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

aainstitutetext: Center for Theoretical Physics and College of Physics, Jilin University
2699 Qianjin St, Changchun 130012, People’s Republic of China
bbinstitutetext: Department of Physics and Astronomy, University of Waterloo
220 University Ave W, Waterloo, ON N2L 3G1, Canada
ccinstitutetext: Perimeter Institute For Theoretical Physics
31 Caroline St N, Waterloo, ON N2L 2Y5, Canada
ddinstitutetext: Department of Physics and Winnipeg Institute for Theoretical Physics, University of Winnipeg
515 Portage Avenue, Winnipeg, Manitoba R3B 2E9, Canada

Complexity, scaling, and a phase transition

Jiayue Yang d    and Andrew R. Frey [email protected] [email protected]
Abstract

We investigate the holographic complexity of CFTs compactified on a circle with a Wilson line, dual to magnetized solitons in AdS4 and AdS5. These theories have a confinement-deconfinement phase transition as a function of the Wilson line, and the complexity of formation acts as an order parameter for this transition. Through explicit calculation, we show that proposed complexity functionals based on volume and action obey a scaling relation with radius of the circle and further prove that a broad family of potential complexity functionals obeys this scaling behavior. As a result, we conjecture that the scaling law applies to the complexity of conformal field theories on a circle in more general circumstances.

Keywords:
AdS-CFT Correspondence

1 Introduction

Holographic complexity is the gravity-side dual of the (circuit) complexity of a gauge theory in the AdS/CFT correspondence, which is itself a measure of the Hilbert space distance of the gauge theory’s state from some reference state. There is by now a large literature studying holographic complexity (see arxiv:1403.5695 ; arXiv:1406.2678 ; arXiv:1509.07876 ; arXiv:1512.04993 for some foundational work and arXiv:2110.14672 for more references and a recent review); because there are continuous families of distance measures and possible reference states, it is clear that there should be many formulations of holographic complexity. Indeed, arXiv:2111.02429 ; arXiv:2210.09647 demonstrated that a large family of functionals in asymptotically AdS spacetimes have the expected behavior of complexity on thermal states.

As a result, it is the change of complexity in parameter space rather than the value that is physically important; often, as in black hole spacetimes, the time derivative of complexity is the quantity of interest. It is also important to understand the variation of complexity with other parameters, even in static situations. To that end, we examine the holographic complexity of a CFT on a circle as a function of the radius and the Wilson line around the circle (of a U(1)U(1) gauge field). The gravitational dual of the compactified CFT with Wilson line is a magnetized generalization of the AdS soliton first discussed by hep-th/9808079 ; hep-th/9803131 ; as in the standard AdS soliton, the periodic direction of the magnetized solitons shrinks at a finite AdS radius arXiv:1205.6998 ; arXiv:1807.07199 ; arXiv:2009.14771 ; arXiv:2104.14572 , meaning that the spacetime has no horizon. Moreover, if the Wilson line is below a critical value, the soliton has negative energy (compared to AdS spacetime), so it is the ground state with those boundary conditions, and the gauge theory exhibits confinement. At larger Wilson lines, the soliton energy becomes positive, so the system has a (zero-temperature) confinement-deconfinement phase transition — empty AdS with a periodic boundary direction and a Wilson line is the ground state.

Reynolds and Ross arXiv:1712.03732 evaluated the holographic complexity of the standard AdS soliton. Here, we extend their work by evaluating the complexity of formation of the magnetized solitons in AdS4 and AdS5, the difference in the complexity between the soliton and AdS with periodic boundary arXiv:1509.07876 ; arXiv:1512.04993 ; arXiv:1610.08063 . Since the ground state with large Wilson line corresponds to periodic AdS, the complexity of formation vanishes in the deconfined phase and acts as an order parameter for the phase transition. To our knowledge, this is the first investigation of complexity as an order parameter in a first-principles confining scenario in holography.111though see Ghodrati:2018hss ; arXiv:1808.08719 for studies of complexity in phenomenological models of confinement As emphasized in arXiv:2104.14572 , the state with critical Wilson line is also supersymmetric, so tuning the Wilson line also allows us to examine the effect of supersymmetry breaking on complexity.

After reviewing the magnetized soliton backgrounds in section 2, we will consider the complexity of formation as calculated both as a volume (of a maximal spatial slice and of a spacetime region) in section 3 and as an action in section 4. In addition to finding the dependence of complexity on the Wilson line, we demonstrate that the density of complexity of formation (per unit volume of the boundary CFT) scales as the inverse (d1)(d-1)st power of the circumference of the boundary circle.222Due to divergences in the UV, we must calculate complexity with a UV cutoff. Since the divergence structure is the same for AdS, the complexity of formation is finite; this scaling emerges as we take the cutoff to infinity. This scaling behavior persists for a large family of possible formulations for holographic complexity, as we discuss in section 5. Reynolds and Ross arXiv:1712.03732 first found this scaling for the standard soliton (and a consistent decrease in complexity with increasing radius for a lattice fermion model); our major results on the scaling are that it factors from the dependence on the Wilson line and a proof that it is universal for any holographic complexity functional obeying a few assumptions.

We summarize our results and make some conjectures in section 6.

2 Magnetized AdS solitons and phase transition

Here, we review the geometry of magnetized AdSd+1 solitons along with the corresponding physics of the dual gauge theory, largely following the results of arXiv:2104.14572 . Suppose that the gauge theory is on a flat spacetime with one spatial coordinate ϕ\phi periodically identified with period Δϕ\Delta\phi. In this case, the gauge theory can have a U(1)U(1) Wilson line around the ϕ\phi direction, and there are three regular solutions of the holographic dual gravity theory with a U(1)U(1) gauge field. (This is a bosonic subsector of gauged supergravity with 8 supercharges in either 4 or 5 dimensions, and we limit our calculations to d=3,4d=3,4.)

The first solution is AdS with a periodic direction and a Wilson line in the bulk. For reference, the metric and gauge field are

ds2=r2l2(dt2+dx2+dϕ2)+l2r2dr2,A=ΦΔϕdϕds^{2}=\frac{r^{2}}{l^{2}}\left(-dt^{2}+d\vec{x}^{2}+d\phi^{2}\right)+\frac{l^{2}}{r^{2}}dr^{2},\quad A=-\frac{\Phi}{\Delta\phi}d\phi (1)

for holonomy Φ\Phi of the boundary gauge field. The other two solutions are generalizations of the AdS soliton; both have metric (for d=3,4d=3,4)

ds2=r2l2(dt2+dx2+f(r)dϕ2)+l2r2f(r)dr2,f(r)1μl2rdQ2l2r2d2ds^{2}=\frac{r^{2}}{l^{2}}\left(-dt^{2}+d\vec{x}^{2}+f(r)d\phi^{2}\right)+\frac{l^{2}}{r^{2}f(r)}dr^{2},\quad f(r)\equiv 1-\frac{\mu l^{2}}{r^{d}}-\frac{Q^{2}l^{2}}{r^{2d-2}} (2)

and gauge field

A=7dQ(1rd21r0d2)dϕ,Q17dr0d2ΦΔϕ.A=\sqrt{7-d}\,Q\left(\frac{1}{r^{d-2}}-\frac{1}{r_{0}^{d-2}}\right)d\phi,\quad Q\equiv\frac{1}{\sqrt{7-d}}\frac{r_{0}^{d-2}\Phi}{\Delta\phi}. (3)

Like the AdS soliton, the bulk spacetimes terminate at r0r_{0}, the largest root of f(r)f(r); for the soliton geometries to be regular at r0r_{0}, μ\mu must yield

Δϕ=4πl2r02f(r0)\Delta\phi=\frac{4\pi l^{2}}{r_{0}^{2}f^{\prime}(r_{0})} (4)

(note that ϕ\phi has units of length, so Δϕ\Delta\phi is a circumference). There are two solutions to f(r0)=0f(r_{0})=0 and (4) for both values of dd we consider:

μ=r02d2Q2l2l2r0d2,r0=2πl2dΔϕ(1±1Φ2Φmax2)\mu=\frac{r_{0}^{2d-2}-Q^{2}l^{2}}{l^{2}r_{0}^{d-2}},\quad r_{0}=\frac{2\pi l^{2}}{d\Delta\phi}\left(1\pm\sqrt{1-\frac{\Phi^{2}}{\Phi_{max}^{2}}}\right) (5)

with Φmax=g(d)πl\Phi_{max}=g(d)\pi l, g(d)=(4610d)/dg(d)=\sqrt{(46-10d)/d}. The shorter soliton, ie, the solution with larger r0r_{0}, goes to the AdS soliton for Q=0Q=0, whereas the longer soliton merges with the periodic AdS solution in that limit. For completeness, we note that the two magnetic solitons have field strength

F=7d(d2)Qrd1drdϕ.F=-\sqrt{7-d}(d-2)\frac{Q}{r^{d-1}}dr\wedge d\phi. (6)

With the usual boundary conditions of fixed geometry and gauge field, the variables of the CFT are the periodicity Δϕ\Delta\phi and Wilson line Φ\Phi.

