This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Second Inner Variations of Energy and Index of Codimension 22 Minimal Submanifolds

Jared Marx-Kuo
(July 18, 2023)
Abstract

We compute the second inner variation of the Abelian Yang–Mills–Higgs and Ginzburg–Landau energies. Given a sequence of critical points with energy measures converging to a codimension 22 minimal submanifold, we use the second inner variation formula to bound the morse index of the submanifold by the index of the critical points. The key tools are the convergence of the energy measures and the stress-energy tensors of solutions to Abelian Yang–Mills–Higgs and Ginzburg–Landau equations.

1 Introduction

There is a long history of approximating minimal submanifolds with solutions to partial differential equations. The Allen–Cahn energy gives one such equation

EAC,ϵ(u)\displaystyle E_{AC,\epsilon}(u) =Mϵ2|u|2+W(u)ϵ\displaystyle=\int_{M}\frac{\epsilon}{2}|\nabla u|^{2}+\frac{W(u)}{\epsilon} (1)
ϵ2Δg(u)\displaystyle\epsilon^{2}\Delta_{g}(u) =u(u21)EAC,ϵ(u)=0\displaystyle=u(u^{2}-1)\iff E_{AC,\epsilon}^{\prime}(u)=0 (2)

Modica–Mortola first established Γ\Gamma-convergence of the Allen–Cahn energy to the perimeter functional [19] and Hutchison–Tonegawa later showed weak convergence of the energy measures to a stationary codimension 11 varifold [14]. Pacard–Ritorè [20] then showed that given a non-degenerate minimal surface, an Allen–Cahn solution can be glued in nearby. Guaraco established a min-max construction for the Allen–Cahn energy [12], and Chodosh–Mantoulidis established curvature estimates and index bounds using the Allen–Cahn equation [8].

These results highlight a deep connection between Allen–Cahn solutions and minimal surfaces. In particular, we are interested in the following: for {uϵ}\{u_{\epsilon}\} a sequence of critical points to equation (1), Le [16] and Gaspar [11] computed the second inner variation of the EAC,ϵ(u)E_{AC,\epsilon}(u) and use it to bound the area Morse index of a limiting minimal surface, Yn1MnY^{n-1}\subseteq M^{n}, by the Allen–Cahn energy morse index of the solutions {uϵ}\{u_{\epsilon}\}. The same result was also proved by Hiesmayr [13] but using different techniques:

Theorem 1.1 (Gaspar 2018, Hiesmayr 2018).

Let MM a closed Riemannian manifold of dimension n3n\geq 3 and {uϵk}\{u_{\epsilon_{k}}\} a sequence of critical points of equation (1) with ϵk0\epsilon_{k}\downarrow 0. Assume that there are positive constants c0c_{0}, E0E_{0}, and a non-negative integer pp such that

lim supksupM|uϵk|c0,lim supkEϵk(uϵk)E0,lim supkm(uϵk)p\limsup_{k}\sup_{M}|u_{\epsilon_{k}}|\leq c_{0},\qquad\limsup_{k}E_{\epsilon_{k}}(u_{\epsilon_{k}})\leq E_{0},\qquad\limsup_{k}m(u_{\epsilon_{k}})\leq p

Then up to a subsequence, for all open subsets UM\sing(Γ)U\subset\subset M\backslash\text{sing}(\Gamma) with Ureg(Γ)U\cap\text{reg}(\Gamma)\neq\emptyset, the eigenvalues {λϵ(U)}\{\lambda_{\ell}^{\epsilon}(U)\}_{\ell} of the linearized Allen–Cahn operator at uϵl|Uu_{\epsilon_{l}}\Big{|}_{U}, Lk=ϵkΔ+W′′(uϵk)/ϵkL_{k}=-\epsilon_{k}\Delta+W^{\prime\prime}(u_{\epsilon_{k}})/\epsilon_{k}, and the eigenvalues {λ(U)}\{\lambda_{\ell}(U)\}_{\ell} of the Jacobi operator of Γ\Gamma acting on normal vector fields supported on UΓU\cap\Gamma satisfy

lim supkλϵk(U)ϵkλ(U)\limsup_{k}\frac{\lambda_{\ell}^{\epsilon_{k}}(U)}{\epsilon_{k}}\leq\lambda_{\ell}(U)

for all \ell. In particular, reg(Γ)\text{reg}(\Gamma) has morse index at most pp.

Many of the results connecting the Allen–Cahn equation to minimal hypersurfaces are now being extended to the codimension 22 case. In this context, the Abelian Yang–Mills–Higgs equations (sometimes referred to as “U(1)U(1)-Yang–Mills–Higgs”) are

ϵ2u\displaystyle\epsilon^{2}\nabla^{*}\nabla u =12(1|u|2)u\displaystyle=\frac{1}{2}(1-|u|^{2})u (3)
ϵ2dF\displaystyle\epsilon^{2}d^{*}F_{\nabla} =iImu,uL\displaystyle=i\text{Im}\langle u,\nabla u\rangle_{L}

and Ginzburg–Landau equations

ϵ2Δu=u(|u|21)\epsilon^{2}\Delta u=u(|u|^{2}-1) (4)

These equations have been used to approximate minimal submanifolds Yn2Y^{n-2}. In the Abelian Yang–Mills–Higgs setting, Jaffe and Taubes [15] initially constructed solutions to equation (3) when M=2M=\mathbb{R}^{2}, concentrating at a set of points {zi}i=1k\{z_{i}\}_{i=1}^{k} as ϵ0\epsilon\to 0. In all dimensions, Badran–Del-Pino [1] [2] have completed a gluing construction, paralleling the result of Pacard–Ritore. Parise–Pigati–Stern [21] have established Γ\Gamma-convergence properties and Pigati–Stern [22] established weak convergence of the energy measures to stationary integral rectifiable (n2)(n-2)-varifolds:

Theorem 1.2 (Pigati-Stern, Thm 1.1).

Let {(uϵ,ϵ)}\{(u_{\epsilon},\nabla_{\epsilon})\} a family of critical points of solutions to equation (3) satisfying

Eϵ(uϵ,ϵ)Λ<E_{\epsilon}(u_{\epsilon},\nabla_{\epsilon})\leq\Lambda<\infty

Then as ϵ0\epsilon\to 0, the energy measures

μϵ:=12πeϵ(uϵ,ϵ)volg\mu_{\epsilon}:=\frac{1}{2\pi}e_{\epsilon}(u_{\epsilon},\nabla_{\epsilon})vol_{g}

converge subsequentially, in duality with C0(M)C^{0}(M), to the weight measure of a stationary integral (n2)(n-2)-varifold VV. Also for all 0δ<10\leq\delta<1,

spt(V)=limϵ0{|uϵ|δ}\text{spt}(V)=\lim_{\epsilon\to 0}\{|u_{\epsilon}|\leq\delta\}

in the Hausdorff topology.

In the Ginzburg–Landau setting, Lin–Riviere [18], Betheul–Brezis–Orlandi [3], and Betheul–Orlandi–Smets [4] showed that solutions concentrate about a collection of codimension 22 minimal submanifolds. This was later refined in the Riemannian setting by Cheng [5], Stern [25] and Pigati–Stern [23]. These authors show that the energy of a sequence of solutions converge to a stationary, but potentially not integral, n2n-2 varifold, plus a diffuse measure on all of MM.

Theorem 1.3 (Betheul–Brezis–Orlandi, Cheng, Stern, Pigati–Stern, Betheul–Orlandi–Smets).

Let (Mn,g)(M^{n},g) a closed manifold with n3n\geq 3 and {uϵ}\{u_{\epsilon}\} a sequence of solutions to equation (4) with

lim supϵ01|logϵ|K(12|duϵ|g2+W(uϵ)ϵ2)𝑑volgϵ<\limsup_{\epsilon\to 0}\frac{1}{|\log\epsilon|}\int_{K}\left(\frac{1}{2}|du_{\epsilon}|_{g}^{2}+\frac{W(u_{\epsilon})}{\epsilon^{2}}\right)dvol_{g_{\epsilon}}<\infty

for all compact KMK\subseteq M. Then, up to a subsequence, the normalized energy densities

μϵ=1|logϵ|(12|duϵ|g2+W(uϵ)ϵ2)\mu_{\epsilon}=\frac{1}{|\log\epsilon|}\left(\frac{1}{2}|du_{\epsilon}|_{g}^{2}+\frac{W(u_{\epsilon})}{\epsilon^{2}}\right)

converge to a radon measure, μ\mu, which decomposes as

μ=|V|+fvolg0\mu=|V|+fvol_{g_{0}}

for ff non-negative.

While the diffuse measure and lack of integrality is differs from the Abelian Yang–Mills–Higgs setting, Cheng [7] and Stern [25] have shown that there exists solutions, {uϵk}\{u_{\epsilon_{k}}\}, for which f=0f=0 and VV is integral. When the minimal submanifold is prescribed, De-Phillipis–Pigati [10] used variational methods to prove the existence of solutions accumulating along codimension 11 and 22 in each of the Allen-Cahn, Abelian Yang–Mills–Higgs, and Ginzburg–Landau settings.

Given the parallels between the Allen-Cahn, Abelian Yang–Mills–Higgs, and Ginzburg–Landau equations in Γ\Gamma-convergence and convergence of the energy measures, one expects to replicate results which pass geometric information between a sequence of solutions, {(uϵ,ϵ)\{(u_{\epsilon},\nabla_{\epsilon}) ( ({uϵ}(\{u_{\epsilon}\}), and the limiting stationary varifold, VV. This is the goal of this paper: to recreate the Morse-Index bound of Le and Gaspar in the context of the Ginzburg–Landau and Abelian Yang–Mills–Higgs equations on Riemannian manifolds.

2 Statement of Results

Let Γ\Gamma be a minimal codimension 22 submanifold with components {Γi}\{\Gamma_{i}\}. Define

μV:=i=1qmidΓin2\mu_{V}:=\sum_{i=1}^{q}m_{i}d\mathcal{H}^{n-2}_{\Gamma_{i}}\\

where mim_{i} are positive integers so that V=v(Γ,μV)V=v(\Gamma,\mu_{V}) defines a stationary integral (n2)(n-2)-varifold. We will assume the following:

Assumption 1.

