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

institutetext: Department of Mathematical Sciences, Durham University, Durham, DH1 3LE, United Kingdom

Branes and Non-Invertible Symmetries

Iñaki García Etxebarria [email protected]
Abstract

𝒩=4\mathcal{N}=4 supersymmetric Yang-Mills theories with algebra 𝔰𝔬(4N)\mathfrak{so}(4N) and appropriate choices of global structure can have non-invertible symmetries. We identify the branes holographically dual to the non-invertible symmetries, and derive the fusion rules for the symmetries from the worldvolume dynamics on the branes.

1 Introduction

The notion of symmetry is undergoing rapid evolution: during the last few years a number of works have convincingly argued that the classical textbook definition of symmetry as a group of transformations acting on local operators can (and should) be extended to include higher form symmetries acting on extended operators Gaiotto:2014kfa , higher groups structures Sharpe:2015mja ; Tachikawa:2017gyf ; Benini:2018reh and more generally higher categorical structures.

The importance of such higher categorical structures in two dimensions has been realised for a long time, where they often appear from discrete gauging Frohlich:2009gb ; Carqueville:2012dk ; Brunner:2013xna ; Bhardwaj:2017xup . A number of recent works have shown that symmetry operators without inverses (and which are therefore not elements of any group, but should rather be thought of in categorical terms) are also very common in higher dimensional theories Gaiotto:2019xmp ; Heidenreich:2021xpr ; Choi:2021kmx ; Kaidi:2021xfk ; Roumpedakis:2022aik ; Bhardwaj:2022yxj ; Arias-Tamargo:2022nlf ; Choi:2022zal ; Choi:2022jqy ; Cordova:2022ieu ; Kaidi:2022uux ; Antinucci:2022eat ; Bashmakov:2022jtl ; Damia:2022bcd ; Bhardwaj:2022lsg ; Lin:2022xod ; Bartsch:2022mpm . In this paper we will focus on one class of theories where such non-invertible symmetries appear: 𝒩=4\mathcal{N}=4 theories with gauge group111In this note we do not aim to analyse fully the mapping from boundary conditions to global structures, so we will ignore the existence of discrete choices of θ\theta angles in some of the theories we discuss Aharony:2013hda . A careful analysis of the mapping from global structures to properties of the holographic duals will be provided in EGEHR . Pin+(4N)\mathrm{Pin}^{+}(4N), Sc(4N)Sc(4N) and PO(4N)PO(4N) Bhardwaj:2022yxj . The details are a little different in the three cases, so in this introduction we will focus on the Sc(4N)Sc(4N) case for concreteness. This theory has three 2-surface symmetry generators, which we will call D2𝖼,e(Σ2)D_{2}^{\mathsf{c},e}(\Sigma_{2}), D2𝗌,m(Σ2)D_{2}^{\mathsf{s},m}(\Sigma_{2}) and their product D2𝖼,e(Σ2)D2𝗌,m(Σ2)D_{2}^{\mathsf{c},e}(\Sigma_{2})D_{2}^{\mathsf{s},m}(\Sigma_{2}). There is additionally a three-surface operator 𝒩(3)\mathcal{N}(\mathcal{M}^{3}). The terms in the fusion algebra involving 𝒩(3)\mathcal{N}(\mathcal{M}^{3}) are

𝒩(3)×𝒩(3)\displaystyle\mathcal{N}(\mathcal{M}^{3})\times\mathcal{N}(\mathcal{M}^{3}) =Σ2,Σ2H2(3;)D2𝖼,e(Σ2)D2𝗌,m(Σ2),\displaystyle=\sum_{\Sigma_{2},\Sigma_{2}^{\prime}\in H_{2}(\mathcal{M}^{3};\mathbb{Z})}D_{2}^{\mathsf{c},e}(\Sigma_{2})D_{2}^{\mathsf{s},m}(\Sigma_{2}^{\prime})\,, (1a)
𝒩(3)×D2𝖼,e(Σ2)\displaystyle\mathcal{N}(\mathcal{M}^{3})\times D_{2}^{\mathsf{c},e}(\Sigma_{2}) =𝒩(3),\displaystyle=\mathcal{N}(\mathcal{M}^{3})\,, (1b)
𝒩(3)×D2𝗌,m(Σ2)\displaystyle\mathcal{N}(\mathcal{M}^{3})\times D_{2}^{\mathsf{s},m}(\Sigma_{2}) =𝒩(3).\displaystyle=\mathcal{N}(\mathcal{M}^{3})\,. (1c)

The right hand side of (1a) is generically a sum of operators, and therefore 𝒩(3)\mathcal{N}(\mathcal{M}^{3}) is not invertible.

All these theories can be obtained from the 𝒩=4\mathcal{N}=4 SO(4N)SO(4N) theories by suitable gaugings of discrete symmetries. Whether we have performed the gauging or not is not visible for a local observer measuring processes on a topologically trivial (but arbitrarily large) neighbourhood of a point. This suggests that the holographic dual of all these theories is the same, which is indeed the case: the holographic dual is in all cases IIB on AdS5×5\mathrm{AdS}_{5}\times{\mathbb{R}\mathbb{P}}^{5}. The different theories arise from different choices of asymptotic behaviour for discrete gauge fields in the bulk, as discussed in related examples in Witten:1998xy ; Aharony:2016kai .

Since all these theories share the same bulk description, it should be possible to describe the non-invertible symmetry generators (in the cases where they are present in the field theory) in terms of objects living on the holographic IIB dual. The goal of this note is to identify these objects, and to derive their fusion rules using IIB techniques.222The techniques we use in our analysis do not require knowledge of the Lagrangian of the boundary SCFT (although the choice of theories to study is certainly informed by the field theory results in Bhardwaj:2022yxj , and we will chose our notation to dovetail the field theory analysis), so they apply equally well to the study of non-Lagrangian theories realised either holographically or via geometric engineering. See Bashmakov:2022jtl for a recent study of non-invertible symmetries in non-Lagrangian theories using a different approach.,333We refer the reader to Damia:2022bcd for a holographic study of a different class of non-invertible defects. Surprisingly, given the perhaps unfamiliar fusion relations (1), it will transpire that the symmetry generators are represented holographically by ordinary branes wrapping torsional cycles in the internal 5{\mathbb{R}\mathbb{P}}^{5}.

In order to explain how this is possible, it is useful to review briefly how the fusion relations (1) are derived in Bhardwaj:2022yxj (see also Kaidi:2021xfk ). We start with the SO(4N)SO(4N) theory, which has a 2\mathbb{Z}_{2} outer automorphism 0-form symmetry and a 2×2\mathbb{Z}_{2}\times\mathbb{Z}_{2} 1-form symmetry. We will denote the background for the 0-form symmetry A1A_{1}, and the backgrounds for the two 2\mathbb{Z}_{2} factors B2mB_{2}^{m} and C2eC_{2}^{e}. There is a cubic ’t Hooft anomaly represented by an anomaly theory with action

iπA1B2eC2m.\begin{split}i\pi\int A_{1}B_{2}^{e}C_{2}^{m}\,.\end{split} (2)

We obtain the Sc(4N)Sc(4N) theory by gauging both 1-form symmetries simultaneously.444There is a discrete choice when gauging, related to the precise way in which we sum over B2mB_{2}^{m} backgrounds. A slightly different choice (related by the outer automorphism) gives the Ss(4N)Ss(4N) global form instead, which also has non-invertible symmetries. The analysis of both cases is essentially identical, so we will focus on the Sc(4N)Sc(4N) case. (The Pin+\mathrm{Pin}^{+} and POPO cases are obtained by gauging other pairs of symmetries involved in the cubic anomaly.) Naively, we would say that the 0-form symmetry is broken due to the cubic anomaly (2). The more precise statement is that due to the anomaly the generator D3(0)(3)D_{3}^{(0)}(\mathcal{M}^{3}) of the 0-form symmetry is not invariant under combined gauge transformations of B2mB_{2}^{m} and C2eC_{2}^{e}. But as argued in Kaidi:2021xfk ; Bhardwaj:2022yxj it is possible to “dress” (or stack) D3(0)(3)D_{3}^{(0)}(\mathcal{M}^{3}) with an anomalous TQFT 𝒯\mathcal{T} depending on B2mB_{2}^{m} and C2eC_{2}^{e}. The combined topological operator is gauge invariant, and survives as a topological operator of the gauged theory. The price to pay is that the fusion rules for 𝒯\mathcal{T} are more involved, and lead to non-invertibility of the dressed operator D3(0)(3)×𝒯D_{3}^{(0)}(\mathcal{M}^{3})\times\mathcal{T} (the details will be reviewed below).

