
Acoustic neuroma is a common benign tumor that is frequently associated with postoperative complications such as facial nerve dysfunction, which greatly affects the physical and mental health of patients. In this paper, clinical data of patients with acoustic neuroma treated with microsurgery by the same operator at Xiangya Hospital of Central South University from June 2018 to March 2020 are used as the study object. Machine learning and SMOTE-ENN techniques are used to accurately predict postoperative facial nerve function recovery, thus filling a gap in auxiliary diagnosis within the field of facial nerve treatment in acoustic neuroma. First, raw clinical data are processed and dependent variables are identified based on clinical context and data characteristics. Secondly, data balancing is corrected using the SMOTE-ENN technique. Finally, XGBoost is selected to construct a prediction model for patients' postoperative recovery, and is also compared with a total of four machine learning models, LR, SVM, CART, and RF. We find that XGBoost can most accurately predict the postoperative facial nerve function recovery, with a prediction accuracy of 90.0% and an AUC value of 0.90. CART, RF, and XGBoost can further select the more important preoperative indicators and provide therapeutic assistance to physicians, thereby improving the patient's postoperative recovery. The results show that machine learning and SMOTE-ENN techniques can handle complex clinical data and achieve accurate predictions.
Citation: Jianing Wang. Prediction of postoperative recovery in patients with acoustic neuroma using machine learning and SMOTE-ENN techniques[J]. Mathematical Biosciences and Engineering, 2022, 19(10): 10407-10423. doi: 10.3934/mbe.2022487
[1] | Yunhe Sheng . Categorification of VB-Lie algebroids and VB-Courant algebroids. Journal of Geometric Mechanics, 2023, 15(1): 27-58. doi: 10.3934/jgm.2023002 |
[2] | Valentin Duruisseaux, Melvin Leok . Time-adaptive Lagrangian variational integrators for accelerated optimization. Journal of Geometric Mechanics, 2023, 15(1): 224-255. doi: 10.3934/jgm.2023010 |
[3] | Alexander Pigazzini, Cenap Özel, Saeid Jafari, Richard Pincak, Andrew DeBenedictis . A family of special case sequential warped-product manifolds. Journal of Geometric Mechanics, 2023, 15(1): 116-127. doi: 10.3934/jgm.2023006 |
[4] | Jacob R. Goodman . Local minimizers for variational obstacle avoidance on Riemannian manifolds. Journal of Geometric Mechanics, 2023, 15(1): 59-72. doi: 10.3934/jgm.2023003 |
[5] | Robert I McLachlan, Christian Offen . Backward error analysis for conjugate symplectic methods. Journal of Geometric Mechanics, 2023, 15(1): 98-115. doi: 10.3934/jgm.2023005 |
[6] | Markus Schlarb . A multi-parameter family of metrics on stiefel manifolds and applications. Journal of Geometric Mechanics, 2023, 15(1): 147-187. doi: 10.3934/jgm.2023008 |
[7] | Marcin Zając . The dressing field method in gauge theories - geometric approach. Journal of Geometric Mechanics, 2023, 15(1): 128-146. doi: 10.3934/jgm.2023007 |
[8] | Álvaro Rodríguez Abella, Melvin Leok . Discrete Dirac reduction of implicit Lagrangian systems with abelian symmetry groups. Journal of Geometric Mechanics, 2023, 15(1): 319-356. doi: 10.3934/jgm.2023013 |
[9] | Xavier Rivas, Daniel Torres . Lagrangian–Hamiltonian formalism for cocontact systems. Journal of Geometric Mechanics, 2023, 15(1): 1-26. doi: 10.3934/jgm.2023001 |
[10] | Jordi Gaset, Arnau Mas . A variational derivation of the field equations of an action-dependent Einstein-Hilbert Lagrangian. Journal of Geometric Mechanics, 2023, 15(1): 357-374. doi: 10.3934/jgm.2023014 |
Acoustic neuroma is a common benign tumor that is frequently associated with postoperative complications such as facial nerve dysfunction, which greatly affects the physical and mental health of patients. In this paper, clinical data of patients with acoustic neuroma treated with microsurgery by the same operator at Xiangya Hospital of Central South University from June 2018 to March 2020 are used as the study object. Machine learning and SMOTE-ENN techniques are used to accurately predict postoperative facial nerve function recovery, thus filling a gap in auxiliary diagnosis within the field of facial nerve treatment in acoustic neuroma. First, raw clinical data are processed and dependent variables are identified based on clinical context and data characteristics. Secondly, data balancing is corrected using the SMOTE-ENN technique. Finally, XGBoost is selected to construct a prediction model for patients' postoperative recovery, and is also compared with a total of four machine learning models, LR, SVM, CART, and RF. We find that XGBoost can most accurately predict the postoperative facial nerve function recovery, with a prediction accuracy of 90.0% and an AUC value of 0.90. CART, RF, and XGBoost can further select the more important preoperative indicators and provide therapeutic assistance to physicians, thereby improving the patient's postoperative recovery. The results show that machine learning and SMOTE-ENN techniques can handle complex clinical data and achieve accurate predictions.
We give a biased introduction to generalised Kähler structures based on physics applications. Generalised Kähler or Calabi-Yau geometry [10,12] arise from the study of supersymmetric sigma models or super-string compactifications. For example, by placing a super-string theory on R1,3×X where X is a Calabi-Yau 3-fold, one can obtain a gauge theory on R1,3 that preserves N=2 supersymmetry. For a sigma model based on mappings from a Riemann surface Σ to a manifold M, Zumino showed that one can obtain N=2 supersymmetry if M is Kähler [26]. But string theory has an important duality called the T-duality (torus duality) that relates the IIA to the IIB super-string and is the local model for mirror symmetry [22]. However the two integral ingredients, the Kähler geometry and T-duality are not compatible. The problem is that Kähler geometry is characterised by the covariant constancy of the complex structure under the Levi-Civita connection. But the T-duality will generate a Neveu–Schwarz 3-form H. This 3-form modifies the Levi-Civita connection, making it torsionful [14,20]. So a more appropriate notion of generalised Kähler structure is formulated as the constancy of complex structure under a torsionful connection ∇H (which we will recall in the next section). As for the sigma models, it turns out that Zumino's result can be generalised [15] to accommodate a bi-hermitian geometry with torsion connection.
The local geometry of generalised Kähler structure (GKS) has been well understood and used to construct supersymmetric sigma models mentioned above [18]. The generalised Kähler potential (GKP) was used to write the action of such models, but the drawback was that the formulation only works at regular points and a more global understanding of the GKP was lacking. A more recent work [3] partially answered this question for GKS of the symplectic type, where the GKP was formulated in terms of Morita equivalence of certain holomorphic Poisson structures.
In [13] Hitchin gave a construction of GKS on Del Pezzo surfaces using a flow that deforms the original Kähler structure on these surfaces. In general, it can be hard to solve a flow equation due to its non-linear nature. In this work we restrict ourselves to the toric Del Pezzo surfaces, where thanks to the toric symmetry, the flow equations greatly simplify. We show that on CP2 the flow is solved by the Weierstrass elliptic functions. To present the solution we use the coordinate system c3,s to parametrise the triangle which is the moment map polygon for CP2, see Figure 1. If y1,2 are the Euclidean coordinates of R2, the triangle is the region 1≥y1,2≥0, y1+y2≤1, we have
yk(c3,s)=c3112−℘(ω2+ω1π(s+2kπ3)) , | (1) |
where ℘ is the Weierstrass elliptic function satisfying
˙℘2=4℘3−g2℘−g3,g2=112−2c3,g3=c36−1216−c23 |
and with half-periods ω1,2 depending on c3. With the c3,s coordinates, the flow is simply a shift
ϕΔt:s→s+πΔtω1 |
with Δt the flow time. Our main result is that (the details are in Proposition 6.6, 6.4 and 6.9)
Theorem 1.1. On CP2, Hitchin's construction gives a Generalised Kähler structure with two toric invariant complex structures. The first one I− is the standard one while I+ is deformed from I− by a flow ϕΔt: I+=(ϕΔt∗)−1I−ϕΔt∗, concretely
(I+)θis=(I−)θis|s+πΔtω1 ,(I+)θic3=(I−)θic3|s+πΔtω1+3˜η1πΔtω21c3(1−27c3)(I−)θis|s+πΔtω1 ,(I+)sθi=(I−)sθi|s+πΔtω1−3˜η1πΔtω21c3(1−27c3)(I−)c3θi|s+πΔtω1 ,(I+)c3θi=(I−)c3θi|s+πΔtω1 |
where the coordinate we use for CP2 are the two angles θ1,2 in addition to the c3,s. The quantity ˜η1 is introduced in (57) and depends only on c3.
The complex structure I+ is valid for arbitrary flow time Δt, but the bi-hermitian metric is positive definite for small but non-zero Δt. The generalised Kähler potential is given by
K=Δt4logc3+3ω18π∫Δtπ/ω10du3∑k=1yk(c3,s+u)logyk(c3,s+u)yk(c3,s) . |
The first term is proportional to the Fubini-Study Kähler potential.
We leave the further exploration of our explicit results, such as the relation to certain integrable models [19] as well as mirror symmetry for a future work.
A generalized Kähler structure (GKS) on a manifold M can be described as a pair of integrable complex structures I± toghether with a Riemannian metric g, which is hermitian with respect to both, and a closed 3-form H such that ∇±I±=0 where ∇± is the Riemannian connection with skew torsion ±H.
In the case of 4-manifolds, suppose we are given (M4,g,I±) with g hermitian with respect to I± and ω±=I∗±g are the hermitian forms. We can then define 3-forms H± as
H±=12dc±ω±=i2(ˉ∂±ω±−∂±ω±) |
where ˉ∂±,∂± are defined using the complex structure I±. It is proved in [1] that if the first Betti number of M4 is even then H+=−H−=H automatically.
Moreover, let ∇± be the connections on the tangent bundle TM with torsion ±H defined as
⟨Z,∇±XY⟩=⟨Z,∇XY⟩±H(X,Y,Z), |
where X,Y,Z are sections of TM and ⟨−,−⟩ denotes the pairing using the metric. It can be shown that ∇±I±=0, (see e.g. [9] Proposition 2). This means that constructing a GKS on a 4-manifold with even Betti number is equivalent to finding I± and a metric that is hermitian with respect to both.