A sequence of solutions, {(uϵk,ϵk)}\{(u_{\epsilon_{k}},\nabla_{\epsilon_{k}})\} or {uϵk}\{u_{\epsilon_{k}}\}, to the Abelian Yang–Mills–Higgs or Ginzburg–Landau equations has bounded total energy, i.e.

lim supkEAYMH,ϵk(uϵk,ϵk)\displaystyle\limsup_{k\to\infty}E_{AYMH,\epsilon_{k}}(u_{\epsilon_{k}},\nabla_{\epsilon_{k}}) <\displaystyle<\infty
lim supkEGL,ϵk(uϵk)\displaystyle\limsup_{k\to\infty}E_{GL,\epsilon_{k}}(u_{\epsilon_{k}}) <\displaystyle<\infty

and converges to VV in that:

12πlimkeAYMH,ϵk(uϵk,ϵk)dVolM\displaystyle\frac{1}{2\pi}\lim_{k\to\infty}e_{AYMH,\epsilon_{k}}(u_{\epsilon_{k}},\nabla_{\epsilon_{k}})dVol_{M} =μV\displaystyle=\mu_{V} (5)
1πlimk1|log(ϵk)|eGL,ϵk(uϵk)dVolM\displaystyle\frac{1}{\pi}\lim_{k\to\infty}\frac{1}{|\log(\epsilon_{k})|}e_{GL,\epsilon_{k}}(u_{\epsilon_{k}})dVol_{M} =μV\displaystyle=\mu_{V} (6)

Regularity aside, theorem 1.2 says that equation (5) is typical. By contrast, theorem 1.3 says that equation (6) is atypical, but Cheng [7] and Stern [25] show that b1(M)=0b_{1}(M)=0 implies f=0f=0 in theorem 1.3. In any case, we assume both convergence assumptions to establish morse index bounds. We state our main theorems here:

Theorem 2.1.

Let {(uϵk,ϵk)}\{(u_{\epsilon_{k}},\nabla_{\epsilon_{k}})\} (resp. {uϵk}\{u_{\epsilon_{k}}\}) be a sequence of critical points for Abelian Yang–Mills–Higgs (resp. Ginzburg–Landau) satisfying assumption (5) (resp. (6)). For XΓ(TM)X\in\Gamma(TM), let Φt:MM\Phi^{t}:M\to M be the corresponding flow, and define {(uϵkt,ϵkt)}\{(u_{\epsilon_{k}}^{t},\nabla_{\epsilon_{k}}^{t})\} (resp. {(uϵt)}\{(u_{\epsilon}^{t})\}). Then

12πlimkd2dt2EAYMH,ϵ(uϵkt,ϵkt)|t=0\displaystyle\frac{1}{2\pi}\lim_{k\to\infty}\frac{d^{2}}{dt^{2}}E_{\text{AYMH},\epsilon}(u_{\epsilon_{k}}^{t},\nabla_{\epsilon_{k}}^{t})\Big{|}_{t=0} =i=1qmi[D2A(Γi)(X,X)+Γi(divNΓi(X)2+12g˙XNΓi2dVolΓi)]\displaystyle=\sum_{i=1}^{q}m_{i}\left[D^{2}A(\Gamma_{i})(X,X)+\int_{\Gamma_{i}}\left(-\text{div}_{N\Gamma_{i}}(X)^{2}+\frac{1}{2}||\dot{g}_{X}||_{N\Gamma_{i}}^{2}dVol_{\Gamma_{i}}\right)\right] (7)
1πlimkd2dt2EGL,ϵ(uϵkt)|t=0\displaystyle\frac{1}{\pi}\lim_{k\to\infty}\frac{d^{2}}{dt^{2}}E_{\text{GL},\epsilon}(u_{\epsilon_{k}}^{t})\Big{|}_{t=0} =i=1qmi[D2A(Γi)(X,X)+Γi(divNΓi(X)2+12g˙XNΓi2dVolΓi)]\displaystyle=\sum_{i=1}^{q}m_{i}\left[D^{2}A(\Gamma_{i})(X,X)+\int_{\Gamma_{i}}\left(-\text{div}_{N\Gamma_{i}}(X)^{2}+\frac{1}{2}||\dot{g}_{X}||_{N\Gamma_{i}}^{2}dVol_{\Gamma_{i}}\right)\right]

Moreover, the integral term in (7) is non-negative.

Remark  We note that Le ([17], theorem 1.5) and Cheng ([6], Proposition 2.6) have done this computation for a sequence of Ginzburg–Landau solutions in the Euclidean and Riemannian setting, respectively. Both authors compute this to investigate stability of Ginzburg-Landau solutions. We also note that Cheng ([6], Proposition 5.2) computes the second inner variation for Abelian Yang-Mills-Higgs in a different form, again to prove stability. We include our own derivations since they emphasize the parallels between the Abelian Yang–Mills–Higgs and Ginzburg–Landau cases, unified by the convergence of the stress energy tensor.

For any UMU\subseteq M with Ureg(V)U\cap\text{reg}(V)\neq\emptyset, let {λp(U)}\{\lambda_{p}(U)\} denote the eigenvalues of the Jacobi operator of VV acting on normal vector fields supported on Ureg(V)U\cap\text{reg}(V). Similarly, for a sequence of critical points {(uϵk,ϵk)}\{(u_{\epsilon_{k}},\nabla_{\epsilon_{k}})\} (resp. {uϵk}\{u_{\epsilon_{k}}\} ) to the Abelian Yang–Mills–Higgs (resp. Ginzburg–Landau) equations, let λAYMH,pϵk(U)\lambda_{AYMH,p}^{\epsilon_{k}}(U) (resp. λGL,pϵk\lambda_{GL,p}^{\epsilon_{k}}) denote the eigenvalues of the linearized Abelian Yang–Mill–Higgs (resp. Ginzburg–Landau) operator at (uϵk,ϵk)|U(u_{\epsilon_{k}},\nabla_{\epsilon_{k}})\Big{|}_{U} (resp. uϵk|Uu_{\epsilon_{k}}\Big{|}_{U}).

Theorem 2.2.

Let {(uϵk,ϵk)}\{(u_{\epsilon_{k}},\nabla_{\epsilon_{k}})\} (resp. {uϵk}\{u_{\epsilon_{k}}\}) be a sequence of critical points for Abelian Yang–Mills–Higgs (resp. Ginzburg–Landau) satisfying assumption 1 and

lim supkm(uϵk,ϵk)\displaystyle\limsup_{k}m(u_{\epsilon_{k}},\nabla_{\epsilon_{k}}) m\displaystyle\leq m
lim supkm(uϵk,ϵk)\displaystyle\limsup_{k}m(u_{\epsilon_{k}},\nabla_{\epsilon_{k}}) m\displaystyle\leq m

Then up to a subsequence, for all UM\sing(V)U\subseteq M\backslash\text{sing}(V) with Ureg(V)U\cap\text{reg}(V)\neq\emptyset

lim supkλAYMH,pϵk(U)\displaystyle\limsup_{k}\lambda_{AYMH,p}^{\epsilon_{k}}(U) 2λp(U)\displaystyle\leq 2\lambda_{p}(U)
lim supkλGL,pϵk(U)\displaystyle\limsup_{k}\lambda_{GL,p}^{\epsilon_{k}}(U) λp(U)\displaystyle\leq\lambda_{p}(U)

In particular, the morse index of the regular part of VV has morse index at most mm.

We can immediately apply this to any of the solutions constructed in [10], [7], [25], [2] to give a lower bound on the morse index of the constructed solutions, {uϵ}\{u_{\epsilon}\} or {(uϵ,ϵ)}\{(u_{\epsilon},\nabla_{\epsilon})\}, in terms of the morse index of the limiting varifold.

The techniques to prove theorems 2.1 2.2 are almost identical to that of Gaspar [11] and Hiesmayr [13], expect we replace the Allen-Cahn equation with the Abelian Yang–Mills–Higgs and Ginzburg–Landau equations. We also substitute proposition 2.2 of Gaspar [11] with knowledge of the limit of the stress-energy tensor for the Abelian Yang–Mills-Higgs and Ginzburg–Landau equations. We also remark that up to the appropriate constant, the integral term in (7) simplifies to that of Gaspar’s ([11], Proposition 3.3) and Le’s ([16], Theorem 1) when the normal bundle is 11-dimensional and XX lies in the normal bundle.

2.1 Acknowledgements

The author would like to thank Otis Chodosh for suggesting the initial idea of the project. The author would also like to thank Daniel Stern and Pedro Gaspar for their time in answering questions about their relevant papers. The author dedicates this paper to his grandmother, Shirley Kuo.

3 Preliminaries

3.1 Riemannian Background

Let (M,g)(M,g) a Riemannian manifold. For XΓ(TM)X\in\Gamma(TM), let Φt\Phi^{t} be the associated flow and define gt=(Φt)(g)g^{t}=(\Phi^{t})^{*}(g). Fix a point qMq\in M and a basis {i}i=1n\{\partial_{i}\}_{i=1}^{n} at qq. Let g˙ij,g¨ij\dot{g}_{ij},\ddot{g}_{ij} denote the first and second derivatives for the metric coefficients evaluated at pp, i.e.

g˙ij=ddtgijt|t=0,g¨ij=d2dt2gijt|t=0\dot{g}_{ij}=\frac{d}{dt}g^{t}_{ij}\Big{|}_{t=0},\qquad\ddot{g}_{ij}=\frac{d^{2}}{dt^{2}}g^{t}_{ij}\Big{|}_{t=0}

Adopt the same notation for g˙ij,g¨ij\dot{g}^{ij},\ddot{g}^{ij} as derivatives of the metric inverse coefficients. We also define

TP(V,W)\displaystyle T_{P}(V,W) :=(g|P)ij(g|P)kiV,kW,j\displaystyle:=(g|_{P})^{ij}(g|_{P})^{k\ell}\langle\nabla_{i}V,\partial_{k}\rangle\langle\nabla_{\ell}W,\partial_{j}\rangle (8)

where PP is any sub-bundle of TMTM and g|Pg|_{P} is the restricted metric. Then we have:

Lemma 3.1.