Coming back to the IIB holographic setup, the main observation of this paper is that D3(0)(3)×𝒯D_{3}^{(0)}(\mathcal{M}^{3})\times\mathcal{T} is precisely the IR limit of the theory on branes wrapping suitable torsional cycles in the holographic dual background. For instance, we will see that in the Sc(4N)Sc(4N) case the non-invertible operator arises from a D3 brane wrapping 3×1\mathcal{M}^{3}\times{\mathbb{R}\mathbb{P}}^{1}, where 15{\mathbb{R}\mathbb{P}}^{1}\subset{\mathbb{R}\mathbb{P}}^{5}. Reducing on the 5{\mathbb{R}\mathbb{P}}^{5} leaves an effective 3-dimensional brane wrapping 3\mathcal{M}^{3} inside the five dimensional bulk, which becomes D3(0)(3)D_{3}^{(0)}(\mathcal{M}^{3}) when pushed to the boundary.

A pleasing consequence of the identification in this note is that anomaly cancellation of the dressed operator follows automatically: the background fields for the symmetries of the theory are given by asymptotic values for the supergravity fields in the IIB dual, and the D3 brane action is necessarily gauge invariant under all possible gauge transformations of these (although the precise way in which this happens is often subtle). Since anomaly cancellation is automatic once we start talking about branes, it is illuminating to understand why non-invertible symmetries appear in the holographic dual without referring to anomalous operators. This is also desirable since the split between the bare D(3)D(\mathcal{M}^{3}) and its “dressing” 𝒯\mathcal{T} is unnatural in terms of the brane theory, particularly once we try to formulate things in the language of K-theory. We provide such an explanation below in terms of incomplete cancellation of induced brane charges due to quantum effects.

Note added

I thank the authors of ABBS for informing me of their related upcoming work, where they give complementary evidence for the identification of non-invertible symmetries with branes in holographic settings, and for agreeing to coordinate submissions.

2 4d 𝒩=4\mathcal{N}=4 𝔰𝔭𝔦𝔫(4N)\mathfrak{spin}(4N) SYM and non-invertibles

The Spin(4N)\mathrm{Spin}(4N) SYM theory has a 2-group structure, with one-form symmetry group555Our conventions are as follows: Spin(4n)\mathrm{Spin}(4n) has two spinor irreps unrelated by complex conjugation, which we denote by “𝗌\mathsf{s}” and “𝖼\mathsf{c}”. 2𝗌\mathbb{Z}_{2}^{\mathsf{s}} acts on 𝖼\mathsf{c}, and leaves 𝗌\mathsf{s} invariant, while 2𝖼\mathbb{Z}_{2}^{\mathsf{c}} acts on 𝗌\mathsf{s} and leaves 𝖼\mathsf{c} invariant. This choice of notation is motivated by consistency with the fact that the diagonal 2\mathbb{Z}_{2} combination, traditionally denoted 2V\mathbb{Z}_{2}^{V}, does not act on the vector. We define Sc(4N)Spin(4N)/2𝗌Sc(4N)\coloneqq\mathrm{Spin}(4N)/\mathbb{Z}_{2}^{\mathsf{s}}.

Γ(1)=2𝗌×2𝖼,\Gamma^{(1)}=\mathbb{Z}_{2}^{\mathsf{s}}\times\mathbb{Z}_{2}^{\mathsf{c}}\,, (3)

and a 0-form symmetry part 2(0)\mathbb{Z}_{2}^{(0)} which is an outer automorphism that acts on the 1-form symmetry by exchanging the two factors: 2𝗌2𝖼\mathbb{Z}_{2}^{\mathsf{s}}\leftrightarrow\mathbb{Z}_{2}^{\mathsf{c}}. We will now construct the topological defects that generate these symmetries in the holographic dual.

This holographic dual is obtained as the near horizon limit of a stack of D3-branes on top of an O3- orientifold, and is given by IIB string theory on AdS5×5\mathrm{AdS}_{5}\times{\mathbb{R}\mathbb{P}}^{5} Witten:1998xy . In general we want to put the field theory on some spin666We will assume for simplicity that neither 4\mathcal{M}^{4} nor any of the submanifolds where we will wrap defects contains torsion in homology. This is not physically required, but it simplifies some of the formulas below. manifold 4\mathcal{M}^{4} different from S4S^{4}, so we will replace AdS5\mathrm{AdS}_{5} by a non-compact manifold X5X^{5} which asymptotically becomes ×4\mathbb{R}\times\mathcal{M}^{4} Witten:1998wy . There is a non-trivial SL(2,)SL(2,\mathbb{Z}) duality fibration over 5{\mathbb{R}\mathbb{P}}^{5}, which acts with the 1SL(2,)-1\in SL(2,\mathbb{Z}) element as we go around the non-trivial generator of π1(5)=2\pi_{1}({\mathbb{R}\mathbb{P}}^{5})=\mathbb{Z}_{2}. (This element can be represented alternatively as ΩFL\Omega F_{L} in worldsheet terms, but with future generalisations in mind we will describe it as an SL(2,)SL(2,\mathbb{Z}) bundle instead.) The 2-form supergravity fields B2B_{2} and C2C_{2} get a sign under this action, and project down to 2\mathbb{Z}_{2} fields on AdS5\mathrm{AdS}_{5}, while C4C_{4} does not get a sign and survives as a continuous field. We will find it useful to work in a democratic formulation, where we also include the B6B_{6} and C6C_{6} fields magnetic dual to B2B_{2} and C2C_{2}. SL(2,)SL(2,\mathbb{Z}) is a gauge symmetry of the theory on the (orientable) space AdS5×5\mathrm{AdS}_{5}\times{\mathbb{R}\mathbb{P}}^{5}, so in order for the action to be well defined we need B6B_{6} and C6C_{6} to also transform with a minus sign under 1SL(2,)-1\in SL(2,\mathbb{Z}).

What this means is that H3H_{3} and F3F_{3} are elements of the cohomology group with local coefficients H3(X5×5;~)H^{3}(X^{5}\times{\mathbb{R}\mathbb{P}}^{5};\widetilde{\mathbb{Z}}) (we refer the reader to appendix 3.H of Hatcher:478079 for details), and similarly their magnetic duals H7H_{7} and F7F_{7} are elements of H7(X5×5;~)H^{7}(X^{5}\times{\mathbb{R}\mathbb{P}}^{5};\widetilde{\mathbb{Z}}). On the other hand F5F_{5} is classified by H5(X5×5;)H^{5}(X^{5}\times{\mathbb{R}\mathbb{P}}^{5};\mathbb{Z}). In what follows we will focus on the structure on 5{\mathbb{R}\mathbb{P}}^{5}, as the SL(2,)SL(2,\mathbb{Z}) bundle is trivial on X5X^{5}. The untwisted cohomology groups of 5{\mathbb{R}\mathbb{P}}^{5} are standard, and the twisted ones can be derived easily from the results in Thom1952 :

H(5,)\displaystyle H^{\ast}({\mathbb{R}\mathbb{P}}^{5},\mathbb{Z}) ={, 0,2, 0,2,}\displaystyle=\{\mathbb{Z}\,,\ 0\,,\ \mathbb{Z}_{2}\,,\ 0\,,\ \mathbb{Z}_{2}\,,\ \mathbb{Z}\} (4)
H(5,~)\displaystyle H^{\ast}({\mathbb{R}\mathbb{P}}^{5},\widetilde{\mathbb{Z}}) ={0,2, 0,2, 0,2}.\displaystyle=\{0\,,\ \mathbb{Z}_{2}\,,\ 0\,,\ \mathbb{Z}_{2}\,,\ 0\,,\ \mathbb{Z}_{2}\}\,.

Similar considerations hold for homology: (p,q)(p,q) 1-branes (such as fundamental strings and D1 branes) are elements of H2(X5×5;~)H_{2}(X^{5}\times{\mathbb{R}\mathbb{P}}^{5};\widetilde{\mathbb{Z}}), (p,q)(p,q) 5-branes are elements of H6(X5×5;~)H_{6}(X^{5}\times{\mathbb{R}\mathbb{P}}^{5};\widetilde{\mathbb{Z}}), and D3 branes are elements of H4(X5×5;)H_{4}(X^{5}\times{\mathbb{R}\mathbb{P}}^{5};\mathbb{Z}). The relevant homology groups are (by Poincaré duality, which holds since 5{\mathbb{R}\mathbb{P}}^{5} is orientable)

H(5,)\displaystyle H_{\ast}({\mathbb{R}\mathbb{P}}^{5},\mathbb{Z}) ={,2, 0,2, 0,}\displaystyle=\{\mathbb{Z}\,,\ \mathbb{Z}_{2}\,,\ 0\,,\ \mathbb{Z}_{2}\,,\ 0\,,\ \mathbb{Z}\} (5)
H(5,~)\displaystyle H_{\ast}({\mathbb{R}\mathbb{P}}^{5},\widetilde{\mathbb{Z}}) ={2, 0,2, 0,2, 0}.\displaystyle=\{\mathbb{Z}_{2}\,,\ 0\,,\ \mathbb{Z}_{2}\,,\ 0\,,\ \mathbb{Z}_{2}\,,\ 0\}\,.