There are two methods to determine the energy density of each solution and therefore the ground state of the CFT with given periodicity and Wilson line. Because it will be useful later, we give the holographic renormalization argument here.333and a calculation following Hawking and Horowitz gr-qc/9501014 in appendix A. We make a Fefferman-Graham expansion fg of the asymptotic form of the bulk metric as

ds2=l2dz2z2+l2z2(γij(0)+n=dznγij(n))dxidxj;ds^{2}=l^{2}\frac{dz^{2}}{z^{2}}+\frac{l^{2}}{z^{2}}\left(\gamma^{(0)}_{ij}+\sum_{n=d}^{\infty}z^{n}\gamma^{(n)}_{ij}\right)dx^{i}dx^{j}; (7)

with z=0z=0 at the boundary (and xi=[t,x,ϕ]x^{i}=[t,\vec{x},\phi] in our case). Generally, the sum can include terms with 1n<d1\leq n<d as well as terms proportional to znln(z)z^{n}\ln(z) for ndn\geq d and dd even, although those terms vanish for the solitons we consider. Apparently

dzz=drrf(r)ln(z/l)=ln(r/l)+μl22drd+\int\frac{dz}{z}=-\int\frac{dr}{r\sqrt{f(r)}}\quad\Rightarrow\quad\ln(z/l)=-\ln(r/l)+\frac{\mu l^{2}}{2dr^{d}}+\cdots (8)

using a large radius expansion. Solving iteratively,

r=l2z(1+μl22dzdl2d+).r=\frac{l^{2}}{z}\left(1+\frac{\mu l^{2}}{2d}\frac{z^{d}}{l^{2d}}+\cdots\right). (9)

In holographic renormalization (see for example hep-th/0002230 ; arxiv:1211.6347 ), the boundary stress tensor is Tij=dld1γij(d)/16πG+\langle T_{ij}\rangle=dl^{d-1}\gamma^{(d)}_{ij}/16\pi G+\cdots, where GG is the (d+1)(d+1)-dimensional Newton constant in the AdS spacetime and the dots are terms that vanish on the soliton solutions. The energy density is therefore Ttt=μ/16πGld1\langle T_{tt}\rangle=-\mu/16\pi Gl^{d-1}. The long soliton (r0r_{0} small) has negative μ\mu and positive energy density always, while the short soliton has negative energy at small Φ\Phi and positive energy when ΦΦS2πl/[3]\Phi\geq\Phi_{S}\equiv 2\pi l/[\sqrt{3}] for d=3d=3 [d=4d=4]. Therefore, the system undergoes a phase transition from the short soliton at small Φ\Phi to periodic AdS for Φ>ΦS\Phi>\Phi_{S}. Since the ground state of the theory is either periodic AdS or the short soliton, we will henceforth always mean the short soliton when we discuss a soliton solution.

Because the soliton solutions cap off smoothly in the infrared rather than having a horizon, they exhibit confinement (like the standard Q=0Q=0 AdS soliton hep-th/9808079 ; hep-th/9803131 ). As a result, the phase transition from tuning the Wilson line Φ\Phi through ΦS\Phi_{S} is a confining/deconfining transition. (arXiv:2104.14572 also emphasized that the short soliton is supersymmetric at Φ=ΦS\Phi=\Phi_{S}.)

As noted above, the usual Dirichlet boundary condition on the gauge field δAμ=0\delta A_{\mu}=0 means that the CFT is defined in terms of fixed Wilson line Φ\Phi; in thermodynamics, this is the grand canonical ensemble. With an additional term in the action for the Maxwell field on the conformal boundary of AdS (or, precisely speaking, on a cutoff surface at fixed large radius rr), the natural variable of the CFT is QQ hep-th/9902170 ; arXiv:2104.14572 . The boundary term in the action means that the variational problem of the gauge field has Neumann boundary conditions (δFμν=0\delta F_{\mu\nu}=0). (Multiplication of the boundary action term by an arbitrary coefficient leads to mixed, or Robin, boundary conditions.) In the Euclidean theory, adding the boundary term gives the theory in the canonical ensemble. We will return to this point later.

3 Volume complexity of formation

The simplest proposals for holographic complexity are given in terms of a volume on the gravity side of the correspondence. The initial proposal for complexity, known as CV or “complexity=volume” arxiv:1403.5695 ; arXiv:1406.2678 , for an asymptotically AdS spacetime is CV=(d1)V/2π2GlC_{V}=(d-1)V/2\pi^{2}Gl, where VV is the volume of a maximal volume slice anchored at a fixed time on the boundary. While we have written the normalization of the complexity with the AdS scale ll, the choice of length scale is ambiguous. To compare the complexity of magnetized solitons with varying Wilson line, we should choose a fixed length scale, but the overall normalization is unimportant, so we choose the AdS scale for simplicity.444The remainder of the normalization constant is chosen so the AdS-Schwarzschild black hole saturates Lloyd’s bound on the time derivative of complexity at late times arXiv:1712.03732 .

All the backgrounds we consider have both time translation and time reversal symmetries, so the maximal volume surface is a constant time surface (we will always choose to measure complexity at boundary time t=0t=0). Because the volume diverges near the conformal boundary, we can integrate out only to a finite radius rmr_{m}. For both the magnetized solitons and periodic AdS, this volume is

V=VxΔϕr0rm𝑑r(rl)d2=VxΔϕd1rmd1r0d1ld2,V=V_{\vec{x}}\Delta\phi\int_{r_{0}}^{r_{m}}dr\left(\frac{r}{l}\right)^{d-2}=\frac{V_{\vec{x}}\Delta\phi}{d-1}\frac{r_{m}^{d-1}-r_{0}^{d-1}}{l^{d-2}}, (10)

with r00r_{0}\to 0 for periodic AdS, where VxV_{\vec{x}} is the volume along the x\vec{x} directions.

To find the complexity of formation of the soliton, we should subtract the corresponding volume of periodic AdS from (10) arXiv:1610.08063 . As a result, periodic AdS has by definition vanishing complexity of formation, and any soliton appears to have negative complexity of formation. There is a subtlety to consider, however. Rather than comparing volumes using the same cut-off radius rmr_{m}, we should compare them using the same Fefferman-Graham coordinate zz as defined in (8). If rmr^{\prime}_{m} is the cut off in periodic AdS and rmr_{m} in the soliton, then (9) implies that

rmd1=rmd1(1+μl2(d1)2drmd+).r_{m}^{d-1}=r_{m}^{\prime d-1}\left(1+\frac{\mu l^{2}(d-1)}{2dr_{m}^{\prime d}}+\cdots\right). (11)

Therefore, the divergent terms in the maximal volumes cancel as rmr_{m}\to\infty, meaning the density of complexity of formation is negative: 𝒞V=r0d1/2π2Gld1\mathcal{C}_{V}=-r_{0}^{d-1}/2\pi^{2}Gl^{d-1} in terms of the soliton radius r0r_{0}. Similarly, the difference in rmr_{m} between solitons with different values of Φ\Phi also does not contribute as we take rmr_{m}\to\infty.555As a note, suppose that we instead choose a cut off rmr^{\prime}_{m} in periodic AdS such that the proper circumference of ϕ\phi at the cut off is the same in both periodic AdS and the soliton. We also see that the difference between rmd1r_{m}^{\prime d-1} and rmd1r_{m}^{d-1} vanishes at infinite cut off.