The following hold

g˙ij\displaystyle\dot{g}_{ij} =iX,j+i,jX\displaystyle=\langle\nabla_{i}X,\partial_{j}\rangle+\langle\partial_{i},\nabla_{j}X\rangle (9)
g˙ij\displaystyle\dot{g}^{ij} =gikg˙kmgmj\displaystyle=-g^{ik}\dot{g}_{km}g^{mj}
g¨ij\displaystyle\ddot{g}_{ij} =iXX,j+jXX,i+2iX,jX2R(X,i,X,j)\displaystyle=\langle\nabla_{i}\nabla_{X}X,\partial_{j}\rangle+\langle\nabla_{j}\nabla_{X}X,\partial_{i}\rangle+2\langle\nabla_{i}X,\nabla_{j}X\rangle-2R(X,\partial_{i},X,\partial_{j})
g¨ij\displaystyle\ddot{g}^{ij} =2girg˙rsgskg˙kmgmjgikg¨kmgmj\displaystyle=2g^{ir}\dot{g}_{rs}g^{sk}\dot{g}_{km}g^{mj}-g^{ik}\ddot{g}_{km}g^{mj}
ddtdetgt|t=0\displaystyle\frac{d}{dt}\sqrt{\det g^{t}}\Big{|}_{t=0} =div(X)detg\displaystyle=\text{div}(X)\sqrt{\det g}
d2dt2detgt|t=0\displaystyle\frac{d^{2}}{dt^{2}}\sqrt{\det g^{t}}\Big{|}_{t=0} =div(XX)Ric(X,X)TTM(X,X)+div(X)2\displaystyle=\text{div}(\nabla_{X}X)-\text{Ric}(X,X)-T_{TM}(X,X)+\text{div}(X)^{2}

See [11], lemma 3.1 for reference. Here we follow the convention

R(V,W,Y,Z)=(WVVW[W,V])Y,ZR(V,W,Y,Z)=\langle(\nabla_{W}\nabla_{V}-\nabla_{V}\nabla_{W}-\nabla_{[W,V]})Y,Z\rangle

We also recall that for XX, a not necessarily normal vector field, that the second variation of area along a submanifold ΣM\Sigma\subseteq M, flowed by Φt\Phi^{t} is given by

D2A|Σ(X,X)=Σ[trTΣR(X,,X,)TTΣ(X,X)+divTΣ(X)2+divTΣ(XX)+TΣNΣX2]𝑑VolΣD^{2}A\Big{|}_{\Sigma}(X,X)=\int_{\Sigma}[-\text{tr}_{T\Sigma}R(X,\cdot,X,\cdot)-T_{T\Sigma}(X,X)+\text{div}_{T\Sigma}(X)^{2}+\text{div}_{T\Sigma}(\nabla_{X}X)+||\nabla_{T\Sigma}^{N\Sigma}X||^{2}]dVol_{\Sigma} (10)

(See [24], 9.4 or [9], section 8.1).

3.2 Ginzburg–Landau and Abelian Yang–Mills–Higgs Equations

The Ginzburg–Landau functional is given by

u\displaystyle u H1(M,)\displaystyle\in H^{1}(M,\mathbb{C})
eGL,ϵ(u)\displaystyle e_{GL,\epsilon}(u) :=12|u|2+W(u)ϵ2\displaystyle:=\frac{1}{2}|\nabla u|^{2}+\frac{W(u)}{\epsilon^{2}}
EGL,ϵ(u)\displaystyle E_{GL,\epsilon}(u) =M12|u|2+W(u)ϵ2\displaystyle=\int_{M}\frac{1}{2}|\nabla u|^{2}+\frac{W(u)}{\epsilon^{2}} (11)
μGL,ϵk\displaystyle\mu_{GL,\epsilon_{k}} :=1π|log(ϵ)|eϵk(uϵk)dVolg\displaystyle:=\frac{1}{\pi|\log(\epsilon)|}e_{\epsilon_{k}}(u_{\epsilon_{k}})dVol_{g}

with critical points of (11) satisfying equation (4).

ϵ2Δu=u(|u|21)\epsilon^{2}\Delta u=u(|u|^{2}-1)

For Abelian Yang–Mills–Higgs, let LL be a complex line bundle over MM, then

u\displaystyle u :ML\displaystyle:M\to L
\displaystyle\nabla :Γ(TML)Γ(L)\displaystyle:\Gamma(TM\otimes L)\to\Gamma(L)
eAYMH,ϵ(u,)\displaystyle e_{AYMH,\epsilon}(u,\nabla) :=|u|2+ϵ2F2+W(u)ϵ2\displaystyle:=|\nabla u|^{2}+\epsilon^{2}||F_{\nabla}||^{2}+\frac{W(u)}{\epsilon^{2}}
EAYMH,ϵ(u,)\displaystyle E_{\text{AYMH},\epsilon}(u,\nabla) :=M|u|2+ϵ2F2+W(u)ϵ2\displaystyle:=\int_{M}|\nabla u|^{2}+\epsilon^{2}||F_{\nabla}||^{2}+\frac{W(u)}{\epsilon^{2}} (12)
μAYMH,ϵk\displaystyle\mu_{AYMH,\epsilon_{k}} :=12πeϵk(uϵk,ϵk)dVolg\displaystyle:=\frac{1}{2\pi}e_{\epsilon_{k}}(u_{\epsilon_{k}},\nabla_{\epsilon_{k}})dVol_{g}

Throughout this paper, we will refer to μGL,ϵk\mu_{GL,\epsilon_{k}} and μAYMH,ϵk\mu_{AYMH,\epsilon_{k}} as the “energy measures.” Here, we follow the weighting and norm convention for 2-forms as in [22]: let {xi}\{x_{i}\} be a local coordinate basis, then

τ\displaystyle\tau =i<jτijdxidxj\displaystyle=\sum_{i<j}\tau_{ij}dx_{i}\wedge dx_{j}
τg2\displaystyle||\tau||_{g}^{2} =gijgklτikτj\displaystyle=g^{ij}g^{kl}\tau_{ik}\tau_{j\ell}

when {i}\{\partial_{i}\} are orthonormal at a point, this gives

τg2=i<jτij2=12i,j=1nτij2||\tau||_{g}^{2}=\sum_{i<j}\tau_{ij}^{2}=\frac{1}{2}\sum_{i,j=1}^{n}\tau_{ij}^{2}

We also define an ϵ\epsilon-weighted inner product on pairs of sections and ii\mathbb{R}-valued one forms

(f,a),(h,b):=f,hL+ϵ2a,bL\langle(f,a),(h,b)\rangle:=\langle f,h\rangle_{L}+\epsilon^{2}\langle a,b\rangle_{L} (13)

Critical points of equation (12) satisfy the coupled system equation (3)

ϵ2u\displaystyle\epsilon^{2}\nabla^{*}\nabla u =12(1|u|2)u\displaystyle=\frac{1}{2}(1-|u|^{2})u
ϵ2dF\displaystyle\epsilon^{2}d^{*}F_{\nabla} =iIm(u,uL)\displaystyle=i\text{Im}\left(\langle u,\nabla u\rangle_{L}\right)

3.2.1 Stress Energy Tensors

For each equation, we define the corresponding stress and stress-energy tensors, following the convention of [22], section 44.

SGL,ϵ\displaystyle S_{GL,\epsilon} =(uϵu)\displaystyle=(\nabla u_{\epsilon}^{*}\nabla u) (14)
SGL,ϵ(V,W)\displaystyle S_{GL,\epsilon}(V,W) :=Vuϵ,Wuϵ\displaystyle:=\langle\nabla_{V}u_{\epsilon},\nabla_{W}u_{\epsilon}\rangle_{\mathbb{C}}
TGL,ϵ\displaystyle T_{GL,\epsilon} :=eϵ(u)gSGL,ϵ\displaystyle:=e_{\epsilon}(u)g-S_{GL,\epsilon}

For Abelian Yang–Mills–Higgs, the gauge group of LL is U(1)U(1). We can write F=iωF_{\nabla}=i\omega for ωΩ2(TM,)\omega\in\Omega^{2}(TM,\mathbb{R}). We then define

(uu)(V,W)\displaystyle(\nabla u^{*}\nabla u)(V,W) :=Vu,WuL\displaystyle:=\langle\nabla_{V}u,\nabla_{W}u\rangle_{L}
(ωω)(V,W)\displaystyle(\omega^{*}\omega)(V,W) :=gijω(V,ei)ω(W,ei)\displaystyle:=g^{ij}\omega(V,e_{i})\omega(W,e_{i})
SAYMH,ϵ\displaystyle S_{AYMH,\epsilon} :=2[uu+ωω]\displaystyle:=2[\nabla u^{*}\nabla u+\omega^{*}\omega]
TAYMH,ϵ\displaystyle T_{AYMH,\epsilon} :=eϵ(u,A)gSAYMH,ϵ\displaystyle:=e_{\epsilon}(u,A)g-S_{AYMH,\epsilon}

From Pigati–Stern [22], we recall the convergence of the stress energy tensor

Proposition 1 (Pigati–Stern, Proposition 6.4).

For a family {(uϵ,ϵ)}\{(u_{\epsilon},\nabla_{\epsilon})\} of solutions to equation (3) with uniform energy bound, after passing to a subsequence {ϵj}0\{\epsilon_{j}\}\to 0, there exists a stationary, rectifiable, integral (n2)(n-2)-varifold, V=v(Σn2,θ)V=v(\Sigma^{n-2},\theta) such that

limϵ0MTAYMH,ϵ(uϵ,ϵ),P=Σθ(x)TxΣ,P(x)𝑑n2\lim_{\epsilon\to 0}\int_{M}\langle T_{AYMH,\epsilon}(u_{\epsilon},\nabla_{\epsilon}),P\rangle=\int_{\Sigma}\theta(x)\langle T_{x}\Sigma,P(x)\rangle d\mathcal{H}^{n-2}

for every PC0(M,Sym(TM))P\in C^{0}(M,\text{Sym}(TM)).

As a result of proposition 1 and theorem 1.2, we see that for a sequence {(uϵk,ϵk)}\{(u_{\epsilon_{k}},\nabla_{\epsilon_{k}})\} of solutions to equation (3), we have

limkSAYMH,ϵ(uϵk,ϵk),P=Σθ(x)NΣx,P𝑑n2\lim_{k\to\infty}\langle S_{AYMH,\epsilon}(u_{\epsilon_{k}},\nabla_{\epsilon_{k}}),P\rangle=\int_{\Sigma}\theta(x)\langle N\Sigma_{x},P\rangle d\mathcal{H}^{n-2} (15)

In the Ginzburg–Landau setting, we have the following from [3], §IX, and Cheng [6] (in the proof of proposition 2.6):

Proposition 2.