What played a crucial role in [3] was the formulation of the GKS in terms of holomorphic Poisson structures. One starts with the Hitchin's Poisson structure
Q:=[I+,I−]g−1. | (2) |
It can be shown that this is of type (0, 2)+(2, 0) with respect to both I±. From Q one can construct two Poisson structures
σ±:=14(I±Q+iQ), | (3) |
which are (2, 0) and holomorphic under I± respectively. The authors of [3] considered the GKS of symplectic type, we refer to [3] for the details of this definition, while we only record what would be important for this work. One has a symplectic form F, with the following relations to I±
I+−I−=−QF, | (4) |
FI++I∗−F=0. | (5) |
The solution to such a non-linear pair of equations is encoded in terms of the Morita equivalence of σ± as holomorphic Poisson structures. We recall its definition following Ping Xu, adapted to the holomorphic setting
Definition 2.1. (Definition 2.1 [24]) The two holomorphic Poisson structures (M,σ±) are Morita equivalent if there is a holomorphic symplectic manifold (X,J,Ω) (called the equivalence bi-module) and two maps s, t
![]() |
with the following properties. The source map s is holomorphic, anti-Poisson, i.e. s∗Ω−1=−σ− and complete, while the target map t is holomorphic, Poisson i.e. t∗Ω−1=σ+ and complete. Besides, the fibre of s,t is connected and simply connected. One requires further that kert∗ be symplectic orthogonal complement to kers∗.
Here completeness means that s∗f generates a complete Hamiltonian vector field (its flow time extendable to (−∞,∞)) if the function f on M generates under σ− a complete vector field. The same goes for σ+ and t∗f. The symplectic orthogonality of kers∗ and kert∗ implies*
*For a Poisson structure π on M, the map π#:T∗M→TM is defined as (π#dg)∘f:={f,g}π where ∘ denotes the action of a vector field on a function by derivation, and f,g∈C∞(M). Then we extend π# to act on 1-forms using C∞-linearity.
t∗(Ω−1)#s∗=s∗(Ω−1)#t∗=0. |
This ensures that the hamiltonian vector field generated by a function of type t∗h with h∈C∞(M) has no effect on the image of the source map (and vice versa). Finally that s and t are holomorphic means that the two complex structures I± are 'unified' as one single complex structure up on X.
The Morita equivalence is an equivalence relation between integrable Poisson manifolds. In particular, integrable Poisson manifolds are self Morita equivalent with self-equivalence bi-module given by the symplectic groupoid, which we will need next.
In the setting of GKS of symplectic type, this Morita equivalence between (M,σ±) can be constructed by deforming a self-Morita equivalence of (M,σ−). Let (X,J0,Ω0) be a holomorphic symplectic groupoid integrating the Poisson structure σ−, i.e. a holomorphic symplectic manifold (X,J0,Ω0) with maps
![]() |
satisfying identical conditions as in Definition 2.1 by replacing I+,σ+ with I−,σ−. One has again a similar orthogonality
t∗(Ω−10)#s∗=s∗(Ω−10)#t∗=0. | (6) |
It is helpful to keep in mind that J0 can be expressed as
J0=(ImΩ0)−1ReΩ0 | (7) |
and that the inverse of Ω0 (as a holomorphic 2-form) is
Ω−10=14((ReΩ0)−1−i(ImΩ0)−1). | (8) |
It turns out that a simple deformation Ω0→Ω=Ω0+t∗F gives us the Morita equivalence between σ±. For simplicity of discussion we assume that t∗F is such that ReΩ0+t∗F remains invertible. First the kernel of Ω defines a distribution ˉL⊂TCX which is integrable since Ω is closed. The invertibility of ReΩ0+t∗F ensures that L∩ˉL=0 and L⊕ˉL=TCX. This means that L is spanned by the (1, 0) vector fields in TCX but under a different complex structure J≠J0. The new complex structure can be expressed in the same way as (7)
J=(ImΩ)−1ReΩ=(ImΩ0)−1(ReΩ0+t∗F)=J0+(ImΩ0)−1t∗F. | (9) |
The source map s remains holomorphic (X,J)→(M,I−), indeed for any V∈TX
s∗JV=s∗J0V+s∗(ImΩ0)−1t∗Ft∗V=I−s∗V |
where the second term drops thanks to (6). In contrast, for the t map we have
t∗JV=t∗(J0+(ImΩ0)−1t∗F)V=I−t∗V+t∗(ImΩ0)−1t∗Ft∗V=(I–QF)t∗V, |
where we have used the fact that t is a Poisson map sending Ω−10 to σ−=(I−Q+iQ)/4. The last equation says that t∗ now intertwines J with a certain automorphism I+ of TM with I+=I–QF=I−(1+I−QF). This gives (4).
Furthermore, that the lhs of (9) squares to −1 gives
J0(ImΩ0)−1t∗F+(ImΩ0)−1t∗FJ0+((ImΩ0)−1t∗F)2=0,⇒ J∗0t∗F+t∗FJ0+t∗F(ImΩ0)−1t∗F=0,Apply t∗⇒I∗−F+FI–FQF=0 |
which is (5). This also implies (I+)2=−1.
The anti-Poisson property of s, i.e. s∗Ω−1=−σ− is unaffected. To check that t∗ sends Ω−1 to σ+, we have
t∗(Ω−1)=14t∗((ReΩ0+t∗F)−1−iImΩ−10)=14(t∗(ReΩ0+t∗F)−1+iQ). |
To evaluate the first term we can use a geometric series expansion [5] and get
t∗(ReΩ0+t∗F)−1=(1+I−QF)−1I−Q |
provided 1+I−QF is invertible. So we get
t∗(Ω−1)=14((1+I−QF)−1I−Q+iQ)=14(I+Q+iQ)=σ+ |
using (4). This shows that t∗ maps Ω−1 to σ+.
The Morita equivalence between σ± encodes the generalised Kähler potential (GKP) as a kind of generating function (see the discussion in Section 4 of [3]). Recall that just as the usual Kähler potential is not a globally defined function, the GKP is defined locally, and so the following discussion is local in nature.
From the previous section, one has a holomorphic symplectic form Ω=Ω0+t∗F, which can be expressed using local Darboux coordinate as Ω=i∑kdPk∧dQk where Qk,Pk are local holomorphic coordinates. The Morita equivalence (X,Ω,J) was constructed out of the symplectic groupoid (X,Ω0,J0) integrating σ−. The space of units is a Lagrangian, denoted L0, with respect to Ω0. Though L0 is no longer Lagrangian under Ω, one still has ImΩ|L0=0 since F is real. Suppose that locally in the P,Q coordinate system L0 is given by Pk=ηk(Q,ˉQ) where η=ηkdQk is a 1-form, then one has
d(iηkdQk)=Ω|L0=t∗F. |
This forces Imiη to be a closed 1-form and hence Reη=−dK locally for some real function K(Q,ˉQ), as a consequence
η=−2∂K. |
The local real function K is the GKP.
As an example, for an actual Kähler manifold (M,ω,I), one has σ±=0 and I+=I−=I. Then the Morita equivalence is given simply by the holomorphic cotangent bundle X=T(1,0)M and Ω0 is the standard holomorphic symplectic form. The maps s, t are the bundle projection π. The holomorphic symplectic form is then deformed to Ω0+π∗ω, for which we can pick the Darboux coordinates as follows. Let (xk,ξk) be the local holomorphic coordinate for M and the fibre coordinate respectively. Let Qk=xk and Pk=ξk−2∂kK with K being the Kähler potential in the usual sense, i.e. 2i∂ˉ∂K=ω. We can easily check
idPk∧dQk=id(ξkdxi−2∂K)=idξk∧dxk−2iˉ∂∂K=Ω0+ω. |
Then the zero section L0=M is given by the condition ξi=0 or Pi=−2∂iK, which is how the GKP is formulated in the last paragraph.
The pervading theme of the construction of the last section is that one starts from a Kähler manifold with I− as its complex structure, and one obtains I+ as a deformation of I−. In [13] Hitchin used a flow to deform I− so that for small flow time, one gets a 4D bi-hermitian metric. Note that in this construction the flow time is small but cannot be zero.
Let us give a quick overview of Hitchin's construction, while some proofs are collected in the next section. One starts with the data (M,I−,g,σ−), that is, a Kähler structure on the surface M plus a choice of a holomorphic Poisson tensor σ− i.e. any element of H0(M,K−1) with K the canonical bundle. The key observation is that the two holomorphic Poisson structures in (3) encoding the GKS share the same imaginary part Q, and so in deforming I− to reach I+ one has to preserve Q. It is thus natural to use a Q-Hamiltonian flow ϕt with ϕ0=id generated by some Hamiltonian h to deform I− so that
I+=(ϕt∗)−1I−ϕt∗ | (10) |
is the new complex structure. The difficulty is to produce a positive definite bi-hermitian metric.
Hitchin picked the hamiltonian for the flow h∼log||σ−||2. Even though h is ill-defined where σ− vanishes, the hamiltonian vector field
V=Q#dh |
is globally defined, as demonstrated in [13]. This is essentially because Q is zero whenever h diverges, rendering the divergence innocuous. Thus V, although not hamiltonian, is Poisson and its flow certainly preserves Q.
Take the inverse of σ−, considered as a meromorphic (2, 0)-form, and denote with ω−=Re(σ−1−). It is surely of type (2, 0)+(0, 2) with respect to I−. Also its pull-back ω+=ϕ∗tω− is by construction of type (2,0)+(0,2) with respect to I+. But we consider its (1, 1)-component with respect to I−
g=−I∗−ω++ω+I−. | (11) |
This 2-tensor turns out to be well defined (even though ω+ is not), and symmetric (1, 1) with respect to both I±. But it may not be positive definite. To ensure its positivity, Hitchin studied the first derivative of g with respect to t and showed that
I∗−˙g|t=0=4i∂ˉ∂h, | (12) |
where ∂, ˉ∂ use the complex structure I−. But the rhs is the curvature of the line bundle K−1, thus if the anti-canonical bundle is ample, then the curvature is positive definite. Since positivity is an open condition, for small but nonzero t one has constructed a bi-hermitian metric.
For clarity we temporarily denote with I=I− and It=I+. The flow ϕt is generated by the vector field V=Q#dh, and so It=I+ given in (10) can be written as
It=I+∫t0dt˙It=I+∫t0dt(ϕt∗)−1(LVI)ϕt∗, |
while LVI can be written as LVI=−Qd(I∗dh) which we prove in Lemma 3.1. Comparing with (4) one has
−QF=It−I=−∫t0dt(ϕt∗)−1(Qd(I∗dh))ϕt∗=−Q∫t0dt(ϕt∗)−1d(I∗dh)ϕt∗, |
since Q is preserved by ϕt. So we get the formula for F
F(t)=d∫t0dtϕ∗t(I∗dh). | (13) |
The first time derivative of F(t) is
˙F(0)=dI∗dh=2i∂ˉ∂h. | (14) |
As for the metric, denote with σ−=σ and σ+=σt=(ϕ−t)∗σ. Let ω+=ωt=Re(σ−1t), and set as in (11)
gt(X,Y)=−ωt(IX,Y)+ωt(X,IY). |
At flow time t=0, ω0 is of type (2, 0)+(0, 2) with respect to I and so g is zero. This is crucial for g to be globally defined for t>0, since the divergence of ωt resides only in the (2, 0)+(0, 2) component. We observe that
gt(X,Y)=−ωt(IX,Y)+ωt(X,IY)=Q−1(IX,ItY)−Q−1(ItX,IY) |
where we used ωt=Re(σ−1t)=−Q−1It and that Q is of type (0,2)+(2,0) w.r.t. I. We see from the last formula that gt is a symmetric tensor and
g=Q−1[I,It] |
c.f. (2). We omit the subscript t in g next. We see a democracy between It and I, thus if g is (1, 1) under I so will it be under It. Finally that g is well-defined comes from the relation It−I=−QF
g=Q−1[I,−QF]=FI−I∗F. |
Taking t-derivative at t=0 and using (14), we get the positivity (12).
