Sign up with your email address to be the first to know about new products, VIP offers, blog features & more.

The Gel’fand-Kapranov-Zelevinski HGS, the D-Brane Superpotential, and the Witten Mirror

The D-brane superpotential, a section of special holomorphic line bundles of the complex and Kähler product moduli space, and hence the correlation-function generator, is pivotal for at least three central reasons in the construction of Calabi-Yau compactifications: first, for allowing M-theory compactifications on Kovalev twisted connected sum {{G}_{2}}-manifolds. Second, for giving rise to Yukawa couplings, and consequently, allows us to determine the vacuum of the low energy N=1 effective theory. And third, to allow for topological CY properties such as Jones-Witten and Ooguri–Vafa invariants. In this post, we will analyze the D-brane superpotential for pseudo-Fermat Calabi-Yau manifolds via mirror symmetry and the Gel’fand-Kapranov-Zelevinski hypergeometric system in the context of Saito’s differential equations of deformation theory of singularities and derive an Aganagic-Klemm-Vafa type mirror symmetry based on the Witten equation. First, let us set the stage. Dp-brane solutions preserving half supersymmetry have general form:

\displaystyle ds_{{Dp}}^{2}={{\Omega }^{{-1}}}\left[ {d{{t}^{2}}-ds_{p}^{2}} \right]-\Omega dx_{{5-p}}^{2}

\displaystyle {{e}^{{2\phi }}}={{\Omega }^{{1-p}}}

\displaystyle F_{{01...pm}}^{A}={{\partial }_{m}}{{H}^{A}}

\displaystyle {{M}_{{Kah}}}^{{AB}}={{\text{I}}_{{\text{Re}{{\text{p}}_{G}}}}}^{{AB}}+2{{\Omega }^{{-1}}}{{H}^{A}}{{H}^{B}}

where the Hamiltonian metaplectic action in the Heisenberg representation on the Dp+1 dimensional worldvolume gives us:

    \[H = {\dot \psi ^{2\pi ik}}{\Im _i} + V_t^{p + 1}\not K + \oint_{p + i} {\delta _k^{{\rm{susy}}}} \left| {_{{B_{{\rm{Bos}}}}}} \right.d\,\Omega {({\phi _{si}})^{p + 1}}{\not H_i} + \lambda {\not H^i}\]

where:

    \[\left\{ {\begin{array}{*{20}{c}}{{\Im _i} = - \not \partial {\phi _{si}}{T_{Dp}}d\,\Omega {{({\phi _{si}})}^{2\pi ik}}}\\{\not K = - {{\not \partial }_i}{{\widetilde E}^i} + {{( - 1)}^{p + 1}}{T_{Dp}}{S^{{\rm{Fer}}}}}\\{{{\not H}_i} = \widetilde P{\alpha _i}\widetilde E_i^\alpha {{\not \partial }_i}{\phi _{si}} + \widetilde E{{\not F}_{ij}}}\\{H = \frac{1}{{2\pi ik}}\left[ {{{\widetilde P}^2} + {{\widetilde E}^i}{{\widetilde E}^j}{G_{ij}} + T_{Dp}^2{e^{ - 2{\phi _{si}}}}{\rm{det}}\left( {{G_{ij}} + {{\not F}_{ij}}} \right)} \right]}\end{array}} \right.\]

with:

    \[S = * {\left( {{{\not R}_{\mu \nu }}{\varepsilon ^{{\rm{Fer}}}}} \right)_p}\]

and:

    \[E_i^\alpha = \delta \int {d\not E_m^\alpha } {\not \partial _i}{\dot X^m}\]

where the Ramond-Ramond gauge-coupling sector is given by the action:

\displaystyle \mathcal{L}_{G}^{{Loc}}=\sum\limits_{{b=1}}^{{N-1}}{{\frac{1}{{2g_{b}^{2}}}}}{{\int{\text{d}}}^{2}}\theta {{W}^{\alpha }}{{W}_{\alpha }}{{\delta }^{2}}\left( {\left( {1-{{e}^{{ib\phi }}}} \right)z} \right)

and generally, the action of a Dp-brane is given by a Dirac-Born-Infeld part coupled to a Chern-Simons WZ part:

\displaystyle S_{{DBI}}^{{cs}}=-{{T}_{p}}\int_{{{\mathcal{W}}'}}{{{{d}^{{p+1}}}}}\xi {{e}^{{-\Phi }}}\Xi -{{T}_{p}}\int_{{{\mathcal{W}}'}}{{C_{F}^{B}}}

with:

\displaystyle \Xi \doteq \sqrt{{{{{\det }}_{{\left[ {a,b} \right]}}}\left( {{{g}_{{ab}}}+{{B}_{b}}+2\pi {\alpha }'{{F}_{{ab}}}} \right)}}

\displaystyle C_{F}^{B}\doteq \text{TrP}\left[ {C\wedge {{e}^{{-B}}}} \right]\wedge {{e}^{{2\pi {\alpha }'F}}}

where P is the worldvolume pullback with p-orientifold action:

\displaystyle {{S}_{{Op}}}={{2}^{{p-4}}}-{{T}_{p}}\int\limits_{{{{\mathcal{W}}_{\parallel }}}}{{{{d}^{{p+1}}}}}\xi {{e}^{{-\phi }}}\left( \Pi \right)-\Psi

with:

\displaystyle \Pi \doteq \sqrt{{-\det P\left[ {{{g}_{{\mu \nu }}}} \right]}}-{{2}^{{2-4}}}

and

\displaystyle \Psi \doteq -{{T}_{p}}\int_{{{{\mathcal{W}}_{\parallel }}}}{{P\left[ {{{C}_{{p+1}}}} \right]}}

where the pullback to the Dp-worldvolume yields the 10-D SYM action:

\displaystyle {{S}_{{YM}}}=\frac{1}{{4g_{{_{{YM}}}}^{2}}}\int{{{{d}^{{10}}}}}x\left[ {\text{Tr}\left( {{{F}_{{\mu \nu }}}{{F}^{{\mu \nu }}}} \right)+2\text{iTr}\left( {\bar{\psi }\,{{\Gamma }^{\mu }}{{D}_{\mu }}\psi } \right)} \right]

with string coupling:

\displaystyle \frac{1}{{{{g}_{s}}}}={{e}^{{-\Phi }}}

and the 10-D SUGRA dimensionally reduced Type-IIB action is:

\displaystyle S_{{DBI}}^{{c{s}'}}=-\frac{{{{T}_{p}}{{{\left( {2\pi {\alpha }'} \right)}}^{2}}}}{{4{{g}_{s}}}}\int_{{{\mathcal{W}}'}}{{{{d}^{{p+1}}}\left( {\hat{F}+\hat{X}} \right)-\frac{{{{T}_{p}}}}{{{{g}_{s}}}}{{V}_{\vartheta }}}}

with:

\displaystyle \hat{F}\equiv {{F}_{{ab}}}{{F}^{{ab}}}

\displaystyle \hat{X}\equiv \frac{2}{{{{{\left( {2\pi {\alpha }'} \right)}}^{2}}}}{{\partial }_{a}}{{X}^{m}}{{\partial }^{a}}{{X}_{m}}

\displaystyle {{V}_{\vartheta }}\equiv V_{{p+1}}^{{WV}}+\vartheta \left( {{{F}^{4}}} \right)

and in the string-frame, the type-IIB SUGRA action is given by:

\displaystyle {{S}_{{IIB}}}={{S}_{{NS}}}+{{S}_{R}}+{{S}_{{CS}}}

with:

\displaystyle {{S}_{{NS}}}=\frac{1}{{2k_{{10}}^{2}}}\int{{{{d}^{{10}}}}}x{{\left( {-{{G}_{{10}}}} \right)}^{{1/2}}}{{e}^{{-2\Phi }}}\left[ {{{R}_{{10}}}+4\left( {{{\partial }^{\mu }}\Phi } \right)\left( {{{\partial }_{\mu }}\Phi } \right)-\frac{1}{2}{{{\left| {{{H}_{{\left( 3 \right)}}}} \right|}}^{2}}} \right]

\displaystyle {{S}_{R}}=-\frac{1}{{4k_{{10}}^{2}}}\int{{{{d}^{{10}}}}}x{{\left( {-{{G}_{{10}}}} \right)}^{{1/2}}}\left[ {{{{\left| {{{F}_{1}}} \right|}}^{2}}+{{{\left| {{{{\tilde{F}}}_{{\left( 3 \right)}}}} \right|}}^{2}}+\frac{1}{2}{{{\left| {{{{\tilde{F}}}_{{\left( 5 \right)}}}} \right|}}^{2}}} \right]

\displaystyle {{S}_{{CS}}}=-\frac{1}{{4k_{{10}}^{2}}}\int{{{{C}_{{\left( 4 \right)}}}\wedge {{H}_{{\left( 3 \right)}}}\wedge {{F}_{{\left( 3 \right)}}}}}

where the Calabi-Yau superpotential is:

\displaystyle W=\int_{Y}{{{{G}_{3}}}}\wedge {{\Omega }_{3}}+\sum\limits_{{i=1}}^{{{{h}^{{1,1}}}}}{{{{A}_{i}}}}\left( {S,U} \right){{e}^{{-{{a}_{i}}{{T}_{i}}}}}

where:

\displaystyle \int_{Y}{{{{G}_{3}}}}\wedge {{\Omega }_{3}}

is the Gukov-Vafa-Witten superpotential stabilization complex term, as well as the axio-dilaton field:

\displaystyle S={{e}^{{-\phi }}}+i{{C}_{0}}

Given the presence of E3-brane instantons, {{T}_{i}} are of Kähler moduli Type-IIB-orbifold class:

\displaystyle {{T}_{i}}={{e}^{{-\phi }}}{{\tau }_{1}}+i{{\rho }_{i}}

with {{\tau }_{i}} being the volume of the divisor {{D}_{i}} and {{\rho }_{i}} the 4-form Ramond-Ramond axion field corresponding to:

\displaystyle {{\tau }_{i}}=\frac{1}{2}\int_{{{{D}_{i}}}}{{J\wedge J}}=\frac{1}{2}{{k}_{{ijk}}}+{{\,}^{j}}{{t}^{k}}

and:

\displaystyle {{\rho }_{i}}=\int_{{{{D}_{i}}}}{{{{C}_{4}}}}

where J is the Kähler form:

\displaystyle J=\sum\limits_{i}{{{{t}_{i}}}}{{\eta }_{i}}

and:

\displaystyle \left\{ {{{\eta }_{i}}} \right\}\in {{H}^{{1,1}}}\left( {Y,\mathbb{Z}} \right)

an integral-form basis and {{k}_{{ijk}}} the associated intersection coefficients. Hence, the Kähler potential is given by:

\displaystyle K=-2\text{In}\left( {\tilde{\mathcal{V}}+\frac{\xi }{{2g_{s}^{{3/2}}}}} \right)-\text{In}\left( {S+\tilde{S}} \right)-\text{In}\left( {-i\int_{Y}{{\Omega \wedge \tilde{\Omega }}}} \right)

with {\tilde{\mathcal{V}}} the Calabi-Yau volume, and in the Einstein frame, is given by:

\displaystyle \mathcal{V}=\frac{1}{{3!}}\int_{Y}{{J\wedge J\wedge J=}}\frac{1}{6}{{k}_{{ijk}}}{{t}^{i}}{{t}^{j}}{{t}^{k}}

The F-term is given by:

\displaystyle {{V}_{F}}={{e}^{K}}\left( {\sum\limits_{{i={{T}_{i}};j={{S}_{j}}}}{{{{K}^{{ij}}}}}{{D}_{i}}W{{D}_{j}}\tilde{W}-3{{{\left| {{{W}_{i}}} \right|}}^{2}}} \right)

with the Large Volume Scenario D-term is given by:

\displaystyle {{V}_{D}}=\sum\limits_{{i=1}}^{N}{{\frac{1}{{\operatorname{Re}\left( {{{f}_{i}}} \right)}}}}{{\left( {\sum\limits_{j}{{Q_{j}^{{\left( i \right)}}{{{\left| {{{\phi }_{j}}} \right|}}^{2}}-{{{\hat{\xi }}}_{i}}}}} \right)}^{2}}

with:

\displaystyle \operatorname{Re}\left( {{{f}_{i}}} \right)\doteq {{e}^{{-\phi }}}\frac{1}{2}\int_{{{{D}_{i}}}}{{J\wedge J-}}{{e}^{{-\phi }}}\int_{{{{D}_{i}}}}{{\text{c}{{\text{h}}_{2}}}}\left( {{{\mathcal{L}}_{i}}-B} \right)

and the Fayet-Illopoulos terms being:

\displaystyle {{\hat{\xi }}_{i}}=-\text{Im}\left( {\frac{1}{{\tilde{\mathcal{V}}}}\int_{Y}{{{{e}^{{-\left( {B+iJ} \right)}}}}}{{\Gamma }_{i}}} \right)

where {{\Gamma }_{i}} are the D7-brane charge-vectors. Then we saw that the Chern-Simons orientifold action gets a Calabi-Yau curvature correction in the form of:

\displaystyle {{S}_{{{{O}_{p}},CS}}}=-{{2}^{{p-4}}}-{{T}_{p}}\int\limits_{{{{\mathcal{W}}^{\prime }}}}{{P\left[ C \right]}}\wedge \Theta

with

\displaystyle \Theta \doteq \sqrt{{\frac{{L\left( {\frac{1}{4}{{{\left( {2\pi } \right)}}^{2}}{\alpha }'T{{\mathcal{W}}^{\prime }}} \right)}}{{L\left( {\frac{1}{4}{{{\left( {2\pi } \right)}}^{2}}{\alpha }'N{{\mathcal{W}}^{\prime }}} \right)}}}}

due to the Gauss–Codazzi equations:

\displaystyle {{S}_{{{{O}_{p}}}}}={{2}^{{p-4}}}-{{T}_{p}}{{\int_{{{\mathcal{W}}'}}{\text{d}}}^{{p+1}}}\xi {{e}^{{-\phi }}}\sqrt{{-\det \left( {P\left[ {{{g}_{{\mu \nu }}}} \right]} \right)}}-{{2}^{{2-4}}}-{{T}_{p}}\int_{{{\mathcal{W}}'}}{{P\left[ {{{C}_{{p+1}}}} \right]}}

and the Ramond-Ramond term being:

\displaystyle {{S}_{{CS}}}=\frac{{{{T}_{p}}}}{2}\int\limits_{{{{\Sigma }_{{p+1}}}}}{{C\wedge \text{Tr}}}\left( {{{e}^{{F/2\pi }}}} \right)

which yields the Type-IIB Calabi-Yau three-fold superpotential:

\displaystyle {{V}_{W}}=\int\limits_{X}{{{{G}_{3}}}}\wedge {{\Omega }_{3}}+\sum\limits_{{i=1}}^{{{{h}^{{1,1}}}}}{{{{A}_{i}}}}\left( {\left( {{{e}^{{-\phi }}}+i{{C}_{0}}} \right),U} \right){{e}^{{-a\left( {{{e}^{{-\phi }}}{{\tau }_{i}}+i{{\rho }_{i}}} \right)}}}

and where the topologically mixed Yang-Mills action is given by:

\displaystyle {{\mathcal{L}}_{{TYM}}}\equiv -\frac{1}{4}e{{\tilde{F}}_{{\mu \nu }}}^{M}{{\tilde{F}}^{{\mu \nu N}}}{{\hat{M}}_{{MN}}}+\kappa {{\mathcal{L}}_{{CS}}}

with the corresponding Chern-Simons action:

\displaystyle {{S}_{{CS}}}=\frac{{{{T}_{p}}}}{2}\int\limits_{{{{\Sigma }_{{p+1}}}}}{{C\wedge \text{ch}}}\left( {\tilde{F}} \right)\wedge \sqrt{{\frac{{\hat{A}\left( {{{R}_{T}}} \right)}}{{\hat{A}\left( {{{R}_{N}}} \right)}}}}

Generically, type-II compactifications on Calabi–Yau manifolds give rise to N=1 low energy effective theories given suitable background fluxes and space-filling D-branes wrapping internal special Lagrangian sub-manifolds and holomorphic divisors determined by period integrals satisfying systems of Picard-Fuchs differential equations where the period integrals of the holomorphic three-form whose cycles have boundaries wrapped by D-branes determine the effective superpotential. The associated superpotential of D-branes wrapping internal cycles of Calabi–Yau manifold X is given as such:

\displaystyle {{\mathcal{W}}_{{Brane}}}=\int_{X}{{\Omega \wedge \text{Tr}}}\left[ {A\wedge \bar{\partial }A+\frac{1}{2}A\wedge A\wedge A} \right]

In the Type-II case, it is a linear sum of the period integrals:

\displaystyle {{\mathcal{W}}_{{Brane}}}\left( {\varphi ,\xi } \right)={{{\hat{N}}}_{a}}{{{\hat{\Pi }}}^{a}}\left( {\varphi ,\xi } \right)

with:

\displaystyle {{{\hat{\Pi }}}^{a}}\left( {\varphi ,\xi } \right)=\int_{{{{\gamma }^{a}}\left( \xi \right)}}{{\Omega \left( \varphi \right)}}

The Ramond-Ramond and Neveu-Schwarz background fluxes {{H}^{{RR}}}{{H}^{{NS}}} yield the flux-superpotential:

\displaystyle {{\mathcal{W}}_{{flux}}}=\int_{X}{{\Omega \wedge H=\int_{X}{{\Omega \wedge \left( {{{H}^{{RR}}}+\tau {{H}^{{NS}}}} \right)}}}}