With an understanding of the cycles that the branes can wrap, it is straightforward to identify the charged operators of the Spin(4N)\mathrm{Spin}(4N) theory Witten:1998xy : the vector Wilson line WVW_{V} is a fundamental string on a point of 5{\mathbb{R}\mathbb{P}}^{5}, the 𝗌\mathsf{s}-spinor Wilson line W𝗌W_{\mathsf{s}} is a D5-brane on 4{\mathbb{R}\mathbb{P}}^{4}, and finally the 𝖼\mathsf{c}-spinor Wilson line W𝖼W_{\mathsf{c}} is the combination of both previous lines: a D5-brane/F1 bound state, again wrapped on 4{\mathbb{R}\mathbb{P}}^{4}. In all cases the branes wrap a surface on X5X^{5} extending to the boundary, where they end on a line.

We can go to the SO(4N)SO(4N) theory by gauging the diagonal factor 2V2𝗌×2𝖼\mathbb{Z}_{2}^{V}\subset\mathbb{Z}_{2}^{\mathsf{s}}\times\mathbb{Z}_{2}^{\mathsf{c}}. The vector line WVW_{V} is unaffected by the gauging, so it survives, but the W𝗌W_{\mathsf{s}} and W𝖼W_{\mathsf{c}} lines are no longer gauge invariant, and become non-genuine (that is, boundaries of surface operators). A non-genuine line HVH_{V} of the Spin(4N)\mathrm{Spin}(4N) theory, with w2𝗌=w2𝖼w_{2}^{\mathsf{s}}=w_{2}^{\mathsf{c}} flux around it, now becomes a genuine line operator in the SO(4N)SO(4N) theory. Holographically this operator corresponds to a D1 brane wrapping a point in H0(5;~)=2H_{0}({\mathbb{R}\mathbb{P}}^{5};\widetilde{\mathbb{Z}})=\mathbb{Z}_{2}.777The simplest derivation of this fact follows from recalling that the SO(4N)SO(4N) field theory is invariant under SL(2,)SL(2,\mathbb{Z}), which maps to an SL(2,)SL(2,\mathbb{Z}) action on the holographic dual. We refer the reader to EGEHR for a systematic analysis. We denote the generators for these two symmetries D2B,e(2)D_{2}^{B,e}(\mathcal{M}^{2}) (acting on fundamental strings) and D2C,m()D_{2}^{C,m}(\mathcal{M}) (acting on D1 branes), and the corresponding background fields B2eB_{2}^{e} and C2mC_{2}^{m}.

Starting from the SO(4N)SO(4N) theory we can gauge various pairs of global symmetries, an operation that, due to the cubic anomaly (2), results in theories with non-invertible symmetries Kaidi:2021xfk ; Bhardwaj:2022yxj :

Pin+(4N):\displaystyle\mathrm{Pin}^{+}(4N)\colon    gauge D3(0)D_{3}^{(0)} and D2C,mD_{2}^{C,m} (6)
Sc(4N):\displaystyle Sc(4N)\colon    gauge D2B,eD_{2}^{B,e} and D2C,mD_{2}^{C,m}
PO(4N):\displaystyle PO(4N)\colon gauge D3(0) and D2B,e.\displaystyle\quad\text{gauge $D_{3}^{(0)}$ and $D_{2}^{B,e}$}\,.

In these expressions D3(0)D_{3}^{(0)}, or more precisely D3(0)(3)D_{3}^{(0)}(\mathcal{M}^{3}) is the generator for the outer automorphism 0-form symmetry of the SO(4N)SO(4N) theory.

We are thus led to the crucial question in this paper: having identified the charged operators in the field theory in terms of the holographic dual, what is the holographic description of the charge operators implementing the global symmetries in the SO(4N)SO(4N) theory?

For concreteness, let us specialise to the holographic dual of the symmetry generator D2C,m(2)D_{2}^{C,m}(\mathcal{M}^{2}) of the SO(4N)SO(4N) theory, measuring how many ’t Hooft lines (mod 2) HVH_{V} are linked by 2\mathcal{M}^{2}, without taking into account the Wilson lines WVW_{V}. Given our identification of lines above, a natural guess would be

D2C,m(2)?eiπ2×4C6,D_{2}^{C,m}(\mathcal{M}^{2})\stackrel{{\scriptstyle?}}{{\to}}e^{i\pi\int_{\mathcal{M}_{2}\times{\mathbb{R}\mathbb{P}}^{4}}C_{6}}\,, (7)

where 2\mathcal{M}^{2} lives on 4\mathcal{M}^{4}, and becomes the symmetry operator when pushed to the boundary. This holonomy certainly measures the number of D1 branes linked by 2\mathcal{M}^{2} (the basic argument is given below in case of the outer automorphism 0-form symmetry), but it cannot be the right answer for a number of reasons. First, we know that in IIB string theory fluxes are not measured by cohomology, but rather K-theory Moore:1999gb ; Freed:2000tt ; Freed:2000ta . A way of capturing the right K-theoretic formula is to phrase the answer in terms of the Wess-Zumino coupling in the D5 brane action:

D2C,m(2)?eWZ(2×4).D_{2}^{C,m}(\mathcal{M}^{2})\stackrel{{\scriptstyle?}}{{\to}}e^{\text{WZ}(\mathcal{M}_{2}\times{\mathbb{R}\mathbb{P}}^{4})}\,. (8)

where Cheung:1997az ; Minasian:1997mm ; Freed:2000ta

WZ(X)=2πiXeF2B2A^(TX)A^(NX)(C0+C2+)\mathrm{WZ}(X)=2\pi i\int_{X}e^{F_{2}-B_{2}}\sqrt{\frac{\hat{A}(TX)}{\hat{A}(NX)}}(C_{0}+C_{2}+\ldots) (9)

A second reason why we expect neither (7) nor (8) to be the full answer is that in string theory there are no local operators, only dynamical objects. So we should aim to represent the charge generator by a dynamical object, and not simply a defect. The dynamical objects that are electrically charged under C6C_{6}, and would arise when fixing the insertion of the defect as a boundary condition, are D5 branes.

While neither argument is conclusive, they both suggest that the holographic description of the symmetry generator is a full D5, pushed to the boundary:888The authors of ABBS provide complementary evidence for the same proposal.

D2C,m(2)D5(2×4).D_{2}^{C,m}(\mathcal{M}^{2})\to\text{D5}(\mathcal{M}_{2}\times{\mathbb{R}\mathbb{P}}^{4})\,. (10)

This ansatz has the additional virtue of restoring the common origin between lines and charge generators, familiar from the formulation of symmetries in terms of relative field theories Freed:2012bs .

An objection one might raise about (10) is that branes are not topological, while charge operators should be. As we will see in a moment, the worldvolume theory on the branes, when reduced to 2X5\mathcal{M}^{2}\subset X^{5}, is a discrete 2\mathbb{Z}_{2} gauge theory. Therefore the potential lack of deformation-invariance coming from the gauge fields on the brane is not an issue. There is still an overall factor of the volume, but it does not couple to the dynamical fields of the field theory on the boundary, so it can be absorbed into a counterterm.

A subtle feature of (10) is that the worldvolume theories on the brane are quantum field theories, so we should sum over them. As we will argue, the sum over worldvolume degrees of freedom provides precisely the minimal anomalous TQFT “dressing” the bare symmetry generator identified in Bhardwaj:2022yxj . This is a very non-trivial test of the identification (10).

