}_{\ge 0}$
be the largest index such that
$\nu (\lambda _{\kappa _i^G})=E^G_i$
. In the future, we omit G from the notation and simply write
$E_i,\kappa _i$
since G will be fixed in each instance and no confusion should occur.
Given
$\alpha \in C$
, a differential form with coefficients in R, we denote by
$(\alpha )_j \in C$
the summand of differential form of degree j in
$\alpha $
. Let
$\Upsilon '$
be an
$\mathbb {R}-$
vector space, let
$\Upsilon "=R, Q,$
or
$\Lambda ,$
and let
$\Upsilon =\Upsilon '\otimes \Upsilon "$
. For
$x\in \Upsilon $
and a monomial
$\lambda \in \Upsilon ",$
denote by
$[\lambda ](x)\in \Upsilon '$
the coefficient of
$\lambda $
in x.
Fix a sababa multiplicative monoid
$G=\{\lambda _j\}_{j=0}^\infty \subset R$
ordered as in (4.7). Let
$l\ge 0.$
Suppose we have
$b_{(l)}\in C$
with
$|b_{(l)}|=1$
, and

Define the obstruction cochains
$o_j\in A^{*}(L)$
for
$j=\kappa _l+1,\ldots ,\kappa _{l+1}$
to be

Lemmas 4.12–4.17 are Lemmas 3.5–3.10 from [Reference Solomon and Tukachinsky46].
Lemma 4.12.
$|o_j|=2-|\lambda _j|$
.
Lemma 4.13.
$do_j=0$
.
Lemma 4.14. If
$|\lambda _j|=2,$
then
$o_j=c_j\cdot 1$
for some
$c_j\in \mathbb {R}.$
If
$|\lambda _j|\ne 2,$
then
$o_j\in A^{>0}(L;\mathbb {R})$
.
Lemma 4.15. If
$|\lambda _j| = 2-n$
and
$(db_{(l)})_n = 0,$
then
$o_j = 0$
.
Lemma 4.16. Suppose for all
$j\in \{\kappa _l+1,\ldots ,\kappa _{l+1}\}$
such that
$|\lambda _j|\ne 2$
, there exist
$b_j\in A^{1-|\lambda _j|}(L;\mathbb {R})$
such that
$(-1)^{|\lambda _j|}db_j=-o_j.$
Then

satisfies

Lemma 4.17. Let
$\zeta \in \mathcal {I}_R C$
. Then
$\mathfrak {m}^{\gamma }(e^\zeta )\equiv 0\pmod {F^{E_0}C}.$
Lemma 4.18. Let
$b_0\in C$
such that

Then there exists a bounding cochain b, such that
-
1.
$b\equiv b_0 \pmod {R^+}$ ,
-
2.
$\int _L b=\int _L b_0$ ,
-
3.
$(b)_j=(b_0)_j$ where
$j\in \{n-1,n\}$ .
Proof. We follow the proof of Proposition 3.4 from [Reference Solomon and Tukachinsky46]. Define
$b_{(0)}:=b_0$
. By Lemma 4.17, the cochain
$b_{(0)}$
satisfies

Moreover,
$|b_{(0)}|=1$
,
$\int _L b_{(0)}=\int _L b_0$
,
$(db_{(0)})_n=0$
, and
$(b_{(0)})_j=(b_0)_j$
where
$j\in \{n-1,n\}$
. Observe also that
$F^{E_0}C \subset R^+C$
. Proceed by induction. Suppose we have
$b_{(l)}\in C$
such that
$|b_{(l)}|=1$
,
$b_{(l)}\equiv b_0 \pmod {R^+C}$
, and

By Lemma 4.13, we have
$do_j=0$
, and by Lemma 4.12, we have
$o_j\in A^{2-|\lambda _j|}(L;\mathbb {R}).$
In order to apply Lemma 4.16, we need to find forms
$b_j\in A^{1-|\lambda _j|}(L;\mathbb {R})$
such that
$(-1)^{|\lambda _j|}db_j=-o_j$
for all
$j\in \{\kappa _l+1,\ldots ,\kappa _{l+1}\}$
such that
$|\lambda _j|\ne 2.$
Since
$F^{E_l}C \subset R^+C,$
we have

It follows that
$o_j=0$
when
$\lambda _j\not \in R^+$
, so we choose
$b_j=0$
. Hence, we may assume
$\lambda _j\in R^+$
. If
$|\lambda _j|= 2-n,$
since
$(db_{(l)})_n =0$
, Lemma 4.15 gives
$o_j = 0,$
so we choose
$b_j = 0.$
If
$2-n < |\lambda _j|<~2,$
then
$0 < |o_j| < n.$
The assumption
$H^{*}(L;\mathbb {R})\simeq H^{*}(S^n;\mathbb {R})$
implies
$[o_j] = 0 \in H^{*}(L;\mathbb {R}),$
so we choose
$b_j$
such that
$(-1)^{|\lambda _j|}db_j = -o_j.$
For other possible values of
$|\lambda _j|,$
degree considerations imply
$o_j = 0,$
so we choose
$b_j =0.$
By Lemma 4.16,
$b_{(l+1)}:=b_{(l)}+\sum _{\substack {\kappa _l+1\le j\le \kappa _{l+1}\\ |\lambda _j|\ne 2}}\lambda _jb_j$
satisfies

Since
$b_j=0$
when
$\lambda _j\not \in R^+$
, it follows that
$b_{(l+1)}\equiv b_{(l)}\equiv b_0 \pmod {R^+C}$
. Since
$b_j = 0$
when
$|\lambda _j| = 2-n$
, it follows that
$(b_{(l+1)})_{n-1}=(b_{(l)})_{n-1}=(b_{(0)})_{n-1}$
, and that
$(db_{(l+1)})_n = (db_{(l)})_n=0$
. In addition,
$b_j = 0$
when
$|\lambda _j| = 1-n$
. Thus,
$(b_{(l+1)})_{n}=(b_{(l)})_{n}=(b_{(0)})_{n},$
and
$\int _L b_{(l+1)}=\int _L b_{(l)}=\int _L b_0$
. The inductive process gives rise to a convergent sequence
$\{b_{(l)}\}_{l=0}^\infty $
where
$b_{(l)}$
is bounding modulo
$F^{E_l}C$
. Taking the limit as l goes to infinity, we obtain


4.6 Straightforward counts
In the following, we will be using the following notation conventions. For
$k\ge 1$
, denote by
$[k]$
the set
$[k]:= \{1,\ldots ,k \}.$
For
$N>0$
, we denote
$t:= (t_1,\ldots ,t_N)$
. Similarly, we denote by
$\alpha $
an N-tuple where all the components are taken from the set
$\mathbb {Z}_{\ge 0}$
; that is,

We denote
$t^\alpha := t_1^{\alpha _1}\cdots t_N^{\alpha _N}$
. Recall the definition of
$P_\beta $
from (1.1).
Theorem 4.19. Let
$A_{j}\in \widetilde {H}^{*}(X,L;\mathbb {R})$
. Let
$\bar {b}\in A^n(L;\mathbb {R})$
such that
$\text {PD}([\bar {b}])=pt$
, and let
$a_{j}$
be a representative of
$A_{j}$
. Let
$\sigma _k = (k-1)!$
for
$k \in \mathbb {Z}_{>0}$
and let
$\sigma _0 = 1.$
Then,

if for every

one of the following two conditions is satisfied:
-
1.
$1-\mu (\tilde {\beta }) +j(n-1) + \sum _{i\in I} (|A_{i}|-2)<0,$ or
-
2.
$1-\mu (\tilde {\beta }) +j(n-1) + \sum _{i\in I} (|A_{i}|-2)\ge n-1.$
Proof. Let
$\Gamma _0,\ldots ,\Gamma _M$
be a basis of
$\widetilde {H}^{*}(X,L;\mathbb {R})$
with
$\Gamma _0 = 1$
. From multilinearity of
${OGW}$
and Proposition 4.6, it suffices to prove the lemma for
$\mathrm {OGW}_{\beta ,k}(\Gamma _0^{\otimes {r_0}}\otimes \cdots \otimes \Gamma _M^{\otimes {r_M}})$
. By the unit axiom 4, we can assume that
$r_0=0$
, and by the zero axiom 5, we can assume that
$\beta \ne \beta _0$
. We have
$\gamma = \sum _{i=1}^M t_i \gamma _i$
, where
$\gamma _i$
is a representative of
$\Gamma _i\in \widetilde {H}^{*}(X,L;\mathbb {R})$
. Let
$b_{0,1,0}\in A^n(L;\mathbb {R})$
be a representative of the Poincaré dual of a point. Since by Proposition 4.3 we have

it follows by Lemma 4.18 that there exists a bounding cochain
$b\in C$
such that
$b\equiv s\cdot b_{0,1,0} \pmod {R^+C}$
, and

Write
$b=s\cdot b_{0,1,0}+\sum T^{\tilde {\beta }}s^j t^\alpha b_{\tilde {\beta },j,\alpha }$
, where

Since

it follows that

Hence, if conditions 1 and 2 are satisfied, then by (4.8) we must have
$b_{\tilde {\beta },j,\alpha }=0$
when

Define

and denote

Since
$b'\equiv b \pmod {JC}$
, it follows that

Hence, it suffices to consider
$\Omega (\gamma ,b')$
instead of
$\Omega (\gamma ,b)$
in the computation of
$\mathrm {OGW}_{\beta ,k}(\Gamma _1^{\otimes {r_1}}\otimes \cdots \otimes \Gamma _M^{\otimes {r_M}})$
. Since

by Proposition 4.3, for every
$m\ge 0$
,
$0\le j\le k$
and
$\tilde {\beta }\in H_2(X,L;\mathbb {Z})$
such that
$\omega (\tilde {\beta })=0$
, we get

So, by Proposition 4.5 for every
$1\le i\le m$
, we get

Hence,

Denote
$l=\sum _{i=1}^Mr_i.$
For
$k=0$
, we get

where
$\mathcal {M}_{0,l}(\beta )=\mathcal {M}^S_{0,l}(\beta ).$
For
$k\ne 0$
, we get

The diffeomorphism of
$\mathcal {M}^S_{k,l}(\beta )$
corresponding to relabeling boundary marked points by a permutation
$\sigma \in S_k$
preserves or reverses orientation depending on
$sgn(\sigma )$
. Let
$\sigma \in S_k$
be a permutation. Denote by
$\mathcal {M}_{\sigma (k),l}(\beta )$
the moduli space obtained from
$\mathcal {M}_{k,l}(\beta )$
by relabeling boundary marked points by
$\sigma $
. So,