Thus, putting the two superpotentials together gives us:

\displaystyle \mathcal{W}\left( {\varphi ,\xi } \right){{\mathcal{W}}_{{brane}}}\left( {\varphi ,\xi } \right)+{{\mathcal{W}}_{{flux}}}\left( \varphi \right)=\sum{{{{N}_{\Sigma }}}}{{\Pi }_{\Sigma }}\left( {\varphi ,\xi } \right)

where {{N}_{\Sigma }} is the D-brane topological charge and {{\Pi }_{\Sigma }}\left( {\varphi ,\xi } \right) is the period-integral over the CY 3-form:

\displaystyle \left\{ {\begin{array}{*{20}{c}} {{{\Pi }_{\Sigma }}\left( {\varphi ,\xi } \right)=\int_{{{{\Gamma }^{\alpha }}\left( \xi \right)}}{{\Omega \left( \varphi \right)}}} \\ {{{\Gamma }^{\alpha }}\left( \xi \right)\in {{H}_{3}}\left( {Y,S,Z} \right)} \end{array}} \right.

and the domain-wall tension is given by:

\displaystyle \mathcal{T}\left( {\varphi ,\xi } \right)=\mathcal{W}\left( {C_{{\left( {\varphi ,\xi } \right)}}^{+}} \right)-\mathcal{W}\left( {C_{{\left( {\varphi ,\xi } \right)}}^{-}} \right)

Now, in the A-model, the D-brane superpotential is defined in terms of closed/open holomorphic correlation-functional coordinates determined by OPE coefficients corresponding to the worldsheet chiral ring:

\displaystyle \mathcal{W}\left( {t,\hat{t}} \right)=\sum\limits_{{\vec{k},\vec{m}}}{{{{G}_{{\vec{k},\vec{m}}}}}}{{q}^{{\text{d}\vec{k}}}}{{\hat{q}}^{{\text{d}\vec{m}}}}=\sum\limits_{{\vec{k},\vec{m}}}{{\sum\limits_{d}{{{{n}_{{\vec{k},\vec{m}}}}}}}}\frac{{{{q}^{{\text{d}\vec{k}}}}{{{\hat{q}}}^{{\text{d}\vec{m}}}}}}{{{{k}^{2}}}}

where we have:

\displaystyle \left\{ {\begin{array}{*{20}{c}} {q={{e}^{{2\pi \text{i}t}}}} \\ {\hat{q}={{e}^{{2\pi \text{i}\hat{t}}}}} \\ {{{n}_{{\vec{k},\vec{m}}}}\doteq \text{Ooguri-Vafa invariant}} \end{array}} \right.

and the following relations follow for the D-brane superpotential and the Ooguri–Vafa invariants:

\displaystyle \frac{{{{W}^{{\left( \pm \right)}}}\left( {z\left( q \right)} \right)}}{{{{\omega }_{0}}\left( {z\left( q \right)} \right)}}=\frac{1}{{{{{\left( {2\pi i} \right)}}^{2}}}}\sum\limits_{{k\,\in \,\ \text{odd}}}{{\sum\limits_{{{{d}_{1}}\ge 0,{{d}_{2}}\,\in \ \ \text{odd}}}{{n_{{{{d}_{1}},{{d}_{2}}}}^{{\left( \pm \right)}}}}}}\frac{{q_{1}^{{k{{d}_{1}}}}q_{2}^{{k{{d}_{2}}/2}}}}{{{{k}^{2}}}}

and:

\displaystyle {{\omega }_{0}}\left( z \right)=\sum\limits_{n}{{c\left( n \right){{z}^{n}}}}\sum\limits_{n}{{\frac{{\prod\nolimits_{j}{{\Gamma \left( {\sum\nolimits_{a}{{l_{{0j}}^{{\left( a \right)}}{{n}_{a}}+1}}} \right)}}}}{{\prod\nolimits_{i}{{\Gamma \left( {\sum\nolimits_{a}{{l_{i}^{{\left( a \right)}}{{n}_{a}}+1}}} \right)}}}}}}{{z}^{n}}

where the mirror-symmetry map is given by:

\displaystyle {{t}_{a}}=\frac{{{{\partial }_{a}}{{\omega }_{0}}}}{{{{\omega }_{0}}}}

We now define a mirror pair of hypersurfaces \left( {X,{{X}^{*}}} \right) in toric ambient spaces \left( {V,{{V}^{*}}} \right) with corresponding fans \left( {\Sigma \left( \Delta \right),\Sigma \left( {{{\Delta }^{*}}} \right)} \right), where the hypersurface defining polynomial is given as such:

\displaystyle \mathcal{P}=\sum\limits_{{i=0}}^{{p-1}}{{{{a}_{i}}}}\prod\limits_{{k=1}}^{4}{{X_{k}^{{\upsilon _{i}^{*},k}}}}\dot{\equiv }\mathcal{P}=\sum\limits_{{i=0}}^{{p-1}}{{{{a}_{i}}}}\prod\limits_{{{{\upsilon }_{j}}\in \,\Delta }}{{x_{j}^{{\left\langle {{{\upsilon }_{j}},\upsilon _{i}^{*}} \right\rangle +1}}}}

It is clear that a GKZ hypergeometric differential system:

\displaystyle \left\{ {\begin{array}{*{20}{c}} {{{{\hat{D}}}_{l}}\Pi \left( a \right)=0\ ,\ \left( {l\in L} \right)} \\ {{{{\hat{Z}}}_{i}}\Pi \left( a \right)=0\ ,\ \left( {i=0,1,...,p} \right)} \\ {\Pi \left( a \right)=\frac{1}{{{{{\left( {2\pi i} \right)}}^{4}}}}\int_{{\left| {{{X}_{k}}} \right|=1}}{{\frac{1}{P}\prod\limits_{{k=1}}^{4}{{\frac{{\text{d}{{X}_{k}}}}{{{{X}_{k}}}}}}}}} \end{array}} \right.