Clearly, if the ansatz (10) is correct, the holographic dual of the operator counting ’t Hooft lines HVH_{V} is the S-dual of (10):

D2B,e(2)NS5(2×4).D_{2}^{B,e}(\mathcal{M}^{2})\to\text{NS5}(\mathcal{M}_{2}\times{\mathbb{R}\mathbb{P}}^{4})\,. (11)

Additionally, the SO(4M)SO(4M) theory has the 0-form parity symmetry discussed above. The point operator charged under this symmetry is known as the Pfaffian operator. As discussed in Witten:1998xy the Pfaffian operator is represented holographically by a D3 brane wrapping the 3{\mathbb{R}\mathbb{P}}^{3} cycle inside 5{\mathbb{R}\mathbb{P}}^{5}, and extending to a point on the boundary. We will refer to this brane as the “Pfaffian brane”.

We now argue that the holographic dual of the generator of this symmetry is

D3(0)(3)D3(3×1).D_{3}^{(0)}(\mathcal{M}^{3})\to\text{D3}(\mathcal{M}^{3}\times{\mathbb{R}\mathbb{P}}^{1})\,. (12)

More precisely, we will show that the Pfaffian brane is charged under this D3 in the Hamiltonian formalism, so we take the boundary to be of the form 3×t\mathcal{M}^{3}\times\mathbb{R}_{t}, with the last component the time direction along the boundary. We choose coordinates so that the endpoint of the Pfaffian operator is at t=0t=0. Now we wrap the putative symmetry generator D3 on 3×1\mathcal{M}^{3}\times{\mathbb{R}\mathbb{P}}^{1}, where 3\mathcal{M}^{3} is at the boundary at t=0t=0. Because the F5F_{5} RR flux is self-dual, the two D3 branes that we have introduced do not commute Gukov:1998kn ; Moore:2004jv ; Freed:2006ya ; Freed:2006yc :

D3(3×1)Pf(pt)=e2πi𝖫(1,3)Pf(pt)D3(3×1)=Pf(pt)D3(3×1)\begin{split}\text{D3}(\mathcal{M}^{3}\times{\mathbb{R}\mathbb{P}}^{1})\operatorname{Pf}(\mathrm{pt})=e^{2\pi i\,\mathsf{L}({\mathbb{R}\mathbb{P}}^{1},{\mathbb{R}\mathbb{P}}^{3})}\operatorname{Pf}(\mathrm{pt})\text{D3}(\mathcal{M}^{3}\times{\mathbb{R}\mathbb{P}}^{1})=-\operatorname{Pf}(\mathrm{pt})\text{D3}(\mathcal{M}^{3}\times{\mathbb{R}\mathbb{P}}^{1})\end{split} (13)

where Pf(pt)\operatorname{Pf}(\mathrm{pt}) is the D3 brane representing the Pfaffian operator, 𝖫(1,3)=12\mathsf{L}({\mathbb{R}\mathbb{P}}^{1},{\mathbb{R}\mathbb{P}}^{3})=\frac{1}{2} is the linking pairing between the given cycles of 5{\mathbb{R}\mathbb{P}}^{5}, and we have used that the given branes intersect at a point on the t=0t=0 spatial slice on X5X^{5}. Equation (13) is the Hamiltonian version of the statement that the Pfaffian operator is charged under D3(3×1)\text{D3}(\mathcal{M}^{3}\times{\mathbb{R}\mathbb{P}}^{1}), as claimed. This discussion can be generalised straightforwardly to show that the branes (10) and (11) do indeed give the expected charges to the WVW_{V} and HVH_{V} lines of the SO(4N)SO(4N) theory, as claimed.

3 TQFT stacking from Wess-Zumino couplings

Our task in this section will be to deduce the non-invertibility of the symmetry generators of the theories in (6) from our assumption that symmetry generators are represented holographically by branes.

3.1 Fluxes and twisted differential cohomology

Our basic tool will be differential cohomology. We refer the reader to Apruzzi:2021nmk for a review of the basic techniques and notation that we use. The analysis in this paper has some novelties with respect to the discussion in Apruzzi:2021nmk , which we now discuss.

The main difference is that we will be working with twisted differential cohomology. The twisted and untwisted cohomology groups of 5{\mathbb{R}\mathbb{P}}^{5} were given in (4) above. The ring structure induced by the cup product for 5{\mathbb{R}\mathbb{P}}^{5} can be obtained by adapting the discussion in Lemma 1 of twisted-cohomology-ring (see also Thom1952 ). It is most easily described by adjoining the twisted and untwisted cohomology groups

𝖧(5)\displaystyle\mathsf{H}^{*}({\mathbb{R}\mathbb{P}}^{5}) =H(5;)H(5;~)\displaystyle=H^{*}({\mathbb{R}\mathbb{P}}^{5};\mathbb{Z})\oplus H^{*}({\mathbb{R}\mathbb{P}}^{5};\widetilde{\mathbb{Z}}) (14)
=[t1,u5]/(2t1,t16,u52).\displaystyle=\mathbb{Z}[t_{1},u_{5}]/(2t_{1},t_{1}^{6},u_{5}^{2})\,.

That is, we have free components of degree 0 and 5, and 2\mathbb{Z}_{2} torsional components of degrees 1 to 5, generated by t1nt_{1}^{n}. In particular, taking an even number of powers of t1t_{1} gives an untwisted class, while taking an odd number of powers gives a twisted one. In what follows we will use the notation t2n+1=t12n+1t_{2n+1}=t_{1}^{2n+1} for twisted classes and u2n=t12nu_{2n}=t_{1}^{2n} together with u5u_{5} for untwisted ones.

We denote by t˘k\breve{t}_{k} a flat differential cohomology class with characteristic class tkt_{k}, which we denote by I(t˘k)=tkI(\breve{t}_{k})=t_{k}, and similarly for u˘k\breve{u}_{k}. We note that I(t˘12)=u2I(\breve{t}_{1}^{2})=u_{2}, and similarly I(t˘14)=u4I(\breve{t}_{1}^{4})=u_{4}, so perfectness of the linking pairing on H2(5;)×H4(5;)=2×2H^{2}({\mathbb{R}\mathbb{P}}^{5};\mathbb{Z})\times H^{4}({\mathbb{R}\mathbb{P}}^{5};\mathbb{Z})=\mathbb{Z}_{2}\times\mathbb{Z}_{2} implies that

5t˘16=5u˘2u˘4=12mod1.\int_{{\mathbb{R}\mathbb{P}}^{5}}\breve{t}_{1}^{6}=\int_{{\mathbb{R}\mathbb{P}}^{5}}\breve{u}_{2}\star\breve{u}_{4}=\frac{1}{2}\mod 1\,. (15)

This equation together with the ring structure (14) will be our workhorses in what follows.

Finally, before moving on to the examples, we need to know how to represent background fluxes in terms of differential cohomology. To lighten notation, in this section we introduce b2B2eb_{2}\coloneqq B_{2}^{e} and c2C2mc_{2}\coloneqq C_{2}^{m}. Recall that the objects charged under these backgrounds are F1 and D1 branes, respectively, so the holographic fluxes encoding these backgrounds are H˘3\breve{H}_{3} and F˘3\breve{F}_{3}, which are asymptotically of the form H˘3=b2t˘1\breve{H}_{3}=b_{2}\star\breve{t}_{1} and F˘3=c2t˘1\breve{F}_{3}=c_{2}\star\breve{t}_{1}. By imposing this asymptotic form we ensure that the charged lines in the field theory acquire the right holonomies, see Garcia-Etxebarria:2019cnb ; Apruzzi:2021nmk for analysis of similar examples. We could also include terms proportional to t˘3\breve{t}_{3} in these expansions, but they would correspond to a change of the gauge algebra to 𝔰𝔬(4N+1)\mathfrak{so}(4N+1) (for F˘3\breve{F}_{3}) of 𝔲𝔰𝔭(4N)\mathfrak{usp}(4N) (for H˘3\breve{H}_{3}) Witten:1998xy so we will not consider these terms further.999When doing this sort of expansion there is an additional subtlety involving topologically trivial differential characters that is discussed at length in Apruzzi:2021nmk . It will not affect our considerations, so we will ignore such terms. Finally, a field theory 0-form symmetry background a1a_{1} for D2(0)D_{2}^{(0)} is represented by F˘5=a1u˘4\breve{F}_{5}=a_{1}\star\breve{u}_{4}. There are additional terms possible in the expansion for F˘5\breve{F}_{5}, we will discuss these below.