Therefore, since
$\mathcal {M}_{k,l}^S(\beta ) = \coprod _{\sigma \in S_k} \mathcal {M}_{\sigma (k),l}(\beta ),$
it follows that

Proof of Theorem 1.6.
By Lemma 5.14 of [Reference Solomon and Tukachinsky46], if
$H^{*}(L;\mathbb {R})\simeq H^{*}(S^n;\mathbb {R})$
, then
$\widehat {H}^{*}(X,L;\mathbb {R})\simeq \widetilde {H}^{*}(X,L;\mathbb {R})$
. Recall that if
$k\ne 0$
or
$\beta \not \in \operatorname {\mathrm {Im}} \varpi $
, then
$\overline {OGW}\!_{\beta ,k}={OGW}_{\beta ,k}$
. Thus, when
$k\ne 0$
or
$\beta \not \in \operatorname {\mathrm {Im}} \varpi $
, Theorem 4.19 yields Theorem 1.6.
5 Computation of basic invariants
In order to use the recursive formula to compute the open Gromov Witten invariants of the Chiang Lagrangian
$L_\triangle ,$
we need the initial values for the recursive formula, which are given by the following theorems. Fix the orientation
$\mathfrak {o}_\triangle $
and the spin structure
$\mathfrak {s}_\triangle $
on
$L_\triangle $
as defined in Section 3.2.
Theorem 5.1.
$\overline {OGW}\!_{1, 1} = -3$
.
Theorem 5.2.
$\overline {OGW}\!_{1, 0}(\Gamma _2) =\frac {1}{4}$
.
Theorem 5.3.
$\overline {OGW}\!_{2, 0}(\Gamma _3) = 1$
.
The proofs appear below.
5.1 Intersections with the compactification divisor
Recall from Section 3.1 the definition of the Chiang Lagrangian
$L_\triangle \subset X_\triangle \simeq \mathbb {C} P^3 \simeq \text {Sym}^3\mathbb {C}P^1.$
Recall also the definitions of the discriminant locus
$Y_\triangle \subset X_\triangle $
and the subvariety
$N_\triangle \subset Y_\triangle .$
Lemma 5.4.
$\mathbb {C}P^3$
is Fano with anticanonical divisor
$Y_{\triangle }$
.
Proof. The proof is part of Section 3.4 in [Reference Smith43].
The following is Lemma 3.1 from [Reference Auroux2].
Lemma 5.5. If L is special Lagrangian in the complement of an anticanonical divisor Y in a compact Kähler manifold X, then the Maslov index of a disk
$u: (D, \partial D) \rightarrow (X, L)$
is given by twice the algebraic intersection number
$[u] \cdot [Y]$
.
The following is Lemma 3.7 from [Reference Smith43].
Lemma 5.6. A clean intersection of a holomorphic disk u with
$N_{\triangle }$
contributes at least 2 to the intersection number
$[u]\cdot [Y_{\triangle }]$
. A non-clean intersection contributes at least 3.
5.2 Axial disks
The following is Definition 2.1 from [Reference Smith43].
Definition 5.7. If X is a complex manifold carrying an action of a compact Lie group K by holomorphic automorphisms, and L is a totally real submanifold which is an orbit of the K-action, then we say
$(X,L)$
is K-homogeneous.
Example 5.8. The pair
$(\mathbb {C}P^3,L_\triangle )$
is
$\text {SU}(2)$
-homogeneous as explained in Section 3.1.
The following is Definition 2.3 from [Reference Smith43].
Definition 5.9. Let
$(X, L)$
be K-homogeneous. If
$u: (D, \partial D) \rightarrow (X, L)$
is a holomorphic disk, and there exists a smooth group homomorphism
$R: \mathbb {R} \rightarrow K$
such that (possibly after reparametrizing u) we have
$u(e^{i\theta } z) = R(\theta )u(z)$
for all
$z \in D$
and all
$\theta \in \mathbb {R}$
, then we say u is axial. Let
$\xi \in \mathfrak {k} = \operatorname {Lie}(K).$
We say that u is axial of type
$\xi $
if
$\dot R(0)$
belongs to the orbit of
$\xi $
under the adjoint action of K.
Thinking of
$\mathbb {C}P^1$
as
$\mathbb {C} \cup \{\infty \}$
, we identify
$\mathbb {C}P^1 \simeq S^2$
by stereographic projection from the north pole. Thus, the complex structure on the
$2$
-sphere is given by left-handed rotation around the outward normal by an angle of
$\pi /2.$
For concreteness, we choose
$\triangle $
to be the triangle with vertices

By abuse of notation, we also denote by
$\triangle $
the point
$[c_1,c_2,c_3]\in \text {Sym}^3\mathbb {C}P^1.$
Let
$\xi _v\in \mathfrak {su}(2)$
be the infinitesimal right-handed rotation about the z axis and let
$\xi _f\in \mathfrak {su}(2)$
be the infinitesimal right-handed rotation about the y axis scaled so that for
$\xi = \xi _v,\xi _f,$
we have
$\{t \in \mathbb {R}: e^{2\pi \xi t}\cdot \triangle = \triangle \} =\mathbb {Z}$
. Thus,
$\xi _v,\xi _f,$
are generators of rotations about the vertex
$c_1$
and the center of the face of the triangle, respectively, as shown in Figure 1.

Figure 1 The choice of
$\xi _v, \xi _f$
for the configuration
$\triangle $
.
Example 5.10 (Axial Maslov 2 disk).
Consider the homomorphism

given by

Let
$u: (D,\partial D) \to (\mathbb {C}P^3,L_\triangle )$
be given by
$u(z)=e^{-i\xi _v\log z}\cdot \triangle $
. See below for more explicit formulas. Then u satisfies
$u(e^{i\theta }z)=R(\theta )u(z),$
so it is an axial disk of type
$\xi _v$
.
Recall we denote by
$c_1,c_2,c_3$
the vertices of the triangle
$\triangle $
. The stereographic projection from the north pole

is given by

We have

Since
$\xi _v$
is the generator of a rotation about the vertex
$c_1$
, the flow of
$\xi _v$
is given by

for some
$\chi \in \mathbb {R}.$
Since
$\xi _v$
is normalized so that

we have
$\chi = \frac {1}{2}$
and

Hence, for
$z\in D$
, we get

Hence, we can write
$u(z)= [0:1:0: -\frac {z}{3}]\in \mathbb {C}P^3.$
We can describe u geometrically as follows. The boundary of u is obtained by the action of
$R(\theta )$
on
$\triangle $
. Call an isosceles triangle narrow if the congruent sides are longer than the base. A point in the interior of u is described by a narrow isosceles triangle on a great circle where the north pole is the apex. For example, the triangle
$c_1d_2d_3$
in Figure 2 represents a point in the interior of u. Note that as
$z\rightarrow 0$
, the vertices
$c_2$
and
$c_3$
move toward the south pole, so u intersects
$Y_{\triangle }$
in a single point. So,
$[u]\cdot [Y_{\triangle }]\ge 1$
. The proof of Lemma 3.8 in [Reference Smith43] shows that equality holds. Hence, by Lemma 5.5, u is a Maslov 2 disk.

Figure 2 A Maslov 2 disk passing through
$\triangle $
.
Example 5.11 (Axial Maslov 4 disk).
Consider the homomorphism

given by

The disk
$u:(D,\partial D) \to (\mathbb {C}P^3,L_\triangle )$
given by
$u(z)=e^{-i\xi _f\log z}\cdot \triangle $
satisfies
$u(e^{i\theta }z)=R(\theta )u(z),$
so it is axial of type
$\xi _f$
.
We can describe u geometrically as follows. The boundary of u is obtained by the action of
$R(\theta )$
on
$\triangle $
. The interior of u is given by all the equilateral triangles on northern lines of latitude if we take
$r:=(0,-1,0)$
for the north pole. For example, the triangle
$s_1s_2s_3$
in Figure 3 represents an interior point of
$u.$
Since u intersects
$N_\triangle $
at the unique point
$[r,r,r]$
, it follows by Lemma 5.6 that
$[u]\cdot [Y_{\triangle }] \geq 2.$
The proof of Lemma 3.8 in [Reference Smith43] shows that equality holds. Hence, by Lemma 5.5, u is a Maslov 4 disk.

Figure 3 A Maslov 4 disk passing through
$\triangle $
.
5.3 Classification of holomorphic disks
The following definition is given in Section 4 from [Reference Smith43].
Definition 5.12. The intersection points of a holomorphic disk
$u:(D,\partial D)\rightarrow (\mathbb {C}P^3, L_\triangle )$
with the compactification divisor
$Y_{\triangle }$
are called poles of u.
The following is Definition 4.6 from [Reference Smith43]. Recall from Section 3.1 the definition of
$\Gamma _\triangle .$
Definition 5.13. A pole germ is the germ (at 0) of a holomorphic map
$u,$
from an open neighborhood of
$0$
in
$\mathbb {C}$
to
$\mathbb {C}P^3,$
such that
$u^{-1}(Y_{\triangle })$
contains
$0$
as an isolated point. More generally, for a Riemann surface
$\Sigma $
and a point
$a\in \Sigma ,$
one can speak of a pole germ at
$a.$
If we do not specify ‘at a’, then we are implicitly working at
$0$
in
$\mathbb {C}.$
We define an equivalence relation on pole germs at a by
$u_1\sim u_2$
if and only if there exists a germ of holomorphic map
$A,$
from a neighborhood of a in
$\Sigma $
to
$\mathrm {SL}(2,\mathbb {C}),$
such that
$u_2=A\cdot u_1.$
The principal part of a pole germ u is its equivalence class
$[u]_a$
under this relation.
We say a pole germ u is of type
$\xi \in \mathfrak {su}(2)$
and order
$k\in \mathbb {Z}_{\ge 1}$
if its principal part is

and
$\xi $
is scaled so that
$\{t\in \mathbb {R}: e^{2\pi \xi t}\in \Gamma _\triangle \}=\mathbb {Z}$
.
Example 5.14. An axial disk of type
$\xi $
has a pole germ of type
$\xi $
and order
$1$
.
The following is Lemma 4.10 from [Reference Smith43].
Lemma 5.15. A pole germ u is of type
$\xi _v$
if and only if
$u(0) \in Y_\triangle \backslash N_\triangle $
.
The following is Corollary 4.13 from [Reference Smith43].
Lemma 5.16. All Maslov index
$2$
holomorphic disks
$u: (D,\partial D)\to (\mathbb {C}P^3,L_\triangle )$
are, up to reparametrization, of the form