Refer to caption
Figure 1: CV complexity of formation per unit volume for d=3d=3 (solid blue) and d=4d=4 (dashed red). Plotted curves are cVG𝒞V(Δϕ/l)d1c_{V}\equiv G\mathcal{C}_{V}(\Delta\phi/l)^{d-1}, so complexity scales with an inverse power of Δϕ\Delta\phi.

Two features of 𝒞V\mathcal{C}_{V} are immediately apparent. First, it scales as an inverse power of the circumference 𝒞VΔϕ(d1)\mathcal{C}_{V}\propto\Delta\phi^{-(d-1)}, and it is discontinuous at the phase transition since r00r_{0}\neq 0 at Φ=ΦS\Phi=\Phi_{S}.666See section 4 for the case that the natural CFT variable is QQ rather than Φ\Phi, however. Note also that the total complexity of formation scales as CVΔϕ(d2)C_{V}\propto\Delta\phi^{-(d-2)}. We show the dimensionless quantity cVG𝒞V(Δϕ/l)d1c_{V}\equiv G\mathcal{C}_{V}(\Delta\phi/l)^{d-1} as a function of Φ/Φmax\Phi/\Phi_{max} in figure 1. Note that cVc_{V} jumps from a negative value to zero at Φ=ΦS\Phi=\Phi_{S}, the supersymmetric value, because of the deconfining phase transition at that Wilson line. While the value of the complexity appears similar at ΦS\Phi_{S} for the two dimensionalities, it is not equal.

The negative value of 𝒞V\mathcal{C}_{V} is noteworthy; arXiv:2109.06883 showed that maximal volume slices in a wide variety of asymptotically AdS spacetimes are always larger than maximal slices in AdS. However, they assume maximal symmetry of the spatial slice of the conformal boundary, which the solitons (and periodic AdS) violate, so their theorems do not apply.

Another proposal for holographic complexity, known as CV2.0, is that C2VWDW/Gl2C_{2}\equiv V_{WDW}/Gl^{2}, where VWDWV_{WDW} is the spacetime volume of the Wheeler–DeWitt (WDW) patch arXiv:1610.02038 . The WDW patch is the spacetime region bounded by future- and past-directed lightsheets emitted from the boundary time slice where the complexity is measured; once again, we choose the length scale in the normalization as the AdS length ll for simplicity and note that the overall normalization of complexity is unimportant physically. In defining the WDW patch, we choose the boundary conditions that both lightsheets satisfy t=0t=0 at the radial cutoff r=rmr=r_{m} as a regulator, rather than taking t=0t=0 as rr\to\infty and cutting off the patch at r=rmr=r_{m}. See Akhavan:2019zax ; Omidi:2020oit for a demonstration that these regulators are equivalent when properly defined in the context of action complexity.

Refer to caption
Figure 2: The future lightsheets tF(r)t_{F}(r) for Φ/Φmax=\Phi/\Phi_{max}=0 (dot-dashed blue), 0.5 (dashed green), 1 (solid red). In this instance, Δϕrm/2=10π/3\Delta\phi\,r_{m}/\ell^{2}=10\pi/3 and d=4d=4. For reference, tF(r)t_{F}(r) for periodic AdS is dotted black.

The lightsheets on the WDW patch boundary are given by tF(r)t_{F}(r) and tP(r)=tF(r)t_{P}(r)=-t_{F}(r), where F,PF,P respectively designate the future and past lightsheets; they are given by dtF/dr=l2/r2f(r)dt_{F}/dr=-l^{2}/r^{2}\sqrt{f(r)} and extended along x\vec{x} and ϕ\phi. For fixed Δϕrm\Delta\phi\,r_{m}, increasing Φ\Phi lengthens the soliton (decreasing r0r_{0} for the short soliton). In addition, tF(r0)t_{F}(r_{0}) is finite but develops a cusp (infinite derivative). See figure 2 for a comparison of tFt_{F} for several values of Φ/Φmax\Phi/\Phi_{max} in d=4d=4 (we choose a small value of Δϕrm\Delta\phi\,r_{m} to emphasize the difference in the lightsheets near the soliton cap). Note that the end of each curve at the left is the end of the spacetime at r0r_{0}. For reference, we also show tF(r)=l2/rl2/rmt_{F}(r)=l^{2}/r-l^{2}/r_{m} for AdS spacetime.

Refer to caption
Figure 3: CV2.0 complexity of formation per boundary volume for d=3d=3 (solid blue) and d=4d=4 (dashed red). Plotted curves are c2G𝒞2(Δϕ/l)d1c_{2}\equiv G\mathcal{C}_{2}(\Delta\phi/l)^{d-1}, so complexity scales with an inverse power of Δϕ\Delta\phi.

Because the WDW patch volume is

VWDW=WDWdd+1xg=2VxΔϕr0rm𝑑r(rl)d1tF(r),V_{WDW}=\int_{WDW}d^{d+1}x\sqrt{-g}=2V_{\vec{x}}\Delta\phi\int_{r_{0}}^{r_{m}}dr\,\left(\frac{r}{l}\right)^{d-1}t_{F}(r), (12)

we expect the complexity of formation to be negative — the increase in tFt_{F} near r0r_{0} is not enough to compensate for the shortening of spacetime. In appendix B, we describe the numerical calculation of 𝒞2\mathcal{C}_{2}, the density of complexity of formation, holding the Wilson line Φ\Phi (rather than QQ) fixed. Direct calculation shows that 𝒞2Δϕ(d1)\mathcal{C}_{2}\propto\Delta\phi^{-(d-1)} as we remove the cutoff rmr_{m}\to\infty, which is notably the same scaling as 𝒞V\mathcal{C}_{V}. We therefore show c2G𝒞2(Δϕ/l)d1c_{2}\equiv G\mathcal{C}_{2}(\Delta\phi/l)^{d-1} in figure 3. Like the CV complexity of formation, the magnitude of c2c_{2} decreases as Φ\Phi increases and jumps to zero at the deconfining phase transition.

4 Action complexity of formation

The action complexity (“complexity=action” or CA) is given by CA=SWDW/πC_{A}=S_{WDW}/\pi, where SWDWS_{WDW} is the action evaluated on the WDW patch, including appropriate terms on the boundary of the patch arXiv:1509.07876 ; arXiv:1512.04993 ; arXiv:1609.00207 . Altogether, this action is SWDW=Sbulk+Sbdy+SjointS_{WDW}=S_{bulk}+S_{bdy}+S_{joint}, where

Sbulk\displaystyle S_{bulk} =\displaystyle= 116πGWDWdd+1xg(R+d(d1)l214FμνFμν),\displaystyle\frac{1}{16\pi G}\int_{WDW}d^{d+1}x\sqrt{-g}\left(R+\frac{d(d-1)}{l^{2}}-\frac{1}{4}F_{\mu\nu}F^{\mu\nu}\right), (13)
Sbdy\displaystyle S_{bdy} =\displaystyle= 18πGF𝑑λdd2x𝑑ϕγ(κF+ΘFln|βlΘF|)\displaystyle-\frac{1}{8\pi G}\int_{F}d\lambda\,d^{d-2}\vec{x}\,d\phi\sqrt{\gamma}\left(\kappa_{F}+\Theta_{F}\ln|\beta l\Theta_{F}|\vphantom{\frac{1}{2}}\right) (14)
+18πGP𝑑λdd2x𝑑ϕγ(κP+ΘPln|βlΘP|),\displaystyle+\frac{1}{8\pi G}\int_{P}d\lambda\,d^{d-2}\vec{x}\,d\phi\sqrt{\gamma}\left(\kappa_{P}+\Theta_{P}\ln|\beta l\Theta_{P}|\vphantom{\frac{1}{2}}\right),
Sjoint\displaystyle S_{joint} =\displaystyle= 18πGFPdd2x𝑑ϕγln|kFkP/2|.\displaystyle-\frac{1}{8\pi G}\int_{F\cap P}d^{d-2}\vec{x}\,d\phi\sqrt{\gamma}\ln\left|k_{F}\cdot k_{P}/2\right|. (15)

The action diverges, so we regulate by integrating for rrmr\leq r_{m} only and subtract the corresponding action of periodic AdS to cut-off radius rmr^{\prime}_{m} related to rmr_{m} by (11); note that the Wilson line does not contribute to the periodic AdS action because the field strength vanishes. The somewhat novel boundary and joint terms are by now thoroughly discussed in the literature, so we relegate a detailed description to appendix C. We note that β\beta is an arbitrary parameter that in principle could affect the complexity; however, it cancels in the complexity of formation when subtracting the periodic AdS action in the rmr_{m}\to\infty limit. (In d=4d=4, there is also a Chern-Simons term for the gauge field in the bulk action, but it vanishes for the soliton solutions.)