with:

\displaystyle \left\{ {\begin{array}{*{20}{c}} {{{{\hat{D}}}_{l}}=\prod\limits_{{{{l}_{i}}>0}}{{{{{\left( {\frac{\partial }{{\partial {{a}_{i}}}}} \right)}}^{{{{l}_{i}}}}}-\prod\limits_{{{{l}_{j}}<0}}{{{{{\left( {\frac{\partial }{{\partial {{a}_{j}}}}} \right)}}^{{-{{l}_{j}}}}}}}}}} \\ {{{{\hat{Z}}}_{j}}=\sum\limits_{{i=0}}^{p}{{\bar{\upsilon }_{{i,j}}^{*}{{\theta }_{{{{a}_{i}}}}}-{{\beta }_{j}}\quad ,\ \left( {j=0,1,...,n} \right)}}} \end{array}} \right.

annihilates the integral periods, and the torus invariant Kähler coordinates of the system are given as such:

\displaystyle {{z}_{a}}={{\left( {-1} \right)}^{{l_{0}^{a}}}}\prod\limits_{j}{{a_{j}^{{l_{j}^{a}}}}}

generating the Mori cone. Now, since we have:

\displaystyle {{\theta }_{a}}={{z}_{a}}{{\partial }_{{{{z}_{a}}}}}

we get:

\displaystyle {{{\mathrm B}}_{{\left\{ {{{l}^{a}}} \right\}}}}\left( {{{z}_{a}};{{\rho }_{a}}} \right)=\sum\limits_{{{{n}_{1}},...,{{n}_{\text{N}}}\in Z_{0}^{+}}}{{\frac{{\Gamma \left( {1-\sum\limits_{a}{{l_{0}^{a}\left( {{{n}_{a}}+{{\rho }_{a}}} \right)}}} \right)}}{{_{{{{\Pi }_{{i>0}}}}}\Gamma \left( {1+\sum\limits_{a}{{l_{i}^{a}\left( {{{n}_{a}}+{{\rho }_{a}}} \right)}}} \right)}}}}\times \prod\limits_{a}{{z_{a}^{{\left( {{{n}_{a}}+{{\rho }_{a}}} \right)}}}}

Now let us consider Saito’s system of DE of deformations of a singularity. Take a sheaf \Omega _{{\not{\wp }/T}}^{p} of germs of relative holomorphic p forms for the natural projection \pi :\not{\wp }={{\text{C}}^{3}}\times T\to T. One consider now sheaves on S with the following properties:

\displaystyle {{\mathcal{H}}^{{\left( 0 \right)}}}={{\varphi }_{*}}\Omega _{{\not{\wp }/T}}^{3}/df\wedge d\left( {{{\varphi }_{*}}\Omega _{{\not{\wp }/T}}^{1}} \right)

\displaystyle {{\mathcal{H}}^{{\left( {-1} \right)}}}={{\varphi }_{*}}\Omega _{{\not{\wp }/T}}^{2}/\left( {d{{f}_{1}}\wedge {{\varphi }_{*}}\Omega _{{\not{\wp }/T}}^{1}+d\left( {{{\varphi }_{*}}\Omega _{{\not{\wp }/T}}^{1}} \right)} \right)

We now take an image of a primitive form \zeta in {{H}^{0}}\left( {S,{{\mathcal{H}}^{{\left( {-1} \right)}}}} \right):

\displaystyle {{\mho }_{0}}\left( a \right)=\operatorname{Re}{{s}_{{\left\{ {{{a}_{1}}+{{f}_{1}}=0} \right\}}}}\left( {\frac{{dU\wedge dV\wedge dW}}{{{{a}_{1}}+{{f}_{1}}\left( {a,U,V,W} \right)}}} \right)

Let us connect the Gel’fand-Leray form to the Saito period integral and the GKZ system. The defining singularity polynomial equation for singularity {{U}^{2}}+{{V}^{2}}+{{W}^{{\mu +1}}} reflecting a monodromy action, is:

\displaystyle \left\{ {\begin{array}{*{20}{c}} {{{f}_{\Sigma }}\left( {a,W} \right)+{{U}^{2}}+{{V}^{2}}={{a}_{1}}+{{f}_{1}}\left( {a,U,V,W} \right)} \\ {{{f}_{1}}\left( {a,U,V,W} \right)={{a}_{2}}W+...+{{a}_{\mu }}{{W}^{{\mu -1}}}+{{W}^{{\mu +1}}}+{{U}^{2}}+{{V}^{2}}} \end{array}} \right.

for a deformation:

\displaystyle {{W}^{{\mu +1}}}+{{U}^{2}}+{{V}^{2}}=0\subset {{\mathbb{C}}^{2}}

where the parameters \left( {{{a}_{1}},{{a}_{2}},...,{{a}_{\mu }}} \right) act as a coordinate system of:

\displaystyle \left\{ {\begin{array}{*{20}{c}} {S\doteq \mathbb{C}\times T} \\ {T:={{\mathbb{C}}^{{\mu -1}}}} \end{array}} \right.

There is then a natural map:

\displaystyle \left\{ {\begin{array}{*{20}{c}} {\varphi :\not{\wp }\to S} \\ {\left( {U,V,W,{{a}_{2}},...{{a}_{\mu }}} \right)\Rightarrow \left( {-{{f}_{1}}\left( {a,U,V,W} \right),{{a}_{2}},...{{a}_{\mu }}} \right)} \end{array}} \right.