for
$A\in \text {SU}(2)$
. In particular, they are all axial.
The following is Corollary 4.14 from [Reference Smith43].
Lemma 5.17. Suppose
$u : (D,\partial D) \to (\mathbb {C}P^3,L_\triangle )$
is a holomorphic disk of Maslov index 4. Then either u has two poles of type
$\xi _v$
and order
$1$
, one pole of type
$\xi _v$
and order
$2$
, or one pole of type
$\xi _f$
and order
$1$
. In the last case, the disk is axial of type
$\xi _f$
.
Recall the notation for moduli spaces of holomorphic disks from Section 4.1. The following is Corollary 4.21 from [Reference Smith43].
Lemma 5.18. Let
$u : (D,\partial D) \to (\mathbb {C}P^3,L_\triangle )$
be an axial disk of type
$\xi _f$
. Then for any
$w \in \operatorname {\mathrm {int}} D$
, the evaluation map

is a submersion at
$[u;w]$
.
Lemma 5.19. Let
$a,b\in \mathbb {C}P^1$
be two sufficiently close points. There exists a unique point
$c\in \mathbb {C}P^1$
such that there exists a holomorphic disk
$u:(D, \partial D)\rightarrow (\mathbb {C}P^3,L_\triangle )$
of Maslov index 2 passing through
$[a, b, c]\in \text {Sym}^3\mathbb {C}P^1 = \mathbb {C}P^3.$
This disk is unique up to reparameterization.
Proof. Following Example 5.10, we call an isosceles triangle narrow if the congruent sides are longer than the base. Since
$a,b,$
are sufficiently close, they cannot belong to an equilateral triangle on a great circle, so a point of the form
$[a,b,c]$
must be in
$\mathbb {C}P^3\setminus L_\triangle .$
Thus, if a J-holomorphic disk
$u : (D,\partial D) \to (\mathbb {C}P^3,L_\triangle )$
passes through a point of the form
$[a,b,c],$
then
$[a,b,c] \in u(\operatorname {\mathrm {int}} D).$
By Lemma 5.16, all J-holomorphic disks
$u: (D,\partial D) \to (\mathbb {C}P^3,L_\triangle )$
of Maslov index
$2$
have the form
$u(z)=A\cdot e^{-i\xi _v\log z}\cdot \triangle .$
For such
$u,$
it follows from the geometric explanation given in Example 5.10 that a point of
$u(\operatorname {\mathrm {int}}{D} )$
is given by the vertices of a narrow isosceles triangle with apex
$A \cdot (0,0,1).$
So, since
$a,b,$
are sufficiently close, if u passes through
$[a,b,c]$
, then
$a,b,c,$
must be the vertices of a narrow isosceles triangle on a great circle with apex
$c = A \cdot (0,1,1).$
We claim that such c is unique. Indeed, the points a and b are sufficiently close, so they determine a unique great circle. In addition, there exists a unique point c on the great circle such that the triangle with vertices
$a,b,c,$
is narrow isosceles. The condition
$c = A \cdot (0,0,1)$
determines A up to the action of
$S^1 = \operatorname {\mathrm {Stab}}_{(0,0,1)} \subset \text {SU}(2).$
This
$S^1$
acts on u by reparameterization.
Lemma 5.20. Let
$p\in N_\triangle .$
There exists an axial disk of type
$\xi _f$
that passes through
$p.$
This disk is unique up to reparameterization.
Proof. Let
$q\in \mathbb {C}P^1$
such that
$p:=[q,q,q].$
Denote
$r:=(0,-1,0)$
. By definition, the general axial disk of type
$\xi _f$
has the form
$u(z)=A\cdot e^{-i\xi _f\log z}\cdot \triangle .$
It follows from Example 5.11 that such a disk intersects
$N_\triangle $
at the unique point
$[A\cdot r,A\cdot r,A\cdot r].$
Let
$A\in \text {SU}(2)$
be a rotation such that
$A\cdot r=q.$
Then
$u(0) = p.$
However, the condition
$A\cdot r = q$
determines A up to the action of
$S^1 = \operatorname {\mathrm {Stab}}_{q} \subset \text {SU}(2).$
This
$S^1$
acts on u by reparameterization.
5.4 Riemann-Hilbert pairs
Definition 5.21. A Riemann-Hilbert pair consists of holomorphic rank n vector bundle
$E\rightarrow D$
over the closed unit disk with a smooth totally real n-dimensional subbundle
$F\subset E|_{\partial D}$
. A spin Riemann-Hilbert pair is a Riemann-Hilbert pair
$(E,F)$
along with an orientation and spin structure on
$F.$
Definition 5.22. Let
$(E,F)$
,
$(E',F')$
be spin Riemann-Hilbert pairs. An isomorphism of Riemann-Hilbert pairs
$(E,F)\rightarrow (E',F')$
consists of
-
1. a biholomorphism
$f:D\rightarrow D$ , and
-
2. a holomorphic isomorphism of bundles
$\phi :E\rightarrow E'$ covering f such that
$\phi |_{\partial D}$ takes F to
$F'$ .
If
$(E,F)$
and
$(E',F')$
are spin Riemann-Hilbert pairs, we say that
$(f,\phi )$
is an isomorphism if additionally,
$\phi |_F : F \to F'$
preserves orientation and spin structure. We may refer to the pair
$(f,\phi )$
by
$\phi $
alone.
Let
$(E,F)$
be a Riemann-Hilbert pair. Since D is contractible, we identify E with the trivial bundle
$\underline {\mathbb {C}}^n$
, and thus, the pair is determined by the subbundle F. A totally real subspace of
$\mathbb {C}^n$
is described by
$A\cdot \mathbb {R}^n$
where
$A\in GL(n,\mathbb {C})$
. Moreover,
$A' \in GL(n,\mathbb {C})$
gives the same totally real subspace if and only if
$A^{-1}A'\in GL(n,\mathbb {R})$
. The fibers of a totally real subbundle
$F \subset \underline {\mathbb {C}}^n$
can be written as
$F_z=A(z)\cdot \mathbb {R}^n$
where
$A(z)\in GL(n,\mathbb {C})$
, so such a family of matrices
$\{A(z)\}_{z\in \partial D}$
determines a Riemann-Hilbert pair.
Let
$F^\kappa \subset \underline {\mathbb {C}}$
denote the totally real subbundle given by
$F^\kappa _z= z^{\kappa /2}\mathbb {R}$
and let
$F^{\kappa _1,\kappa _2}\subset \underline {\mathbb {C}}^2$
denote the totally real subbundle given by
$F^{\kappa _1,\kappa _2}_z= z^{\kappa _1/2}\mathbb {R}\oplus z^{\kappa _2/2}\mathbb {R}.$
The following is Theorem 1 from [Reference Oh39].
Theorem 5.23. Any Riemann-Hilbert pair
$(E,F)$
splits as a direct sum of Riemann-Hilbert pairs

where
$\kappa _i\in \mathbb {Z}$
.
If the boundary real subbundle F is orientable, which is equivalent to having even total index, by Theorem 5.23, we can write

where
$\kappa _i$
and
${\kappa _1}_i+{\kappa _2}_i$
are even numbers. Hence, each summand is orientable.
For a Riemann-Hilbert pair
$(E,F),$
let

denote the Cauchy-Riemann operator determined by the holomorphic structure on E. Let
$\det \bar {\partial }_{(E,F)}$
denote the Fredholm determinant. Abbreviate

The following is a special case of Proposition 2.8 from [Reference Solomon45].
Theorem 5.1. Let
$(E,F)$
be a spin Riemann-Hilbert pair. Then
$\det \bar {\partial }_{(E,F)}$
admits a canonical orientation. If
$\phi :(E,F)\rightarrow (E',F')$
is an isomorphism of spin Riemann-Hilbert pairs, then the induced isomorphism

preserves the canonical orientation. Furthermore, the canonical orientation varies continuously in a family of Cauchy-Riemann operators.
The following is a consequence of the proof of a special case of Lemma 8.1 from [Reference Solomon45].
Lemma 5.24. Suppose

is a short exact sequence of real vector bundles over a base
$B.$
Then an orientation on any two of
$V, V' , V"$
naturally induces an orientation on the third. A spin structure on any two of
$V, V' , V"$
naturally induces a spin structure on the third.
Suppose now that
$V',V"$
are oriented and spin and equip
$V = V' \oplus V"$
with the induced orientation and spin structure. If
$\xi ^{\prime }_1,\ldots ,\xi ^{\prime }_{k'}$
(resp.
$\xi ^{\prime \prime }_1,\ldots ,\xi ^{\prime \prime }_{k"})$
is an oriented frame of
$V'$
(resp.
$V"$
) that lifts to the associated spin bundle, then
$\xi ^{\prime }_1 \oplus 0,\ldots ,\xi ^{\prime }_{k'} \oplus 0, 0 \oplus \xi ^{\prime \prime }_1,\ldots ,0\oplus \xi ^{\prime \prime }_{k"}$
is an oriented frame of
$V' \oplus V"$
that lifts to the associated spin bundle.
In light of Lemma 5.24, the direct sum of spin Riemann-Hilbert pairs is naturally a spin Riemann-Hilbert pair. The following can be deduced from the proof of Proposition 8.4 from [Reference Solomon45].
Lemma 5.25. Let
$(E,F)$
be a spin Riemann-Hilbert pair. If
$(E,F)$
splits as a direct sum of spin Riemann-Hilbert pairs
$(E',F')\oplus (E",F")$
, then the isomorphism

is orientation-preserving.
Remark 5.26. An oriented real line bundle has a canonical spin structure. So, a one-dimensional Riemann-Hilbert pair with an orientation on the real bundle determines a spin Riemann-Hilbert pair. We will use this implicitly below.
The proof of the following lemma is given as a part of the proof of Theorem C.4.1 in [Reference McDuff and Salamon38].
Lemma 5.27. If
$\kappa \ge -1$
, then
$\ker \bar {\partial }_{\kappa }$
can be written explicitly as

and
$\operatorname {\mathrm {Coker}}(\bar {\partial }_{\kappa }) = 0.$
In particular,
$\dim \ker \bar {\partial }_{\kappa } = \kappa +1.$
Below, we often use the fact that if
$(E,F)$
is a Riemann-Hilbert pair with
$\operatorname {\mathrm {Coker}} \bar {\partial }_{(E,F)} = 0,$
then an orientation of
$\det \bar {\partial }_{(E,F)}$
is the same as an orientation of
$\ker \bar {\partial }_{(E,F)}.$
Lemma 5.28. Evaluation at 1 defines an orientation-preserving isomorphism