Like the boundary terms (14) for gravitational degrees of freedom, the WDW patch action includes a boundary term for the Maxwell field

ΔSbdy=ν16πGF𝑑λdd2x𝑑ϕγkFμAνFμν+ν16πGP𝑑λdd2x𝑑ϕγkFμAνFμν\Delta S_{bdy}=\frac{\nu}{16\pi G}\int_{F}d\lambda\,d^{d-2}\vec{x}\,d\phi\sqrt{\gamma}k_{F}^{\mu}A^{\nu}F_{\mu\nu}+\frac{\nu}{16\pi G}\int_{P}d\lambda\,d^{d-2}\vec{x}\,d\phi\sqrt{\gamma}k_{F}^{\mu}A^{\nu}F_{\mu\nu} (16)

with ν\nu an arbitrary constant arXiv:1901.00014 . This is the lightlike analog of the term on the AdS conformal boundary that changes the natural variable of the CFT. However, as especially emphasized by the “complexity==anything” program arXiv:2111.02429 ; arXiv:2210.09647 , there is no logical connection between ν\nu and the boundary conditions at r=rmr=r_{m}. In other words, ν\nu can take any value regardless of whether Φ\Phi or QQ is the natural variable of the CFT. Also, since the complexity is given by the action evaluated on shell, ΔSbdy\Delta S_{bdy} is equivalent to a multiple of the bulk Maxwell action (in the absence of sources) arXiv:1901.00014 . As a result, we can take the prefactor of FμνFμνF_{\mu\nu}F^{\mu\nu} to (2ν1)/4(2\nu-1)/4 rather than 1/4-1/4 in (13) rather than adding an additional boundary term.

The Wilson line in the CFT has several effects on the action complexity (evaluated at fixed periodicity Δϕ\Delta\phi). As noted previously, increasing Φ\Phi lengthens the soliton (decreasing r0r_{0}), tending to increase the magnitudes of SbulkS_{bulk} and SbdyS_{bdy}. The change in the shape of tF(r)t_{F}(r) additionally affects the expansion ΘF,P\Theta_{F,P} of the lightsheets at the WDW patch boundary, primarily in the interior region. The field strength itself gives contributes to the Lagrangian density (including a positive semidefinite contribution to the curvature for d3d\geq 3); this contribution is negative for small ν\nu but positive for ν>1/(d1)\nu>1/(d-1).

Refer to caption
(a) d=3d=3
Refer to caption
(b) d=4d=4
Figure 4: CA complexity of formation per unit boundary volume for d=3,4d=3,4 as labeled with ν=0\nu=0 (solid blue), ν=1/(d1)\nu=1/(d-1) (dashed red), and ν=1\nu=1 (dot-dashed green). Plotted curves are cA=G𝒞A(Δϕ/l)d1c_{A}=G\mathcal{C}_{A}(\Delta\phi/l)^{d-1} as a function of the Wilson line Φ\Phi.

Using the grand canonical ensemble variables Δϕ,Φ\Delta\phi,\Phi appropriate to standard boundary conditions on AdS, the density of complexity of formation 𝒞A\mathcal{C}_{A} scales as Δϕ(d1)\Delta\phi^{-(d-1)}, like the volume complexities 𝒞V,𝒞2\mathcal{C}_{V},\mathcal{C}_{2}, as we show in appendix C. Given that 𝒞A\mathcal{C}_{A} is considerably more intricate to calculate than either volume complexity, it may be surprising that it obeys the same simple scaling relation. Note that this scaling does not hold in the canonical ensemble in which the complexity is evaluated as a function of QQ. In that case, the scaling property of the action implies that 𝒞AΔϕ(d1)\mathcal{C}_{A}\propto\Delta\phi^{-(d-1)} along lines of constant QΔϕd1Q\Delta\phi^{d-1} (see equations (3,5)).

We show cAG𝒞A(Δϕ/l)d1c_{A}\equiv G\mathcal{C}_{A}(\Delta\phi/l)^{d-1} as a function of Φ\Phi in figure 4. In contrast to 𝒞V\mathcal{C}_{V} and 𝒞2\mathcal{C}_{2}, 𝒞A\mathcal{C}_{A} is positive semi-definite (equal zero in the deconfined phase). In the absence of the lightsheet boundary term for the Maxwell field (ν=0\nu=0), the complexity is maximized with no Wilson line and decreases toward Φ=ΦS\Phi=\Phi_{S} (solid blue curve). For ν=1/(d1)\nu=1/(d-1), the FμνFμνF_{\mu\nu}F^{\mu\nu} term cancels, so the corresponding (dashed red) curve shows only the geometric effect of the Wilson line (increased soliton length, altered WDW patch shape) on the complexity. Finally, for ν=1\nu=1 (dot-dashed green curve), the d=3d=3 𝒞A\mathcal{C}_{A} curve increases at Φ=0\Phi=0 and reaches a maximum for Φ<ΦS\Phi<\Phi_{S}. In this case, the (positive) contribution of the FμνFμνF_{\mu\nu}F^{\mu\nu} term increases with Φ\Phi, but the overall scaling with r0d1r_{0}^{d-1} which decreases with Φ\Phi eventually dominates. For d=4d=4 and ν=1\nu=1, cAc_{A} decreases with increasing Φ\Phi.

5 Scaling with circumference in general complexity

In the previous two sections, we have seen that the complexity of formation density 𝒞\mathcal{C} for magnetized AdS solitons scales as 1/Δϕd11/\Delta\phi^{d-1} when 𝒞\mathcal{C} is written as a function of the grand canonical variables Δϕ,Φ\Delta\phi,\Phi and the UV cutoff goes to infinity for the CV, CV2.0, and CA holographic complexity proposals. The details of these complexity measures and some of their properties (such as positivity or negativity on the soliton) are sufficiently different that it is reasonable to suspect a general principle rather than a coincidence at work.

Specifically, arXiv:2210.09647 argued that a large family of functionals of the form

C=1Gl2dd+1xg𝒢+1Gl+ddσh++1GlddσhC=\frac{1}{Gl^{2}}\int_{\mathcal{M}}d^{d+1}x\sqrt{-g}\,\mathcal{G}+\frac{1}{Gl}\int_{\mathcal{M}^{+}}d^{d}\sigma\sqrt{h}\,\mathcal{F}^{+}+\frac{1}{Gl}\int_{\mathcal{M}^{-}}d^{d}\sigma\sqrt{h}\,\mathcal{F}^{-} (17)

all satisfy basic properties expected of holographic complexity. Here \mathcal{M} is a bulk region in asymptotically AdS spacetime with ±\mathcal{M}^{\pm} respectively spacelike surfaces at the future and past portions of the boundary of \mathcal{M} (σ\sigma are worldvolume coordinates on ±\mathcal{M}^{\pm}) and 𝒢,±\mathcal{G},\mathcal{F}^{\pm} are dimensionless scalar functions of the metric (and curvatures). Since the magnetized solitons contain matter, we generalize to allow 𝒢,±\mathcal{G},\mathcal{F}^{\pm} to be functions of the gauge potential and field strength as well. \mathcal{M} itself optimizes a possibly different functional of the same form. The boundaries of ±\mathcal{M}^{\pm} form a joint at the time slice on the AdS boundary (more precisely, at the cutoff radius) where the CFT complexity is to be evaluated; arXiv:2210.09647 implies that additional contributions to CC on this joint are included, possibly in a manner that cancels boundary terms from the ±\mathcal{M}^{\pm} integrals (see appendix C for example). Further, there are limits of the optimization procedure choosing \mathcal{M} that leads to lightlike ±\mathcal{M}^{\pm}, so the WDW patch is a possible region \mathcal{M}.