Lemma 3.1.
LVI=−Qd(I∗dh). | (15) |
Proof. Let U be any (0, 1) vector field we compute (LVI)U=[V,IU]−I[V,U]=(−i−I)[V,U]. To evaluate the rhs we pick a local holomorphic function f so that df=∂f
⟨(LVI)U,df⟩=−2i⟨[V,U],df⟩=−2i(V∘U∘f−U∘V∘f)=2iU∘V∘f=2iU∘Q(df,dh)=2iU∘Q(∂f,∂h). |
As Q(2,0) and f are holomorphic the U derivative passes them by and gives
⟨(LVI)U,df⟩=Q(∂f,ιUdI∗dh)=−⟨df,(QdI∗dh)U⟩. |
The case for U being (1,0) is similar.
In this section we combine the two constructions of Section 2 and Section 3.1. Let (X,Ω0) be holomorphic symplectic groupoid integrating (M,I−,σ−), then the flow ϕt generated by the vector field V can be lifted up to ˆϕt on X generated by
ˆV=−((ImΩ0)−1)#(t∗dh), |
where the minus sign comes from (8) so that ˆV covers V. The lifted flow ˆϕt satisfies
t∘ˆϕt=ϕt∘t,s∘ˆϕt=s, | (16) |
recall that the symplectic orthogonality (6) ensures that the image of s remains fixed, hence the second relation.
We have an expected result
Lemma 4.1. (Proposition 7.2 in [3])
Ωt=ˆϕ∗tΩ0=Ω0+t∗F. | (17) |
Proof. We let Ωt=ˆϕ∗tΩ0 be the flow of Ω0, then
˙Ωt=ˆϕ∗tLˆVΩ0=ˆϕ∗tLˆV((ImΩ0)J0))=(ImΩ0)ˆϕ∗tLˆVJ0 |
One has again the relation LˆVJ0=(ImΩ0)−1dJ∗0t∗dh (here we understand the 2-form as a map from T to T∗ and a bi-vector as a map from T∗ to T) similar to Lemma 3.1. Thus
˙Ωt=ˆϕ∗tdJ∗0t∗dh=ˆϕ∗tt∗dI∗dh=t∗ϕ∗tdI∗dh |
where we used (16). Continuing
Ωt−Ω0=∫t0dt˙Ωt=∫t0t∗ϕ∗tdI∗dh=t∗F |
where F was given in (13).
Next we look for the Darboux coordinates for Ωt. Pictorially the flow looks like Figure 2. If we have found the Darboux coordinates for Ω0
i∑kdPk∧dQk=Ω0 |
then these will pull back to
Qt=ˆϕ∗tQ,Pt=ˆϕ∗tP |
and are subsequently the Darboux coordinates of Ωt since we have shown (17). From this we also see that the new complex structure Jt on X induced from the deformed holomorphic symplectic structure Ωt is related to J0 by the flow ˆϕt.
The space of units L0 is flown to Lt. Suppose that L0 is described under Qt,Pt chart as
L0:={Pt,k=ηk(t,Qt,ˉQt)} |
for some (1, 0)-form η. Then as discussed earlier
idη=Ωt|L0=ˆϕ∗tΩ0|L0=(Ω0+t∗F)|L0=t∗F|L0 | (18) |
showing that Im(iη) is closed and one has a local real function K(Q,ˉQ) with iη=−2i∂K.
There is in general no formula for finding the primitive for iη, apart from its flow equation.
Lemma 4.2. Assuming that at t=0 the Lagrangian L0 is described by P=0, then K satisfies the flow equation
˙K=−(t∗h)|L0. |
Proof. Let Ω0=i∑kdPk∧dQk as above and L0={Pi=0}. We flow L0 with a Hamiltonian vector field ˆV=−((ImΩ0)−1)#dH for some function H(Q,ˉQ,P,ˉP). For small t, L0 is described by the graph of a function K(t,Qt,ˉQt): L0={Pt,j=−2∂QjtK}. Take the t-derivative of the equation Pt,j=−2∂QjtK (we temporarily suppress the subscript t)
lhs=˙Pj=2∂QjH|P=−2∂QK , |
rhs=−2ddt∂QjK=−2∂Qj∂tK−2(∂Qk∂QjK)(−2∂PkH)−2(∂ˉQk∂QjK)(−2∂ˉPkH)|P=−2∂QK |
where one can use (8) for the imaginary part of Ω0. Note that the lhs differentiates H first and then sets P=−2∂QK, so if we combine it with the second and third term on the rhs
2∂QjH(Q,ˉQ,−2∂QK,−2∂ˉQK)=−2∂Qj∂tK⇒ ∂tK=−H(Q,ˉQ,−2∂QK,−2∂ˉQK). |
Setting H=t∗h, we get the result.
In the toric case later, we will be able to solve the flow equation for K somewhat more explicitly.
Hitchin construction applies to Del Pezzo surfaces. Although we will perform the computation on CP2 certain steps of the construction, for example the description of the symplectic groupoid, are valid more generally for the toric Del Pezzo surfaces. For this reason we recall general facts in this section.
A good textbook for this is Fulton [8], and we shall also make use of the explicit parametrisation of toric manifolds due to Guillemin [11].
Our focus of the toric geometry is more on its Kähler aspect, and so it is more appropriate to use the moment map polytope description (the fan description is more adapted for the algebraic aspect).
Definition 5.1. A polytope Δ⊂Rd is the convex hull of a finite number of points in Rd, called the vertices. A polytope Δ⊂Rd is said to be Delzant if for each vertex p one has
1. there are exactly d edges meeting at p;
2. each edge meeting in p is of the form p+tu with u∈Zd and t≥0 (rationality);
3. the d edges meeting in p form a Z-basis for Zd. (smoothness)
Theorem 5.2. (Delzant [6]) There is a 1-1 correspondence between symplectic toric manifolds (up to Td equivariant symplectomorphisms) and Delzant polytopes.
Given a toric symplectic manifold M, the image of the moment map for the torus action is a polytope and it is Delzant. While from a Delzant polytope, one can construct a toric symplectic manifold MΔ through symplectic reduction or symplectic cut [16], the Delzant condition ensures the smoothness. It is useful to keep in mind the gist of the construction: one can visualise MΔ as a Td fibration over the interior of Δ, however on a face Δa with normal →va the torus action
→vai∂∂ϕi |
degenerates. Here we use ϕi to denote the d angle-coordinates of Td. Condition 3 above says that, each vertex has a neighbourhood of type Cd, and up to an SL(d,Z) the ϕi are the standard angle coordinates of Cd.
For the remainder of the paper, we let d=2 and all polygons are assumed Delzant. We denote the primitive normal to face Δa as →va. In fact, each face corresponds to an embedded CP1 in MΔ. It is a toric invariant divisor, and all toric invariant divisors are generated (not freely) by the faces. Suppose that the polygon is described by
Δ={→y∈R2|⟨→y,→va⟩+λa≥0}, |
where →y are the coordinates of R2. Then from [11] the Kähler class is expressed as
[ω]=∑aλa[a]. |
Since the divisors [a] are not independent and have n relations among them, this expansion is not unique.
Definition 5.3. A smooth surface M is called Del Pezzo if it is birational to CP2 and the anti-canonical class is ample. The self intersection d=K⋅K is called the degree of the Del Pezzo, here K is the canonical divisor.
Except for CP1×CP1, the other Del Pezzo's are obtained from blowing up up to 8 points in general position. But we are interested in toric Del Pezzo's, so one can only blow up toric fixed points on CP2. The polygon corresponding to CP2 is a triangle, see Figure 5.1. There are only three toric invariant points located at the three corners of the triangle. We are left with the following polygons
![]() |
Let →y=(y1,y2) parametrise Δ⊂R2, and let θ1,θ2 be the angle coordinates of T2. We have
Proposition 5.4. (Guillemin [11]) Let MΔ be the toric symplectic manifold corresponding to Δ, with normals →va. The symplectic form of MΔ reads
ω=∑idyi∧dθi. |
Further the complex structure and metric read
J=∑ijGij∂θi⊗dyj−Gij∂yi⊗dθj, | (19) |
g=∑ijGijdyi⊗dyj+Gijdθi⊗dθj, | (20) |
where
G=12∑a(⟨→va,→y⟩+λa)log(⟨→va,→y⟩+λa),Gi=∂yiG,Gij=∂yi∂yjG=12∑a→vai→vaj⟨→va,→y⟩+λa |
and GijGjk=δik. The complex coordinates can be expressed as
ξi=eiθi+Gi. | (21) |
Recall from the discussion below Theorem 5.2 that each corner of the polygon corresponds to a C2 patch, covering the whole M.
![]() |
At a corner with normals →u1,→u2, one can take combination of (21) to give local complex coordinates for the C2 patch
z1=ξu221ξ−u212,z2=ξ−u121ξu112. | (22) |
For later use we record how the complex coordinates change from one local patch to another.
Lemma 5.5. Suppose at the two corners of Δ with normals →ua and →va respectively, one has complex coordinates {za,a=1,2} and {˜za,a=1,2} defined as in (22) using {→ua} and {→va} respectively, then over the open intersection where both coordinate systems are valid, one has the transformation
˜za=∏bzAabb,Aab=[→u1×→v2→u2×→v2→v1×→u1→v1×→u2],detA=1. | (23) |
The proof is a direct calculation. Of course in the derivation, we used →u1×→u2=1=→v1×→v2 since Δ is Delzant.
A consequence of (23) is
Lemma 5.6. The unique (up to a constant multiple) toric invariant holomorphic Poisson tensor on a toric surface has local expression
σ=iz1z2∂∂z1∧∂∂z2. | (24) |
This expression is defined on the C2 patch corresponding to a corner of Δ with normals →u1,→u2, but is in fact globally defined.
Proof. We need only check that if one makes a coordinate transform as in (23), the local expression of σ remains unchanged, but this follows from detA=1.
A straightforward calculation gives
Lemma 5.7. When written in the action-angle coordinates, the holomorphic Poisson tensor σ reads
σ=i(12G1i∂yi−i2∂θ1)∧(12G2j∂yj−i2∂θ2). | (25) |
Its norm computed with the metric (20) is
||σ||2=12det−1(Gij). | (26) |
The determinant det(Gij) can be evaluated as
4det(Gij)=(∑ava1va1→y⋅→va+λa)(∑bvb2vb2→y⋅→vb+λb)−(∑ava1va2→y⋅→va+λa)2. |
The concrete expression for CP2 is given in Section 6.