Proof. This follows from the definition of the orientation on
$\det (\partial _0)$
given in Section 2 of [Reference Solomon45].
The following is Lemma A.2.1 from [Reference Smith44].
Lemma 5.29. Let
$(E,F)$
be a rank 1 Riemann-Hilbert pair of Maslov index
$2$
. Evaluation at
$0$
and
$1$
defines an isomorphism

Moreover, for any choice of orientation on
$F,$
this isomorphism is orientation-preserving if the codomain is oriented by the complex structure on
$E_0$
and the orientation on
$F_1$
.
We choose an orientation on
$F^2$
such that the frame z is positively oriented. By Remark 5.26, this choice of orientation determines a spin structure on
$F^2.$
Lemma 5.30. The orientation on
$\ker \bar {\partial }_2$
is determined by the ordered basis

Proof. By Lemma 5.27, we have

Consider the map

from Lemma 5.29. Orient the fiber
$\underline {\mathbb {C}}_0$
by its complex structure and the fiber
$F^2_1 = \mathbb {R} \subset \underline {\mathbb {C}}_1$
by the basis
$1$
. By Lemma 5.29, the map
$f_2$
is orientation-preserving. Since

is an oriented basis, it follows that the basis

determines the orientation on
$\ker \bar {\partial }_2.$
We choose the orientation on
$F^{1,1}$
such that the frame

is positively oriented. In addition, we choose the spin structure on
$F^{1,1}$
such that this frame can be lifted to the associated double cover of the frame bundle.
Lemma 5.31. Evaluation at zero

defines an isomorphism. If the codomain is oriented by the complex structure on
$\underline {\mathbb {C}}^2_0,$
this isomorphism is orientation-preserving.
Proof. By Lemma 5.27, the map
$f_{1,1}$
defines an isomorphism. We degenerate
$(\underline {\mathbb {C}}^2,F^{1,1})$
to
$(\underline {\mathbb {C}}^2,F^{0,2})$
in order to determine the orientation on
$\ker \bar {\partial }_{1,1}$
. Consider the family of loops

given by

We claim that the family of boundary conditions
$F(t)$
given by
$F(t)_z = A_t(z) \cdot \mathbb {R}^2$
is a degeneration from
$F^{1,1}$
to
$F^{0,2}.$
Indeed,

where the matrix

is an invertible real matrix for every
$z\in \partial D.$
The family of frames given by the columns of
$A_t$
gives rise to a continuous family of orientations and spin structures on
$F(t)$
. The orientation on
$F(1)$
coincides with the orientation on
$F^{1,1}.$
The orientation on
$F(0)$
coincides with the direct sum orientation
$F^{0,2} = F^0 \oplus F^2.$
The spin structure on
$F(1)$
coincides with the spin structure on
$F^{1,1}$
and the spin structure on
$F(0)$
coincides with the direct sum spin structure on
$F^{0,2}$
as in Lemma 5.24. Let

Then
$B(t,z)$
is an invertible holomorphic matrix on D for every
$t\in (0,1]$
, and the matrices satisfy

Hence,
$B(t,z)$
is an isomorphism between
$(\underline {\mathbb {C}}^2, F(1))$
and
$(\underline {\mathbb {C}}^2, F(t))$
for every
$t\in (0,1]$
. Denote by
$\bar {\partial }(t)$
the family of the Cauchy-Riemann operators obtained from
$(\underline {\mathbb {C}}^2, F(t))$
for
$t\in (0,1]$
. Note that
$\ker \bar {\partial }(1)=\ker \bar {\partial }_{1,1}$
. By Lemma 5.27, the vectors

form a basis for
$\ker \bar {\partial }_{1,1}.$
So, the following vectors form a basis for
$\ker \bar {\partial }(t)$
:


Let

For
$i=2,3,4,$
define
$u_{i,t}(z):=tw_{i,t}(z)$
when
$t>0,$
and
$u_{i,0}(z):=\lim _{t\rightarrow 0}tw_{i,t}(z).$
In addition, define
$u_{1,t}(z):=w_{1,t}(z).$
Note that the bases

determine the same orientation on
$\ker \bar {\partial }(t)$
for
$t\in (0,1]$
. We have

By Lemma 5.30, the basis
$\{-u_{3,0},u_{2,0},u_{4,0}\}$
determines the orientation on
$\ker \bar {\partial }_2$
, and by Lemma 5.28, the vector
$u_{1,0}(z)$
determines the orientation on
$\ker \bar {\partial }_0.$
Hence, by Lemma 5.25, the basis

determines the orientation on
$\ker \bar {\partial }(0)$
. Continuity implies that
$\{u_{1,t}, u_{2,t}, u_{3,t}, u_{4,t}\}$
is positively oriented for every
$t\in [0,1],$
and so is
$\{v_{1,t}, v_{2,t}, v_{3,t}, v_{4,t}\}$
for
$t\in (0,1]$
. Hence, the basis

is positively oriented. Therefore,
$f_{1,1}$
is orientation-preserving.
The orientation convention for
$\partial D$
is the counter-clockwise orientation. For a point
$z\in \partial D$
, we identify
$T_z\partial D$
with
$iz\mathbb {R},$
so
$iz$
is a positively oriented basis. The following definition is the orientation convention for
$\mathrm {PSL}(2,\mathbb {R})$
in [Reference Fukaya, Oh, Ohta and Ono14].
Definition 5.32. Let
$z_0, z_1, z_2\in \partial D$
be three distinct points in anticlockwise order. Consider the embedding


The orientation on
$\mathrm {PSL}(2,\mathbb {R})$
is determined such that F is orientation-reversing.
Consider the Lie algebra
$\mathfrak {psl}(2,\mathbb {R}).$
Let

be a basis for this algebra.
Lemma 5.33.
$\{\eta _1,\eta _2,\eta _3\}$
is a positively oriented basis of
$\mathfrak {psl}(2,\mathbb {R}).$
Proof. Let
$\mathbb {H}$
denote the closed upper half plane and let
$\overline {\mathbb {H}}$
denote the compactification by adding a point at infinity. Since there is a biholomorphism
$D\rightarrow \overline {\mathbb {H}}$
, we can identify the range of the map F from Definition 5.32 with
$\partial \overline {\mathbb {H}}\times \partial \overline {\mathbb {H}} \times \partial \overline {\mathbb {H}}$
. Choose

The orientation of

is determined by the basis

We have

Similarly,

Hence,

Since

it follows that F is orientation-reversing. Thus, by Definition 5.32, the basis
$\{\eta _1, \eta _2, \eta _3\}$
is positively oriented.
Lemma 5.34. The action of
$\mathrm {PSL}(2,\mathbb {R})$
on D induces an isomorphism

which is orientation-reversing.
Proof. Identify D with the one-point compactified upper half plane
$\overline {\mathbb {H}}$
and identify the Riemann-Hilbert pair
$(\underline {\mathbb {C}},F^2)$
with the Riemann-Hilbert pair
$(T\overline {\mathbb {H}},T\partial \overline {\mathbb {H}}).$
The latter identification preserves the canonical orientation on the determinant line by Proposition 5.1. The map
$\mathcal {D}$
is given by

Let
$ev:=ev_i\oplus ev_0$
, where
$ev_i$
and
$ev_0$
are the evaluation maps at i and
$0$
, respectively. By Lemma 5.29,
$\mathcal {D}$
is orientation reversing if and only if
$ev\circ \mathcal {D}$
is orientation reversing. Consider the points
$0,i\in \overline {\mathbb {H}}$
. The basis

is positively oriented basis of
$T_i\overline {\mathbb {H}}\oplus T_0\partial \overline {\mathbb {H}}$
. By the calculation of the previous lemma, we have

Since

it follows that
$\mathcal {D}$
is orientation-reversing.
5.5 Riemann-Hilbert pairs and holomorphic disks
Let
$(X,J)$
be an n-dimensional complex manifold, and let
$L\subset X$
be a smooth totally real n-dimensional submanifold. A holomorphic disk
$u:(D,\partial D)\rightarrow (X,L)$
gives rise to a holomorphic vector bundle
$u^{*}TX\rightarrow D$
and a smooth totally real subbundle
$u|_{\partial D}^{*}TL\subset u^{*}TX|_{\partial D}$
. Thus, we obtain a Riemann-Hilbert pair
$(E_u,F_u)$
associated to
$u.$
5.5.1 A useful example
Lemma 5.35. Let
$u:(D,\partial D)\rightarrow (\mathbb {C}P^3,L_\triangle )$
be the holomorphic disk given by
$u(z)=e^{-i\xi _v\log z}\cdot \triangle .$
Let
$\theta _1,\theta _2 \in \mathfrak {su}(2)$
be such that
$\xi _v,\theta _1,\theta _2$
form a basis of
$\mathfrak {su}(2).$
Equip
$L_\triangle $
with the orientation and spin structure arising from the trivialization of
$TL_\triangle $
given by the infinitesimal action of
$\mathfrak {su}(2)$
and the basis
$\xi _v,\theta _1,\theta _2.$
Let

Then, the sections

form a holomorphic frame. Moreover, there is an isomorphism of spin Riemann-Hilbert pairs

given by

Proof. Recalling Example 5.10, we see that the kernel of the infinitesimal action of
$\mathfrak {sl}(2,\mathbb {C}) $
at
$u(0)$
is spanned by
$\xi _v$
. Given this, Smith (Reference Smith44, Appendix A.3) shows that the sections (5.1) form a holomorphic frame for
$E.$
We split
$(E,F)$
by


It follows that the formula (5.2) gives the desired isomorphism
$\Psi .$
Recall the choices of the spin structures and orientations of
$F^2$
and
$F^0$
given in Section 5.4. By Lemma 5.24, the frame

of
$F^2\oplus F^0\oplus F^0$
is positively oriented and can be lifted to the associate double cover of the frame bundle. Therefore, the isomorphism
$\Psi $
preserves orientation and spin structure.
5.5.2 Orientation convention for disk moduli spaces
Let
$(X,\omega )$
be a symplectic manifold with
$\omega $
-tame (integrable) complex structure
$J,$
and let
$L \subset X$
be a Lagrangian submanifold. Let
$\widetilde {\mathcal {M}}(\beta )$
denote the space of parameterized J-holomorphic maps
$u: (D^2,\partial D^2) \to (X,L)$
such that
$u_*([D^2,\partial D^2]) = \beta $
. Each
$u \in \widetilde {\mathcal {M}}(\beta )$
determines a Riemann-Hilbert pair
$(E_u,F_u) = (u^{*}TX,u^{*}TL)$
as explained above. For the following discussion, we assume that the linear Cauchy-Riemann operator
$\bar \partial _{(E_u,F_u)} : \Gamma ((D^2,\partial D^2),(E_u,F_u)) \to \Gamma (D^2,\Omega ^{0,1}(E_u))$
is surjective for every u, so
$\widetilde {\mathcal {M}}(\beta )$
is a smooth manifold and there is a canonical isomorphism