3.2 Non-invertibles in Sc(4N)Sc(4N) from D3 branes

We start with the case of the Sc(4N)Sc(4N) theory, where following the analysis in Kaidi:2021xfk ; Bhardwaj:2022yxj , we expect to get the topological defect that is non-invertible from the generator of the 0-form symmetry in the SO(4N)SO(4N) theory. We identified this generator above with the D3-brane wrapped on 3×1=3×S1\mathcal{M}^{3}\times{\mathbb{R}\mathbb{P}}^{1}=\mathcal{M}^{3}\times S^{1}. The worldvolume U(1)U(1) field on the D3 brane is odd under 1SL(2,)-1\in SL(2,\mathbb{Z}), because it is a trivialisation of B2B_{2}, which is odd. Therefore it takes values in the twisted cohomology group H2(3×1;~)H^{2}(\mathcal{M}^{3}\times{\mathbb{R}\mathbb{P}}^{1};\widetilde{\mathbb{Z}}). We have H(1;~)={0,2}H^{*}({\mathbb{R}\mathbb{P}}^{1};\widetilde{\mathbb{Z}})=\{0,\mathbb{Z}_{2}\}.

When computing the path integral on the D3, the field strength F2F_{2} on the brane will induce D1 charge due to the Wess-Zumino term (9), while the magnetic field strength F2DF_{2}^{D} will induce F1 charge. The Wess-Zumino action written in terms of the electric variable F2F_{2} is

SD3,e=2πi3×1F˘5+F˘3(˘2)+F˘1(12˘2˘2+124e˘),S_{\text{D3},e}=2\pi i\int_{\mathcal{M}^{3}\times{\mathbb{R}\mathbb{P}}^{1}}\breve{F}_{5}+\breve{F}_{3}\star(\breve{\mathscr{F}}_{2})+\breve{F}_{1}\star(\frac{1}{2}\breve{\mathscr{F}}_{2}\star\breve{\mathscr{F}}_{2}+\frac{1}{24}\breve{e})\,, (16)

with ˘2=F˘2B˘2\breve{\mathscr{F}}_{2}=\breve{F}_{2}-\breve{B}_{2}, F˘1\breve{F}_{1} a differential cohomology uplift of C0C_{0}, and e˘\breve{e} is a differential cohomology uplift of the Euler class of 3×1\mathcal{M}^{3}\times{\mathbb{R}\mathbb{P}}^{1}. The term proportional to F˘1\breve{F}_{1} will be relevant only for analysis of anomalies in the space of coupling constants, which we do not analyse in this note (although this is certainly an interesting direction to explore further).

We also need to consider couplings of the form F˘3B˘2\breve{F}_{3}\star\breve{B}_{2}. As argued in Freed:2000ta (elaborating on results of Taylor:2000za ; Alekseev:2000ch ), these couplings are not present when measuring the actual K-theory charges, which is what we are ultimately interested in, so we will simply set B˘2\breve{B}_{2} to 0. A more careful treatment of this issue would be desirable, but given that inclusion of these background fields would not change our conclusions (since they would provide overall invertible prefactors on the brane action in any case, even if we included them), we will postpone a more careful treatment of this point to future work.

With these simplifications taken into account, the relevant part of the Wess-Zumino action for the D3 becomes

SD3,e=2πi3×1F˘5+F˘3F˘2.S_{\text{D3},e}=2\pi i\int_{\mathcal{M}^{3}\times{\mathbb{R}\mathbb{P}}^{1}}\breve{F}_{5}+\breve{F}_{3}\star\breve{F}_{2}\,. (17)

The first term is the differential cohomology avatar of the naive guess 3×1C4\int_{\mathcal{M}^{3}\times{\mathbb{R}\mathbb{P}}^{1}}C_{4} for the flux operator in the field theory. F˘5\breve{F}_{5} is even under the 1SL(2,)-1\in SL(2,\mathbb{Z}) action, so its general decomposition is of the form F˘5=Nu˘5+a1u˘4+a3u˘2+N1˘\breve{F}_{5}=N\star\breve{u}_{5}+a_{1}\star\breve{u}_{4}+a_{3}\star\breve{u}_{2}+N\star\breve{1}. Here NN are the number of units of RR 5-form flux on the 5{\mathbb{R}\mathbb{P}}^{5}, and we have used that F5F_{5} is self-dual to relate the components of degrees 5 and 0.

In terms of this decomposition we have an effective operator in AdS5\mathrm{AdS}_{5} of the form

D(3)=exp(2πi3×1F˘5)=exp(2πi3×5F˘5u˘4)=exp(πi3a3),\begin{split}D(\mathcal{M}^{3})&=\exp\left(2\pi i\int_{\mathcal{M}^{3}\times{\mathbb{R}\mathbb{P}}^{1}}\breve{F}_{5}\right)=\exp\left(2\pi i\int_{\mathcal{M}^{3}\times{\mathbb{R}\mathbb{P}}^{5}}\breve{F}_{5}\star\breve{u}_{4}\right)\\ &=\exp\left(\pi i\int_{\mathcal{M}^{3}}a_{3}\right)\,,\end{split} (18)

where in the second equality we have used Poincaré duality on 5{\mathbb{R}\mathbb{P}}^{5} to relate 1{\mathbb{R}\mathbb{P}}^{1} to u4u_{4}, and in the third used that the only non-trivial pairing in 5{\mathbb{R}\mathbb{P}}^{5} appearing after the expansion of F˘5\breve{F}_{5} is (15). This is the expected formula for the operator measuring discrete electric flux for the outer automorphism symmetry in the SO(4N)SO(4N) theory.

The second term is the more interesting one for our purposes. As explained above, field theory backgrounds for the symmetry D2C,mD_{2}^{C,m} are described holographically by fluxes with asymptotic form F˘3=c2t˘1\breve{F}_{3}=c_{2}\star\breve{t}_{1}. Similarly, we can expand F˘2=γ1t˘1\breve{F}_{2}=\gamma_{1}\star\breve{t}_{1}. We then have (using the formulas for integration on products reviewed in Apruzzi:2021nmk )

2πi3×1F˘3F˘2=2πi3c2γ11t˘1t˘1=πi3c2γ12\pi i\int_{\mathcal{M}^{3}\times{\mathbb{R}\mathbb{P}}^{1}}\breve{F}_{3}\star\breve{F}_{2}=2\pi i\int_{\mathcal{M}^{3}}c_{2}\gamma_{1}\int_{{\mathbb{R}\mathbb{P}}^{1}}\breve{t}_{1}\star\breve{t}_{1}=\pi i\int_{\mathcal{M}^{3}}c_{2}\gamma_{1} (19)

where in the last step we have again used the fact that the linking pairing is perfect, so

1t˘1t˘1=12mod1.\int_{{\mathbb{R}\mathbb{P}}^{1}}\breve{t}_{1}\star\breve{t}_{1}=\frac{1}{2}\mod 1. (20)

So far we have considered the charge induced on a D3 due to the gauge field strength F2F_{2}. The computation above shows that it induces an effective coupling on 3\mathcal{M}^{3} to the background for D2C,mD_{2}^{C,m}. By IIB S-duality, this implies that a dual field strength F2D=ϕ1t˘1F_{2}^{D}=\phi_{1}\star\breve{t}_{1} induces a coupling of the form

πi3b2ϕ1\begin{split}\pi i\int_{\mathcal{M}^{3}}b_{2}\phi_{1}\end{split} (21)

on the effective operator on AdS5\mathrm{AdS}_{5}. The same result can be obtained from the effective action presented in the magnetic variables obtained in Kimura:1999jb .

In elementary terms, the two couplings (19) and (21) that we have just derived can be understood as encoding the well known facts that worldvolume flux on the D3 induces D1 charge, and magnetic worldvolume flux F1 charge. Recall that the D1 and F1 are the charged objects in the SO(4N)SO(4N) theory before gauging their corresponding symmetries. After gauging, they will become the symmetry generators for the dual magnetic symmetries in the Sc(4N)Sc(4N) theory (at least if our general philosophy of identifying branes with symmetries is correct). So what we have just shown, is that when doing the path integral on the D3 we will have to sum over insertions of the symmetry generators for the 1-forms of the theory. This is certainly suggestive that condensations Gaiotto:2019xmp ; Choi:2022zal ; Roumpedakis:2022aik are going to enter the picture after gauging.

The precise details are nevertheless somewhat subtle. In general, when performing the path integral the standard prescription is that we choose whether we formulate the theory in terms of electric or magnetic variables, and then sum over the specified variables only. From this point of view the two couplings (19) and (21) seem somewhat at odds, and it is not clear which one we should choose. What saves the day is that this standard prescription has to be subtly modified whenever the cohomology groups where the electric and magnetic fluxes live contain torsional components. In this case, as originally pointed out by Moore:2004jv ; Freed:2006ya ; Freed:2006yc , the electric and magnetic flux operators do not commute. As shown in Apruzzi:2021nmk (see also Camara:2011jg for a different derivation of the same result) this flux non-commutativity leads to the existence of a discrete gauge theory when the theory is compactified on the space with torsion. The argument, adapted to the system at hand, goes as follows.