According to the recipe for constructing the Morita equivalence (see Section 2.1), we need to find (X,Ω0) integrating σ which is given in (24). One can use the general construction given in [25] for Poisson structures twisted by a solution of the classical Yang–Baxter equation. For the case we are interested in there is an easy description that we discuss in this section. We start from an example on C2.
Example 5.8. (See Theorem 3.7 of [17]) On C2 with complex coordinates x,y, we have a holomorphic Poisson tensor
π=xy∂x∧∂y. | (27) |
The pervading trick in finding the integrating object is to notice that π is non-degenerate except when xy=0. This means that (X,Ω) would look like a pair groupoid X∼U×U in an open dense U={xy≠0}=(C∗)2, with
Ω−1=−π|(x,y)+π|(ˆx,ˆy) |
where x,y are the source coordinates and ˆx,ˆy those of the target.
To extend from U to the whole of C2, we change coordinates to (x,y,u,v) with (x,y)∈U and (u,v)∈C2. They are related to the old ones as
ˆx/x=euy,ˆy/y=evx. | (28) |
This trade is meant to reflect the fact that if xy=0 then ˆx=x,ˆy=y necessarily and u,v,x,y are a more fundamental set of coordinates.† Rewriting Ω−1 in the (x,y,u,v) coordinates
†Note that u,v in (28) are fixed by (x,ˆx,y,ˆy) up to a Z×Z worth of shifts due to the exponential. So one should really call the symplectic groupoid at this stage the path groupoid, with Z×Z reflecting the fundamental group of C∗×C∗. We thank the referee for pointing this out to us.
−Ω−1=−∂u∧∂y−uv∂u∧∂v−vy∂v∧∂y+∂v∧∂x+ux∂u∧∂x+xy∂x∧∂y. |
This is invertible, in fact detΩ=1 everywhere. Its inverse is
Ω=uvdx∧dy+vydx∧du−dx∧dv+dy∧du−uxdy∧dv−xydu∧dv. | (29) |
The next step is to globalise this construction by regarding u,v as the fibre coordinates of some vector bundle to take into account the non-trivial coordinate change.
Theorem 5.9. Let M be a smooth toric surface associated with a Delzant polygon Δ. The holomorphic Poisson structure σ (24) can be integrated into a symplectic groupoid (X,Ω0). The space X is diffeomorphic to the holomorphic cotangent bundle of M. If (za,ξa) are the local base and fibre coordinates, then Ω0 reads
iΩ0=−ξ1ξ2dz1∧dz2−ξ1z2dz1∧dξ2+dz1∧dξ1+dz2∧dξ2+ξ2z1dz2∧dξ1−z1z2dξ1∧dξ2 . | (30) |
Its inverse Π=Ω−10 reads
−iΠ=∂ξ1∧∂z1+∂ξ2∧∂z2−ξ1ξ2∂ξ1∧∂ξ2+ξ1z2∂ξ1∧∂z2−ξ2z1∂ξ2∧∂z1−z1z2∂z1∧∂z2 . |
Let g=(z1,z2;ξ1,ξ2) be a point in X, the source, target maps read
g=(z1,z2;ξ1,ξ2),s(g)=(z1,z2),t(g)=(z1eξ2z2,z2e−ξ1z1). | (31) |
Let h=(z1eξ2z2,z2e−ξ1z1;η1,η2) be another point with s(h)=t(g) then
h∘g=(z1,z2;ξ1+η1eξ2z2,ξ2+η2e−ξ1z1). |
Proof. We only need to prove that the local model ex.5.8 globalises under the coordinate change rule Lemma 5.5. If za→˜za as in (23), then as fibre coordinates of T∗M
ξa=∑b˜zbzaAba˜ξb. |
This means the combination zaξa (no sum) transforms linearly
zaξa=∑b˜zb˜ξbAba. |
Note that Aab is an SL(2,Z) matrix, but recall that for all SL(2,C) matrices A one has
ϵ−1ATϵ=A−1,ϵ=[01−10] |
This shows that
ϵ[z1ξ1z2ξ2]=[z2ξ2−z1ξ1] |
transform the same way as logza, thus the target (z1eξ2z2,z2e−ξ1z1) transform the same way as (z1,z2). This shows that the t map is globally defined.
Using the coordinate transform for T∗C2 it is straightforward, though tedious, to check that Ω0 is globally well-defined.
Lemma 5.10. The Darboux coordinates for the holomorphic symplectic form (30) are
q1=z1eξ2z2/2,q2=z2e−ξ1z1/2,p1=ξ1e−ξ2z2/2,p2=ξ2eξ1z1/2,iΩ0=dq1∧dp1+dq2∧dp2. |
Under these coordinates, the source, target maps are
s(q1,q2,p1,p2)=(q1e−p2q2/2,q2ep1q1/2) ,t(q1,q2,p1,p2)=(q1ep2q2/2,q2e−p1q1/2) . |
The proof is a direct computation.
Remark 5.11. From the Darboux coordinates it is easy to check
{t∗f1,s∗f2}Π=0,f1,2∈C∞(M) |
which leads to (6).
Lemma 5.10 also gives a simpler check that the symplectic form (30) is well-defined globally. Using the same argument as above, the qa,pb transform as
˜qa=∏b(qb)Aab,˜pa=(˜qa)−1∑b(pbqb)(A−1)ba. |
Therefore using logqa and paqa as coordinates, it is clear
∑ad˜qa∧d˜pa=∑adlog˜qa∧d(˜qa˜pa)=∑adlogqa∧d(qapa)=∑adqa∧dpa. |
We put together the ingredients of previous sections and construct a GKS in explicit terms expressed with action-angle coordinates. The polygon is a triangle with normals →v1=[1;0], →v2=[0;1] and →v3=[−1;−1]. We fix the Kähler class by fixing λ1=0=λ2 and λ3=1 so that [ω]=[3].
![]() |
From the general formula Proposition 5.4, we have
G=123∑i=1yilogyi,Gij=12[1−y2y1y31y31y31−y1y2y3],Gij=2[y1(1−y1)−y1y2−y1y2y2(1−y2)]. |
where we have let y3=1−y1−y2 which turns out quite convenient later on. Now
4det(Gij)=1y1y2y3⇒||σ||2=2y1y2y3 |
from (26). This is completely expected since σ vanishes on a cubic curve, which in the toric invariant case is the product of three lines, corresponding to each of the three faces Δa={ya=0}, a=1,2,3.
The complex coordinates (21) are
z1=eiθ1+(logy1−logy3)/2=(y1/y3)1/2eiθ1,z2=eiθ2+(logy2−logy3)/2=(y2/y3)1/2eiθ2, | (32) |
they are in fact the standard inhomogeneous coordinates for CP2.
The Hitchin Poisson tensor is the imaginary part of σ
Q=4Imσ=4y1y2y3∂y1∧∂y2−∂θ1∧∂θ2. | (33) |
We pick the Hamiltonian h as
h=−14log(y1y2y3). |
The pre-factor is for convenience and can always be absorbed by redefining t. This gives a vector field
V=Q#dh=y1(y2−y3)∂∂y1+y2(y3−y1)∂∂y2 |
and so the flow equation reads
˙y1(t)=y1(t)(y2(t)−y3(t)),and cyc perm (1→2→3). | (34) |
It pays to keep the symmetry between y1,y2,y3 and so we define the symmetric polynomials
c1=y1+y2+y3=1,c2=y1y2+y2y3+y3y1,c3=y1y2y3. |
It is obvious that c3 is a conserved quantity under the flow since h∼logc3.
Proposition 6.1. The flow equation (34) is solved by
yk(t)=c31/12−℘(t+kt∞) | (35) |
with ℘ the Weierstrass elliptic function with ω1∈R and ω2∈iR as the half periods of a square lattice in C (Figure 3) while t∞=(2/3)ω1.
Finally t has constant imaginary part −iω2 and arbitrary real part‡.
‡We made this awkward choice to avoid clutter in later formulae. A more pedantic way would be to write t as t+ω2 for real t. But as far as the flow is concerned one only cares about shifts of t, which is still real.
Proof. We can solve from y2+y3=1−y1, y2y3=c3/y1 that
y2−y3=±((1−y1)2−4c3/y1)1/2 |
and hence
(˙y1)2=(y1)2((1−y1)2−4c3/y1)=y1(y1(1−y1)2−4c3). |
The rhs is a quartic polynomial with a root at y1=0, which hints at changing variables y1=1/u so one can turn the quartic into a cubic
(˙u)2=u2−2u+1−4c3u3. |
Further letting ℘=1/12−c3u gives
˙℘2=4℘3−g2℘−g3=4℘3−(112−2c3)℘−(c36−1216−c23). | (36) |
The solution is the celebrated Weierstrass ℘ function. It is a doubly periodic meromorphic function with period lattice 2Λ generated by 2ω1,2ω2∈C. It is a convention that the lattice is 2Λ while ω1,ω2 are the half periods. We compute the modular invariant (46) (see the appendix), which shows that the lattice is a square lattice and ω1 is real while ω2 is purely imaginary. For such a lattice, the real loci are the lattice lines of Λ. We have the solution
yi(t)=c31/12−℘(t+ti) |
where ti is real and fixed using the initial data. In order that yi be real, the t must have imaginary part equal to 0 or ω2. It cannot be the former option, since then t+ti can reach the point t∞ where ℘(t∞)=1/12 (see Lemma A.2) and yi is unbounded. For the latter option, the solution is bounded and periodic with period 2ω1. Indeed starting from 0+ (where ℘=+∞) going along straight lines via ω1, ω1+ω2, ω2 and reaching i0+ (where ℘=−∞), the function ℘ is monotonically decreasing. As 1/12 is reached at (2/3)ω1, one has ℘<1/12 along z=ω2. This is reviewed in Section A.1.
To fix the initial values ti, we can analytically continue the solution down to the real axis and then approach t+t3=t∞ from the right, where y3→∞. But the conditions ∑yi=1, ∏yi=c3 forces one of y1,y2 to go to 0− and −∞ respectively. We pick the option y1→−∞ and y2→0− first. In order for this to happen one has to have
t1=t3−2t∞mod2Λ∼t3+t∞,t2=t3−t∞. |
Indeed, this way t+t1 approaches −t∞ from the right giving y1→−∞, and t+t2 approaches 0 from the right giving y2→0−.
The other option would simply swap t1 and t2, to show that this is not the case, we look at Figure 4. The figure shows the contour of fixed c3, and the flow goes clockwise along this contour, as can be seen from (34). The three red lines cut the triangle into six regions depending on the relative sizes of yi. With the first option, pick, say, t3=0, i.e. one starts the flow at y1=y2>y3 (the red dot). Increasing t, one has y3,y1 increase while y2 decreases until one reaches t=ω2+t∞/2, where y1>y3=y2. This means one moves down into the east triangle and reaches the boundary with the southeast triangle at t=ω2+t∞/2. However, had one picked the other option, one would swap y1,y2 and one would be moving up into the north east triangle. This does not agree with the direction of the flow.