Now we focus on the sheaf \displaystyle \Omega _{{\not{\wp }/T}}^{p} of germs of holomorphic p-forms for the projection \displaystyle \pi :\not{\wp }={{\mathbb{C}}^{3}}\times T\to T of form:

\displaystyle {{\mathcal{H}}^{{\left( 0 \right)}}}={{\varphi }_{*}}\Omega _{{\not{\wp }/T}}^{3}/d{{f}_{1}}\wedge d\left( {{{\varphi }_{*}}\Omega _{{\not{\wp }/T}}^{1}} \right)

\displaystyle {{\mathcal{H}}^{{\left( {-1} \right)}}}={{\varphi }_{*}}\Omega _{{\not{\wp }/T}}^{2}/\left( {d{{f}_{1}}\wedge {{\varphi }_{*}}\Omega _{{\not{\wp }/T}}^{1}+d\left( {{{\varphi }_{*}}\Omega _{{\not{\wp }/T}}^{1}} \right)} \right)

Hence, we can re-write the image of the primitive form as such:

\displaystyle {{\mho }_{0}}\left( a \right)=\operatorname{Re}{{s}_{{\left\{ {{{a}_{1}}+{{f}_{1}}=0} \right\}}}}\left( {\frac{{dU\wedge dV\wedge dW}}{{a+{{f}_{1}}\left( {a,U,V,W} \right)}}} \right)

Now the Saito’s system is characterized as a set of differential equations satisfied by the PF:

\displaystyle {{P}_{{ij}}}{{\mho }_{0}}\left( a \right)=\left\{ {\frac{{{{\partial }^{2}}}}{{{{\partial }_{{{{a}_{i}}}}}{{\partial }_{{{{a}_{j}}}}}}}-{{\nabla }_{{\frac{\partial }{{{{\partial }_{{{{a}_{i}}}}}}}}}}\frac{\partial }{{{{\partial }_{{{{a}_{j}}}}}}}-\left( {\frac{\partial }{{{{\partial }_{{{{a}_{i}}}}}}}*\frac{\partial }{{{{\partial }_{{{{a}_{j}}}}}}}} \right)\frac{\partial }{{{{\partial }_{{{{a}_{1}}}}}}}} \right\}{{\mho }_{0}}\left( a \right)=0

\displaystyle Q\left( {\frac{\partial }{{{{\partial }_{{{{a}_{i}}}}}}}} \right){{\mho }_{0}}\left( a \right)=\left\{ {w\left( {\frac{\partial }{{{{\partial }_{{{{a}_{i}}}}}}}} \right)\frac{\partial }{{{{\partial }_{{{{a}_{1}}}}}}}-N\left( {\frac{\partial }{{{{\partial }_{{{{a}_{i}}}}}}}} \right)+\frac{3}{2}\frac{\partial }{{{{\partial }_{{{{a}_{i}}}}}}}} \right\}{{\mho }_{0}}\left( a \right)=0

We can now prove the following:

With {{\beta }_{0}}\left( a \right),...,{{\beta }_{\mu }}\left( a \right) the roots of {{\psi }_{0}}\left( W \right):={{a}_{1}}+{{a}_{2}}W+...+{{a}_{\mu }}{{W}^{{\mu -1}}}+{{W}^{{\mu +1}}} satisfying {{\beta }_{0}}\left( a \right)+...+{{\beta }_{\mu }}\left( a \right)=0, the space of the solutions to Saito’s system is generated by:

\displaystyle 1,{{\beta }_{0}}\left( a \right)-{{\beta }_{1}}\left( a \right),...,{{\beta }_{{\mu -1}}}\left( a \right)-{{\beta }_{\mu }}\left( a \right)

with a singularity at the discriminant locus:

\displaystyle div{{\psi }_{0}}\left( a \right)=\prod\nolimits_{{1\le i,j\le \mu }}{{{{{\left( {{{\beta }_{i}}\left( a \right)-{{\beta }_{j}}\left( a \right)} \right)}}^{2}}=0}}

and where the monodromy group about the discriminant is isomorphic to the order-\mu symmetric group {{S}_{\mu }} among the roots {{\beta }_{i}}\left( a \right)s.

Note that now we have a defining equation:

\displaystyle {{f}_{\Sigma }}\left( {a,W} \right)+{{U}^{2}}+{{V}^{2}}

where we take the canonical torus action {{\left( {{{\mathbb{C}}^{*}}} \right)}^{2}} on \left( {{{a}_{1}},...,{{a}_{{\mu +2}}}} \right):

\displaystyle \left\{ {\begin{array}{*{20}{c}} {{{a}_{i}}\mapsto \lambda {{a}_{i}}} \\ {{{a}_{i}}\mapsto {{\lambda }^{{i-1}}}{{a}_{i}}} \end{array}} \right.\quad \left( {\lambda \in {{\mathbb{C}}^{*}}} \right)

The Saito system now naturally relates to a Gel’fand-Kapranov-Zelevinski system with canonical period integral forms:

\displaystyle \int_{C}{{{{\mho }_{0}}\left( a \right)}}={{\int_{C}{{\operatorname{Re}s}}}_{{{{f}_{\Sigma }}\left( {a,W} \right)+{{U}^{2}}+{{V}^{2}}=0}}}\left( {\frac{{dW\wedge dU\wedge dV}}{{{{f}_{\Sigma }}\left( {a,W} \right)+{{U}^{2}}+{{V}^{2}}}}} \right)

integration is over homology 2-cycle curves on:

\displaystyle {{f}_{\Sigma }}\left( {a,W} \right)+{{U}^{2}}+{{V}^{2}}=0\subset {{\mathbb{C}}^{3}}

and:

\displaystyle {{f}_{\Sigma }}\left( {a,W} \right)={{a}_{1}}+{{a}_{2}}W+...+{{a}_{{\mu +2}}}{{W}^{{\mu +1}}}

and up to homotopic rescaling, the period integral satisfies a Gel’fand-Kapranov-Zelevinski system. Two central properties follow. The period integral satisfies:

\displaystyle \left\{ {\begin{array}{*{20}{c}} {{{\square }_{{l\in L}}}\int_{C}{{{{\mho }_{0}}\left( a \right)}}=0} \\ {{{{{\not{Z}}'}}_{{i,(i=1,2)}}}\int_{C}{{{{\mho }_{0}}\left( a \right)}}=0} \end{array}} \right.

with respect to the Saito lattice-operator, with:

\displaystyle \left( {\begin{array}{*{20}{c}} {{{{{\not{Z}}'}}_{1}}} \\ {{{{{\not{Z}}'}}_{2}}} \end{array}} \right)=\left( {\begin{array}{*{20}{c}} {{{\theta }_{1}}+{{\theta }_{2}}+...+{{\theta }_{{\mu +2}}}} \\ {{{\theta }_{2}}+2{{\theta }_{3}}+...+\left( {\mu +1} \right){{\theta }_{{\mu +2}}}+1} \end{array}} \right)

and the system is rank-{\mu +1} irreducible and the \mu-independent solutions are given by:

\displaystyle \left\{ {\begin{array}{*{20}{c}} {{{\beta }_{0}}\left( a \right)-{{\beta }_{1}}\left( a \right),...,{{\beta }_{{\mu -1}}}\left( a \right)-{{\beta }_{\mu }}\left( a \right)} \\ {{{\beta }_{i}}\left( a \right)\in Roots\left[\!\left[ {\psi \left( W \right)={{a}_{1}}+{{a}_{2}}W+...+{{a}_{{\mu +2}}}{{W}^{{\mu +1}}}=0} \right]\!\right]} \end{array}} \right.

So the root \beta \left( a \right) of \psi \left( W \right) satisfies:

\displaystyle \frac{{\partial \beta }}{{\partial {{a}_{i}}}}=-\frac{{{{\beta }^{{i-1}}}}}{{{\psi }'\left( \beta \right)}}

\displaystyle \frac{{{{\partial }^{2}}\beta }}{{\partial {{a}_{i}}\partial {{a}_{j}}}}=\frac{1}{{{\psi }'\left( \beta \right)}}\frac{d}{{dx}}{{\left( {\frac{{{{x}^{{i+j-2}}}}}{{{\psi }'\left( x \right)}}} \right)}_{{{{\mid }_{{x=\beta }}}}}}

Hence, the independent solutions of the Gel’fand-Kapranov-Zelevinski system:

\displaystyle \left\{ {\begin{array}{*{20}{c}} {{{{\hat{D}}}_{l}}\Pi \left( a \right)=0\ ,\ \left( {l\in L} \right)} \\ {{{{\hat{Z}}}_{i}}\Pi \left( a \right)=0\ ,\ \left( {i=0,1,...,p} \right)} \end{array}} \right.

are given by:

\displaystyle 1\ ,\ \log {{\beta }_{0}}\left( a \right)-\log {{\beta }_{1}}\left( a \right)\ ,\ ...\ ,\ \log {{\beta }_{{\mu -1}}}\left( a \right)-\log {{\beta }_{\mu }}\left( a \right)

whose monodromy transformations properties are given by:

\displaystyle \left\{ {\begin{array}{*{20}{c}} {\omega \left( {x;\frac{J}{{2\pi i}}} \right)=1+\sum\nolimits_{{k=1}}^{\mu }{{\omega \left( a \right){{J}_{k}}}}} \\ {2\pi i{{\omega }_{k}}\left( x \right)=\log {{\beta }_{{k-1}}}\left( a \right)-\log {{\beta }_{k}}\left( a \right)} \end{array}} \right.\quad ,\quad \left( {k=1,...,\mu } \right)

Hence,

\displaystyle \log {{\beta }_{k}}-\log {{\beta }_{{k-1}}}{{\ }_{{{{,}_{{\ \left( {k=1,...\mu } \right)}}}}}}

get annihilated by the operators {{\hat{\not{Z}}}_{1}},{{\hat{\not{Z}}}_{2}}. Moreover, we have the following relationship holding:

\displaystyle \frac{{{{\partial }^{2}}}}{{{{\partial }_{{{{a}_{i}}}}}{{\partial }_{{{{a}_{j}}}}}}}\log \beta =\frac{{{{\partial }^{2}}}}{{{{\partial }_{{{{a}_{k}}}}}{{\partial }_{{{{a}_{l}}}}}}}\log \beta \quad ,\quad \text{if}\quad i+j=k+l

given that the following holds:

\displaystyle \frac{{{{\partial }^{2}}}}{{{{\partial }_{{{{a}_{i}}}}}{{\partial }_{{{{a}_{j}}}}}}}\log \beta =-\frac{1}{{{{\beta }^{2}}}}\frac{{{{\beta }^{{i+j-2}}}}}{{{{{\left( {{\psi }'\left( \beta \right)} \right)}}^{2}}}}+\frac{1}{\beta }\frac{1}{{{\psi }'\left( \beta \right)dx}}\left( {\frac{{{{x}^{{i+j-2}}}}}{{{\psi }'\left( x \right)}}} \right)\left| {_{{x=\beta }}} \right.

By recursion, we can derive:

\displaystyle {{\prod\limits_{i}{{\left( {\frac{\partial }{{{{\partial }_{{{{a}_{i}}}}}}}} \right)}}}^{{l_{i}^{+}}}}\log \beta ={{\prod\limits_{i}{{\left( {\frac{\partial }{{{{\partial }_{{{{a}_{i}}}}}}}} \right)}}}^{{l_{i}^{-}}}}\log \beta

\displaystyle {{\square }_{l}}\log \beta =0

Now since by definition we have:

\displaystyle \left\{ {\begin{array}{*{20}{c}} {2\pi i{{\omega }_{k}}\left( x \right)\sim \log \beta =0} \\ {{{x}_{k}}=\frac{{{{a}_{k}}{{a}_{{k+2}}}}}{{a_{{k+1}}^{2}}}} \end{array}} \right.

the {{{\omega }_{k}}\left( x \right)}‘s are solutions to the Gel’fand-Kapranov-Zelevinski hypergeometric system, with:

\displaystyle \psi \left( W \right)={{a}_{1}}+{{a}_{2}}W+...+{{a}_{{\mu +2}}}{{W}^{{\mu +1}}}=\prod\limits_{k}{{\left( {{{\lambda }_{k}}W+1} \right)}}