We will argue here that a general complexity functional as in (17) leads to a density for complexity of formation 𝒞r0d1Δϕ(d1)\mathcal{C}\propto r_{0}^{d-1}\propto\Delta\phi^{-(d-1)} for the magnetized soliton backgrounds described in section 2, when the soliton is considered a function of the grand canonical variables Δϕ,Φ\Delta\phi,\Phi.

To make this argument, we assume that the complexity of empty periodic AdS is C=(aVxΔϕ/G)(rm/l)d1C=(aV_{\vec{x}}\Delta\phi/G)(r_{m}/l)^{d-1} when evaluated with a cutoff radius rmr_{m}, where aa is a numerical constant. This is the behavior seen in the CV, CV2.0, and CA proposals and is the strongest divergence possible for the bulk term in AdS since curvature invariants are constant and the embedding t±(r)t_{\pm}(r) of ±\mathcal{M}^{\pm} satisfies |rt±(r)|l2/r2|\partial_{r}t_{\pm}(r)|\leq l^{2}/r^{2} for spacelike and lightlike surfaces. The lack of subleading divergences (or a leading divergence of a lower power) means that we can write this term as

C=aVxΔϕG[(r0l)d1+(d1)r0rm𝑑rrd2],C=\frac{aV_{\vec{x}}\Delta\phi}{G}\left[\left(\frac{r_{0}}{l}\right)^{d-1}+(d-1)\int_{r_{0}}^{r_{m}}dr\,r^{d-2}\right], (18)

which is suitable for subtraction from the soliton’s complexity.

To proceed, we rewrite the metric (2) and gauge field (3,6) in terms of rescaled coordinates r~=r/r0\tilde{r}=r/r_{0} and t~=r0t/l\tilde{t}=r_{0}t/l, which arXiv:1712.03732 introduced:777We use these coordinates to rewrite integrals in dimensionless form in appendices B,C.

ds2\displaystyle ds^{2} =\displaystyle= r~2(dt~2+r02l2dx2+r02l2f(r~)dϕ2)+l2r~2f(r~)dr~2,f(r~)=11Q~2r~dQ~2r~2d2,\displaystyle\tilde{r}^{2}\left(-d\tilde{t}^{2}+\frac{r_{0}^{2}}{l^{2}}d\vec{x}^{2}+\frac{r_{0}^{2}}{l^{2}}f(\tilde{r})d\phi^{2}\right)+\frac{l^{2}}{\tilde{r}^{2}f(\tilde{r})}d\tilde{r}^{2},\quad f(\tilde{r})=1-\frac{1-\tilde{Q}^{2}}{\tilde{r}^{d}}-\frac{\tilde{Q}^{2}}{\tilde{r}^{2d-2}},
A\displaystyle A =\displaystyle= 7dr0Q~(1r~d21)dϕ,F=7d(d2)r0Q~r~d1dr~dϕ,\displaystyle\sqrt{7-d}\,r_{0}\tilde{Q}\left(\frac{1}{\tilde{r}^{d-2}}-1\right)d\phi,\quad F=-\sqrt{7-d}(d-2)\,\frac{r_{0}\tilde{Q}}{\tilde{r}^{d-1}}d\tilde{r}\wedge d\phi,
Q~2\displaystyle\tilde{Q}^{2} \displaystyle\equiv d2/47dg(d)2(ΦΦmax)2(1+1Φ2/Φmax2)2.\displaystyle\frac{d^{2}/4}{7-d}g(d)^{2}\left(\frac{\Phi}{\Phi_{max}}\right)^{2}\left(1+\sqrt{1-\Phi^{2}/\Phi_{max}^{2}}\right)^{-2}. (19)

Because we are comparing the density of complexity of formation (ie, complexity per volume VxΔϕV_{\vec{x}}\Delta\phi) across different values of Δϕ\Delta\phi and Φ\Phi, we cannot rescale the x\vec{x} and ϕ\phi coordinates. Likewise, the UV cutoff is at a fixed value of the un-rescaled radial coordinate r=rmr=r_{m} corresponding to fixed Fefferman-Graham coordinate.

In these coordinates, the tensor components gμν,Aμ,Fμνg_{\mu\nu},A_{\mu},F_{\mu\nu} contain precisely one factor of r0r_{0} for each leg along x\vec{x} or ϕ\phi. It is straightforward to check that the same is true of the Riemann tensor with all indices lowered. The Christoffel symbols are also such that covariant derivatives in the x,ϕ\vec{x},\phi directions add a factor of r0r_{0} but those in the t~,r~\tilde{t},\tilde{r} directions do not (this argument also relies on translation invariance in the x,ϕ\vec{x},\phi directions). As a result, any bulk scalar function 𝒢\mathcal{G} constructed from the metric, the Riemann tensor, the gauge field, or their covariant derivatives is independent of r0r_{0} — the inverse metric components needed for contractions cancels them. In addition, on either spatial surface ±\mathcal{M}^{\pm}, the normal vector nμn_{\mu} can have components only in the t~,r~\tilde{t},\tilde{r} directions due to translational invariance, and its only nonzero covariant derivatives have only those legs also. As a result, the extrinsic curvature has no factors of r0r_{0}. Similar arguments apply to the curvature κ\kappa and expansion Θ\Theta if ±\mathcal{M}^{\pm} is lightlike. Therefore, assuming ±\mathcal{F}^{\pm} are constructed only from the metric, gauge field, and those curvatures, they contain no explicit factors of r0r_{0} either. Altogether, any functional CC of the form (17) is proportional to r0d1r_{0}^{d-1}, with all those factors arising from determinants of the metric, except for a dependence on the region of integration.

Under these assumptions, the density of complexity of formation takes the form

𝒞=1G(r0l)d1{a+1rm/r0𝑑r~[(t~+(r~)t~(r~))𝔤(r~)+𝔣+(r~;t~+(r~))+𝔣(r~;t~(r~))]},\mathcal{C}=\frac{1}{G}\left(\frac{r_{0}}{l}\right)^{d-1}\left\{-a+\int_{1}^{r_{m}/r_{0}}d\tilde{r}\left[\left(\tilde{t}_{+}(\tilde{r})-\tilde{t}_{-}(\tilde{r})\right)\mathfrak{g}(\tilde{r})+\mathfrak{f}^{+}(\tilde{r};\tilde{t}_{+}(\tilde{r}))+\mathfrak{f}^{-}(\tilde{r};\tilde{t}_{-}(\tilde{r}))\right]\right\}, (20)

where 𝔤,𝔣±\mathfrak{g},\mathfrak{f}^{\pm} are the result of evaluating 𝒢,±\mathcal{G},\mathcal{F}^{\pm} on the soliton (with the AdS complexity subtracted), and t~±(r~)\tilde{t}_{\pm}(\tilde{r}) describe the surfaces ±\mathcal{M}^{\pm}. The dependence of 𝔣±\mathfrak{f}^{\pm} on t~±(r~)\tilde{t}_{\pm}(\tilde{r}) includes dependence on the derivatives of that embedding function. Here we give the case that ±\mathcal{M}^{\pm} are spacelike, but the timelike case is similar (see appendix C for the example of CA complexity). From the above discussion, 𝔤,𝔣±\mathfrak{g},\mathfrak{f}^{\pm} must be independent of r0r_{0}, so the only possible dependence on Δϕ\Delta\phi is through the upper limit of integration or implicitly in t~±\tilde{t}_{\pm}. However, t~±\tilde{t}_{\pm} are given by optimization of a functional of the form (17), in which dependence on r0r_{0} scales out of the integrand. Therefore, t~±\tilde{t}_{\pm} depends on r0r_{0} only through the boundary condition t~±(r~=rm/r0)=0\tilde{t}_{\pm}(\tilde{r}=r_{m}/r_{0})=0. However, 𝒞\mathcal{C} is by design convergent in the rmr_{m}\to\infty limit, so the integral cannot depend on r0r_{0} in that limit. The overall scaling is therefore 𝒞r0d1Δϕ(d1)\mathcal{C}\propto r_{0}^{d-1}\propto\Delta\phi^{-(d-1)}.