Remark 6.2. It is instructive to see that the solution satisfies ∏yi=c3. Indeed, the product
(112−℘(z))(112−℘(z−t∞))(112−℘(z+t∞)) |
is doubly periodic without poles or zeros: for example, the first factor has a double pole at z=0, but the other two both have a single zero at z=0. Thus the product must be a constant, which can be obtained by setting z at a convenient point. We choose z=w1 and in Lemma A.4 we show that the constant is c23.
The solution involves the inversion of ℘ function, we would like to avoid that by expressing elliptic functions with Weierstrass zeta functions. These are recalled briefly in the appendix.
Lemma 6.3. One can express yi as Weierstrass zeta functions (where t∞=2/3ω1)
y3(t)=ζ(t−t∞)−ζ(t+t∞)+2ζ(t∞),y1(t)=ζ(t)−ζ(t−t∞)+12−ζ(t∞),y2(t)=ζ(t+t∞)−ζ(t)+12−ζ(t∞). |
Proof. We use the inversion formula (48.2 [7])
℘′(v)℘(u)−℘(v)=ζ(u−v)−ζ(u+v)+2ζ(v). |
This is a special case of a broader formula that expresses any elliptic function using the Weierstrass ζ function, see Theorem 6.2 [7]. We show in Lemma A.2 that ℘(t∞)=1/12, and ℘′(t∞)=−c3, the result for y3 follows. For y1, one would get from the same formula that y1(t)=ζ(t)−ζ(t+2t∞)+2ζ(t∞). Then one uses the periodicity ζ(z+2ωi)=ζ(z)+2ηi and also ζ(t∞)=1/6+2/3η1 from Lemma A.5.
Proposition 6.4. The GKP reads
K=t4log(c3)+38∫t0dt∑iyitlogyityi0 | (37) |
where yi(z) is given in (35) and we write yi0=yi(t0), yit=yi(t+t0).
Recall from Proposition 6.1 that for fixed c3, the value of t fixes the yi's. Here t0 is chosen to correspond to the initial values of yi in the flow through (35).
Proof. § In the proof, we exploit the fact that in an open dense subset, we can use the source and target coordinates z,ˆz to parametrise X, making the flow easier to write down. With these coordinates, the space of units is L0={z=ˆz}. The Darboux coordinates Q,P for Ω0 are related to z,ˆz as
§ We thank the referee for the suggestion about summarising the relations between the various coordinates that improved the readability of the proof.
ˆz1z1=eP2Q2,z1ˆz1=(Q1)2,ˆz2z2=e−P1Q1,z2ˆz2=(Q2)2 | (38) |
as can be read from Lemma 5.10.
Referring to Figure 2, under the flow ˆϕt, the Q,P coordinates pullback to Qt,Pt, the Darboux coordinates for Ωt. As the flow is generated by the function t∗h, only ˆz will experience this flow: (z,ˆz)⇝(z,ˆzt). The coordinates Qt,Pt are related to z,ˆzt in the same way as (38), only with the addition of a subscript t. Also L0 is described in the z,ˆzt coordinates by setting ˆzt=zt, where zt is the flow of z under ϕt generated by h.
Now to the GKP, Ωt is written as
Ωt=idηt=id∑kPtkdQkt. |
We know that Reηt is closed when restricted to L0. The goal is to find its local primitive. As laid out above, to do this efficiently, we use the source and target coordinates z and ˆzt to parametrise X
ηt=∑kPtkdQkt=∑k(PtkQkt)dlogQkt=−12logˆz2tz2dlog(z1ˆz1t)+12logˆz1tz1dlog(z2ˆz2t)=12(logˆz[1tdlogˆz2]t−logz[1dlogz2])+12d(logˆz1tlogz2−logz1logˆz2t) |
where [a⋯b] means anti-symmetrisation of indices a,b.
The second term in ηt is already exact, we focus on the first term and its real part
Re(logˆz[1tdlogˆz2]t−logz[1dlogz2])=log|ˆz[1t|dlog|ˆz2]t|−log|z[1|dlog|z2]|−argˆz[1tdargˆz2]t+argz[1dargz2] . |
Now restrict Reηt to L0 by setting ˆzt=zt, and keeping in mind that zt and z have the same phase as can be seen from the flow equation (34) and (32)
8Reη|L0=log|z[1t|2dlog|z2]t|2−log|z[1|2dlog|z2]|2+d(log|z1t|2log|z2|2−log|z1|2log|z2t|2) . | (39) |
We focus on finding the primitive of the first term. We denote by
ξ=log|z[1|2dlog|z2]|2 |
and the first term on the rhs of (39) is just ϕ∗tξ−ξ. We have
ϕ∗tξ−ξ=∫t0dt˙ϕ∗tξ=∫t0dtϕ∗tLVξ=∫t0dtϕ∗t(ιVd+dιV)ξ. |
We can evaluate explicitly
ιVξ=log|z1|2(1−3y1)+log|z2|2(1−3y2)=logy1y3(1−3y1)+logy2y3(1−3y2)=3∑i=1(1−3ya)logya ,ιVdξ=2ιV(dlog|z1|2∧dlog|z2|2)=2(3y2−1)dlogy2y3+2(3y1−1)dlogy1y3=2(3y2−1)y2y3(y2dy1+(1−y1)dy2)+2(3y1−1)y1y3(y1dy2+(1−y2)dy1) ,=2dy1y1−y3y1y3+2dy2y2−y3y2y3=−2dlog(y1y2y3) . |
In the computation we have used
|z1|2=y1y3,V∘|z1|2=y1y3(3y2−1),|z2|2=y2y3,V∘|z2|2=y2y3(1−3y1), |
where ∘ denotes the action of a vector field on a function by derivation. Altogether
ϕ∗tξ−ξ=d∫t0(−3∑iyitlogyit−log(y10y20y30)), |
recall that the product y1y2y3 is a conserved quantity under the flow.
The second term of (39) is already exact, but we can write it also as an integral. Let
g(t)=log|z1t|2log|z2|2−log|z1|2log|z2t|2. |
Now use the same trick by differentiating and integrating
g(t)−g(0)=∫t0(3y2t−1)logy20y30+(3y1t−1)logy10y30=∫t03∑i=1(3yit−1)logyi0. |
Putting everything together
8Reηt|L0=d∫t0dt(−3∑iyitlogyityi0−2log(y10y20y30)). |
Remembering the convention that Reηt|L0=−dK, we get the GKP (37). Note the leading term of (37) agrees with Lemma 4.2.
Remark 6.5. First, the generalised Kähler potential is not a global function, indeed our expression (37) is ill-defined whenever yi=0. However the ill-defined term resides only in the leading term which is proportional to the Kähler potential of the Fubini-Study metric. The correction term can be extended to yi=0 because the log always contains the combination yit/yi0.
Second, just knowing the GKP is not enough, one needs to know the local complex coordinates Qt,ˉQt on Lt. Again in the complement of the anti-canonical divisor where σ vanishes, we have
Q1t=(y10y1ty30y3t)1/2eiθ1,Q2t=(y20y2ty30y3t)1/2eiθ2. |
These can be read off from the relation between z,ˆz and Q in (38) and how z is related with y-coordinates given in (32). Note when t=0, they revert back to the standard complex coordinates.
Lastly, unfortunately we did not manage to perform the integral in the GKP. A possible strategy is to relate σ to the theta functions as in (54) and use the infinite product formula for the latter to deal with the logarithm.
Our flow ϕt is solved by elliptic functions, so to compute e.g. the pullback map ϕ∗t, one needs to differentiate ϕt with respect to the initial points y1,20, where the notation is as in Proposition 6.4. These initial values determine c3 and so affect the flow in two ways
1. through c3 dependence of the shift t∞=2/3ω1 given in (56)
2. through the dependence of ℘,ζ,σ on c3 (via g2 and g3 as in (45)) given in (55).
To compute these derivatives, it is beneficial to preserve the symmetry between y1,2,3. So we opt to use the coordinates c3 and the t to parametrise the triangle base of CP2, see Figure 4, where c3 determines the contours and t parametrises each contour. However, the t-variable has a c3-dependent period 2ω1, so we re-scale t in order that all the contours be of period 2π
t=ω2+sω1π. | (40) |
This has a small price since s is not a good variable when c3=0 or 1/27 where the ω1 or ω2 goes to infinity. To summarise, the c3 and s coordinates are the 'polar' coordinates of the triangle, with c3∈[0,1/27] being the radius while s∈[0,2π] being the angle coordinate.
The initial values y10,y20 of the flow will be expressed in terms of the new coordinates c3,s0. We compute the following partial derivatives in the appendix (see Lemma A.10)
∂c3yi(s)=3(c3−27c23)(16((yi)2−yi)−3c3yi+2c3+yi(yi+1−yi+2)ς(s+2π3(i−3))) |
where the index i is taken mod 3 and the function ς(z) is defined as
ς(z)=ζ(ω2+zω1π)−η2−zπη1. |
It is real of period 2π for real z. See Definition A.7 for further properties of this function. With these we have
Proposition 6.6. In the c3,s coordinate system, the Hitchin Poisson structure (33) reads
Q=4c3πω1∂c3∧∂s−∂θ1∧∂θ2. |
The standard complex structure reads
⟨dθi,J∂s⟩=[Jθ1sJθ2s]=ω12π[3y2−11−3y1],⟨dc3,J∂θi⟩=[Jc3θ1Jc3θ2]=4πc3ω1[−Jθ2sJθ1s],⟨dθi,J∂c3⟩=[Jθ1c3Jθ2c3]=14c3(1−27c3)[1+3y2−18y1y2−6ς(s)(1−3y2)1+3y1−18y1y2+6ς(s)(1−3y1)],⟨ds,J∂θi⟩=[Jsθ1Jsθ2]=4c3πω1[Jθ2c3−Jθ1c3] |
where yk(s) read
yk(s)=c3112−℘(ω2+ω1π(s+2πk3)) . |
Proof. The first statement is trivial, since h=−1/4logc3 is the hamiltonian that generates the flow ∂t. Nonetheless, we can calculate it explicitly. The angular part of Q is not affected, while the radial part is
4c3∂y1∧∂y2=4c3∂(c3,s)∂(y1,y2)∂c3∧∂s |
and the Jacobian can be computed using the explicit expressions ∂c3yi(s) given above. It turns out that the Jacobian equals
∂(c3,s)∂(y1,y2)=πω1 |
which reflects the fact that when ω1=0 (which happens when c3=1/27, the contour in Figure 4 is the centre point), the s-coordinate is not defined, indeed the s variable is like the angle coordinate of the contours in the triangle.