Thus, Proposition 5.1 gives a canonical orientation of
$\widetilde {\mathcal {M}}(\beta ).$
We discuss now the relevant conventions concerning the orientations of the moduli spaces
$\mathcal {M}_{k+1,l}(\beta )$
of unparameterized stable J-holomorphic maps with marked points. The following definition is Convention 8.2.1 from [Reference Fukaya, Oh, Ohta and Ono14].
Definition 5.36. Let G be an oriented Lie group with a smooth, proper, free right action on an oriented manifold M. For
$p \in M,$
let
$\varphi _p : G \to M$
be given by
$g \mapsto p\cdot g.$
Let
$\pi : M \to M/G$
be the quotient map. Split the short exact sequence

to obtain an isomorphism

The quotient orientation on
$M/G$
is determined by the condition that the preceding isomorphism preserves orientation for all
$p \in M$
.
The following definition is based on [Reference Fukaya, Oh, Ohta and Ono14, p. 698].
Definition 5.37. Let
$U \subset \widetilde {\mathcal {M}}(\beta ) \times (\partial D^2)^{k+1} \times (\operatorname {\mathrm {int}} D^2)^l$
denote the open subset where the marked points are pairwise disjoint and the cyclic ordering on the boundary marked points given by the orientation on
$\partial D^2$
induced from the complex orientation of
$D^2$
agrees with the order of the labels. Thus, points of U are tuples
$(u,z,w)$
where

An automorphism of the disk
$\psi \in \mathrm {PSL}_2(\mathbb {R})$
acts on U by

The orientation of
$\mathcal {M}_{k+1,l}(\beta )$
is determined by the quotient orientation on the open subset
$U/PSL_2(\mathbb {R}) \subset \mathcal {M}_{k+1,l}(\beta ).$
5.6 Computation of
$\overline {OGW}_{1, 1}$
Recall from Section 4 that we denote by
$\mathcal {M}_{k+1,l}(\beta )$
the moduli space of holomorphic disks
$u:(D, \partial D)\rightarrow (\mathbb {C}P^3,L_\triangle )$
representing a class
$\beta \in H_2(\mathbb {C}P^3,L_\triangle ;\mathbb {Z})$
with
$k+1$
boundary points and l interior points. We denote elements in
$\mathcal {M}_{k+1,l}(\beta )$
by
$[u;z_0,\ldots ,z_k,w_1,\ldots ,w_l]$
, where
$z_i\in \partial D$
and
$w_i\in \mathrm {int}D$
. Consider the moduli space
$\mathcal {M}_{1,0}(1)$
, and the evaluation map


The following lemma is part of Proposition 4.2 from [Reference Smith43].
Lemma 5.38. The evaluation map
$evb_0:\mathcal {M}_{1,0}(1)\rightarrow L_\triangle $
is a covering map of degree
$\pm 3$
.
Proof of Theorem 5.1.
By Theorem 1.6, we have

where
$\text {PD}([\bar {b}])=pt$
. Hence,
$\overline {OGW}\!_{1,1}$
is determined by the degree of the map
$evb_0$
, and Lemma 5.38 gives
$\overline {OGW}\!_{1,1}=\mp 3,$
where the sign depends on whether
$evb_0$
preserves or reverses orientation.
We show that
$evb_0$
preserves orientation. Choose
$\theta _1,\theta _2$
that together with
$\xi _v$
form a basis of
$\mathfrak {su}(2).$
Equip
$L_\triangle $
with the orientation and spin structure arising from the trivialization of
$TL_\triangle $
given by the infinitesimal action of
$\mathfrak {su}(2)$
and the basis
$\xi _v,\theta _1,\theta _2.$
The orientation of this basis does not affect the calculation below, in accordance with the orientation axiom Proposition 2.2 8.
Let
$u:(D,\partial D)\rightarrow (\mathbb {C}P^3,L_\triangle )$
be a holomorphic disk of Maslov index 2. Let

By Lemma 5.16, we can write
$u(z)=A\cdot e^{-i\xi _v \log z}\cdot \triangle $
, where
$A\in \text {SU}(2)$
. We may assume that A is the identity since acting by
$A^{-1}$
does not change the isomorphism class of
$(E,F)$
.
Let
$z\in \partial D$
and let
$\zeta \in T_z\partial D$
denote the unit vector in the direction of the orientation. By Definition 5.37, the oriented tangent space of
$\mathcal {M}_{1,0}(1)$
at
$[u;z]$
is given by

By Proposition 5.1 and Lemma 5.35, we have an orientation-preserving isomorphism

By Lemma 5.27, we have an orientation-reversing isomorphism

The linearization of the action of
$\mathrm {PSL}_2(\mathbb {R})$
in Definition 5.37 composed with the projection on
$\ker \bar \partial _2$
gives the map
$\mathcal {D}:\mathfrak {psl}(2,\mathbb {R})\rightarrow \ker \bar {\partial }_2$
of Lemma 5.34. So, by Definition 5.36 and Lemmas 5.34 and 5.27, we have an orientation-preserving isomorphism

given by

Abbreviate

Thus, by Lemmas 5.25 and 5.28, the orientation on
$T_{[u:z]}\mathcal {M}_{1,0}(1)$
is given by the basis
$\bar {\theta }_1, \bar {\theta }_2, \bar {\zeta }$
. In order to show that
$evb_0$
preserves orientation, it suffices to show that

preserves orientation when
$z=1$
. The tangent vector
$\zeta \in T_1 \partial D$
is represented by the path
$t \mapsto e^{it}.$
So,

Since

it follows that
$evb_0$
preserves orientation. Therefore,
$\overline {OGW}\!_{1,1}=-3$
.
5.7 Computation of
$\overline {OGW}_{1, 0}(\Gamma _2)$
By Theorem 1.6, we have

where
$\gamma _2$
is a a differential form representing
$\Gamma _2 \in H^4(\mathbb {C}P^3, L_\triangle; \mathbb {R})$
. By Poincaré-Lefschetz duality, we have
$H^4(\mathbb {C}P^3, L_\triangle; \mathbb {R}) \simeq H_2(\mathbb {C}P^3 \backslash L_\triangle ;\mathbb {R})$
. Hence, our strategy for computing
$\overline {OGW}\!_{1,0}(\Gamma _2)$
is to find a complex curve in
$\mathbb {C}P^3\setminus L_\triangle $
representing the Poincaré-Lefschetz dual to
$\Gamma _2,$
and determine how many holomorphic disks intersect it.
Any complex subvariety
$\Upsilon \subset \mathbb {C}P^3 \backslash L_\triangle $
of complex dimension 2 represents
$k\text {PD}([\omega ])$
where
$k \in \mathbb {Z}$
is the degree of
$\Upsilon $
(i.e., the number of intersection points with a generic line). Consider complex subvarieties
$\Upsilon _1, \Upsilon _2$
of degree
$k_1, k_2$
, respectively, in general position. Then
$\Upsilon _1 \cap \Upsilon _2$
represents
$k_1k_2\text {PD}([\omega ^2])$
, and thus,
$\text {PD}(\Gamma _2) = [\Upsilon _1 \cap \Upsilon _2] / (k_1k_2)$
.
We consider the case where
$\Upsilon _1, \Upsilon _2$
are small perturbations of the anticanonical divisor
$Y_\triangle $
. For
$i = 1, 2$
, let
$g_i \in \text {SU}(2)$
be lifts of rotations by arbitrary small angles
$\epsilon _i$
about different axes. Write

Note that
$\Upsilon _i\subset \mathbb {C}P^3\setminus L_\triangle $
since
$\epsilon _i$
are small.
Theorem 5.2.
$\deg \Upsilon _i = 4$
Proof. Write
$a=[a_0:a_1]$
,
$b=[b_0: b_1]\in \mathbb {C}P^1$
. Identifying
$\mathbb {C}P^3$
with the projectivization of the space of homogeneous polynomials of degree
$3$
in two variables, we can write

where
$(g_i)_{kl}$
are the components of the matrix
$g_i.$
Hence, we have an embedding


where
$f_i^j$
are homogeneous polynomials of bidegree
$(2,1),$
such that
$\Upsilon _i=\operatorname {\mathrm {Im}} f_i$
.
Let
$H_1$
and
$H_2$
be hyperplanes in
$\mathbb {C}P^3$
given by

The intersection
$H_1\cap H_2$
is a generic line, so we have

The preimages
$f_i^{-1}(H_1)$
and
$f_i^{-1}(H_2)$
are the vanishing sets of the polynomials

of bidegree
$(2,1)$
. Denote by
$\pi _1, \pi _2 : \mathbb {C}P^1 \times \mathbb {C}P^1 \rightarrow \mathbb {C}P^1$
the projection maps. Let

So,

The polynomials
$p^h_i, p^k_i,$
define sections of the line bundle
$\mathcal {O}(2,1).$
So,

Proof of Theorem 5.2.
Let
$\gamma _2$
be a a differential form representing
$\Gamma _2 \in H^4(\mathbb {C}P^3, L_\triangle; \mathbb {R})$
. We show below that
$evi_1$
is transverse to
$\Upsilon _1 \cap \Upsilon _2.$
Thus, by Theorem 1.6 and Poincaré duality, we have

So, we need to count Maslov 2 disks passing through
$\Upsilon _1 \cap \Upsilon _2$
with sign given by the orientation of the moduli space
$\mathcal {M}_{0,1}(1)$
and the complex orientation of the normal bundle to
$\Upsilon _1 \cap \Upsilon _2.$
We represent a point of
$\Upsilon _i$
by three dots with an arrow between two of them labeled by
$g_i$
as follows.