6 Discussion and conjecture

To summarize, we have evaluated several proposals for holographic complexity for a 3- or 4-dimensional CFT on a circle (times Minkowski spacetime) and a U(1)U(1) gauge field, which corresponds to magnetized AdS soliton backgrounds through gauge/gravity duality. One immediate observation is that “complexity=volume” proposals (both CV and CV2.0) have negative complexity of formation, while the complexity of formation for the “complexity=action” proposal is positive. As a result, we speculate that the reference states for the CV and CV2.0 proposals lie in a different region of the CFT’s Hilbert space than the reference states for CA complexity proposals (which may vary with varying boundary conditions for the gauge field). Another interpretation is that the volume and action complexity proposals have similar reference states but very different distance measures, though that seems less likely to lead to the observed sign change.

As emphasized by arXiv:2104.14572 , the compactified CFT exhibits two behaviors near the critical Wilson line value Φ=ΦS\Phi=\Phi_{S}: supersymmetry breaking for ΦΦS\Phi\neq\Phi_{S} and a transition from the magnetized soliton (confining) phase to the periodic AdS (deconfined) phase. The latter effect seems to be more important for the complexity; the complexity of formation acts like an order parameter, vanishing in the deconfined phase and changing discontinuously at the transition. This is perhaps unsurprising given the nature of the phase transition with periodic AdS as the ground state in the deconfined phase. On the other hand, the degree of supersymmetry breaking does not seem to influence the complexity; 𝒞\mathcal{C} is constant for Φ>ΦS\Phi>\Phi_{S} (by definition) but |𝒞||\mathcal{C}| grows generically as Φ\Phi decreases below ΦS\Phi_{S}. (With further decreases in Φ\Phi, |𝒞||\mathcal{C}| decreases again in some cases.)

A striking result is that the dependence of the complexity of formation on the grand canonical variables Δϕ,Φ\Delta\phi,\Phi factorizes, with the density behaving as 𝒞=h(Φ)/Δϕd1\mathcal{C}=h(\Phi)/\Delta\phi^{d-1}, extending the original observation of arXiv:1712.03732 for the soliton without magnetic field to magnetized solitons (in d=3,4d=3,4 only, however). This scaling law holds for the three specific forms of holographic complexity that we calculated; we further showed that the general holographic complexity functional as advocated by the “complexity=anything” program arXiv:2111.02429 ; arXiv:2210.09647 has the same scaling, under some mild assumptions. Therefore, we make a series of progressively stronger conjectures; the first is simply that this scaling extends to all magnetic AdS solitons of any dimensionality dd. Next, we conjecture that the density of complexity of formation for any dd-dimensional holographic field theory on a circle scales as the inverse (d1)(d-1)st power of the circumference, when the theory is written in terms of grand canonical variables. A stronger conjecture is that this scaling is true for the complexity of formation for any conformal field theory on a circle (or possibly even any quantum field theory).

A critical point in our arguments is the finiteness of the complexity of formation, which allows us to remove the UV cutoff (ie, take rmr_{m}\to\infty). This is manifestly true since the soliton is asymptotically AdS, so the complexity of the soliton has the same divergence structure as periodic AdS. However, for the complexity of formation to scale with the circumference as conjectured, we assumed that the complexity of AdS itself must be a pure divergence rmd1\propto r_{m}^{d-1}, the strongest possible divergence. If we presuppose that the density of complexity of formation should scale as Δϕ(d1)\Delta\phi^{-(d-1)}, this gives us instead an additional requirement for a functional to describe holographic complexity (beyond the switchback effect and linear growth at late time in black hole backgrounds). Since this seems like an unnatural condition for theories with a scale, it may indicate that complexity may not have a simple scaling behavior in non-conformal theories.

Next, we note that the Casimir energy density of a dd-dimensional field theory on one finite dimension scales as dd powers of the inverse length of that dimension. In fact, we can see this behavior in the holographic energy density as reviewed in section 2. It is intriguing to consider whether there is a deeper connection between the scaling of the Casimir energy and of complexity of formation. On the other hand, both scaling laws may simply arise because both are given by integration over field modes in the noncompact directions of the field theory.

Finally, it is worth considering the connection of our results to those of Andrews:2019hvq regarding the complexity of another type of soliton in AdS5. There are two major distinctions from our work: first, the solitons considered in Andrews:2019hvq have positive energy so are never the ground state; second, they are asymptotic to AdS5 in global coordinates, so the compact boundary dimension is instead one of the angular directions of the boundary S3S^{3} and cannot have arbitrary circumference. As a result, it is difficult to make a direct comparison with our work. Nonetheless, Andrews:2019hvq observe that the complexity of formation for those solitons also obeys a scaling law, in their case with the thermodynamic volume of the soliton. Since AlBalushi:2020rqe observes a similar scaling for black holes, it would be interesting in the future to determine what kind of relationship there may be between that scaling law and the one we have discussed here.

Acknowledgements.
JY would like to thank Robert Mann, Niayesh Afshordi, Haijun Wang, and Wencong Gan for encouragement in pursuing this project. The work of ARF was supported by the Natural Sciences and Engineering Research Council of Canada Discovery Grant program, grants 2020-00054. The work of JY was supported by the Mitacs Globalink program.

Appendix A Energy density from curvature

An alternative method to determine the energy density of asymptotically AdS backgrounds is to compare the extrinsic curvature of a boundary surface at a cutoff radius rmr_{m} to that of one with identical geometry in pure AdS, as in gr-qc/9501014 .

Specifically, we choose a surface at fixed tt and large rr. The energy of the spacetime is

E=18πGN(KK0)E=-\frac{1}{8\pi G}\int N(K-K_{0}) (21)

with the integral over the surface. Here, NN is the lapse, KK is the extrinsic curvature of the surface in the spacetime in question (the solitons in our case), and K0K_{0} the curvature of the surface in empty AdS. The curvature is given by K=nμμAK=n^{\mu}\partial_{\mu}A, where nμn^{\mu} is the radial unit vector and AA is the area of the fixed t,rt,r surface.

For the soliton backgrounds with a surface at r=rmr=r_{m}, the lapse is N=rm/lN=r_{m}/l, the radial unit vector has only one nontrivial component nr=rmf(rm)/ln^{r}=r_{m}\sqrt{f(r_{m})}/l, and the area of the surface is (in the rmr_{m}\to\infty limit)

A=VxΔϕ(rml)d1f(rm)NK=VxΔϕld+1[(d1)rmd+(1d2)μl2],A=V_{\vec{x}}\Delta\phi\left(\frac{r_{m}}{l}\right)^{d-1}\sqrt{f(r_{m})}\quad\Rightarrow\quad\int NK=\frac{V_{\vec{x}}\Delta\phi}{l^{d+1}}\left[(d-1)r_{m}^{d}+\left(1-\frac{d}{2}\right)\mu l^{2}\right], (22)

with VxV_{\vec{x}} as in the complexity. NK0NK_{0} is the same with μ=0\mu=0, except we must take Δϕf(rm)Δϕ\Delta\phi\to\sqrt{f(r_{m})}\Delta\phi (changing the asymptotic periodicity), so the two surfaces have the same proper size at the cutoff radius rmr_{m}, as emphasized in hep-th/9808079 . In the end, we find E=VxΔϕμ/16πGld1E=-V_{\vec{x}}\Delta\phi\mu/16\pi Gl^{d-1}, consistent with the holographic stress tensor.