For the complex structure, since we have
J=Gij∂θi⊗dyj−Gij∂yi⊗dθj. |
We only need to focus on the first term, since the second is the inverse of the first. Evaluating the derivatives we get
⟨dθi,J∂s⟩=12[1−y2y1y31y31y31−y1y2y3][y1(y2−y3)y2(y3−y1)]ω1π=ω12π[3y2−11−3y1], |
while
⟨dθi,J∂c3⟩=12[1−y2y1y31y31y31−y1y2y3][∂c3y1∂c3y2]=14c3(1−27c3)[1+3y2−18y1y2−6ς(1−3y2)1+3y1−18y1y2+6ς(1−3y1)]. |
The components ⟨ds,J∂θi⟩ and ⟨dc3,J∂θi⟩ are fixed by demanding J2=−1, giving
⟨dc3,J∂θi⟩=−2c3[1−3y11−3y2]T ,⟨ds,J∂θi⟩=−π(1−27c3)ω1[−1−3y1+18y1y2−6ς(s)(1−3y1)1+3y2−18y1y2−6ς(s)(1−3y2)]T. |
Remark 6.7. With the explicit Poisson vector field Q given above, the flow ϕt is simply
ϕt:(c3,s)→(c3,s+tπω1). |
It is also crucial to remember that the flow time is measured in t, not in s. Even though s and t differ only by a factor of ω1, the half period depends on c3 in a quite complicated way (56). As a concrete instance where this can cause confusion we state the next corollary.
Corollary 6.8. The two complex structures of the generalised Kähler structure of CP2 are explicitly given by I−(c3,s)=J(c3,s) and
(I+)θis=Jθis(c3,s+πΔtω1) ,(I+)θic3=Jθic3(c3,s+πΔtω1)+Jθis(c3,s+πΔtω1)3˜η1πΔtω21c3(1−27c3) ,(I+)sθi=Jsθi(c3,s+πΔtω1)−Jc3θi(c3,s+πΔtω1)3˜η1πΔtω21c3(1−27c3) ,(I+)c3θi=Jc3θi(c3,s+πΔtω1) |
where Δt is the flow time and J(c3,s) is given in the proposition above. The terms proportional to Δt above comes from the c3-derivative of s+πΔt/ω1.
Next let us investigate the 2-form F more closely.
Proposition 6.9. For fixed flow time Δt, the 2-form F in (4) reads
F=−(I∗+dθ1−I∗−dθ1)∧dθ2+[1↔2] |
with I± given in the last corollary.
This result provides an explicit check for (4).
Proof. Fixing Δt we have from (13)
F=d∫Δt0dtϕ∗t(J∗dh)=12d∫Δt0dtϕ∗t((1−3y1)dθ1+(1−3y2)dθ2)=−32d∫Δt0dtϕ∗t(yi∧dθi) . |
where the factor 3 is expected since ddch is proportional to the curvature of the canonical class which is 3 times the Kähler class. The integral
Yi=∫Δt0dtϕ∗tyi |
can be worked out using Lemma 6.3 and the fact that the integral of ζ is logσ, see (52).
Y1=logσ(ω2+ω1sπ+Δt)σ(ω2+ω1sπ−t∞)σ(ω2+ω1sπ+Δt−t∞)σ(ω2+ω1sπ)+Δt(12−ζ(t∞)) ,Y2=logσ(ω2+ω1sπ+Δt+t∞)σ(ω2+ω1sπ)σ(ω2+ω1sπ+Δt)σ(ω2+ω1sπ+t∞)+Δt(12−ζ(t∞)) | (41) |
And so F is simply
F=−322∑i=1dYi∧dθi . |
It is straightforward to compute dYi i.e. ∂c3Yi,∂sYi, though rather tedious. So we record only one intermediate step
∂c3(logσ(ω2+ω1sπ)σ(ω2+ω1sπ−t∞)+sω1π(12−ζ(t∞))) =1c3(1−27c3)(12(y2−y3)−sπ˜η1+(3y1−1)ς(s+2π/3)) =−23Jθ2c3(s)−s˜η1πc3(1−27c3) ,∂c3(logσ(ω2+ω1sπ+t∞)σ(ω2+ω1sπ)+sω1π(12−ζ(t∞))) =1c3(1−27c3)(12(y3−y1)−sπ˜η1+(3y2−1)ς(s−2π/3)) =23Jθ1c3(s)−s˜η1πc3(1−27c3) |
where ˜ηi=ηi+ωi(1/12−3c3) and we have discarded any s-independent term.
The reason we can disregard such terms is that to get ∂c3Yi we need only take the difference of the above expression at s+πΔt/ω1 and at s, so the s-independent terms drop. However we must not forget the c3-dependence in the shift πΔt/ω1, which produces 3yiΔt˜η1/(c3(1−27c3)ω1). This is another instance where the remark 6.7 is important.
∂c3Y1=−23Jθ2c3(s+πΔt/ω1)+23Jθ2c3(s)+(3y1−1)˜η1Δtω1c3(1−27c3)=−23Jθ2c3(s+πΔt/ω1)+23Jθ2c3(s)−Jθ2s2π˜η1Δtω21c3(1−27c3)=−23(I+−I−)θ2c3 ,∂c3Y2=23Jθ1c3(s+πΔt/ω1)−23Jθ1c3(s)+(3y2−1)˜η1Δtω1c3(1−27c3)=23Jθ1c3(s+πΔt/ω1)−23Jθ1c3(s)+Jθ1s2π˜η1Δtω21c3(1−27c3)=23(I+−I−)θ1c3 |
where we have used cor.6.8. The s-derivatives are much easier
∂sY1=ω1π(y1(s+πΔt/ω1)−y1(s))=−23(I+−I−)θ2s ,∂sY2=ω1π(y2(s+πΔt/ω1)−y2(s))=23(I+−I−)θ1s . |
Now the assembling is straightforward
−F=(−I++I−)θ2sds∧dθ1−(I+−I−)θ2c3dc3∧dθ1+(I+−I−)θ1sds∧dθ2+(I+−I−)θ1c3dc3∧dθ2 . |
Our primary source of reference is the book [7].
Given a lattice Λ=spanZ⟨ω1,ω2⟩, the function ℘(z) is double periodic with period 2Λ. It can be expressed as a sum
℘(u)=℘(u|2Λ)=u−2+∑ω∈Λ′(1(u−2ω)2−1(2ω)2) | (42) |
where Σ′ means excluding the origin in the sum. The series is absolutely and uniformly convergent, once this is secured the double periodicity is immediate since one can shift the summation.
The function ℘ satisfies the well-known differential equation
(℘′)2=4℘3−g2℘−g3. | (43) |
This function is even with a double pole with zero residue at u=0+2Λ. This fact plus Liouville's theorem show that the equation ℘(u)=c has two roots (in one fundamental region). In particular, the three half periods ω1,ω2,ω3=ω1+ω2 are the only points mod 2Λ where ℘′=0, if we denote
ei=℘(ωi) ,c1(ei)=e1+e2+e3=0, c2(ei)=e1e2+e2e3+e3e1=−g24, c3(ei)=e1e2e3=g34 | (44) |
then the equation ℘(u)−ei=0 has a double root at u=ωi.
From (36), the value of g2,g3 corresponding to our lattice is
g2=112−2c3,g3=c36−1216−c23,c3∈[0,1/27]. | (45) |
The discriminant of the cubic 4x3−g2x−g3 is
Δ=g32−27g23=c33(1−27c3). |
One can compute the modular invariant
j(τ)=1728g32Δ=(−1+24c3)3c33(−1+27c3)∈[1728,∞) | (46) |
with the lowest value assumed at c3=(1/12−√3/36).
If the j-invariant is between (1,∞), then the modular parameter τ corresponds to a square lattice (see Figure 5), which turns degenerate when j(τ)→+∞. For us this happens at c3=0 or c3=1/27, i.e. either yi=0 (at the ith face of the polygon) or y1,2,3=1/3 at the very centre.
We assume from now on a square lattice ω1∈R>0 and ω2∈iR>0. In such case the real locus of ℘ is along the lines Rez=0,ω1 or Imz=0,−iω2 mod 2Λ, which is easy to see from the summation (42).
Lemma A.1. (Section 19 [7]) Along the segment [0,ω1] along the real axis, the function ℘ goes from +∞ to e1 monotonically, while along the segment [0,ω2] along the imaginary axis, the function ℘ goes from −∞ to e2 monotonically. Further
e1>e3>e2. |
Proof. As ℘′=0 only at ω1,ω2,ω3=ω1+ω2 mod 2Λ, the monotonicity will be clear once we figure out the relative sizes of e1,2,3.
From the pole u−2 in (42), one knows that going from 0 to ω1 along the x-axis ℘ decreases from +∞ to e1. Similarly going from 0 to ω2 along the y-axis ℘ increases from −∞ to e2.
Suppose e2>e1, then there is a value c∈[e1,e2] where ℘(u)=c has one root u1∈[0,ω1] and another u2∈[0,ω2]. These are the only solutions. But u1+u2 should be 0mod(2Λ) by Theorem 1.5 [7], but this is impossible since u1 is real and u2 is imaginary. Thus e1>e2.
By considering a path from ω1 to ω3 then to ω2, if, say, e3>e1>e2, then again there is c∈[e1,e3] and (only) two roots to ℘=c one in [ω1,ω3], and another in [ω2,ω3]. This is again impossible using the same argument.
We will need the value of ℘ at tertiary periods.
Lemma A.2.
℘(23ω1)=112,℘′(23ω1)=−c3. |
Proof. The four values
℘(23ωi),i=1,2,3,4,ω4=ω1−ω2 |
are the four roots of
x4−12g2x2−g3x−148g22=0. | (47) |
For the proof see 15.1 [7]. This is true for arbitrary g2,g3, but if we plug in the concrete expression (45), one gets
(x−112)((x+136)3+(c3−127)(x+136)+c23−c39+236)=0 . |
Thus one of ℘(2/3ωi) must be the root 1/12. But ℘(2/3ωi) is not real for i=3,4, so we have either ℘(2/3ω1)=1/12 with 1/12>e1>e3>e2, or ℘(2/3ω2)=1/12 with e1>e3>e2>1/12. But the last case would force e1+e2+e3>1/4, but the sum should be zero. So we have ℘(2/3ω1)=1/12. Note the case e1>1/12>e2 is impossible, since this would mean that ℘=1/12 must have a two roots [ω2,ω2+2ω1] or [ω1,ω1+2ω2], but 2/3ω1,2 are not in these segments.
As for the statement about the derivative, we use (43) to get ℘′(2ω1/3))2=c23. We need to take the minus sign from Lemma A.1.