Hence, a point of
$\Upsilon _1 \cap \Upsilon _2$
is represented by three dots with two arrows between them labeled by
$g_1$
and
$g_2$
. There are 6 types up to continuous deformations:






More explicitly,

Since each of
$\Theta _1,\ldots ,\Theta _4$
is the image of a map
$\mathbb {C}P^1 \to \mathbb {C}P^3$
of degree 3, it follows that
$\deg \Theta _1 = \ldots = \deg \Theta _4 =3.$
We claim that each of
$\Theta _5, \Theta _6$
is a union of two lines. Indeed, both
$g_2g_1$
and
$g_2^{-1}g_1$
have two eigenvectors, so there are two choices for
$a\in \mathbb {C}P^1$
. Each such choice a gives a line
$\{[a, g_1(a), c]|\ c\in \mathbb {C}P^1 \}.$
Hence,
$\deg \Theta _5 = \deg \Theta _6 = 2$
. Therefore,
$\sum _{i=1}^6 \deg \Theta _i=16$
. By Bezout’s theorem and Proposition 5.2, we have

Since the degrees coincide, it follows that each
$\Theta _i$
occurs with multiplicity
$1$
in the intersection
$\Upsilon _1 \cap \Upsilon _2.$
In particular, the intersection is generically transverse.
By Lemma 5.16, all Maslov 2 disks are axial of the form
$u(z)=A\cdot e^{-i\xi _v\log z}\cdot \triangle $
. Let
$p_1, p_2, p_3 \in \mathbb {C}P^1$
such that
$[p_1, p_2, p_3]\in \text {Sym}^3 \mathbb {C}P^1$
is in the image of a Maslov index
$2$
disk. As in Example 5.10, there exist
$i,j\in \{1,2,3\}$
such that
$d(p_i,p_j)\ge e$
, where e is the distance between two vertices in an equilateral triangle on a great circle in
$\mathbb {C}P^1$
. By choosing
$\epsilon _i$
sufficiently small,
$\Theta _1,\ldots , \Theta _4$
can be brought arbitrarily close to the locus
$N_\triangle $
where all 3 points coincide. This rules out a Maslov index 2 disk passing through them. Hence, the number of Maslov 2 disks through
$\Upsilon _1 \cap \Upsilon _2$
is equal to the number of Maslov
$2$
disks through
$\Theta _5$
and
$\Theta _6.$
Since a and
$g_1(a)$
are sufficiently close, it follows by Lemma 5.19 that one such disk passes through each of the lines making up
$\Theta _5, \Theta _6,$
for a total of four disks.
Next, we determine the sign with which each disk passing through
$\Theta _5$
and
$\Theta _6$
contributes to
$\# evi_1^{-1}(\Upsilon _1 \cap \Upsilon _2).$
The treatment of
$\Theta _5$
and
$\Theta _6$
is parallel, so we focus on
$\Theta _6$
. Let
$a \in \mathbb {C}P^1$
be one of the two solutions of
$g_2^{-1}g_1(a) = a$
as in the above formula for
$\Theta _6.$
Denote by
$\Theta _6^a$
the corresponding component of
$\Theta _6.$
Let
$[u;w]\in \mathcal {M}_{0,1}(1)$
with
$u(w) \in \Theta _6^a$
. Let
$(\nu _{\Theta _6})_{u(w)}:= T_{u(w)}\mathbb {C}P^3/ T_{u(w)}\Theta _6$
be the normal space to
$\Theta _6$
at
$u(w)$
equipped with the complex orientation. The differential
${devi_1}_{[u;w]}$
induces a linear map

We show
$\overline {devi_1}_{[u;w]}$
is an orientation-reversing isomorphism.

Figure 4 The triangle
$u(w)$
.
By Lemma 5.16, we can write
$u(z)=A\cdot e^{-i\xi _v\log z}\cdot \triangle $
. By applying an appropriate rotation to
$\mathbb {C}P^1 = S^2$
, which does not affect orientations, we may assume that a and
$g_1(a)$
are positioned in the
$xz$
plane symmetrically about the south pole as in Figure 4. It follows from Example 5.10 that we may take A to be the identify matrix. We begin by finding an oriented basis for
$T_{[u;w]}\mathcal {M}_{0,1}(1)$
. Let

We choose
$\theta _1,\theta _2 \in \mathfrak {su}(2)$
to be lifts of infinitesimal right-handed rotations around the x-axis and the y-axis, respectively. Thus, the trivialization of
$TL_\triangle $
given by the infinitesimal action of
$\mathfrak {su}(2)$
and the basis
$\xi _v,\theta _1,\theta _2,$
determines the orientation and the spin structure of
$L_\triangle $
given in Section 3.2. Let
$w\in \operatorname {\mathrm {int}} D$
. The tangent space of
$\mathcal {M}_{0,1}(1)$
at
$[u;w]$
is given by

Consider the canonical identification
$\mathbb {C}\simeq T_w D$
. We denote by
$\hat {x}$
and
$\hat {y}$
the vectors in
$T_w D$
that correspond to
$1$
and i in
$\mathbb {C}$
, respectively. Hence, the basis
$\{\hat {x}, \hat {y}\}$
determines the orientation on
$T_w D$
that induced by its complex structure. By Proposition 5.1 and Lemmas 5.25, 5.28, 5.34 and 5.35, we have an orientation-reversing isomorphism

given by

Abbreviate

Thus,
$\{-\bar \theta _1, \bar \theta _2,\bar {x}, \bar {y}\}$
is an oriented basis for
$T_{[u;w]}\mathcal {M}_{0,1}(1)$
.
Next, we compute
$devi_1$
on
$\bar x, \bar y.$
Indeed,

and

It follows that

To determine whether
$\overline {devi_1}_{[u;w]}$
preserves orientation, it is helpful to work in a holomorphic coordinate chart on
$\text {Sym}^3(\mathbb {C}P^1)$
. Let

be given by

Recall the stereographic projection p from Example 5.10. Observe that

Since
$a \neq g_1(a)$
, we may choose an open
$U \subset \mathbb {C}^3$
containing
$(p(a),p(g_1(a)),0)$
such that
$\hat \psi = \psi |_U$
is a biholomorphism onto its image. Then

Observe that the normal bundle to
$\hat \Theta _6^a$
is canonically identified with
$\mathbb {C}^2 \times \{0\} \subset \mathbb {C}^3.$
Let
$\pi : \mathbb {C}^3 \to \mathbb {C}^2$
be the projection onto the first two factors, and let

Since
$\hat \psi $
is biholomorphic and thus orientation preserving, it suffices to determine whether the map

preserves orientation.
We proceed to compute
$d\widehat {evi_1}$
on the oriented basis
$\{-\bar \theta _1,\bar \theta _2,\bar x,\bar y\}$
of
$T_{[u;w]}\mathcal {M}_{0,1}(1)$
. By Example 5.10, we have

Since a lies in the
$xz$
plane, it follows that
$p(a)\in \mathbb {R}$
, and consequently,
$w \in \mathbb {R}_{>0}.$
Recall the map
$\varphi ^t_{\xi _v}$
from Example 5.10. For
$z \in \mathbb {C},$
we have

So, by equations (5.4) and (5.5), we obtain

However, we can see by stereographically projecting Figure 4 or by direct calculation that

Since

it follows that the basis

is not complex oriented. Thus,
$d\widehat {evi_1}$
and consequently also
$\overline {devi_1}$
reverse orientation.
Therefore, each of the 4 Maslov 2 disks passing through
$\Upsilon _1\cap \Upsilon _2$
contributes to the signed count
$\# evi_1^{-1}(\Upsilon _1 \cap \Upsilon _2)$
with sign
$-1$
. By Proposition 5.2 and equation (5.3), we obtain

5.8 Computation of
$\overline {OGW}_{2, 0}(\Gamma _3)$
Proof of Theorem 5.3.
By Theorem 1.6, we have

where
$\gamma _3$
is a representative of
$\Gamma _3$
. Since
$\Gamma _3$
is Poincaré dual to a point, it follows that

where
$p \in \mathbb {C} P^3$
is a regular value of
$evi_1$
.
Let
$p \in N_\triangle $
. First, we show that p is a regular value of
$evi_1$
and compute
$\#evi_1^{-1}(p)$
up to sign. We claim that any Maslov 4 disk u that passes through p is an axial disk of type
$\xi _f.$
Indeed, by Lemma 5.4,
$Y_{\triangle }$
is an anticanonical divisor, so by Lemma 5.5, the Maslov index of u is given by twice the algebraic intersection number
$[u] \cdot [Y_{\triangle }]$
. Hence,
$[u] \cdot [Y_{\triangle }] = 2$
. Thus, Lemma 5.6 implies that u intersects
$Y_{\triangle }$
only at p. Hence, by Lemma 5.15, a pole germ of u is not of type
$\xi _v$
. So, by Lemma 5.17, u is an axial disk of type
$\xi _f$
as claimed. Thus, Lemma 5.18 guarantees that p is a regular value of
$evi_1$
. Since by Lemma 5.20 there exists a unique axial disk of type
$\xi _f$
of Maslov index
$4$
that passes through
$p,$
it follows that
$\# evi_1^{-1}(p)=\pm 1$
.
It remains to show that in fact
$\# evi_1^{-1}(p)= -1$
. Let
$r:=(0,-1,0),$
and choose

Let
$u:(D,\partial D)\rightarrow (\mathbb {C}P^3,L_\triangle )$
be the J-holomorphic disk of Maslov index 4 that intersects
$N_\triangle $
at p. As shown above, u is an axial disk of type
$\xi _f$
, so by Example 5.11, we have
$u(z)= e^{-i\xi _f \log z}\cdot \triangle $
. Let

We construct frames for E and F as follows. Let
$\alpha \in \mathfrak {su}(2)$
be an infinitesimal right-handed rotation about the
$-x$
axis, and let
$\beta \in \mathfrak {su}(2)$
be an infinitesimal right-handed rotation about the z axis. Recall that
$\xi _f$
is an infinitesimal right-handed rotation about the y axis, so
$\xi _f, \alpha , \beta $
is a basis of
$\mathfrak {su}(2)$
corresponding to infinitesimal right-handed rotations about a right-handed set of orthogonal axes as in the definition of the orientation
$\mathfrak {o}_\triangle $
and spin structure
$\mathfrak {s}_\triangle $
on
$L_\triangle $
given in Section 3.2. Since
$u(0)=p,$
and we have taken the complex structure on
$\mathbb {C}P^1$
given by left-handed rotation around the outward normal by an angle of
$\pi /2,$
it follows that
$i\beta \cdot u(0)= -\alpha \cdot u(0).$
See Figure 5. Thus, the kernel of the infinitesimal action of
$\mathfrak {sl}(2,\mathbb {C})$
at
$u(0)$
is spanned by
$\xi _f$
and by
$\alpha +i\beta $
. So, it follows from the argument of Smith [Reference Smith44, Appendix A.3] that the following sections give a holomorphic frame of
$E:$

Taking
$\mathbb {C}$
-linear combinations, we see that

is also a holomorphic frame of
$E.$
Since
$\xi _f, \alpha , \beta $
is a basis of
$\mathfrak {su}(2)$
, the sections

give a frame of for
$F.$
This frame is
$\mathfrak {o}_\triangle $
oriented and can be lifted to the double cover of the frame bundle of F associated to
$\mathfrak {s}_\triangle $
because of how we have chosen
$\alpha ,\beta $
.