Rather than match the proper size of the cutoff surfaces in the solition and periodic AdS spacetimes, we could instead take cutoff surfaces at the same Fefferman-Graham coordinate zz, as in the complexity of formation. Amusingly, this gives the same result for the energy for the magnetized solitons, though it is not a general feature of this formalism for the energy of asymptotically AdS spacetime.

Appendix B Spacetime volume calculations

We determine tF(r)=tP(r)t_{F}(r)=-t_{P}(r) by numerical integration with t=0t=0 at r=rmr=r_{m}; it turns out that the Q=0Q=0 case can be written in terms of hypergeometric functions, but that does not generalize. It is straightforward to see that the r0tF/lr_{0}t_{F}/l vs r/r0r/r_{0} curve falls in a two-parameter family depending on Δϕrm\Delta\phi\,r_{m} and Φ/Φmax\Phi/\Phi_{max}; the key point to note is that f(r)f(r) is independent of Δϕ\Delta\phi when written in terms of a dimensionless radial variable r~=r/r0\tilde{r}=r/r_{0} for fixed Φ\Phi. Then

tF(r)=l2r0r~rm/r0dr~r~2f(r~)lr0t~F(r~),t_{F}(r)=\frac{l^{2}}{r_{0}}\int_{\tilde{r}}^{r_{m}/r_{0}}\frac{d\tilde{r}^{\prime}}{\tilde{r}^{\prime 2}\sqrt{f(\tilde{r}^{\prime})}}\equiv\frac{l}{r_{0}}\tilde{t}_{F}(\tilde{r}), (23)

where t~F\tilde{t}_{F} depends on the circumference only through the boundary condition that t~F=0\tilde{t}_{F}=0 at r~=rm/r0Δϕrm\tilde{r}=r_{m}/r_{0}\propto\Delta\phi\,r_{m}. We evaluate tF(r)t_{F}(r) numerically.

To find the complexity of formation, we also need the WDW patch volume for periodic AdS, which is

VAdS=2VxΔϕ0rm𝑑r(rl)d1(l2rl2rm)=2VxΔϕl2d(d1)(rml)d1.V_{AdS}=2V_{\vec{x}}\Delta\phi\int_{0}^{r^{\prime}_{m}}dr\left(\frac{r}{l}\right)^{d-1}\left(\frac{l^{2}}{r}-\frac{l^{2}}{r_{m}}\right)=\frac{2V_{\vec{x}}\Delta\phi\,l^{2}}{d(d-1)}\left(\frac{r^{\prime}_{m}}{l}\right)^{d-1}. (24)

As before, the difference between rmr^{\prime}_{m} and rmr_{m} does not change VAdSV_{AdS} as we allow rmr_{m}\to\infty based on equation (11).

Since we must evaluate VWDWV_{WDW} numerically, we write VAdSV_{AdS} as

VAdS=2VxΔϕld{ld1(r0l)d1+r0rm𝑑r(rl)d2}V_{AdS}=\frac{2V_{\vec{x}}\Delta\phi\,l}{d}\left\{\frac{l}{d-1}\left(\frac{r_{0}}{l}\right)^{d-1}+\int_{r_{0}}^{r_{m}}dr\left(\frac{r}{l}\right)^{d-2}\right\} (25)

and subtract integrands. The boundary density of the complexity of formation is

𝒞2\displaystyle\mathcal{C}_{2} =\displaystyle= 2G{1d(d1)(r0l)d1+r0rm𝑑r(rl)d1[tF(r)l21d1r]}\displaystyle\frac{2}{G}\left\{-\frac{1}{d(d-1)}\left(\frac{r_{0}}{l}\right)^{d-1}+\int_{r_{0}}^{r_{m}}dr\left(\frac{r}{l}\right)^{d-1}\left[\frac{t_{F}(r)}{l^{2}}-\frac{1}{d}\frac{1}{r}\right]\right\} (26)
=\displaystyle= 2G(r0l)d1{1d(d1)+1rm/r0𝑑r~r~d1[t~F(r~)l1d1r~]}.\displaystyle\frac{2}{G}\left(\frac{r_{0}}{l}\right)^{d-1}\left\{-\frac{1}{d(d-1)}+\int_{1}^{r_{m}/r_{0}}d\tilde{r}\,\tilde{r}^{d-1}\left[\frac{\tilde{t}_{F}(\tilde{r})}{l}-\frac{1}{d}\frac{1}{\tilde{r}}\right]\right\}.

We see that the factor in braces depends on Δϕ\Delta\phi only through the upper limit on the integral and the boundary condition on t~F\tilde{t}_{F}, and it only appears in the combination rmΔϕr_{m}\Delta\phi. As a result, the quantity in braces is independent of Δϕ\Delta\phi in the limit that we remove the UV cutoff rmr_{m}\to\infty. Then 𝒞2Δϕ(d1)\mathcal{C}_{2}\propto\Delta\phi^{-(d-1)} due to the overall prefactors of r0r_{0}.

To evaluate the numerical integral, we verified that it changes less than 1% between values rmΔϕ=400l2r_{m}\Delta\phi=400l^{2} and 500l2500l^{2} for Φ=0\Phi=0, which has the smallest ratio rm/r0r_{m}/r_{0} at any Δϕ\Delta\phi. We therefore use rmΔϕ=500l2r_{m}\Delta\phi=500l^{2}.

Appendix C Action calculations

Here we give details of our calculation of the action in equations (13,14,15). λ\lambda is the lightlike parameter along each lightsheets on the future and past boundaries, with λ\lambda increasing into the future in each case. The vectors kF,kPdx/dλk_{F},k_{P}\equiv dx/d\lambda are the future-directed lightlike vectors along the corresponding lightsheets, and κF,P\kappa_{F,P} is defined by kμμkν=κkνk^{\mu}\nabla_{\mu}k^{\nu}=\kappa k^{\nu}. γij\gamma_{ij} is the induced metric on the spatial slices at fixed λ\lambda along the lightsheet, and ΘF,ΘP\Theta_{F},\Theta_{P} are the expansions of the lightsheets defined by the logarithmic derivative of γ\sqrt{\gamma} with respect to λ\lambda. The ΘF,P\Theta_{F,P} terms are counterterms needed to ensure reparameterization invariance of SbdyS_{bdy}, and β\beta is an arbitrary parameter that sets the length scale of the counterterm. If we take β\beta the same for both lightsheets and for all backgrounds, we will see that β\beta cancels in the complexity of formation.

Again, see figure 2 for a sample of tF(r)t_{F}(r) at fixed Δϕrm\Delta\phi\,r_{m} and a variety of Φ/Φmax\Phi/\Phi_{max}. The main difference in the shape of the lightsheets is due to the reduction in r0r_{0} as Φ\Phi increases; tF(1/r1/rm)t_{F}\to(1/r-1/r_{m}) at large radius, as in periodic AdS. It is also straightforward to see that κF,P=0\kappa_{F,P}=0 if we choose λ\lambda such that dr/dλ=αF,Pf(r)dr/d\lambda=\mp\alpha_{F,P}\sqrt{f(r)} for arbitrary positive dimensionless constants αF,P\alpha_{F,P}. These derivatives give the μ=t\mu=t and rr components of kF,Pμk_{F,P}^{\mu}. The induced metric for spatial slices of either lightsheet gives γ=f(r)(r/l)d1\sqrt{\gamma}=\sqrt{f(r)}(r/l)^{d-1}.

Because dr/dλdr/d\lambda has opposite sign on the past and future lightsheets, we can see that their contributions to SbdyS_{bdy} are identical when written as an integral over radius. For notational convenience, we define

ΘF,P(r)=drdλ(d1r+12f(r)dfdr)drdλΘ(r).\Theta_{F,P}(r)=\frac{dr}{d\lambda}\left(\frac{d-1}{r}+\frac{1}{2f(r)}\frac{df}{dr}\right)\equiv\frac{dr}{d\lambda}\Theta(r)\ . (27)

The identity γΘ=dγ/dr\sqrt{\gamma}\Theta=d\sqrt{\gamma}/dr allows us to convert several boundary terms to joint terms, and we perform an additional integration by parts after adding and subtracting ln(r/l)\ln(r/l) to cancel a logarithmic divergence between the joint and boundary terms. All terms including αF,P\alpha_{F,P} cancel between SbdyS_{bdy} and SjointS_{joint}.