Lemma A.3. Letting t∞=2ω1/3, we have
℘(t∞/2)=136c3(27c3+6e1−72c3e1+72e21−1),℘(ω2+t∞)=136c3(27c3+6e2−72c3e2+72e22−1),℘′(t∞/2)=−(3c3+℘(t∞/2)−112),℘′(ω2+t∞)=3c3+℘(ω2+t∞)−112. |
Proof. Apply the addition formula to t∞+(−ω1)
℘(t∞/2)=℘(t∞−ω1)=14(℘′(t∞))2(℘(t∞)−e1)2−℘(t∞)−e1. |
As everything on the rhs is known, we get
℘(t∞/2)=14c23(1/12−e1)2−1/12−e1. | (48) |
To simplify (48), we note that g2,3 are real, the ei's are algebraic over R. By some elementary field theory, for any polynomial f(ei), one can express 1/f(ei) as a polynomial of ei. For example
11/12−ei=4c23(e2i+ei/12+c3/2−1/72). |
Applying this to (48) we get
℘(t∞/2)=136c3(27c3+6e1−72c3e1+72e21−1). |
As for the derivative, one can use directly (34), (35). Since ℘(t∞/2)=1/12−c3/y1(−t∞/2) and so ℘′(t∞/2)=c3(y2−y3)/y1|t=−t∞/2. But at −t∞/2, one has y1=y3 and y2=1−2y1
℘′(t∞/2)=c3y1(1−3y1)=c3y1−3c3=112−℘(t∞/2)−3c3. |
The calculation for ℘(ω2+t∞) unfolds by using the addition formula on ω2+t∞, while for ℘′(ω2+t∞) we have ℘′(ω2+t∞)=c3(y2−y3)/y1|t=ω2. At ω2, one has y1=y2 and y3=1−2y1
℘′(ω2+t∞)=c3y1(3y1−1)=−c3y1+3c3=−112+℘(ω2+t∞)+3c3. |
Lemma A.4. Assuming (45), then
(1/12−℘(t∞/2))2(1/12−℘(ω1))=c23. |
Proof. This follows from the explicit values of ℘(t∞/2) from Lemma A.3 and a direct calculation.
The ζ function is defined as
ζ(z)=1z+∫z0(1z2−℘(z))dz,℘′(z)=−ζ(z). | (49) |
This definition is independent of the choice of the integral path, since ℘ has only double pole with zero residue at z=0mod2Λ.
Unlike ℘, the function ζ is an odd function, and it is not periodic. One has instead (see Theorem 6.1 [7])
ζ(z+2ω1)=ζ(z)+2η1,ζ(z+2ω2)=ζ(z)+2η2,η1ω2−η2ω1=iπ/2. | (50) |
Lemma A.5. We have the value of ζ at some special points
ζ(t∞)=16+23η1,ζ(2t∞)=−16+43η1ζ(ω2+t∞)=−1c3(2e22+e2/6+c3−1/36)+η2+16+23η1, |
where η1,2=ζ(ω1,2).
Proof. To compute ζ(t∞), we use the duplication formula
ζ(2t∞)−2ζ(t∞)=12℘″(t∞)℘′(t∞)=−12 |
where we computed ℘″ by differentiating (℘′)2=4℘3−g2℘−g3:
2℘′℘″=12℘2℘′−g2℘′⇒℘″=6℘2−12g2. |
But since 2t∞=2ω1−t∞, one gets ζ(2t∞)=2η1+ζ(−t∞)=2η1−ζ(t∞) as ζ is odd. Thus (2η1−ζ(t∞))−2ζ(t∞)=−1/2.
For the last statement, use again the addition formula
ζ(ω2+t∞)=12c3e2−1/12+η2+ζ(t∞)=−2c3(e22+e2/12+c3/2−1/72)+η2+16+23η1 . |
Definition A.6. (54.1 [7]) Define a new function
ζ∗(z)=ζ(z)−η1ω1z. |
This is odd and periodic along the real axis.
Definition A.7. The function
ς(z)=ζ(ω2+zω1π)−η2−zπη1. |
is real of period 2π for real z. It satisfies the shift relations
ς(z+2π3)=ς(z)+y2(t)−13,ς(z−2π3)=ς(z)−y1(t)+13,t:=ω2+zω1π. |
Proof. For the reality of ς at real z, we note that η1 is real and so we need only check the reality for the first two terms in the expression of ς(z), which can be computed from the addition formula
ζ(zω1π)=12℘′(ω2+zω1π)+℘′(ω2)℘(ω2+zω1π)−℘(ω2)+ζ(ω2+zω1π)−ζ(ω2). | (51) |
The lhs and the first term on the rhs of (51) are real for real z, showing that the remainder of the rhs are also real for real z, as desired.
The periodicity follows from that of ζ∗ defined earlier. For the shift relations, we only show one of them,
ζ(t+t∞)=12℘′(t)+c3℘(t)−1/12+ζ(t)+ζ(t∞)=−12(y1−y2+y3)+ζ(t)+ζ(t∞)⇒ς(z+2π3)−ς(z)=y2(t)−13. |
Since ζ has a simple pole at 0mod2Λ, if one integrates ζ, one gets a branching behaviour similar to that of a log. Indeed ∫ζ is the logarithm of an analytic function
σ(z)=zexp∫z0(ζ(z)−1z)dz. | (52) |
It follows trivially that
σ′(z)σ(z)=ζ(z). | (53) |
Clearly σ is not periodic
σ(z+pω1+qω2)σ(z)=(−1)pq+p+qe2(pη1+qη2)(z+pω1+qω2). |
In fact, from this non-periodicity, one can relate σ function to the more familiar theta functions according to 18.10.8 [2]
σ(z)=2ω1πexp(η1z22ω1)ϑ1(πz/2ω1)ϑ′1(0) ,ϑ1(z)=ϑ1(z;τ)=2q1/4∑n≥0(−1)nqn(n+1)sin(2n+1)z,q=eiπτ,τ=ω2ω1,Δ>0. | (54) |
The ϑ1 function is the solution to the heat equation with τ serving as the time, and so one has the important property
−iπ∂2ϑ1(z;τ)∂z2=4∂ϑ1(z;τ)∂τ |
which helps us to compute the derivation of σ function with respect to the modular parameter τ, and also g2,g3.
From the relation between σ and ϑ1 plus the heat equation satisfied by ϑ1, one can derive
Lemma A.8. (18.6.23-18.6.24 [2]) The derivative of σ with respect to g2,3 reads
∂logσ∂g2=116(g32−27g23)(4g22(zζ−1)+36g3(℘−ζ2)−3g2g3z2),∂logσ∂g3=18(g32−27g23)(z2g22+12ζ2g2−12℘g2−36g3(zζ−1)). |
The same expression for ℘,ζ can be obtained by taking derivatives and using (49) and (52).
In our situation g2,3 depend only on c3, and ∂c3=(−2∂g2+(1/6−2c3)∂g3). We record for the readers convenience
∂c3logσ(z)=−c2396γ(144℘+(24c3−1)z2−144ζ2+(864c3−24)(ζz−1)) , ∂c3℘(z)=c2312γ(6℘(1−36c3)+3℘′(12ζ+z−36c3z)+72℘2+24c3−1),∂c3ζ(z)=−c2348γ(72℘′+(ζ−℘z)(432c3−12)+144℘ζ+(24c3−1)z) | (55) |
where γ=g32−27g23=c33(1−27c3).
Lemma A.9. We have
∂c3[ηiωi]=1c3(1−27c3)[14(1−36c3)148(1−24c3)−3−14(1−36c3)][ηiωi],i=1,2 | (56) |
Note the matrix above is traceless and so it will preserve the condition (50).
Proof. For ω1, since t∞=2/3ω1, we can differentiate ℘(t∞)=1/12 and get ∂c3℘|t=t∞+℘′(t∞)∂c3t∞=0, then we get from (55)
∂c3t∞=1c3∂c3℘|t=t∞=c312(g32−27g23)(12(1−36c3)−3c3(12ζ(t∞)+t∞−36c3t∞)+12+24c3−1) |
and we get the result.
For η1, one can differentiate ζ(t∞)=1/6+2/3η1, while for ω2 and η one differentiates ℘(ω2)=e2 and the second equation of Lemma A.5.
One can define the combination
˜ηi=ηi+ωi(112−3c3), | (57) |
then (56) reads
∂c3˜ηi=−2ωi,∂c3ωi=−3˜ηic3(1−27c3). | (58) |
In particular ω1 satisfies ∂c3(c3(1−27c3)∂c3ω1)=6ω1 and the solution is given by the hyper-geometric functions
ω1∼F(13;23;1;27c3). |
We will not dig further into the hyper-geometric functions here, but the solution to (56) bears heavy resemblance to the torus fibration picture of the Seiberg-Witten solution to the N=2 super-symmetric gauge theory [21] in that (50) works as the Dirac quantisation condition for the electric and magnetic charge and both quantities are solved by the hyper-geometric functions. We list however the values of ˜η1,2 and ω1,2 at c3=0,1/27 for completeness
c3=0,ω1=∞,ω2=iπ,˜η1=12,η2=−iπ12, | (59) |
c3=127,ω1=√3π,ω2=i∞,η1=√3π36,˜η2=−i√36. | (60) |
If we write c3=ϵ then ω1∼ϵ−1/2 in the first line, or if we write 1/27−c3=ϵ then ω2∼iϵ−1/4 for the second line.
Finally we compute the c3 derivative of the functions yi. We have demonstrated in the main text that, in order to exploit the symmetry between yi, we should use the c3 and t coordinate to parametrise the contour in Figure 4. In particular we introduce the time coordinate s
t=ω2+sω1π, | (61) |
so that s∈R and is of period 2π regardless of c3.
Lemma A.10. Holding s in (61) fixed, and differentiate with respect to c3
∂c3yi=3(c3−27c23)(16((yi)2−yi)−3c3yi+2c3+yi(yi+1−yi+2)ς(s+2π3(i−3))) |
where the index i is taken mod 3 and ς(s) is given in Definition A.7.
By using the shift relation in Definition A.7, one can check that ∂c3(y1+y2+y3)=0 and ∂c3(y1y2y3)=1.
Proof. We hold s fixed when taking the partial derivative with respect to c3.
∂c3(y3(t))=∂c3c31/12−℘(t)+˙y3(t)∂c3t=y3(t)c3+c3(1/12−℘)2∂c3℘|t=ω2+sω1/π+y3(y1−y2)∂c3t |
One plugs in (55) for ∂c3℘ and (56) for ∂c3ωi,
![]() |
We replace the last term with the function ς given in Definition A.7. The computation for y1,2 are similar since they are obtained from y3 by shifting s→s±2/3π.
The Jacobian can be computed as
∂(c3,s)∂(y1,y2)=πω1. |
Finally we perform an amusing calculation of the area of the triangle Δ={y1,2∈[0,1],y1+y2≤1}
∫Δdy1∧dy2=∫Δω1πdc3∧ds=2∫1/270ω1dc3(58)=−˜η1|1/270=12. |
This is admittedly a pretentious way of doing some grade school level maths, though a reassuring check of the extensive computations involving elliptic functions.
We thank Maxim Zabzine for help over many issues in the draft. Our paper has some overlap with the work of Francis Bischoff and Marco Gualtieri, and we thank the former for sharing with us his thesis [4] and both for many interesting discussions. The work of J. Q is supported in part by the grant "Geometry and Physics" from the Knut and Alice Wallenberg foundation.
All authors declare no conflicts of interest in this paper.