For a sequence {uϵk}\{u_{\epsilon_{k}}\} with uniformily bounded energy converging to a stationary varifold (i.e. h0=0h_{0}=0 in the context of theorem 1.3),

limkMTGL,ϵk,P\displaystyle\lim_{k\to\infty}\int_{M}\langle T_{GL,\epsilon_{k}},P\rangle =Σθ(x)TxΣ,P(x)𝑑Σn2\displaystyle=\int_{\Sigma}\theta(x)\langle T_{x}\Sigma,P(x)\rangle d\mathcal{H}^{n-2}_{\Sigma} (16)
limkMSGL,ϵk,P\displaystyle\lim_{k\to\infty}\int_{M}\langle S_{GL,\epsilon_{k}},P\rangle =Σθ(x)NxΣ,P(x)𝑑Σn2\displaystyle=\int_{\Sigma}\theta(x)\langle N_{x}\Sigma,P(x)\rangle d\mathcal{H}^{n-2}_{\Sigma} (17)

for all PC0(M,Sym(TM))P\in C^{0}(M,\text{Sym}(TM)).

Remark  In the context of our assumptions 1, we note that the density θ(x)\theta(x) will be 2π2\pi\mathbb{Z} (resp. π\pi\mathbb{Z}) valued in accordance with the normalization in equation (5) (resp. (6)).

4 Computation of first inner variation

In the Ginzburg–Landau setting, define

ut=(Φt)(u)u^{t}=(\Phi^{-t})^{*}(u)

and we have

EGL,ϵ(ut)\displaystyle E_{GL,\epsilon}(u^{t}) =M12|ut|2+W(ut)ϵ2dV\displaystyle=\int_{M}\frac{1}{2}|\nabla u^{t}|^{2}+\frac{W(u^{t})}{\epsilon^{2}}dV
=M[12|ut|2Φt+W(u)ϵ2](Φt)(dVol)\displaystyle=\int_{M}\left[\frac{1}{2}|\nabla u^{t}|^{2}\circ\Phi^{t}+\frac{W(u)}{\epsilon^{2}}\right](\Phi^{t})^{*}(dVol)

Closely mimicking Gaspar [11], §3, let {eit}\{e_{i}^{t}\} be an ONB at yy. We compute

|ut|2|y\displaystyle|\nabla u^{t}|^{2}\Big{|}_{y} =i=1ndut(eit)2\displaystyle=\sum_{i=1}^{n}du^{t}(e_{i}^{t})^{2}
=i=1n(du|Φt(y))(Φt(eit))2\displaystyle=\sum_{i=1}^{n}\left(du\Big{|}_{\Phi^{-t}(y)}\right)(\Phi^{-t}_{*}(e_{i}^{t}))^{2}
=i=1n(du)(vit)2|Φt(y)\displaystyle=\sum_{i=1}^{n}\left(du\right)(v_{i}^{t})^{2}\Big{|}_{\Phi^{-t}(y)}
=i=1ngt(gtu,vit)2|Φt(y)\displaystyle=\sum_{i=1}^{n}g^{t}\left(\nabla^{g^{t}}u,v_{i}^{t}\right)^{2}\Big{|}_{\Phi^{-t}(y)}
|ut|2Φt(x)\displaystyle\implies|\nabla u^{t}|^{2}\circ\Phi^{t}(x) =i=1ngt(gtu,vit)2|x\displaystyle=\sum_{i=1}^{n}g^{t}\left(\nabla^{g^{t}}u,v_{i}^{t}\right)^{2}\Big{|}_{x}

Here, we’ve defined vit=Φt|y(eit)v_{i}^{t}=\Phi^{-t}_{*}\Big{|}_{y}(e_{i}^{t}), which is a vector at Φt(y)\Phi^{-t}(y). Moreover, we’ve noted that vitv_{i}^{t} is an ONB at Φt(y)\Phi^{-t}(y) with respect to the metric gt=(Φt)(g)g^{t}=(\Phi^{t})^{*}(g). We then composed the whole expression with Φt\Phi^{t} to have everything evaluated at the fixed point xx. Now let {i}\{\partial_{i}\} be an arbitrary (time independent!) basis at xx, then we have

|ut|2Φt(x)=|gtu|2|x=(gt)ijuiuj|\nabla u^{t}|^{2}\circ\Phi^{t}(x)=|\nabla^{g^{t}}u|^{2}\Big{|}_{x}=(g^{t})^{ij}u_{i}u_{j}

We can then compute using equation (9)

ddtEGL,ϵ(ut)|t=0\displaystyle\frac{d}{dt}E_{GL,\epsilon}(u^{t})\Big{|}_{t=0} =M[12g˙,SGL,ϵ+eGL,ϵ(u)div(X)]𝑑Vol\displaystyle=\int_{M}\left[-\frac{1}{2}\langle\dot{g},S_{GL,\epsilon}\rangle+e_{GL,\epsilon}(u)\text{div}(X)\right]dVol (18)
=M[SGL,ϵ,X+eGL,ϵ(u)g,X]\displaystyle=\int_{M}\left[-\langle S_{GL,\epsilon},\nabla X\rangle+e_{GL,\epsilon}(u)\langle g,\nabla X\rangle\right]
=MTGL,ϵ,X\displaystyle=\int_{M}\langle T_{GL,\epsilon},\nabla X\rangle

having used symmetry of SGL,ϵS_{GL,\epsilon}. This vanishes exactly when uϵu_{\epsilon} is a solution to (4), reflecting the fact that the stress energy tensor is divergence free.

In the Abelian Yang–Mills–Higgs setting, define

ut=(Φt)(u),t=(Φt)()u^{t}=(\Phi^{-t})^{*}(u),\qquad\nabla^{t}=(\Phi^{-t})^{*}(\nabla)

so that

EAYMH,ϵ(ut,At)\displaystyle E_{AYMH,\epsilon}(u^{t},A^{t}) =M[|tut|2+ϵ2|Ft|2+W(ut)ϵ2]𝑑Vol\displaystyle=\int_{M}\left[|\nabla^{t}u^{t}|^{2}+\epsilon^{2}|F_{\nabla^{t}}|^{2}+\frac{W(u^{t})}{\epsilon^{2}}\right]dVol
=M[|tut|2Φt+ϵ2|Ft|2Φt+W(u)ϵ2](Φt)(dVol)\displaystyle=\int_{M}\left[|\nabla^{t}u^{t}|^{2}\circ\Phi^{t}+\epsilon^{2}|F_{\nabla^{t}}|^{2}\circ\Phi^{t}+\frac{W(u)}{\epsilon^{2}}\right](\Phi^{t})^{*}(dVol)

The analogous computation to equation (18) gives

ddt(|tut|2Φt(x))\displaystyle\frac{d}{dt}\left(|\nabla^{t}u^{t}|^{2}\circ\Phi^{t}(x)\right) =g˙,uu\displaystyle=-\langle\dot{g},\nabla u^{*}\nabla u\rangle
ddt(|Ft|2Φt(x))\displaystyle\frac{d}{dt}\left(|F_{\nabla^{t}}|^{2}\circ\Phi^{t}(x)\right) =g˙,FF\displaystyle=-\langle\dot{g},F_{\nabla}^{*}F_{\nabla}\rangle
ddtEAYMH,ϵ(ut,t)|t=0\displaystyle\frac{d}{dt}E_{AYMH,\epsilon}(u^{t},\nabla^{t})\Big{|}_{t=0} =MeAYMH,ϵ(u,)div(X)g˙,SAYMH,ϵ\displaystyle=\int_{M}e_{\text{AYMH},\epsilon}(u,\nabla)\text{div}(X)-\langle\dot{g},S_{AYMH,\epsilon}\rangle
=MTAYMH,ϵ,X\displaystyle=\int_{M}\langle T_{AYMH,\epsilon},\nabla X\rangle

5 Computation of second inner variation

In the Ginzburg–Landau case, we compute using (9):

d2dt2(|ut|2Φt)(x)|t=0\displaystyle\frac{d^{2}}{dt^{2}}\left(|\nabla u^{t}|^{2}\circ\Phi^{t}\right)(x)\Big{|}_{t=0} =g¨ijuiuj|x\displaystyle=\ddot{g}^{ij}u_{i}u_{j}\Big{|}_{x}
=2(g˙g˙)g¨,uu\displaystyle=\langle 2(\dot{g}\circ\dot{g})-\ddot{g},\nabla u^{*}\nabla u\rangle

here,

(g˙g˙)ij:=g˙ikgkg˙j(\dot{g}\circ\dot{g})_{ij}:=\dot{g}_{ik}g^{k\ell}\dot{g}_{\ell j}

So that

d2dt2EGL,ϵ(uϵt)|t=0\displaystyle\frac{d^{2}}{dt^{2}}E_{\text{GL},\epsilon}(u^{t}_{\epsilon})\Big{|}_{t=0} =M12d2dt2(|tut|2Φt)𝑑Vol+ddt(|tut|2Φt)|t=0ddt(Φt)(dVolg)|t=0+eϵ(u)d2dt2(Φt)(dVolg)|t=0\displaystyle=\int_{M}\frac{1}{2}\frac{d^{2}}{dt^{2}}\left(|\nabla^{t}u^{t}|^{2}\circ\Phi^{t}\right)dVol+\frac{d}{dt}\left(|\nabla^{t}u^{t}|^{2}\circ\Phi^{t}\right)\Big{|}_{t=0}\cdot\frac{d}{dt}(\Phi^{t})^{*}(dVol_{g})\Big{|}_{t=0}+e_{\epsilon}(u)\frac{d^{2}}{dt^{2}}(\Phi^{t})^{*}(dVol_{g})\Big{|}_{t=0}
=M122(g˙g˙)g¨,SGL,ϵ𝑑VolM\displaystyle=\int_{M}\frac{1}{2}\langle 2(\dot{g}\circ\dot{g})-\ddot{g},S_{GL,\epsilon}\rangle dVol_{M}
+Mg˙,SGL,ϵdiv(X)dVol\displaystyle+\int_{M}-\langle\dot{g},S_{GL,\epsilon}\rangle\text{div}(X)dVol
+MeGL,ϵ(uϵ)[div(XX)Ric(X,X)TTM(X,X)+div(X)2]𝑑Vol\displaystyle+\int_{M}e_{\text{GL},\epsilon}(u_{\epsilon})\left[\text{div}(\nabla_{X}X)-\text{Ric}(X,X)-T_{TM}(X,X)+\text{div}(X)^{2}\right]dVol