Our initial theory is four dimensional U(1)U(1) Maxwell theory on the D3, compactified on 3×1\mathcal{M}^{3}\times{\mathbb{R}\mathbb{P}}^{1}. We will present a Hamiltonian quantisation analysis, so we assume that 3=𝒩2×\mathcal{M}^{3}=\mathcal{N}^{2}\times\mathbb{R}, and we identify the last component with the time direction.101010A Lagrangian derivation will appear in GEH . The spatial slice is of the form 𝒩2×1\mathcal{N}^{2}\times{\mathbb{R}\mathbb{P}}^{1}. There is a non-trivial SL(2,)SL(2,\mathbb{Z}) duality bundle along the 1=S1{\mathbb{R}\mathbb{P}}^{1}=S^{1} direction inherited from the 5{\mathbb{R}\mathbb{P}}^{5} background, with holonomy 1-1, which induces a (F2,F2D)(F2,F2D)(F_{2},F_{2}^{D})\to(-F_{2},-F_{2}^{D}) transformation of the worldvolume gauge field. Therefore, just as in the IIB background itself, the worldvolume gauge fields on the D3 are valued in twisted cohomology. In particular H1(1;~)=2H^{1}({\mathbb{R}\mathbb{P}}^{1};\widetilde{\mathbb{Z}})=\mathbb{Z}_{2}, which justifies the statement above that there is torsion in this problem.

Consider the operators Φe(at1)\Phi_{e}(a\otimes t_{1}), Φm(bt1)\Phi_{m}(b\otimes t_{1}) that measure electric and magnetic fluxes on the torsional sector. They are associated with flat, topologically non-trivial elements of TorH2(𝒩2×1;~)\operatorname{Tor}H^{2}(\mathcal{N}^{2}\times{\mathbb{R}\mathbb{P}}^{1};\widetilde{\mathbb{Z}}) Freed:2006ya ; Freed:2006yc , which in our case are all of the form at1a\otimes t_{1}, where aH1(𝒩2;)a\in H^{1}(\mathcal{N}^{2};\mathbb{Z}) and t1t_{1} is the generator of H1(1;~)H^{1}({\mathbb{R}\mathbb{P}}^{1};\widetilde{\mathbb{Z}}). Alternatively, using Poincaré duality, we can view these operators as the holonomy of the twisted fluxes F˘2\breve{F}_{2} and F˘2D\breve{F}_{2}^{D} on cycles α×pt~\alpha\times\widetilde{\mathrm{pt}} and β×pt~\beta\times\widetilde{\mathrm{pt}}, where α\alpha and β\beta are Poincaré dual to aa and bb in 𝒩2\mathcal{N}^{2}, and pt~\widetilde{\mathrm{pt}} the (twisted) point in H0(𝒩2;~)H_{0}(\mathcal{N}^{2};\widetilde{\mathbb{Z}}), which is Poincaré dual on 1{\mathbb{R}\mathbb{P}}^{1} to t1t_{1}. So we have

Φe(at1)=exp(2πiαγ1pt~t˘1)=exp(2πiαγ11t˘12)=exp(πiαγ1)\begin{split}\Phi_{e}(a\otimes t_{1})=\exp\left(2\pi i\int_{\alpha}\gamma_{1}\int_{\widetilde{\mathrm{pt}}}\breve{t}_{1}\right)=\exp\left(2\pi i\int_{\alpha}\gamma_{1}\int_{{\mathbb{R}\mathbb{P}}^{1}}\breve{t}_{1}^{2}\right)=\exp\left(\pi i\int_{\alpha}\gamma_{1}\right)\end{split} (22)

and similarly

Φm(bt1)=exp(πiβϕ1).\begin{split}\Phi_{m}(b\otimes t_{1})=\exp\left(\pi i\int_{\beta}\phi_{1}\right)\,.\end{split} (23)

Now, it follows from the general analysis of Moore:2004jv ; Freed:2006ya ; Freed:2006yc that

Φe(at1)Φm(bt1)=(1)𝒩2abΦm(bt1)Φe(at1),\begin{split}\Phi_{e}(a\otimes t_{1})\Phi_{m}(b\otimes t_{1})=(-1)^{\int_{\mathcal{N}^{2}}ab}\Phi_{m}(b\otimes t_{1})\Phi_{e}(a\otimes t_{1})\,,\end{split} (24)

or equivalently, formulating everything in terms of homology on 𝒩2\mathcal{N}^{2} (and abusing notation slightly):

Φe(α)Φm(β)=(1)αβΦm(β)Φe(α).\begin{split}\Phi_{e}(\alpha)\Phi_{m}(\beta)=(-1)^{\alpha\cdot\beta}\Phi_{m}(\beta)\Phi_{e}(\alpha)\,.\end{split} (25)

These commutation relations are precisely those of a 2\mathbb{Z}_{2} theory. We can represent this theory by a gauge theory on the two fields γ1\gamma_{1}, ϕ1\phi_{1} with action Banks:2010zn

S2=πi3γ1δϕ1.\begin{split}S_{\mathbb{Z}_{2}}=\pi i\int_{\mathcal{M}^{3}}\gamma_{1}\delta\phi_{1}\,.\end{split} (26)

We have identified the fields appearing in the Lagrangian with γ1\gamma_{1} and ϕ1\phi_{1} since these are precisely the fields whole holonomies are measured by the operators in the theory, by construction.

Assembling all the pieces together, we find that the effective partition function on the D3, seen as an 3-surface dynamical object on X5X^{5}, is (up to an overall normalisation)

𝒩(3)=D3(0)(3)𝒟γ1𝒟ϕ1exp(πi3γ1δϕ1+c2γ1+b2ϕ1).\begin{split}\mathcal{N}(\mathcal{M}^{3})=D_{3}^{(0)}(\mathcal{M}^{3})\cdot\int\mathcal{D}\gamma_{1}\mathcal{D}\phi_{1}\exp\left(\pi i\int_{\mathcal{M}^{3}}\gamma_{1}\delta\phi_{1}+c_{2}\gamma_{1}+b_{2}\phi_{1}\right)\,.\end{split} (27)

The path integral over γ1,ϕ1\gamma_{1},\phi_{1} is the remnant of the U(1)U(1) YM path integral in this torsional setting. This is precisely the non-invertible operator found in Bhardwaj:2022yxj .

Fusion rules

Now that we have a full description of the symmetry defect, including its TQFT sector, we can derive the fusion rules for the extended operators in the Sc(4N)Sc(4N) theory, in particular showing that 𝒩(3)\mathcal{N}(\mathcal{M}^{3}) is a non-invertible operator of the Sc(4N)Sc(4N) theory. Since the TQFT that comes out of the brane dynamics is identical to the one conjectured in Bhardwaj:2022yxj , the rest of our derivation of the fusion rules can proceed exactly as in that paper (and the similar analysis in Kaidi:2021xfk ). We include the details of the argument for completeness and convenience for the reader, and then offer some comments reinterpreting some of the features of the computation from a brane perspective.

Consider first the fusion of two copies of 𝒩(3)\mathcal{N}(\mathcal{M}^{3}). Each defect comes with its own 2\mathbb{Z}_{2} TQFT, so we have two sets of dynamical fields:

𝒩(3)×𝒩(3)=𝒟γ1𝒟ϕ1𝒟γ1𝒟ϕ1exp(πi3γ1δϕ1+γ1δϕ1+c2(γ1+γ1)+b2(ϕ1+ϕ1)).\begin{split}\mathcal{N}(\mathcal{M}^{3})\times\mathcal{N}(\mathcal{M}^{3})=\int\mathcal{D}\gamma_{1}\mathcal{D}\phi_{1}\mathcal{D}\gamma_{1}^{\prime}\mathcal{D}\phi_{1}^{\prime}\exp\left(\pi i\int_{\mathcal{M}^{3}}\gamma_{1}\delta\phi_{1}+\gamma_{1}^{\prime}\delta\phi_{1}^{\prime}+c_{2}(\gamma_{1}+\gamma_{1}^{\prime})+b_{2}(\phi_{1}+\phi_{1}^{\prime})\right)\,.\end{split} (28)

Switching to new variables γ1\gamma_{1}, γ^1γ1+γ1\hat{\gamma}_{1}\coloneqq\gamma_{1}+\gamma_{1}^{\prime}, ϕ1\phi_{1}, ϕ^1ϕ1+ϕ1\hat{\phi}_{1}\coloneqq\phi_{1}+\phi_{1}^{\prime}, the action becomes