Figure 5 The rotations
$\alpha $
and
$\beta $
at p.
Let
$\Psi : E \to \underline {\mathbb {C}}^3$
be the isomorphism given by

Since

we see that
$\Psi :(E,F) \to (\underline {\mathbb {C}}^3,F^{2} \oplus F^{1,1}).$
Recall the choices of spin structures and orientations on
$F^2$
and
$F^{1,1}$
given in Section 5.4. By Lemma 5.24, the frame

of
$F^2\oplus F^{1,1}$
is positively oriented and can be lifted to the spin double cover of the frame bundle. Hence, the isomorphism
$\Psi $
preserves orientation and spin structure. Thus, by Lemma 5.25, we obtain an orientation preserving isomorphism

where
$w\in \mathrm {int}D$
. By Lemma 5.34, this gives an orientation-reversing isomorphism

It suffices to show that

reverses orientation when
$w=0$
. Denote by
$f_D:T_0D\rightarrow \mathbb {C}$
the canonical isomorphism, and recall that
$f_{1,1}: \ker \bar {\partial }_{1,1}\rightarrow \underline {\mathbb {C}}^2_0$
is the evaluation map at zero. In the following commutative diagram of isomorphisms, the left vertical arrow reverses orientation, and the right vertical arrow preserves orientation.

So, it suffices to show that
$f_{1,1}\oplus f_D$
preserves orientation. Since
$f_D:T_0D\rightarrow \mathbb {C}$
preserves orientation, and by Lemma 5.31 the map
$f_{1,1}$
preserves orientation, it follows that
$f_{1,1}\oplus f_D$
preserves orientation as desired.
6 Recursions
Proof of Theorem 1.1.
Recall
$\Delta _i=[\omega ^i]\in H^{*}(X;\mathbb {R}),$
for
$i=0,\ldots ,3.$
So, by the definition of
$g^{ij}$
given in Section 2.6, it follows that
$g^{ij}=\delta _{i,3-j}.$
In order to derive recursion 1, let
$I = \{j_2,\ldots ,j_l\}.$
Apply
$\partial _s^{k-1} \partial _{t_I}$
=
$\partial _s^{k-1} \partial _{t_{j_2}} \ldots \partial _{t_{j_l}}$
to equation (2.4) with
$v={j_1-1}$
,
$w=1$
. We consider the coefficients of
$T^{\beta }$
and evaluate at
$s=t_j=0$
. Using the closed zero axiom Proposition 2.16, we single out instances of
$\overline {OGW}\!_{\beta , k}(\Gamma _{j_1},\Gamma _I)$
and obtain

Substituting the expressions in (2.4) gives the required recursion.
In order to derive recursion 2, let
$I = \{j_1,\ldots ,j_l\}.$
Apply
$\partial _s^{k-2} \partial _{t_I}$
=
$\partial _s^{k-2} \partial _{j_1} \ldots \partial _{j_l}$
to equation (2.4) with
$v=2$
,
$w=2$
. We consider the coefficients of
$T^{\beta +2}$
and evaluate at
$s=t_j=0$
. We obtain

Substituting the expressions in (2.4) gives the required recursion.
Recursion 3 is derived in the same way as recursion (a) in Theorem 10 from [Reference Solomon and Tukachinsky48].
Lemma 6.1.
$\overline {OGW}\!_{2,0}(\Gamma _2, \Gamma _2)=\frac {35}{64}.$
Proof. In Theorems 5.1–5.3, we computed

By recursion 3 of Theorem 1.1, we get

By the open divisor axiom Proposition 2.26, we get

In order to prove Corollary 1.3 we will need the following lemma.
Lemma 6.2. Assume
$|\Gamma _{i_j}|\ge 4.$
If
$\overline {OGW}\!_{1,k}(\Gamma _{i_1},\ldots ,\Gamma _{i_l})\ne 0$
. Then
$(k,l)=(1,0)$
or
$(k,l)=(0,1)$
and
$|\Gamma _{i_1}|=4$
.
Proof. By the open degree axiom Proposition 2.23 and Lemma 3.3,

Both summands are positive when
$k>1.$
Hence, equality can hold only if
$(k,l)=(1,0),$
or
$(k,l)=(0,1)$
and
$|\Gamma _{i_1}|=4$
.
Proof of Corollary 1.3.
By Theorem 2.3, invariants with interior constraints in
$\Gamma _{\diamond }$
are computable in terms of invariants with interior constraints of the form
$\Gamma _j=[\omega ^j]$
. Furthermore, by the open unit and divisor axioms Proposition 2.24,6, we may assume that
$|\Gamma _j| \geqslant 4$
. Finally, assume for convenience that interior constraints are written in ascending degree order. It follows from the open degree axiom Proposition 2.23 that for any
$\beta ,$
there are only finitely many values of
$k,l,$
for which there may be nonzero invariants with constraints of the above type. Hence, we give a process for computing
$\overline {OGW}\!_{\beta , k}(\Gamma _{i_1}, \ldots , \Gamma _{i_l})$
which is inductive on
$(\beta , k, l, i_1)$
with respect to the lexicographical order on
$\mathbb {Z}^{\oplus 4}_{\geqslant 0}$
.
Consider a triple
$(\beta , k, l)$
with
$ k + l < 2 $
. If
$\beta =0$
, by the open zero axiom Proposition 2.25, all invariants vanish. For
$\beta =1,2$
all possible values have been computed explicitly in Theorem 1.2. Indeed, if
$\beta =1$
, this follows from Lemma 6.2. If
$\beta =2,$
we have

Equality can only hold if
$k=0$
,
$l=1$
and
$|\Gamma _{i_1}|=6$
. Similarly, the open degree axiom implies that for
$\beta \geqslant 3$
, the only invariants that do not vanish have
$k+l \geqslant 2$
.
Consider a triple
$(\beta , k, l)$
with
$ k + l \geqslant 2 $
. If
$l \geqslant 2$
, by Theorem 1.1 3 and the open divisor and zero axiom, we can express the invariant
$\overline {OGW}\!_{\beta , k}(\Gamma _{i_1}, \ldots , \Gamma _{i_l})$
as a combination of invariants
$\overline {OGW}\!_{\beta ',k'}(\Gamma _{j_1},\ldots ,\Gamma _{j_{l'}})$
that either have
$\beta ' < \beta ,$
or
$\beta ' = \beta , k' = k$
and
$l'< l,$
or
$\beta ' = \beta , k' = k, l'=l$
and
$ j_1 < i_1$
. Thus, the invariant is reduced to invariants with data of smaller lexicographical order, known by induction.
Note that by Lemma 6.1, we have
$\overline {OGW}\!_{2,0}(\Gamma _2,\Gamma _2)\ne 0$
. So, if
$k \geqslant 2$
, by Theorem 1.1 2, we can express the required invariant in terms of invariants that are either of smaller degree or have equal degree and fewer boundary marked points. Indeed, in formula 2, the closed zero axiom Proposition 2.13 and the open zero axiom imply that all the products involving invariants with degree
$ \beta +2$
vanish. By Lemma 3.4, the map
$\varpi $
is given by multiplication by
$4.$
This and Lemma 6.2 imply that products involving invariants with degree
$\beta +1 $
do not occur.
If
$k \geqslant 1$
and
$ l\geqslant $
1, by Theorem 1.1 1, the open zero axiom and the closed zero axiom, we can express the required invariant in terms of invariants that are of smaller degree.
7 Small relative quantum cohomology
In this section, we compute the small relative quantum cohomology of
$(\mathbb {C}P^3,L_\triangle ).$
Lemma 7.1.
$\overline {OGW}\!_{2,2}=-\frac {5}{4}.$
Proof. By recursion 1 of Theorem 1.1, we get

In Theorem 5.1, we computed
$\overline {OGW}\!_{1,1}=-3.$
Thus, by the open divisor axiom Proposition 2.26, we get
$\overline {OGW}\!_{2,1}(\Gamma _2)=-\frac {9}{16}.$
By recursion 2 of Theorem 1.1, we get

Lemma 6.1 gives
$\overline {OGW}\!_{2,0}(\Gamma _2,\Gamma _2)=\frac {35}{64},$
and the number of lines through a point and two lines is
${GW}_1(\Delta _2,\Delta _2, \Delta _3)=1.$
So, by the open unit axiom Proposition 2.24, we get

Recall the definition of
$\Gamma _\diamond $
from Section 2.7. By Lemma 3.4, we have
$H_2(\mathbb {C}P^3, L_\triangle ;\mathbb {Z})=\mathbb {Z},$
and the map

is given by
$m\mapsto 4m.$
Hence, since
$g^{ij}=\delta _{i,3-j}$
, it follows that the relative small quantum product (2.6) for
$(X,L)=(\mathbb {C}P^3, L_\triangle )$
is given by

Proof of Theorem 1.5.
We claim that
$QH^{*}(\mathbb {C}P^3, L_\triangle )$
is generated as a ring by
$1,T, \Gamma _1, \Gamma _\diamond .$
Indeed, by the open degree axiom Proposition 2.2 3 and the closed degree axiom Proposition 2.13, we obtain


So, by the closed zero axiom Proposition 2.1 6, the open divisor axiom Proposition 2.2 6 and Theorem 5.2, we get

and

Consider the ring
$\mathbb {R}[x,y][[q^{1/4}]]/I$
with

Let

be the ring homomorphism defined by

Hence,

In order to show the map
$\phi $
is well defined, it suffices to show that the generators of the ideal I are sent to
$0.$
Using the closed degree axiom, the wall crossing formula Theorem 2.3 and the open degree axiom, we obtain

The number of lines through two points and a plane is
${GW}_1(\Delta _1,\Delta _3,\Delta _3)=1$
. Hence, by Theorem 5.3, Theorem 5.1 and the open divisor axiom, we get

So,

By the wall crossing formula Theorem 2.3 and the open degree axiom, we get

Hence, by Lemma 7.1, we get
$\phi (y^2+\frac {5}{4}q^{1/2}y)=0.$
Using the wall crossing formula Theorem 2.3, the open degree and divisor axioms, and Theorem 5.1, we get