In the Abelian Yang–Mills–Higgs case, we proceed analogously:

d2dt2EAYMH,ϵ(u,A)|t=0\displaystyle\frac{d^{2}}{dt^{2}}E_{\text{AYMH},\epsilon}(u,A)\Big{|}_{t=0} =M122(g˙g˙)g¨,SAYMH,ϵ𝑑VolM\displaystyle=\int_{M}\frac{1}{2}\langle 2(\dot{g}\circ\dot{g})-\ddot{g},S_{AYMH,\epsilon}\rangle dVol_{M}
+M12g˙,SAYMH,ϵdiv(X)dVol\displaystyle+\int_{M}-\frac{1}{2}\langle\dot{g},S_{AYMH,\epsilon}\rangle\text{div}(X)dVol
+MeAYMH,ϵ(uϵ)[div(XX)Ric(X,X)TTM(X,X)+div(X)2]\displaystyle+\int_{M}e_{\text{AYMH},\epsilon}(u_{\epsilon})\left[\text{div}(\nabla_{X}X)-\text{Ric}(X,X)-T_{TM}(X,X)+\text{div}(X)^{2}\right]

We now reduce the second inner variation, using our lemmas about the stress-energy tensor.

Proof of theorem 2.1:
From lemma 2 and equation (6), we have that

limϵ0d2dt2EGL,ϵ(ut)|t=0\displaystyle\lim_{\epsilon\to 0}\frac{d^{2}}{dt^{2}}E_{\text{GL},\epsilon}(u^{t})\Big{|}_{t=0} =πi=1qmiΓi(12[2trNΓi(g˙g˙)trNΓi(g¨)]trNΓi(g˙)div(X)\displaystyle=\pi\sum_{i=1}^{q}m_{i}\int_{\Gamma_{i}}\Big{(}\frac{1}{2}\left[2\text{tr}_{N\Gamma_{i}}(\dot{g}\circ\dot{g})-\text{tr}_{N\Gamma_{i}}(\ddot{g})\right]-\text{tr}_{N\Gamma_{i}}(\dot{g})\text{div}(X)
+[div(XX)Ric(X,X)TTM(X,X)+divTM(X)2])dVol\displaystyle\qquad+[\text{div}(\nabla_{X}X)-\text{Ric}(X,X)-T_{TM}(X,X)+\text{div}_{TM}(X)^{2}]\Big{)}dVol

For P,WTMP,W\subseteq TM subbundles and XΓ(TM)X\in\Gamma(TM), let

PWX2:=trPΠWX,ΠWX||\nabla_{P}^{W}X||^{2}:=\text{tr}_{P}\langle\Pi^{W}\nabla_{\cdot}X,\Pi^{W}\nabla_{\cdot}X\rangle

At a point xMx\in M, if {vi}i=1k\{v_{i}\}_{i=1}^{k} is an orthonormal basis for PxP_{x} and {wj}j=1\{w_{j}\}_{j=1}^{\ell} orthonormal for WxW_{x}, the above becomes

PWX2|x\displaystyle||\nabla_{P}^{W}X||^{2}\Big{|}_{x} =i=1dΠWviX,ΠWviX|x\displaystyle=\sum_{i=1}^{d}\langle\Pi^{W}\nabla_{v_{i}}X,\Pi^{W}\nabla_{v_{i}}X\rangle\Big{|}_{x}
=i=1dj=1viX,wj2|x\displaystyle=\sum_{i=1}^{d}\sum_{j=1}^{\ell}\langle\nabla_{v_{i}}X,w_{j}\rangle^{2}\Big{|}_{x}

We compute in an orthonormal basis:

trNΓi(g˙g˙)\displaystyle\text{tr}_{N\Gamma_{i}}(\dot{g}\circ\dot{g}) =i=n1n(g˙g˙)ii\displaystyle=\sum_{i=n-1}^{n}(\dot{g}\circ\dot{g})_{ii}
=i=n1nj=1ng˙ij2\displaystyle=\sum_{i=n-1}^{n}\sum_{j=1}^{n}\dot{g}_{ij}^{2}
=i=n1nj=1n[iX,j2+jX,i2+2iX,jjX,i]\displaystyle=\sum_{i=n-1}^{n}\sum_{j=1}^{n}[\langle\nabla_{i}X,\partial_{j}\rangle^{2}+\langle\nabla_{j}X,\partial_{i}\rangle^{2}+2\langle\nabla_{i}X,\partial_{j}\rangle\langle\nabla_{j}X,\partial_{i}\rangle]
=NΓiTMX2+TMNΓiX2+2trNΓiXX,\displaystyle=||\nabla_{N\Gamma_{i}}^{TM}X||^{2}+||\nabla_{TM}^{N\Gamma_{i}}X||^{2}+2\text{tr}_{N\Gamma_{i}}\langle\nabla_{\nabla_{\cdot}X}X,\cdot\rangle
trNΓi(g¨)\displaystyle\text{tr}_{N\Gamma_{i}}(\ddot{g}) =i=n1n2iXX,i+2iX,iX2R(X,i,X,i)\displaystyle=\sum_{i=n-1}^{n}2\langle\nabla_{i}\nabla_{X}X,\partial_{i}\rangle+2\langle\nabla_{i}X,\nabla_{i}X\rangle-2R(X,\partial_{i},X,\partial_{i})
=2divNΓi(XX)+2NΓiTMX22trNΓiR(X,,X,)\displaystyle=2\text{div}_{N\Gamma_{i}}(\nabla_{X}X)+2||\nabla_{N\Gamma_{i}}^{TM}X||^{2}-2\text{tr}_{N\Gamma_{i}}R(X,\cdot,X,\cdot)
trNΓi(g˙)\displaystyle\text{tr}_{N\Gamma_{i}}(\dot{g}) =2i=n1niX,i\displaystyle=2\sum_{i=n-1}^{n}\langle\nabla_{i}X,\partial_{i}\rangle
=2divNΓi(X)\displaystyle=2\text{div}_{N\Gamma_{i}}(X)

In sum

limϵ0d2dt2EGL,ϵ(ut)|t=0\displaystyle\lim_{\epsilon\to 0}\frac{d^{2}}{dt^{2}}E_{\text{GL},\epsilon}(u^{t})\Big{|}_{t=0} =πi=1qmiΓi([||TMNΓiX||2+2trNΓiXX,divNΓi(XX)+trNΓiR(X,,X,)]\displaystyle=\pi\sum_{i=1}^{q}m_{i}\int_{\Gamma_{i}}\Big{(}\left[||\nabla_{TM}^{N\Gamma_{i}}X||^{2}+2\text{tr}_{N\Gamma_{i}}\langle\nabla_{\nabla_{\cdot}X}X,\cdot\rangle-\text{div}_{N\Gamma_{i}}(\nabla_{X}X)+\text{tr}_{N\Gamma_{i}}R(X,\cdot,X,\cdot)\right]
2divNΓi(X)div(X)\displaystyle\qquad\quad-2\text{div}_{N\Gamma_{i}}(X)\text{div}(X)
+[div(XX)Ric(X,X)TTM(X,X)+divTM(X)2])dVolΓi\displaystyle\qquad\quad+[\text{div}(\nabla_{X}X)-\text{Ric}(X,X)-T_{TM}(X,X)+\text{div}_{TM}(X)^{2}]\Big{)}dVol_{\Gamma_{i}}

With some cancellation, we group

trNΓiR(X,,X,)Ric(X,X)\displaystyle\text{tr}_{N\Gamma_{i}}R(X,\cdot,X,\cdot)-\text{Ric}(X,X) =trTΓiR(X,,X,)\displaystyle=-\text{tr}_{T\Gamma_{i}}R(X,\cdot,X,\cdot)
divTM(X)22divNΓi(X)div(X)\displaystyle\text{div}_{TM}(X)^{2}-2\text{div}_{N\Gamma_{i}}(X)\text{div}(X) =divTΓi(X)2divNΓi(X)2\displaystyle=\text{div}_{T\Gamma_{i}}(X)^{2}-\text{div}_{N\Gamma_{i}}(X)^{2}
div(XX)divNΓi(XX)\displaystyle\text{div}(\nabla_{X}X)-\text{div}_{N\Gamma_{i}}(\nabla_{X}X) =divTΓi(XX)\displaystyle=\text{div}_{T\Gamma_{i}}(\nabla_{X}X)
TMNΓiX2\displaystyle||\nabla_{TM}^{N\Gamma_{i}}X||^{2} =TΓiNΓiX2+NΓiNΓiX2\displaystyle=||\nabla_{T\Gamma_{i}}^{N\Gamma_{i}}X||^{2}+||\nabla_{N\Gamma_{i}}^{N\Gamma_{i}}X||^{2}
NΓiNΓiX2+2trNΓiXX,TTM(X,X)\displaystyle||\nabla_{N\Gamma_{i}}^{N\Gamma_{i}}X||^{2}+2\text{tr}_{N\Gamma_{i}}\langle\nabla_{\nabla_{\cdot}X}X,\cdot\rangle-T_{TM}(X,X) =TTΓi(X,X)+12g˙NΓi2\displaystyle=-T_{T\Gamma_{i}}(X,X)+\frac{1}{2}||\dot{g}||_{N\Gamma_{i}}^{2}