To simplify the bulk contribution, we use the trace of the Einstein equation, which gives the well-known identity

R=d(d+1)l2+14d3d1FμνFμν.R=-\frac{d(d+1)}{l^{2}}+\frac{1}{4}\frac{d-3}{d-1}F_{\mu\nu}F^{\mu\nu}. (28)

The Maxwell term without the additional boundary term (ν=0\nu=0) is however more negative, so the field strength gives an overall negative contribution to the bulk action. With the Maxwell boundary term, the terms proportional to (Fμν)2(F_{\mu\nu})^{2} give a positive contribution for ν>1/(d1)\nu>1/(d-1).

After simplification, the total soliton action is therefore

Sbulk+Sbdy+Sjoint\displaystyle S_{bulk}+S_{bdy}+S_{joint} =\displaystyle= VxΔϕ8πGr0rmdr(rl)d1{2f(r)[1r+Θ(r)(ln(rΘ(r))+12ln(f(r)))]\displaystyle\frac{V_{\vec{x}}\Delta\phi}{8\pi G}\int_{r_{0}}^{r_{m}}dr\,\left(\frac{r}{l}\right)^{d-1}\left\{2\sqrt{f(r)}\left[\frac{1}{r}+\Theta(r)\left(\ln(r\Theta(r))+\frac{1}{2}\ln(f(r))\right)\right]\right. (29)
tF(r)[2dl2+(1ν(d1))(d2)2d1r02d4Φ2Δϕ21r2d2]}\displaystyle\left.-t_{F}(r)\left[\frac{2d}{l^{2}}+\left(1-\nu(d-1)\vphantom{\frac{1}{2}}\right)\frac{(d-2)^{2}}{d-1}\frac{r_{0}^{2d-4}\Phi^{2}}{\Delta\phi^{2}}\frac{1}{r^{2d-2}}\right]\right\}
+VxΔϕ4πG(rml)d1f(rm)ln(β).\displaystyle+\frac{V_{\vec{x}}\Delta\phi}{4\pi G}\left(\frac{r_{m}}{l}\right)^{d-1}\sqrt{f(r_{m})}\ln(\beta).

Note that the arbitrary constants αF,P\alpha_{F,P} cancel between boundary and joint terms. Because f(r)1f(r)\to 1 and rΘd1r\Theta\to d-1 at large rr, the divergences are the same as empty periodic AdS. Therefore, we remove the divergences by subtracting the action of periodic AdS with cut-off radius rmr^{\prime}_{m} as in (11), which gives the complexity of formation. Since rm=rm(1+𝒪(1/rmd))r_{m}=r^{\prime}_{m}(1+\mathcal{O}(1/r_{m}^{\prime d})), we take rm=rmr^{\prime}_{m}=r_{m}; see further discussion below. To aid in numerical cancellation of divergences, we write the action of periodic AdS as

SAdS=VxΔϕ4πG{(rml)d1ln(β)+ln(d1)[(r0l)d1+r0rm𝑑r(d1l)(rl)d2]}.S_{AdS}=\frac{V_{\vec{x}}\Delta\phi}{4\pi G}\left\{\left(\frac{r_{m}}{l}\right)^{d-1}\ln(\beta)+\ln(d-1)\left[\left(\frac{r_{0}}{l}\right)^{d-1}+\int_{r_{0}}^{r_{m}}dr\,\left(\frac{d-1}{l}\right)\left(\frac{r}{l}\right)^{d-2}\right]\right\}. (30)

The β\beta-dependent terms cancel for rmr_{m}\to\infty because f1𝒪(rmd)f-1\sim\mathcal{O}(r_{m}^{-d}), so we henceforth drop them.

A particular rescaling of the integration variable as described in arXiv:1712.03732 is useful both for numerical calculation of the action and for proving scaling properties of the complexity. The function ff written in terms of r~=r/r0\tilde{r}=r/r_{0} is independent of r0r_{0} and therefore of Δϕ\Delta\phi (after solving for μ\mu and QQ in terms of Δϕ\Delta\phi and Φ\Phi), as discussed in appendix B. Similarly, t~F(r~)r0tF/l\tilde{t}_{F}(\tilde{r})\equiv r_{0}t_{F}/l depends on Δϕ\Delta\phi only through the boundary condition at rm/r0rmΔϕr_{m}/r_{0}\propto r_{m}\Delta\phi. Then the total action is

S=Sbulk+Sbdy+SjointSAdS=VxΔϕ8πG(r0l)d1[ln(d1)+I(rm/r0,Φ)],S=S_{bulk}+S_{bdy}+S_{joint}-S_{AdS}=\frac{V_{\vec{x}}\Delta\phi}{8\pi G}\left(\frac{r_{0}}{l}\right)^{d-1}\left[-\ln(d-1)+I(r_{m}/r_{0},\Phi)\vphantom{\frac{1}{2}}\right], (31)

where I(rm/r0,Φ)I(r_{m}/r_{0},\Phi) is the integral

I(rm/r0,Φ)\displaystyle I(r_{m}/r_{0},\Phi) =\displaystyle= 1rm/r0dr~r~d1{2f(r~)[1r~+Θ~(r~)(ln(r~Θ~(r~))+12ln(f(r~)))]\displaystyle\int_{1}^{r_{m}/r_{0}}d\tilde{r}\,\tilde{r}^{d-1}\left\{2\sqrt{f(\tilde{r})}\left[\frac{1}{\tilde{r}}+\tilde{\Theta}(\tilde{r})\left(\ln(\tilde{r}\tilde{\Theta}(\tilde{r}))+\frac{1}{2}\ln(f(\tilde{r}))\right)\right]\right. (32)
t~F(r~)l[2d+(1ν(d1))d2(d2)2d1g(d)24r~2d2Φ2/Φmax2(1+1Φ2/Φmax2)2]\displaystyle\left.-\frac{\tilde{t}_{F}(\tilde{r})}{l}\left[2d+\left(1-\nu(d-1)\vphantom{\frac{1}{2}}\right)\frac{d^{2}(d-2)^{2}}{d-1}\frac{g(d)^{2}}{4\tilde{r}^{2d-2}}\frac{\Phi^{2}/\Phi_{max}^{2}}{\left(1+\sqrt{1-\Phi^{2}/\Phi_{max}^{2}}\right)^{2}}\right]\right.
2ln(d1)d1r~}\displaystyle\left.-2\ln(d-1)\frac{d-1}{\tilde{r}}\right\}

and Θ~(r~)=r0Θ=(d1)/r~(df/dr~)/2f\tilde{\Theta}(\tilde{r})=r_{0}\Theta=(d-1)/\tilde{r}-(df/d\tilde{r})/2f is the dimensionless expansion. A key property to note is that, when written in terms of the Wilson line Φ\Phi, II also depends on Δϕ\Delta\phi only through the limit of integration rm/r0rmΔϕr_{m}/r_{0}\propto r_{m}\Delta\phi. But the subtraction of SAdSS_{AdS} means that II converges as the upper limit goes to infinity — to a value independent of Δϕ\Delta\phi. As a result, when we remove the cut off, the density of complexity of formation is 𝒞A(Δϕ/l)(d1)\mathcal{C}_{A}\propto(\Delta\phi/l)^{-(d-1)} for standard boundary conditions with the scaling coming from the prefactor of r0d1r_{0}^{d-1}.

Since the complexity of formation is the finite rmr_{m}\to\infty value, we have evaluated (32) for Φ=0\Phi=0, where r0r_{0} is largest, and checked where the integral has converged sufficiently. In both d=3d=3 and d=4d=4, I(rm/r0,0)I(r_{m}/r_{0},0) changes by less than 1% between rmΔϕ=400l2r_{m}\Delta\phi=400l^{2} and 500l2500l^{2} for Φ=0\Phi=0 (which is the shortest soliton). We therefore use rmΔϕ=500l2r_{m}\Delta\phi=500l^{2} for all numerical calculations of the complexity.

References