Now we interpret the Chern-Simons part of the D-brane action as naturally inducing the charge vectors Q_{i}^{a}/\upsilon _{i}^{b} defining the Calabi-Yau condition, and consider the following operators:

\displaystyle {{\mathcal{L}}_{{{{Q}^{a}}}}}={{\prod\limits_{{Q_{i}^{a}>0}}{{\left( {\frac{\partial }{{\partial {{a}_{i}}}}} \right)}}}^{{Q_{i}^{a}}}}-{{\prod\limits_{{Q_{i}^{a}<0}}{{\left( {\frac{\partial }{{\partial {{a}_{i}}}}} \right)}}}^{{-Q_{i}^{a}}}}

\displaystyle {{{\tilde{\not{Z}}}}_{b}}=\sum\limits_{{i=0}}^{n}{{\upsilon _{i}^{b}}}{{a}_{i}}\frac{\partial }{{\partial {{a}_{i}}}}-{{{\hat{\beta }}}_{b}}

characterizing the Gel’fand-Kapranov-Zelevinski hypergeometric {\hat{\beta }}-system:

\displaystyle \left\{ {\begin{array}{*{20}{c}} {{{\mathcal{L}}_{{{{Q}^{a}}}}}\tilde{\pi }\left( a \right)=0} \\ {{{{\tilde{\not{Z}}}}_{b}}\tilde{\pi }\left( a \right)=0} \end{array}} \right.

Now, for the Calabi-Yau manifolds that arise as hypersurfaces in toric varieties, the above Gel’fand-Kapranov-Zelevinski equations are analytically related by a Witten deformation to the above Picard-Fuchs equations. Under the change of variables:

\displaystyle {{z}_{a}}=\prod\limits_{i}{{{{{\left( {{{a}_{i}}} \right)}}^{{Q_{i}^{a}}}}}}

and:

\displaystyle \tilde{\pi }\left( a \right)=\frac{1}{{{{a}_{0}}}}\pi \left( z \right)

the corresponding Gel’fand-Kapranov-Zelevinski system obtained allows us to derive the open-string instanton effects exactly as in the bulk Gel’fand-Kapranov-Zelevinski for the closed-string case. Take the following operators:

\displaystyle {{\mathcal{L}}_{{{{Q}^{a}}}}}={{\prod\limits_{{Q_{i}^{a}>0}}{{\left( {\frac{\partial }{{\partial {{a}_{i}}}}} \right)}}}^{{Q_{i}^{a}}}}-{{\prod\limits_{{Q_{i}^{a}<0}}{{\left( {\frac{\partial }{{\partial {{a}_{i}}}}} \right)}}}^{{-Q_{i}^{a}}}}

\displaystyle {{\mathcal{L}}_{{{{q}^{\alpha }}}}}={{\prod\limits_{{q_{i}^{\alpha }>0}}{{\left( {\frac{\partial }{{\partial {{a}_{i}}}}} \right)}}}^{{q_{i}^{\alpha }}}}-{{\prod\limits_{{q_{i}^{\alpha }<0}}{{\left( {\frac{\partial }{{\partial {{a}_{i}}}}} \right)}}}^{{-q_{i}^{\alpha }}}}

\displaystyle {{{\tilde{\not{Z}}}}_{{b\beta }}}=\sum\limits_{{i=0}}^{n}{{\upsilon _{i}^{\beta }}}{{a}_{i}}\frac{\partial }{{\partial {{a}_{i}}}}-{{{\hat{\beta }}}_{\beta }}

which is the Gel’fand-Kapranov-Zelevinski HGS that naturally corresponds to the boundary toric variety, and under the c.o.v.:

\displaystyle \left\{ {\begin{array}{*{20}{c}} {{{z}_{a}}=\prod\limits_{i}{{{{{\left( {{{a}_{i}}} \right)}}^{{Q_{i}^{a}}}}}}} \\ {{{s}_{\alpha }}=\prod\limits_{i}{{{{{\left( {{{a}_{i}}} \right)}}^{{q_{i}^{\alpha }}}}}}} \end{array}} \right.

yields an open/closed bulk/boundary duality supporting Witten mirror flops and flips in the dual character-torus lattice of the Calabi-Yau:

\displaystyle {{\mathcal{L}}_{{boundary}}}{{/}^{{MS}}}{{\mathcal{L}}_{{bulk}}}

For the Aganagic-Klemm-Vafa system, the solution corresponding to the D-brane superpotential are given as so:

\displaystyle {{\partial }_{{\hat{\omega }}}}W\left( {\hat{\omega }} \right)=\sum\limits_{{\vec{k},m}}{{a\left( {\vec{k},m} \right){{z}^{{\vec{k}}}}}}{{s}^{m}}

where W is the Witten superpotential satisfying the Witten equation:

\bar{\partial }{{u}_{i}}+\frac{{\bar{\partial }\bar{W}}}{{\partial {{u}_{i}}}}=0

whose Picard-Lefschetz virtual fundamental cycles:

\displaystyle \left\{ {\begin{array}{*{20}{c}} {{{{\left[ {{\bar{\mathcal{W}}}'_{{g,k}}^{{rig}}\left( {{{k}_{{{{j}_{1}}}}},...,{{k}_{{{{j}_{k}}}}}} \right)} \right]}}^{{vir}}}} \\ {\deg \quad 2\left( {\left( {{{{\hat{c}}}_{W}}-3} \right)\left( {g-1} \right)+k-\sum\limits_{i}{{{{\iota }_{{{{\gamma }_{i}}}}}}}} \right)-\sum\limits_{i}{{{{N}_{{{{\gamma }_{i}}}}}}}} \end{array}} \right.

naturally induce Witten mirror moduli flops and flips bulk/boundary duality associated to the Aganagic-Klemm-Vafa disk instantons defined by the Gel’fand-Kapranov-Zelevinski system.

It can be shown that all flips and flops in the dual character-torus lattice of the Calabi-Yau moduli space corresponding to an M-theory compactification arise naturally in this way.