For TT as in equation (8). This gives

limϵ0d2dt2EGL,ϵ(ut)|t=0\displaystyle\lim_{\epsilon\to 0}\frac{d^{2}}{dt^{2}}E_{\text{GL},\epsilon}(u^{t})\Big{|}_{t=0} =πi=1qmiΓi(trTΓiR(X,,X,)+divTΓi(X)2+divTΓi(XX)+TΓiNΓiX2TTΓi(X,X))\displaystyle=\pi\sum_{i=1}^{q}m_{i}\int_{\Gamma_{i}}\Big{(}-\text{tr}_{T\Gamma_{i}}R(X,\cdot,X,\cdot)+\text{div}_{T\Gamma_{i}}(X)^{2}+\text{div}_{T\Gamma_{i}}(\nabla_{X}X)+||\nabla_{T\Gamma_{i}}^{N\Gamma_{i}}X||^{2}-T_{T\Gamma_{i}}(X,X)\Big{)}
+(divNΓi(X)2+12g˙NΓi2)\displaystyle\quad+\Big{(}-\text{div}_{N\Gamma_{i}}(X)^{2}+\frac{1}{2}||\dot{g}||_{N\Gamma_{i}}^{2}\Big{)}
=πimi[D2A|Γi(X,X)+Γi(divNΓi(X)2+12g˙NΓi2)𝑑Vol]\displaystyle=\pi\sum_{i}m_{i}\left[D^{2}A\Big{|}_{\Gamma_{i}}(X,X)+\int_{\Gamma_{i}}\Big{(}-\text{div}_{N\Gamma_{i}}(X)^{2}+\frac{1}{2}||\dot{g}||_{N\Gamma_{i}}^{2}\Big{)}dVol\right]

having used equation (10). Note that the error term is non-negative: let {1,2}\{\partial_{1},\partial_{2}\} a basis for NΓiN\Gamma_{i} which is orthonormal when restricted to Γi\Gamma_{i}. We have:

divNΓi(X)2+12g˙NΓi2\displaystyle-\text{div}_{N\Gamma_{i}}(X)^{2}+\frac{1}{2}||\dot{g}||_{N\Gamma_{i}}^{2} =i,j=n1niX,ijX,j+12[iX,j+jX,i]2\displaystyle=\sum_{i,j=n-1}^{n}-\langle\nabla_{i}X,\partial_{i}\rangle\langle\nabla_{j}X,\partial_{j}\rangle+\frac{1}{2}[\langle\nabla_{i}X,\partial_{j}\rangle+\langle\nabla_{j}X,\partial_{i}\rangle]^{2}
=[n1X,n1nX,n]2+[n1X,n+nX,n1]2\displaystyle=[\langle\nabla_{n-1}X,\partial_{n-1}\rangle-\langle\nabla_{n}X,\partial_{n}\rangle]^{2}+[\langle\nabla_{n-1}X,\partial_{n}\rangle+\langle\nabla_{n}X,\partial_{n-1}\rangle]^{2}
=14[g˙n1,n1g˙n,n]2+g˙n1,n2\displaystyle=\frac{1}{4}[\dot{g}_{n-1,n-1}-\dot{g}_{n,n}]^{2}+\dot{g}_{n-1,n}^{2}
=[n1X,n1nX,n]2+[n1X,n+nX,n1]2\displaystyle=[\langle\nabla_{n-1}X,\partial_{n-1}\rangle-\langle\nabla_{n}X,\partial_{n}\rangle]^{2}+[\langle\nabla_{n-1}X,\partial_{n}\rangle+\langle\nabla_{n}X,\partial_{n-1}\rangle]^{2} 0\displaystyle\geq 0

We note that the error term vanishes when iX|γ=0\nabla_{i}X\Big{|}_{\gamma}=0, e.g. if X|Γi=f1(s)1+f2(s)2X\Big{|}_{\Gamma_{i}}=f^{1}(s)\partial_{1}+f^{2}(s)\partial_{2}, where ss is a coordinate on Γi\Gamma_{i}, and 1,2\partial_{1},\partial_{2} are normal vector fields corresponding to a parallel frame on Γi\Gamma_{i}. Such a vector field can be extended to a vector field on MM via a bump function in a tubular neighborhood of Γi\Gamma_{i} (see [11], Appendix). We also note that this is the same error term in Cheng [6], Proposition 2.6, as well as Le [17], Theorem 1.5.

The reduction of limϵ0d2dt2EAYMH,ϵ(ut,t)\lim_{\epsilon\to 0}\frac{d^{2}}{dt^{2}}E_{AYMH,\epsilon}(u^{t},\nabla^{t}) is identical with the same result (but a normalizing factor of 2π2\pi) since the reduction only depends on the limit of the energy measure, the stress energy tensor, and derivatives of gtg^{t}.

6 Proof of Theorem 2.2

This section takes strong inspiration from theorem AA in [11], §4, and Lemma 3.12 [13].

Recall assumption 1 so that

μV=imidΓin2\mu_{V}=\sum_{i}m_{i}d\mathcal{H}^{n-2}_{\Gamma_{i}}

is the measure associated to our limiting varifold, VV. As in [13] (main theorem) and [11] (theorem A), fix a set UM\sing(V)U\subset\subset M\backslash\text{sing}(V). Define

QV(X)=D2V(X)=i=1qmiΓi|X|2Ric(X,X)|A(,),X|2Q_{V}(X)=D^{2}V(X)=\sum_{i=1}^{q}m_{i}\int_{\Gamma_{i}}|\nabla^{\perp}X|^{2}-\text{Ric}(X,X)-|\langle A(\cdot,\cdot),X\rangle|^{2}

for X={Xi}X=\{X_{i}\} a collection of normal vector fields to each Γi\Gamma_{i}. Given a sequence of {uϵk}\{u_{\epsilon_{k}}\} solutions to equation (4) and {(uϵk,ϵk)}\{(u_{\epsilon_{k}},\nabla_{\epsilon_{k}})\} solutions to equations (3) satisfying our assumptions 1, we have

limkμGL,ϵk\displaystyle\lim_{k\to\infty}\mu_{GL,\epsilon_{k}} =μV\displaystyle=\mu_{V}
limkμAYMH,ϵk\displaystyle\lim_{k\to\infty}\mu_{AYMH,\epsilon_{k}} =μV\displaystyle=\mu_{V}

in a weak sense. Let X~\tilde{X} be an extension of XX such that NΓiX~|Γi=0\nabla_{N\Gamma_{i}}\tilde{X}\Big{|}_{\Gamma_{i}}=0 (see, e.g. [11], Appendix for details). Let {uϵkt}\{u^{t}_{\epsilon_{k}}\} (resp. {(uϵkt,ϵkt)}\{(u^{t}_{\epsilon_{k}},\nabla^{t}_{\epsilon_{k}})\}) be the corresponding family of functions (resp. functions and connections) after pulling back our sequence of Ginzburg–Landau (resp. Abelian Yang–Mills–Higgs) solutions by Φt\Phi^{t}, the flow associated to X~\tilde{X} of on MM. Then we have

12πlimkd2dt2EAYMH,ϵ(uϵkt,ϵkt)|t=0=1πlimkd2dt2EGL,ϵ(uϵkt)|t=0=QV(X)\frac{1}{2\pi}\lim_{k\to\infty}\frac{d^{2}}{dt^{2}}E_{\text{AYMH},\epsilon}(u^{t}_{\epsilon_{k}},\nabla^{t}_{\epsilon_{k}})\Big{|}_{t=0}=\frac{1}{\pi}\lim_{k\to\infty}\frac{d^{2}}{dt^{2}}E_{\text{GL},\epsilon}(u^{t}_{\epsilon_{k}})\Big{|}_{t=0}=Q_{V}(X)

Denote

Qϵ,GL(X~)\displaystyle Q_{\epsilon,GL}(\tilde{X}) :=d2dt2EGL,ϵ(uϵkt)|t=0=EGL,ϵ′′(uϵk)(uϵk,X~,uϵk,X~)\displaystyle:=\frac{d^{2}}{dt^{2}}E_{\text{GL},\epsilon}(u^{t}_{\epsilon_{k}})\Big{|}_{t=0}=E_{\text{GL},\epsilon}^{\prime\prime}(u_{\epsilon_{k}})(\langle\nabla u_{\epsilon_{k}},\tilde{X}\rangle,\langle\nabla u_{\epsilon_{k}},\tilde{X}\rangle)
Qϵ,AYMH(X~)\displaystyle Q_{\epsilon,AYMH}(\tilde{X}) :=d2dt2EAYMH,ϵ(uϵkt,ϵkt)|t=0=EAYMH,ϵ′′(uϵk,ϵk)(ϵk,X~uϵk,Fϵk(X~,))\displaystyle:=\frac{d^{2}}{dt^{2}}E_{\text{AYMH},\epsilon}(u^{t}_{\epsilon_{k}},\nabla^{t}_{\epsilon_{k}})\Big{|}_{t=0}=E_{\text{AYMH},\epsilon}^{\prime\prime}(u_{\epsilon_{k}},\nabla_{\epsilon_{k}})(\nabla_{\epsilon_{k},\tilde{X}}u_{\epsilon_{k}},F_{\nabla_{\epsilon_{k}}}(\tilde{X},\cdot))

We now recall the variational definition of λpϵ(U)\lambda_{p}^{\epsilon}(U), of the second variation operators for Ginzburg–Landau and Abelian Yang–Mill–Higgs.