[1] |
J. Halliday, S. A. Rutherford, M. G. McCabe, D. G. Evans, An update on the diagnosis and treatment of vestibular schwannoma, Expert Rev. Neurother., 18 (2018), 29–39. https://doi.org/10.1080/14737175.2018.1399795 doi: 10.1080/14737175.2018.1399795
![]() |
[2] |
D. Starnoni, R. T. Daniel, C. Tuleasca, M. George, M. Levivier, M. Messerer, Systematic review and meta-analysis of the technique of subtotal resection and stereotactic radiosurgery for large vestibular schwannomas: a "nerve-centered" approach, Neurosurg. Focus, 44 (2018), E4. https://doi.org/10.3171/2017.12.FOCUS17669 doi: 10.3171/2017.12.FOCUS17669
![]() |
[3] |
B. Acs, M. Rantalainen, J. Hartman, Artificial intelligence as the next step towards precision pathology, J. Intern. Med., 288 (2020), 62–81. https://doi.org/10.1111/joim.13030 doi: 10.1111/joim.13030
![]() |
[4] |
J. Goecks, V. Jalili, L. Heiser, J. Gray, How machine learning will transform biomedicine, Cell, 181 (2020), 92–101. https://doi.org/10.1016/j.cell.2020.03.022 doi: 10.1016/j.cell.2020.03.022
![]() |
[5] |
G. S. Handelman, H. K. Kok, R. V. Chandra, A. H. Razavi, M. J. Lee, H. Asadi, eDoctor: machine learning and the future of medicine, J. Intern. Med., 284 (2018), 603–609. https://doi.org/10.1111/joim.12822 doi: 10.1111/joim.12822
![]() |
[6] | M. M. Hasan, M. A. Alam, W. Shoombuatong, H. W. Deng, B. Manavalan, H. Kurata, NeuroPred-FRL: an interpretable prediction model for identifying neuropeptide using feature representation learning, Briefings Bioinf., 22 (2021). https://doi.org/10.1093/bib/bbab167 |
[7] |
A. Hoshino, H. S. Kim, L. Bojmar, K. E. Gyan, M. Cioffi, J. Hernandez, et al., Extracellular vesicle and particle biomarkers define multiple human cancers, Cell, 18 (2020), 1044–1061. https://doi.org/10.1016/j.cell.2020.07.009 doi: 10.1016/j.cell.2020.07.009
![]() |
[8] | B. Koo, J. K. Rhee, Prediction of tumor purity from gene expression data using machine learning, Briefings Bioinf., 22 (2021). https://doi.org/10.1093/bib/bbab163 |
[9] | H. Luo, Q. Zhao, W. Wei, L. Zheng, S. Yi, G. Li, et al., Circulating tumor DNA methylation profiles enable early diagnosis, prognosis prediction, and screening for colorectal cancer, Sci. Transl. Med., 12 (2020). https://doi.org/10.1126/scitranslmed.aax7533 |
[10] |
L. Huang, L. Wang, X. Hu, S. Chen, Y. Tao, H. Su, et al., Machine learning of serum metabolic patterns encodes early-stage lung adenocarcinoma, Nat. Commun., 11 (2020), 3556. https://doi.org/10.1038/s41467-020-17347-6 doi: 10.1038/s41467-020-17347-6
![]() |
[11] |
M. K. Abd Ghani, M. A. Mohammed, N. Arunkumar, S. A. Mostafa, D. A. Ibrahim, M. K. Abdullah, et al., Decision-level fusion scheme for nasopharyngeal carcinoma identification using machine learning techniques, Neural Comput. Appl., 32 (2020), 625–638. https://doi.org/10.1007/s00521-018-3882-6 doi: 10.1007/s00521-018-3882-6
![]() |
[12] | P. Achilli, C. Magistro, M. A. A. E. Aziz, G. Calini, C. L. Bertoglio, G. Ferrari, et al., Modest agreement between magnetic resonance and pathological tumor regression after neoadjuvant therapy for rectal cancer in the real world, Int. J. Cancer, (2022), 1–8. https://doi.org/10.1002/ijc.33975 |
[13] | Z. M. Zhuang, Z. B. Yang, S. X. Zhuang, A. N. J. Raj, Y. Yuan, R. Nersisson, Multi-features-based automated breast tumor diagnosis using ultrasound image and support vector machine, Comput. Intell. Neurosci., 2021 (2021). https://doi.org/10.1155/2021/9980326 |
[14] | M. M. Ghiasi, S. Zendehboudi, Application of decision tree-based ensemble learning in the classification of breast cancer, Comput. Biol. Med., 128 (2021). https://doi.org/10.1016/j.compbiomed.2020.104089 |
[15] | A. Moncada-Torres, M. C. van Maaren, M. P. Hendriks, S. Siesling, G. Geleijnse, Explainable machine learning can outperform Cox regression predictions and provide insights in breast cancer survival, Sci. Rep., 11 (2021). https://doi.org/10.1038/s41598-021-86327-7 |
[16] | C. Yang, X. W. Huang, Y. Li, J. F. Chen, Y. Y. Lv, S. X. Dai, et al., Prognosis and personalized treatment prediction in TP53-mutant hepatocellular carcinoma: an in silico strategy towards precision oncology, Briefings Bioinf., 22 (2021). https://doi.org/10.1093/bib/bbaa164 |
[17] | Y. Q. Wu, N. Jiao, R. X. Zhu, Y. D. Zhang, D. F. Wu, A. J. Wang, et al., Identification of microbial markers across populations in early detection of colorectal cancer, Nat. Commun., 12 (2021). https://doi.org/10.1038/s41467-021-23265-y |
[18] | L. Zhang, H. X. Ai, W. Chen, Z. M. Yin, H. Hu, J. F. Zhu, et al., CarcinoPred-EL: Novel models for predicting the carcinogenicity of chemicals using molecular fingerprints and ensemble learning methods, Sci. Rep., 7 (2017). https://doi.org/10.1038/s41598-017-02365-0 |
[19] |
A. Tahmassebi, G. J. Wengert, T. H. Helbich, Z. Bago-Horvath, S. Alaei, R. Bartsch, et al., Impact of machine learning with multiparametric magnetic resonance imaging of the breast for early prediction of response to neoadjuvant chemotherapy and survival outcomes in breast cancer patients, Invest. Radiol., 54 (2019), 110–117. https://doi.org/10.1097/RLI.0000000000000518 doi: 10.1097/RLI.0000000000000518
![]() |
[20] | J. Li, Z. Shi, F. Liu, X. Fang, K. Cao, Y. H. Meng, et al., XGBoost classifier based on computed tomography radiomics for prediction of tumor-infiltrating CD8+T-Cells in patients with pancreatic ductal adenocarcinoma, Front. Oncol., 11 (2021). https://doi.org/10.3389/fonc.2021.671333 |
[21] |
K. Thedinga, R. Herwig, A gradient tree boosting and network propagation derived pan-cancer survival network of the tumor microenvironment, iScience, 25 (2021), 103617. https://doi.org/10.1016/j.isci.2021.103617 doi: 10.1016/j.isci.2021.103617
![]() |
[22] |
W. Tang, H. Zhou, T. H. Quan, X. Y. Chen, H. N. Zhang, Y. Lin, et al., XGboost prediction model based on 3.0T diffusion kurtosis imaging improves the diagnostic accuracy of MRI BiRADS 4 masses, Front. Oncol., 12 (2022), 833680. https://doi.org/10.3389/fonc.2022.833680 doi: 10.3389/fonc.2022.833680
![]() |
[23] |
B. Fu, P. Liu, J. Lin, L. Deng, K. Hu, H. Zheng, Predicting invasive disease-free survival for early stage breast cancer patients using follow-up clinical data, IEEE Trans. Biomed. Eng., 66 (2019), 2053–2064. https://doi.org/10.1109/TBME.2018.2882867 doi: 10.1109/TBME.2018.2882867
![]() |
[24] |
S. L. Li, X. J. Zhang, Research on orthopedic auxiliary classification and prediction model based on XGBoost algorithm, Neural Comput. Appl., 32 (2020), 1971–1979. https://doi.org/10.1007/s00521-019-04378-4 doi: 10.1007/s00521-019-04378-4
![]() |
[25] |
Y. M. Hsiao, C. L. Tao, E. Y. Chuang, T. P. Lu, A risk prediction model of gene signatures in ovarian cancer through bagging of GA-XGBoost models, J. Adv. Res., 30 (2021), 113–122. https://doi.org/10.1016/j.jare.2020.11.006 doi: 10.1016/j.jare.2020.11.006
![]() |
[26] |
B. Krawczyk, Learning from imbalanced data: open challenges and future directions, Prog. Artif. Intell., 5 (2016), 221–232. https://doi.org/10.1007/s13748-016-0094-0 doi: 10.1007/s13748-016-0094-0
![]() |
[27] | S. Fotouhi, S. Asadi, M. W. Kattan, A comprehensive data level analysis for cancer diagnosis on imbalanced data, J. Biomed. Inf., 90 (2019). https://doi.org/10.1016/j.jbi.2018.12.003 |
[28] |
N. V. Chawla, K. W. Bowyer, L. O. Hall, W. P. Kegelmeyer, SMOTE: synthetic minority over-sampling technique, J. Artif. Intell. Res., 6 (2001), 321–357. https://doi.org/10.1613/jair.953 doi: 10.1613/jair.953
![]() |
[29] |
G. Batista, R. C. Prati, M. C. Monard, A study of the behavior of several methods for balancing machine learning training data, ACM SIGKDD Explor. Newsl., 6 (2004), 20–29. https://doi.org/10.1145/1007730.1007735 doi: 10.1145/1007730.1007735
![]() |
[30] | S. Fotouhi, S. Asadi, M. W. Kattan, A comprehensive data level analysis for cancer diagnosis on imbalanced data, J. Biomed. Inf., 90 (2019). https://doi.org/10.1016/j.jbi.2018.12.003 |
[31] | X. Huang, T. Y. Cao, L. Z. Q. Chen, J. P. Li, Z. H. Tan, B. J. M. Xu, et al., Novel insights on establishing machine learning-based stroke prediction models among hypertensive adults, Front. Cardiovasc. Med., 9 (2022). https://doi.org/10.3389/fcvm.2022.901240 |
[32] | L. Breiman, Random forest, Mach. Learn., 45 (2001), 5–32. https://doi.org/10.1023/A:1010933404324 |
[33] | T. Q. Chen, C. Guestrin, XGBoost: a scalable tree boosting system, in Proceedings of the 22nd ACM SIGKDD International Conference on Knowledge Discovery and Data Mining, (2016), 785–794. https://doi.org/10.1145/2939672.2939785 |
[34] |
M. Sokolova, G. Lapalme, TA systematic analysis of performance measures for classification tasks, Inf. Process. Manage., 45 (2009), 427–437. https://doi.org/10.1016/j.ipm.2009.03.002 doi: 10.1016/j.ipm.2009.03.002
![]() |
[35] |
A. P. Bradley, The use of the area under the ROC curve in evaluation of machine learning algorithms, Pattern Recognit., 30 (1997), 1145–1159. https://doi.org/10.1016/S0031-3203(96)00142-2 doi: 10.1016/S0031-3203(96)00142-2
![]() |
1. | Jian Qiu, Quantisation via Branes and Minimal Resolution, 2024, 405, 0010-3616, 10.1007/s00220-024-05137-w |