𝒩(3)×𝒩(3)=𝒟γ1𝒟ϕ1𝒟γ^1𝒟ϕ^1exp(πi3γ^1δϕ^1+γ^1δϕ1+γ1δϕ^1+c2γ^1+b2ϕ^1).\begin{split}\mathcal{N}(\mathcal{M}^{3})\times\mathcal{N}(\mathcal{M}^{3})=\int\mathcal{D}\gamma_{1}\mathcal{D}\phi_{1}\mathcal{D}\hat{\gamma}_{1}\mathcal{D}\hat{\phi}_{1}\exp\left(\pi i\int_{\mathcal{M}^{3}}\hat{\gamma}_{1}\delta\hat{\phi}_{1}+\hat{\gamma}_{1}\delta\phi_{1}+\gamma_{1}\delta\hat{\phi}_{1}+c_{2}\hat{\gamma}_{1}+b_{2}\hat{\phi}_{1}\right)\,.\end{split} (29)

We can integrate ϕ1\phi_{1} and γ1\gamma_{1} out, which imposes δγ^1=δϕ^1=0\delta\hat{\gamma}_{1}=\delta\hat{\phi}_{1}=0, so γ^1δϕ^1=0\hat{\gamma}_{1}\delta\hat{\phi}_{1}=0. We then have

𝒩(3)×𝒩(3)=𝒟γ^1𝒟ϕ^1exp(πi3c2γ^1+b2ϕ^1).\begin{split}\mathcal{N}(\mathcal{M}^{3})\times\mathcal{N}(\mathcal{M}^{3})=\int\mathcal{D}\hat{\gamma}_{1}\mathcal{D}\hat{\phi}_{1}\exp\left(\pi i\int_{\mathcal{M}^{3}}c_{2}\hat{\gamma}_{1}+b_{2}\hat{\phi}_{1}\right)\,.\end{split} (30)

Poincaré dualising γ^1\hat{\gamma}_{1} and ϕ^1\hat{\phi}_{1} to Γ,ΦH2(3;)\Gamma,\Phi\in H_{2}(\mathcal{M}^{3};\mathbb{Z}), this can be rewritten as

𝒩(3)×𝒩(3)=Γ,ΦH2(3;)D2𝖼,e(Γ)D2𝗌,m(Φ)\begin{split}\mathcal{N}(\mathcal{M}^{3})\times\mathcal{N}(\mathcal{M}^{3})=\sum_{\Gamma,\Phi\in H_{2}(\mathcal{M}^{3};\mathbb{Z})}D_{2}^{\mathsf{c},e}(\Gamma)D_{2}^{\mathsf{s},m}(\Phi)\end{split} (31)

where D2𝖼,eD_{2}^{\mathsf{c},e} and D2𝗌,mD_{2}^{\mathsf{s},m} are the 1-form symmetry generators of the Sc(4N)Sc(4N) theory. (The notation is explained below.) So 𝒩\mathcal{N} is indeed a non-invertible defect in the Sc(4N)Sc(4N) theory, since the right hand side is a sum of operators.

This was the derivation in Bhardwaj:2022yxj . Holographically, the physical meaning of the computation can be understood as follows. We have argued that the defect 𝒩(3)\mathcal{N}(\mathcal{M}^{3}) corresponds to a D3 wrapping 3×1\mathcal{M}^{3}\times{\mathbb{R}\mathbb{P}}^{1}, including its quantum dynamics. The effect of the quantum dynamics is to sum over induced charges, which in this case means summing over D3/F1 and D3/F1 bound states. (The precise way in which this sum happens involves, as shown above, a 2\mathbb{Z}_{2} gauge theory.) If there was no sum, but only a fixed induced charge (the trivial one, say), then taking the square would lead to a complete annihilation of the 2\mathbb{Z}_{2} charges, and therefore a trivial operator. Since there is a sum involved some of the cross-terms in the square of the sum will lead to incomplete annihilations, leaving a sum over F1 and D1 insertions along the worldvolume of the D3. The D3 charge is always there no matter the induced charge, and disappears, so only the sum over D1 and F1 insertions remains.

In order to show that this physical process does indeed produce (31), all we need to verify is that the symmetry generators of the Sc(4N)Sc(4N) theory are the F1 and D1. This is immediate, since they are the genuine lines in the SO(4N)SO(4N) theory, and we are gauging the symmetry they are charged under, so they become the magnetic symmetry generators in the gauged theory. It is also instructive to derive it from the Spin(4N)\mathrm{Spin}(4N) starting point. The 𝗌\mathsf{s}-spinor Wilson line W𝗌W_{\mathsf{s}}, given by a D5 brane wrapped on 4{\mathbb{R}\mathbb{P}}^{4}, is neutral under 2𝗌\mathbb{Z}_{2}^{\mathsf{s}} (recall our conventions from footnote 5), which is the symmetry that we gauge to go to Sc(4N)Sc(4N). So the corresponding charge operator, the D1 on pt~\widetilde{\mathrm{pt}}, survives as a charge operator on the Sc(4N)Sc(4N) theory. We have denoted it above by D2𝖼,eD_{2}^{\mathsf{c},e}. On the other hand the 𝖼\mathsf{c}-spinor and vector Wilson lines are not invariant, due to the presence of fundamental strings in them, which are not invariant under 2𝗌\mathbb{Z}_{2}^{\mathsf{s}}. So after gauging 2𝗌\mathbb{Z}_{2}^{\mathsf{s}} the fundamental string on a twisted point in 5{\mathbb{R}\mathbb{P}}^{5} becomes the second (magnetic) symmetry generator in the Sc(4N)Sc(4N) theory, which we have denoted above by D2𝗌,mD_{2}^{\mathsf{s},m}.

We are finally left with the task of determining the fusion of 𝒩(3)\mathcal{N}(\mathcal{M}^{3}) with the one-form symmetry generators D2𝖼,e(Γ)D_{2}^{\mathsf{c},e}(\Gamma) and D2𝗌,m(Φ)D_{2}^{\mathsf{s},m}(\Phi). Consider for example D2𝖼,e(Γ)D_{2}^{\mathsf{c},e}(\Gamma). We have just argued that it corresponds to a D1 brane on Γ×pt~\Gamma\times\widetilde{\mathrm{pt}}. Fusing it with 𝒩(3)\mathcal{N}(\mathcal{M}^{3}), which involves a sum over induced D1 branes wrapping the Poincaré dual PD[γ1]×pt~\operatorname{PD}[\gamma_{1}]\times\widetilde{\mathrm{pt}} to γ1\gamma_{1} amounts to shifting γ1γ1+PD[Γ]\gamma_{1}\to\gamma_{1}+\operatorname{PD}[\Gamma] in (27). But this can clearly be reabsorbed in a change of variables, giving back 𝒩(3)\mathcal{N}(\mathcal{M}^{3}). So

𝒩(3)×D2𝖼,e(Γ)=𝒩(3).\begin{split}\mathcal{N}(\mathcal{M}^{3})\times D_{2}^{\mathsf{c},e}(\Gamma)=\mathcal{N}(\mathcal{M}^{3})\,.\end{split} (32)

An identical argument shows

𝒩(3)×D2𝗌,m(Φ)=𝒩(3).\begin{split}\mathcal{N}(\mathcal{M}^{3})\times D_{2}^{\mathsf{s},m}(\Phi)=\mathcal{N}(\mathcal{M}^{3})\,.\end{split} (33)

We have shown that the 𝒩(3)\mathcal{N}(\mathcal{M}^{3}) operators of the Sc(4N)Sc(4N) theory are non-invertible, and are represented holographically by D3 branes. A small puzzle remains: our starting point was that the bulk of the holographic dual was the same for all global forms, so the same D3 brane appears in the bulk of all theories with the same local dynamics, including theories that are not expected to have non-invertible symmetries. The reason that the D3 does not lead to non-invertible symmetries in some cases has to do with boundary conditions (as it should, as this is the only thing that is different in the various cases). Consider for instance the SO(4N)SO(4N) theory, where the D3 on 1{\mathbb{R}\mathbb{P}}^{1} is also a symmetry operator, implementing the outer automorphism. As we push to the boundary, we obtain an operator of the form (27), but with a crucial difference: the IIB B2B_{2} and C2C_{2} fields have a Dirichlet boundary condition in this case, so they are not dynamical but instead they provide backgrounds for the global 1-form symmetries for the SO(4N)SO(4N) theory. So the term

𝒟γ1𝒟ϕ1exp(πi3γ1δϕ1+C2γ1+B2ϕ1)\begin{split}\int\mathcal{D}\gamma_{1}\mathcal{D}\phi_{1}\exp\left(\pi i\int_{\mathcal{M}^{3}}\gamma_{1}\delta\phi_{1}+C_{2}\gamma_{1}+B_{2}\phi_{1}\right)\end{split} (34)

in (27) (where we have capitalised B2B_{2} and C2C_{2} to indicate that now they are fixed background fields) does not depend on any dynamical field in the SO(4N)SO(4N) theory, so it is essentially trivial as an operator of the SO(4N)SO(4N) theory (it can be taken out of the path integral). In this case it is consistent to split it off from the invertible part D3(0)(3)D_{3}^{(0)}(\mathcal{M}^{3}), which can meaningfully be considered in isolation.