λGL,pϵ(U)\displaystyle\lambda_{GL,p}^{\epsilon}(U) =infdimE=psupϕE\0EGL,ϵ(uϵ)′′(ϕ,ϕ)ϕL2(M)2\displaystyle=\inf_{\dim E=p}\sup_{\phi\in E\backslash 0}\frac{E_{\text{GL},\epsilon}(u_{\epsilon})^{\prime\prime}(\phi,\phi)}{||\phi||_{L^{2}(M)}^{2}}
λAYMH,pϵ(U)\displaystyle\lambda_{AYMH,p}^{\epsilon}(U) =infdimE=psup(ϕ,a)E\0EAYMH,ϵ(uϵ,ϵ)′′((ϕ,a),(ϕ,a))(ϕ,a)L2(M)2\displaystyle=\inf_{\dim E=p}\sup_{(\phi,a)\in E\backslash 0}\frac{E_{\text{AYMH},\epsilon}(u_{\epsilon},\nabla_{\epsilon})^{\prime\prime}((\phi,a),(\phi,a))}{||(\phi,a)||_{L^{2}(M)}^{2}}

In λGL,p\lambda_{GL,p}, EE consists of functions ϕ:M\phi:M\to\mathbb{C}, whereas in λAYMH,p\lambda_{AYMH,p}, EE consists of sections, ϕΓ(L)\phi\in\Gamma(L), and ii\mathbb{R} valued one forms, aΩ1(TM,i)a\in\Omega^{1}(TM,i\mathbb{R}). We similarly define λp(U)\lambda_{p}(U) for the second variation of area

λp(U)\displaystyle\lambda_{p}(U) =infdimE=psupXE\0i=1qΓjX2trTΓi(R(X,,X,))|A(,),X|2XL2(Ureg(V))2\displaystyle=\inf_{\dim E=p}\sup_{X\in E\backslash 0}\frac{\sum_{i=1}^{q}\int_{\Gamma_{j}}||\nabla^{\perp}X||^{2}-\text{tr}_{T\Gamma_{i}}(R(X,\cdot,X,\cdot))-|\langle A(\cdot,\cdot),X\rangle|^{2}}{||X||_{L^{2}(U\cap\text{reg}(V))}^{2}}
=infdimE=psupXE\0QV(X)XL2(V)2\displaystyle=\inf_{\dim E=p}\sup_{X\in E\backslash 0}\frac{Q_{V}(X)}{||X||_{L^{2}(V)}^{2}} (19)

for

XL2(V)2=i=1qmiΓi|X|2𝑑Γin2||X||_{L^{2}(V)}^{2}=\sum_{i=1}^{q}m_{i}\int_{\Gamma_{i}}|X|^{2}d\mathcal{H}^{n-2}_{\Gamma_{i}}

Note that equation (19) holds by normalizing each of X|ΓiX\Big{|}_{\Gamma_{i}} by mi\sqrt{m_{i}}, as noted by Hiesmayr [13], §3.2.

Proof of theorem 2.2:
Given δ>0\delta>0, there is EE, a pp-dimensional space spanned by X1,,XpX_{1},\dots,X_{p}, and {ci}i=1p\{c_{i}\}_{i=1}^{p} so that

max{ci}Sp1QV(iciXi)iciXiL2(V)2λp(W)+δ\max_{\{c_{i}\}\in S^{p-1}}\frac{Q_{V}(\sum_{i}c_{i}X_{i})}{||\sum_{i}c_{i}X_{i}||^{2}_{L^{2}(V)}}\leq\lambda_{p}(W)+\delta (20)

Let {X~i}\{\tilde{X}_{i}\} be normal, parallel extensions of each XiX_{i} to vector fields on all of TUTU with compact support and NΓiX~j|Γi=0\nabla^{N\Gamma_{i}}\tilde{X}_{j}\Big{|}_{\Gamma_{i}}=0 for all i,ji,j. For any c={ci}Sp1c=\{c_{i}\}\in S^{p-1} we will denote

cX~:=i=1pciX~ic\cdot\tilde{X}:=\sum_{i=1}^{p}c_{i}\tilde{X}_{i}

For arbitrary b={bi}b=\{b_{i}\}, consider the maps

FGL,ϵk\displaystyle F_{GL,\epsilon_{k}} :Sp1H01(U)\displaystyle:S^{p-1}\to H^{1}_{0}(U)
bX~\displaystyle b\cdot\tilde{X} bX~,uϵk\displaystyle\mapsto\langle b\cdot\tilde{X},\nabla u_{\epsilon_{k}}\rangle
FAYMH,ϵk\displaystyle F_{AYMH,\epsilon_{k}} :Sp1H01(U,L)×H01(Ω1(TU,L))\displaystyle:S^{p-1}\to H^{1}_{0}(U,L)\times H^{1}_{0}(\Omega^{1}(TU,L))
bX~\displaystyle b\cdot\tilde{X} (ϵk,bX~uϵk,Fϵk(bX~,))\displaystyle\mapsto(\nabla_{\epsilon_{k},b\cdot\tilde{X}}u_{\epsilon_{k}},F_{\nabla_{\epsilon_{k}}}(b\cdot\tilde{X},\cdot))

Note that for ϵ\epsilon sufficiently small (or kk sufficiently large), the maps FGL,ϵkF_{GL,\epsilon_{k}} and FAYMH,ϵkF_{AYMH,\epsilon_{k}} are injective. To see this, suppose not, then there exists a sequence of {ck=(c1k,,cpk)}\{c^{k}=(c_{1}^{k},\dots,c_{p}^{k})\} such that

FGL,ϵk(ckX~)=0F_{GL,\epsilon_{k}}(c^{k}\cdot\tilde{X})=0

Since {ck}Sp1\{c^{k}\}\subseteq S^{p-1}, then up to relabelling it with a subsequence, we have ckcc^{k}\to c. However, the weak convergence of the stress energy tensor in equation (15) gives

0\displaystyle 0 =limkFGL,ϵk(ckX~),FGL,ϵk(ckX~)\displaystyle=\lim_{k\to\infty}\langle F_{GL,\epsilon_{k}}(c^{k}\cdot\tilde{X}),F_{GL,\epsilon_{k}}(c^{k}\cdot\tilde{X})\rangle
=limk(ckX~)(ckX~),SGL,ϵk\displaystyle=\lim_{k\to\infty}\langle(c^{k}\cdot\tilde{X})\otimes(c^{k}\cdot\tilde{X}),S_{GL,\epsilon_{k}}\rangle
=j=1qmjΓjtrNΓi((cX~)(cX~))\displaystyle=\sum_{j=1}^{q}m_{j}\int_{\Gamma_{j}}\text{tr}_{N\Gamma_{i}}((c\cdot\tilde{X})\otimes(c\cdot\tilde{X}))
=j=1qmjΓjΠNΓi(cX~)2\displaystyle=\sum_{j=1}^{q}m_{j}\int_{\Gamma_{j}}||\Pi^{N\Gamma_{i}}(c\cdot\tilde{X})||^{2}

But this immediately implies cX~=0c\cdot\tilde{X}=0, a contradiction because cSp1c\in S^{p-1}. The same computation works for Abelian Yang–Mills–Higgs: suppose we have a sequence {ck}\{c^{k}\} such that

FAYMH,ϵk(ckX~)=0F_{AYMH,\epsilon_{k}}(c^{k}\cdot\tilde{X})=0

Using the inner product on pairs of sections and one forms (13) and the convergence in equation (17), we have

0\displaystyle 0 =limkFAYMH,ϵk(ckX~),FAYMH,ϵk(ckX~)\displaystyle=\lim_{k\to\infty}\langle F_{AYMH,\epsilon_{k}}(c^{k}\cdot\tilde{X}),F_{AYMH,\epsilon_{k}}(c^{k}\cdot\tilde{X})\rangle
=12limk(cX~)(cX~),SAYMH,ϵk\displaystyle=\frac{1}{2}\lim_{k\to\infty}\langle(c\cdot\tilde{X})\otimes(c\cdot\tilde{X}),S_{AYMH,\epsilon_{k}}\rangle
=12j=1qmjΓjtrNΓi((cX~)(cX~))\displaystyle=\frac{1}{2}\sum_{j=1}^{q}m_{j}\int_{\Gamma_{j}}\text{tr}_{N\Gamma_{i}}((c\cdot\tilde{X})\otimes(c\cdot\tilde{X}))
=12j=1qmjΓjΠNΓi(cX~)2\displaystyle=\frac{1}{2}\sum_{j=1}^{q}m_{j}\int_{\Gamma_{j}}||\Pi^{N\Gamma_{i}}(c\cdot\tilde{X})||^{2}

Thus, the spaces of

WGL,ϵk,p\displaystyle W_{GL,\epsilon_{k},p} ={bX~,uϵk|bSp1}\displaystyle=\{\langle b\cdot\tilde{X},\nabla u_{\epsilon_{k}}\rangle\;|\;b\in S^{p-1}\}
WAYMH,ϵk,p\displaystyle W_{AYMH,\epsilon_{k},p} ={(ϵk,bX~uϵk,Fϵk(bX~,))|bSp1}\displaystyle=\{(\nabla_{\epsilon_{k},b\cdot\tilde{X}}u_{\epsilon_{k}},F_{\epsilon_{k}}(b\cdot\tilde{X},\cdot))\;|\;b\in S^{p-1}\}

are both pp-dimensional linear subspaces of H01(U)H_{0}^{1}(U). By definition, we then have

λGL,pϵk(U)\displaystyle\lambda_{GL,p}^{\epsilon_{k}}(U) supcSp1QGL,ϵk(cX~)cX~,uϵkL2(M)2\displaystyle\leq\sup_{c\in S^{p-1}}\frac{Q_{GL,\epsilon_{k}}(c\cdot\tilde{X})}{||\langle c\cdot\tilde{X},\nabla u_{\epsilon_{k}}\rangle||_{L^{2}(M)}^{2}} (21)
λAYMH,pϵk(U)\displaystyle\lambda_{AYMH,p}^{\epsilon_{k}}(U) supbSp1QAYMH,ϵk(bX~)(ϵk,cX~uϵk,Fϵk(bX~,))L2(M)2\displaystyle\leq\sup_{b\in S^{p-1}}\frac{Q_{AYMH,\epsilon_{k}}(b\cdot\tilde{X})}{||(\nabla_{\epsilon_{k},c\cdot\tilde{X}}u_{\epsilon_{k}},F_{\nabla_{\epsilon_{k}}}(b\cdot\tilde{X},\cdot))||_{L^{2}(M)}^{2}} (22)

We define

μGL,p(U)\displaystyle\mu_{GL,p}(U) :=lim supkλGL,pϵk(U)\displaystyle:=\limsup_{k\to\infty}\lambda_{GL,p}^{\epsilon_{k}}(U)
μAYMH,p(U)\displaystyle\mu_{AYMH,p}(U) :=lim supkλAYMH,pϵk(U)\displaystyle:=\limsup_{k\to\infty}\lambda_{AYMH,p}^{\epsilon_{k}}(U)

Up to taking a subsequence, we can replace lim sup\limsup with lim\lim in the above. Now for each kk, choose ckc^{k} and bkb^{k} which achieves the sup\sup in equations (21) and (22). This means for our choice of δ>0\delta>0, there exists KK such that for k>Kk>K there exists {ck}\{c^{k}\} and {bk}\{b^{k}\} so that