So,
$\phi (xy-\frac {3}{4}q^{1/4}y)=0.$
Thus, the map
$\phi $
is well defined. It is surjective because its image generates
$QH^{*}(\mathbb {C}P^3, L_\triangle )$
by equation (7.1).
Abbreviate

We think of M and N as modules over the local ring
$\mathbb {R}[[q^{1/4}]].$
Consider the induced map

It follows from the definition of the ideal I that
$M/mM$
is a real vector space with basis
$1,y,x,x^2,x^3$
. Since
$\bar {\phi }$
is surjective and

it follows that
$\bar {\phi }$
is injective. Since
$\ker \phi /m\ker \phi \subset \ker \bar {\phi }=0,$
it follows that
$m\ker \phi =\ker \phi .$
Since
$\mathbb {R}[[q^{1/4}]]$
is a local ring with maximal ideal m, Nakayama’s lemma gives
$\ker \phi =0.$
Therefore,
$\phi $
is an isomorphism.
8 Dependence of invariants on left inverse
In this section, we review in greater detail the dependence of the invariants
$\overline {OGW}\!_{\beta ,k}$
on the map
$P_{\mathbb {R}}.$
We quantify this dependence in Proposition 8.1, from which we obtain Proposition 1.1 as a special case. Finally, we prove Corollary 1.4.
In the present section, we continue in the setting of Section 2.1. In particular,
$L\subset X$
is a connected spin Lagrangian submanifold with
$H^{*}(L;\mathbb {R}) \simeq H^{*}(S^n;\mathbb {R})$
and
$[L] = 0 \in H_n(X;\mathbb {R}).$
Recall the definition of the map
$P_{\mathbb {R}}: \widehat {H}^{n+1}(X,L;\mathbb {R}) \to H^n(L;\mathbb {R})$
from Section 2.1. In this section, we extend
$P_{\mathbb {R}}$
to a map
$P_{\mathbb {R}}: \widehat {H}^{*}(X,L;\mathbb {R}) \to H^n(L;\mathbb {R})$
by setting it to zero outside
$\widehat {H}^{n+1}(X,L;\mathbb {R})$
. We use Poincaré duality to identify
$H^n(L;\mathbb {R})\simeq \mathbb {R}.$
Let
$\rho : \widehat {H}^{*}(X,L;\mathbb {R})\rightarrow H^{*}(X;\mathbb {R})$
denote the natural map. Consider
$P_{\mathbb {R}},P_{\mathbb {R}}^{\prime }$
with associated invariants
$\overline {OGW}\!_{\beta ,k}$
and
$\overline {OGW}\!\,'\!\!_{\beta ,k}$
, respectively. The long exact sequence of the pair
$(X,L)$
implies there exists a unique map

such that
$\mathfrak {p}_{\mathbb {R}}\circ \rho =P_{\mathbb {R}}-P_{\mathbb {R}}^{\prime }.$
Let
$\Delta _0,\ldots ,\Delta _N\in H^{*}(X;\mathbb {R})$
be a basis. Recall that

and
$g^{ij}$
denotes the inverse matrix.
Theorem 8.1. Let
$A_1,..,A_l\in \widehat {H}^{*}(X,L;\mathbb {R})$
.

Remark 8.1. Proposition 8.1 continues to hold in the more general setting of [Reference Solomon and Tukachinsky48] without change. We have formulated it only for the case of L spin and
$H^{*}(L;\mathbb {R}) \simeq H^{*}(S^n;\mathbb {R})$
to streamline the exposition since the Chiang Lagrangian satisfies these assumptions.
In order to prove Proposition 8.1, we will need a number of results from [Reference Solomon and Tukachinsky48], which we now summarize. The proof appears toward the end of this section. Recall the definition of the rings
$Q_W$
and
$R_W$
from Section 2.4. Define

by

This is a map of complexes when
$R_W$
is equipped with the trivial differential. The cone
$C(\mathfrak {i})$
is the complex with underlying graded
$Q_W$
module
$A^{*}(X;Q_W)\oplus R_W[-n-1]$
and differential

Note that if
$[L]=0\in H_n(X;\mathbb {R}),$
then
$\operatorname {\mathrm {Coker}}\mathfrak {i}\simeq R_W[-n-1]$
. Thus, we consider the following commutative diagram with exact rows and columns, which is taken from Section 4.4 in [Reference Solomon and Tukachinsky48] except that
$\operatorname {\mathrm {Coker}}\mathfrak {i}$
is replaced with
$R_W[-n-1].$

Here,
$\bar a : Q_W \to R_W$
is the inclusion and
$\bar q : R_W \to R_W/Q_W$
is the quotient map. Let

be a left inverse to the map
$\bar {x}$
from the diagram (8.1) satisfying the following two conditions. The first condition is that

This condition and the exactness of the diagram (8.1) imply that there exists a unique
$P_Q:\widehat {H}^{*}(X,L;Q_W)\rightarrow Q_W$
such that the following diagram commutes:

The second condition is that there exists
$P_{\mathbb {R}} : \widehat {H}^{*}(X,L;\mathbb {R}) \rightarrow \mathbb {R}$
, such that

The following is Lemma 4.10 from [Reference Solomon and Tukachinsky48].
Lemma 8.2.
$P_Q\circ \bar {y}_Q=\operatorname {\mathrm {Id}}$
.
The following is Lemma 4.11 from [Reference Solomon and Tukachinsky48].
Lemma 8.3. Let
$l:\widehat {H}^{*}(X,L;\mathbb {R})\rightarrow \mathbb {R}$
satisfy
$l\circ y=\operatorname {\mathrm {Id}}.$
There exists a unique choice of
$P:H^{*}(C(\mathfrak {i}))\rightarrow R_W$
satisfying conditions (8.2) and (8.4) such that
$l=P_{\mathbb {R}}$
. Moreover,
$\ker P= a(\ker P_Q)$
.
Let
$\Psi \in C(\mathfrak {i})$
be the relative potential defined in Section 1.3.1 in [Reference Solomon and Tukachinsky48]. The definition of
$\overline {\Omega }$
given in Section 1.3.3 in [Reference Solomon and Tukachinsky48] is

This together with (2.2) and Lemma 8.3 makes precise the dependence of the invariants
$\overline {OGW}\!_{\beta ,k}$
on
$P_{\mathbb {R}}.$
To quantify this dependence, we recall another lemma.
Denote by
$\rho ^{*}: Q_U\rightarrow Q_W$
the inclusion map. Recall the map
$\pi $
from diagram (8.1). The following is Lemma 5.9 from [Reference Solomon and Tukachinsky48]
Lemma 8.4.
$\pi (\Psi )=\rho ^{*}(\nabla \Phi ).$
Proof of Proposition 8.1.
Apply Lemma 8.3 to obtain maps
$P,P,'$
corresponding to
$P_{\mathbb {R}},P_{\mathbb {R}}^{\prime },$
that are left inverses to
$\bar {x}$
and satisfy conditions (8.2) and (8.4). Denote by
$P_Q, P_Q^{\prime }$
the corresponding maps from diagram (8.3). Since

the maps
$P-P'$
,
$P_Q-P_Q^{\prime },$
factor through
$H^{*}(X;Q_W)$
. Consequently, there exist unique maps
$\mathfrak {p},\mathfrak {p}_Q,$
such that the following diagram commutes.

Indeed, since

then
$(P_Q-P_Q^{\prime })(\ker \rho _Q)=0.$
By diagram (8.1),
$\rho _Q$
is surjective. Thus, for every
$\tilde {\eta }\in \rho _Q^{-1}(\eta )$
, we define
$\mathfrak {p}_Q(\eta )= (P_Q-P_Q^{\prime })(\tilde {\eta })$
. Hence,
$\mathfrak {p}_Q$
is determined by
$P_Q$
and
$P_Q^{\prime }.$
A similar argument applies for
$\mathfrak {p}.$
Since

it follows by the uniqueness of
$\mathfrak {p}_Q$
that

Let
$\overline {OGW}\!$
,
$ \overline {OGW}\!'$
be the Gromov-Witten invariants correspond to P and
$P'$
, respectively, and let
$\overline {\Omega }$
,
$\overline {\Omega }'$
be the corresponding superpotentials.
From multilinearity of
$\overline {OGW}\!_{\beta ,k}$
, it suffices to prove this proposition for
$A_i=\Gamma _i.$
By (8.5), diagram (8.6), and Lemma 8.4, we get

Recall that
$[T^\beta ]$
denotes the coefficient of
$T^\beta .$
So,

By (2.2), we have

Since
$\mathfrak {p}_Q=\mathfrak {p}_{\mathbb {R}}\otimes \operatorname {\mathrm {Id}}_Q$
and
$\nabla \Phi =g^{ij}\Delta _j\partial _i\Phi ,$
it follows that

which completes the proof.
Proof of Proposition 1.1.
Lemma 3.4 asserts that
$\varpi $
is given by multiplication by
$4$
. Recall the choice of the basis
$\Gamma _i \in \widehat H^{*}(\mathbb {C}P^3,L_\triangle ;\mathbb {R})$
from Section 1. Since
$P_{\mathbb {R}}(\Gamma _i)\ne 0$
only if
$i=2,$
it follows that
$\mathfrak {p}_{\mathbb {R}}(\Delta _i)\ne 0$
only if
$i=2.$
Hence, since
$g^{ij}=\delta _{i,3-j}$
, it follows by Proposition 8.1, Lemma 3.5 and the closed divisor axiom Proposition 2.15 that

Proof of Corollary 1.4.
Assume there exists a map
$P^{\prime }_{\mathbb {R}}$
such that
$\overline {OGW}\!_{\beta ,k}'$
vanishes where
$k=0$
and
$\beta \in \operatorname {\mathrm {Im}}\varpi .$
Hence, by Proposition 1.1, we get

where
$\beta \in \operatorname {\mathrm {Im}}\varpi .$
By Table 2, we have

Since the number of lines through two points is
${GW}_{1}(\Delta _3,\Delta _3)=1,$
we get

However, the number of lines through a point and two lines is
${GW}_{1}(\Delta _2,\Delta _2,\Delta _3)=1,$
so

which is a contradiction.
Acknowledgements
The authors would like to thank J. Evans, O. Kedar, C.-C. M. Liu, J. Smith and S. Tukachinsky for helpful conversations. The authors were partially funded by ISF grants 569/18 and 1127/22.
Competing interests
The authors have no competing interest to declare.