3.3 4d PO(4N)PO(4N) and Pin+\mathrm{Pin}^{+} non-invertibles

The other two theories with non-invertible symmetries in (6) can be analysed in a very similar way.

Let us start with the PO(4N)PO(4N) case. Here we gauge D3(0)D_{3}^{(0)} and D2B,eD_{2}^{B,e}, so we expect the non-invertible 2-surface operator to be associated with D2C,mD_{2}^{C,m}, which we argued above is given by a D5 brane wrapping 45{\mathbb{R}\mathbb{P}}^{4}\subset{\mathbb{R}\mathbb{P}}^{5}. We will need the twisted and untwisted cohomology groups of 4{\mathbb{R}\mathbb{P}}^{4}, these are

H(4;)={,0,2,0,2}H(4;~)={0,2,0,2,}.\begin{split}H^{*}({\mathbb{R}\mathbb{P}}^{4};\mathbb{Z})&=\{\mathbb{Z},0,\mathbb{Z}_{2},0,\mathbb{Z}_{2}\}\\ H^{*}({\mathbb{R}\mathbb{P}}^{4};\widetilde{\mathbb{Z}})&=\{0,\mathbb{Z}_{2},0,\mathbb{Z}_{2},\mathbb{Z}\}\,.\end{split} (35)

(The second line follows from analysing the twisted Gysin sequence in Thom1952 .) As above, F˘2\breve{F}_{2} is in the twisted sector, so it expands as γ1t˘1\gamma_{1}\otimes\breve{t}_{1}, but its magnetic dual F˘4D\breve{F}_{4}^{D} is now untwisted: this is needed to be able to write a kinetic term on the twisted 4{\mathbb{R}\mathbb{P}}^{4}. It therefore has an expansion of the form F˘4D=ϕ41+ϕ2u˘2+ϕ0u˘4\breve{F}_{4}^{D}=\phi_{4}\star 1+\phi_{2}\star\breve{u}_{2}+\phi_{0}\star\breve{u}_{4}.

In the electric frame the action on the D5 is of the form

S=2×4F˘7+F˘2F˘5+(12F˘22+124e˘)F˘3+\begin{split}S=\int_{\mathcal{M}^{2}\times{\mathbb{R}\mathbb{P}}^{4}}\breve{F}_{7}+\breve{F}_{2}\star\breve{F}_{5}+\left(\frac{1}{2}\breve{F}_{2}^{2}+\frac{1}{24}\breve{e}\right)\star\breve{F}_{3}+\ldots\end{split} (36)

where the missing terms are proportional to F˘1\breve{F}_{1}, so we will ignore them. The term proportional to F˘22F˘3\breve{F}_{2}^{2}\star\breve{F}_{3} does not contribute for degree reasons, as it goes as t˘13\breve{t}_{1}^{3}. The curvature term e˘F˘3\breve{e}\star\breve{F}_{3} could in principle contribute, but it does not depend on the electric field so it will not enter our considerations. We are left with the first two terms. The first one does clearly contribute, and leads to the expected “naive” 2-surface holonomy operator on 4\mathcal{M}^{4}, entirely analogously to the discussion around (18). The second term is also interesting. Given our expansion of F˘5\breve{F}_{5} above, there is a single non-vanishing contribution of the form

2πi2×4(γ1t˘1)(Nu˘5+a1u˘4+a3u˘2+N1˘)=πi2γ1a1,\begin{split}2\pi i\int_{\mathcal{M}^{2}\times{\mathbb{R}\mathbb{P}}^{4}}(\gamma_{1}\star\breve{t}_{1})\star(N\star\breve{u}_{5}+a_{1}\star\breve{u}_{4}+a_{3}\star\breve{u}_{2}+N\star\breve{1})=\pi i\int_{\mathcal{M}^{2}}\gamma_{1}a_{1}\,,\end{split} (37)

where we have used that 4{\mathbb{R}\mathbb{P}}^{4} is Poincaré dual to t1t_{1} in 5{\mathbb{R}\mathbb{P}}^{5}. This is the statement that worldvolume flux F2F_{2} induces D3 charge. The magnetic flux F4DF_{4}^{D} will induce F1 charge (by a generalisation of the analysis in Kimura:1999jb ), via a coupling of the form

2πi2×4F˘4DH˘3=πi2ϕ0b2,\begin{split}2\pi i\int_{\mathcal{M}^{2}\times{\mathbb{R}\mathbb{P}}^{4}}\breve{F}_{4}^{D}\star\breve{H}_{3}=\pi i\int_{\mathcal{M}^{2}}\phi_{0}b_{2}\,,\end{split} (38)

where we have used the expansion H˘3=b2t˘1\breve{H}_{3}=b_{2}\star\breve{t}_{1} as above.

All that remains is to obtain the prescription for how to sum over electric and magnetic fluxes. As above, flux non-commutativity can be used to argue that there is an effective 2\mathbb{Z}_{2} gauge theory on 2\mathcal{M}^{2} with action

iπ2γ1δϕ0.\begin{split}i\pi\int_{\mathcal{M}^{2}}\gamma_{1}\delta\phi_{0}\,.\end{split} (39)

The only new subtlety in this derivation comes from the fact that on 4{\mathbb{R}\mathbb{P}}^{4}, being non-orientable, the perfect torsional pairing is between a twisted class t˘1\breve{t}_{1} and an untwisted one u˘4\breve{u}_{4}. An easy way to verify the existence of such a coupling is to use Poincaré duality on 5{\mathbb{R}\mathbb{P}}^{5}:

4t˘1u˘4=5t˘16=12mod1.\begin{split}\int_{{\mathbb{R}\mathbb{P}}^{4}}\breve{t}_{1}\breve{u}_{4}=\int_{{\mathbb{R}\mathbb{P}}^{5}}\breve{t}_{1}^{6}=\frac{1}{2}\mod 1\,.\end{split} (40)

Putting all these terms together we obtain the topological action

STFTPO(4N)=iπ2γ1δϕ0+ϕ0b2+γ1a1\begin{split}S_{TFT}^{PO(4N)}=i\pi\int_{\mathcal{M}^{2}}\gamma_{1}\delta\phi_{0}+\phi_{0}b_{2}+\gamma_{1}a_{1}\end{split} (41)

which is precisely the action proposed in Bhardwaj:2022yxj . The fusion algebra can be derived as above.

Finally, in the 4d Pin+(4N)\mathrm{Pin}^{+}(4N) SYM theory we gauge D3(0)D_{3}^{(0)} and D2C,mD_{2}^{C,m}, so the non-invertible surface defects are realised as NS5-branes on 4×2{\mathbb{R}\mathbb{P}}^{4}\times\mathcal{M}^{2}. The worldvolume theory on the NS5 is just as on the D5, but the gauge fields couple to the S-dual supergravity fields. We can therefore write down the answer immediately from (41):

STFTPin+(4N)=iπ2γ1δϕ0+ϕ0c2+γ1a1.\begin{split}S_{TFT}^{\mathrm{Pin}^{+}(4N)}=i\pi\int_{\mathcal{M}^{2}}\gamma_{1}\delta\phi_{0}+\phi_{0}c_{2}+\gamma_{1}a_{1}\,.\end{split} (42)
Acknowledgements.
I thank Saghar Hosseini for related discussions, and Michele Del Zotto, Ben Heidenreich and Sakura Schäfer-Nameki for initial collaboration and discussions. This work is supported by the Simons Foundation via the Simons Collaboration on Global Categorical Symmetries, and by the STFC consolidated grant ST/T000708/1. I would also like to thank the Perimeter Institute, where this work was initiated during the 2022 workshop on Global Categorical Symmetries. Research at Perimeter Institute is supported in part by the Government of Canada through the Department of Innovation, Science and Economic Development Canada and by the Province of Ontario through the Ministry of Colleges and Universities.

References