μGL,p(U)δ\displaystyle\mu_{GL,p}(U)-\delta QGL,ϵk(ckX~)ckX~,uϵkL2(M)2\displaystyle\leq\frac{Q_{GL,\epsilon_{k}}(c^{k}\cdot\tilde{X})}{||\langle c^{k}\cdot\tilde{X},\nabla u_{\epsilon_{k}}\rangle||_{L^{2}(M)}^{2}}
μAYMH,p(U)δ\displaystyle\mu_{AYMH,p}(U)-\delta QAYMH,ϵk(bkX~)(ϵk,bkX~uϵk,Fϵk(bkX~,))L2(M)2\displaystyle\leq\frac{Q_{AYMH,\epsilon_{k}}(b^{k}\cdot\tilde{X})}{||(\nabla_{\epsilon_{k},b^{k}\cdot\tilde{X}}u_{\epsilon_{k}},F_{\nabla_{\epsilon_{k}}}(b^{k}\cdot\tilde{X},\cdot))||_{L^{2}(M)}^{2}}

Now taking another subsequence ckcc^{k}\to c, bkbb^{k}\to b and using the convergence of the second inner variation in equation 7, we have

limkQGL,ϵk(ckX~)\displaystyle\lim_{k\to\infty}Q_{GL,\epsilon_{k}}(c^{k}\cdot\tilde{X}) =πQV(cX~)\displaystyle=\pi Q_{V}(c\cdot\tilde{X})
limkQAYMH,ϵk(bkX~)\displaystyle\lim_{k\to\infty}Q_{AYMH,\epsilon_{k}}(b^{k}\cdot\tilde{X}) =2πQV(bX~)\displaystyle=2\pi Q_{V}(b\cdot\tilde{X})

But by convergence of stress tensors (1, 2), we also have

limkckX~,uϵkL2(M)2\displaystyle\lim_{k\to\infty}||\langle c^{k}\cdot\tilde{X},\nabla u_{\epsilon_{k}}\rangle||_{L^{2}(M)}^{2} =limk(ckX~)(ckX~),SGL,ϵk\displaystyle=\lim_{k\to\infty}\langle(c^{k}\cdot\tilde{X})\otimes(c^{k}\cdot\tilde{X}),S_{GL,\epsilon_{k}}\rangle
=πj=1qmjΓjtrNΓj(cX~)2\displaystyle=\pi\sum_{j=1}^{q}m_{j}\int_{\Gamma_{j}}\text{tr}_{N\Gamma_{j}}\left(c\cdot\tilde{X}\right)^{2}
=πj=1qmjΓjcX~2\displaystyle=\pi\sum_{j=1}^{q}m_{j}\int_{\Gamma_{j}}||c\cdot\tilde{X}||^{2}
=πcX~L2(V)2\displaystyle=\pi||c\cdot\tilde{X}||_{L^{2}(V)}^{2}
limk(ϵk,bkX~uϵk,Fϵk(bkX~,))L2(M)2\displaystyle\lim_{k\to\infty}||(\nabla_{\epsilon_{k},b^{k}\cdot\tilde{X}}u_{\epsilon_{k}},F_{\nabla_{\epsilon_{k}}}(b^{k}\cdot\tilde{X},\cdot))||_{L^{2}(M)}^{2} =12limk(ckX~)(ckX~),SAYMH,ϵk\displaystyle=\frac{1}{2}\lim_{k\to\infty}\langle(c^{k}\cdot\tilde{X})\otimes(c^{k}\cdot\tilde{X}),S_{AYMH,\epsilon_{k}}\rangle
=122πbX~L2(V)2\displaystyle=\frac{1}{2}2\pi||b\cdot\tilde{X}||_{L^{2}(V)}^{2}
=πbX~L2(V)2\displaystyle=\pi||b\cdot\tilde{X}||_{L^{2}(V)}^{2}

This tells us that for kk sufficiently large and using equation (20):

μGL,p(U)δ\displaystyle\mu_{GL,p}(U)-\delta QV(cX~)cX~L2(V)2λp(U)+δ\displaystyle\leq\frac{Q_{V}(c\cdot\tilde{X})}{||c\cdot\tilde{X}||_{L^{2}(V)}^{2}}\leq\lambda_{p}(U)+\delta
μAYMH,p(U)δ\displaystyle\mu_{AYMH,p}(U)-\delta 2QV(bX~)bX~L2(V)22λp(U)+δ\displaystyle\leq\frac{2Q_{V}(b\cdot\tilde{X})}{||b\cdot\tilde{X}||_{L^{2}(V)}^{2}}\leq 2\lambda_{p}(U)+\delta

We note that the factor of 22 is an artifact of the convention of equation (12), i.e. it would not appear if we considered half of the Abelian Yang–Mills–Higgs energy instead. Since δ\delta is arbitrary, we conclude

μGL,p(U)\displaystyle\mu_{GL,p}(U) λp(U)\displaystyle\leq\lambda_{p}(U)
μAYMH,p(U)\displaystyle\mu_{AYMH,p}(U) 2λp(U)\displaystyle\leq 2\lambda_{p}(U)

Recalling that

Ind(reg(V))\displaystyle\text{Ind}(\text{reg}(V)) =supU#{p𝒩|λp(U)<0}\displaystyle=\sup_{U}\#\{p\in\mathcal{N}\;|\;\lambda_{p}(U)<0\}
Ind(reg(V))\displaystyle\implies\text{Ind}(\text{reg}(V)) lim supkIndGL(uϵk)M\displaystyle\leq\limsup_{k\to\infty}\text{Ind}_{GL}(u_{\epsilon_{k}})\leq M
Ind(reg(V))\displaystyle\implies\text{Ind}(\text{reg}(V)) lim supkIndAYMH(uϵk,ϵk)M\displaystyle\leq\limsup_{k\to\infty}\text{Ind}_{AYMH}(u_{\epsilon_{k}},\nabla_{\epsilon_{k}})\leq M

We see that the morse index of the varifold is bounded above by the upper bound on the morse index of our sequence of solutions as ϵk0\epsilon_{k}\to 0.

References

  • [1] Marco Badran and Manuel del Pino. Entire solutions to 4-dimensional ginzburg-landau equations and codimension 2 minimal submanifolds. arXiv preprint arXiv:2205.15099, 2022.
  • [2] Marco Badran and Manuel del Pino. Solutions of the ginzburg-landau equations concentrating on codimension-2 minimal submanifolds. arXiv preprint arXiv:2211.03131, 2022.
  • [3] F Bethuel, H Brezis, and Giandomenico Orlandi. Asymptotics for the ginzburg–landau equation in arbitrary dimensions. Journal of Functional Analysis, 186(2):432–520, 2001.
  • [4] Fabrice Bethuel, Giandomenico Orlandi, and Didier Smets. Convergence of the parabolic ginzburg-landau equation to motion by mean curvature. Annals of mathematics, pages 37–163, 2006.
  • [5] Da Rong Cheng. Asymptotics for the ginzburg-landau equation on manifolds with boundary under homogeneous neumann condition. Journal of Functional Analysis, 278(4):108364, 2020.
  • [6] Da Rong Cheng. Instability of solutions to the ginzburg–landau equation on sn and cpn. Journal of Functional Analysis, 279(8):108669, 2020.
  • [7] DR Cheng. Geometric variational problems: Regular and singular behavior. PhD thesis, PhD thesis, Stanford University, 2017.
  • [8] Otis Chodosh and Christos Mantoulidis. Minimal surfaces and the allen–cahn equation on 3-manifolds: Index, multiplicity, and curvature estimates. Annals of Mathematics, 191(1):213–328, 2020.
  • [9] Tobias H Colding and William P Minicozzi. A course in minimal surfaces, volume 121. American Mathematical Soc., 2011.
  • [10] Guido De Philippis and Alessandro Pigati. Non-degenerate minimal submanifolds as energy concentration sets: a variational approach. arXiv preprint arXiv:2205.12389, 2022.
  • [11] Pedro Gaspar. The second inner variation of energy and the morse index of limit interfaces. The Journal of Geometric Analysis, 30(1):69–85, 2020.
  • [12] Marco AM Guaraco. Min–max for phase transitions and the existence of embedded minimal hypersurfaces. Journal of Differential Geometry, 108(1):91–133, 2018.
  • [13] Fritz Hiesmayr. Spectrum and index of two-sided allen–cahn minimal hypersurfaces. Communications in Partial Differential Equations, 43(11):1541–1565, 2018.
  • [14] John E Hutchinson and Yoshihiro Tonegawa. Convergence of phase interfaces in the van der waals-cahn-hilliard theory. Calculus of Variations and Partial Differential Equations, 10:49–84, 2000.
  • [15] A Jaffe and CH Taubes. Monopoles and vortices, 1980.
  • [16] Nam Q Le. On the second inner variation of the allen-cahn functional and its applications. Indiana University Mathematics Journal, pages 1843–1856, 2011.
  • [17] Nam Q Le. On the second inner variations of allen–cahn type energies and applications to local minimizers. Journal de Mathématiques Pures et Appliquées, 103(6):1317–1345, 2015.
  • [18] Fang-Hua Lin and Tristan Riviere. A quantization property for static ginzburg-landau vortices. Communications on Pure and Applied Mathematics: A Journal Issued by the Courant Institute of Mathematical Sciences, 54(2):206–228, 2001.
  • [19] L Modica-S Mortola. Un esempio di γ\gamma–convergenza. Boll. UMI, 14:285–299, 1977.
  • [20] Frank Pacard and Manuel Ritoré. From constant mean curvature hypersurfaces to the gradient theory of phase transitions. Journal of Differential Geometry, 64(3):359–423, 2003.
  • [21] Davide Parise, Alessandro Pigati, and Daniel Stern. Convergence of the self-dual u(1)u(1)-yang-mills-higgs energies to the (n2)(n-2)-area functional. arXiv preprint arXiv:2103.14615, 2021.
  • [22] Alessandro Pigati and Daniel Stern. Minimal submanifolds from the abelian higgs model. Inventiones mathematicae, 223:1027–1095, 2021.
  • [23] Alessandro Pigati and Daniel Stern. Quantization and non-quantization of energy for higher-dimensional ginzburg–landau vortices. arXiv preprint arXiv:2204.06491, 2022.
  • [24] Leon Simon et al. Lectures on geometric measure theory. The Australian National University, Mathematical Sciences Institute, Centre …, 1983.
  • [25] Daniel Stern. Existence and limiting behavior of min-max solutions of the ginzburg–landau equations on compact manifolds. Journal of Differential Geometry, 118(2):335–371, 2021.