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

The X-ray transform on a generic family of smooth curves

Yang Zhang
Abstract.

We study the X-ray transform over a generic family of smooth curves in 2\mathbb{R}^{2} with a Riemannian metric gg. We show that the singularities cannot be recovered from local data in the presence of conjugate points, and therefore artifacts may arise in the reconstruction. We perform numerical experiments to illustrate the results.

1. Introduction

In this work, we consider the integral transform over a generic family of smooth curves with conjugate points. Let M2M\subset\mathbb{R}^{2} be a bounded domain with a Riemannian metric gg. We consider a family Γ\Gamma of smooth curves satisfying the following properties.

  • (A1)

    For each (x,v)TM0(x,v)\in TM\setminus 0, there is exactly one unique curve γ(t)Γ\gamma(t)\in\Gamma passing xx in the direction of vv. We can assume γ(0)=x\gamma(0)=x and γ˙(0)=μ(x,v)v\dot{\gamma}(0)=\mu(x,v)v with a smooth function μ(x,v)>0\mu(x,v)>0 and denote it by γx,v(t)\gamma_{x,v}(t).

  • (A2)

    Suppose γx,v(t)\gamma_{x,v}(t) depends on (x,v)(x,v) smoothly, and thus we assume it solves a second order ODE

    Dtγ˙x,v(t)=G(γx,v,γ˙x,v),D_{t}\dot{\gamma}_{x,v}(t)=G(\gamma_{x,v},\dot{\gamma}_{x,v}),

    where DtD_{t} is the covariant derivative along γ˙\dot{\gamma} and the generator GG is a smooth map invariantly defined on TMTM with the transformation law

    G~i(x~i,y~i)=x~ixjGj(xi,yi), where y~i=x~ixjyj,\widetilde{G}^{i}(\tilde{x}^{i},\tilde{y}^{i})=\frac{\partial\tilde{x}^{i}}{\partial x^{j}}G^{j}(x^{i},y^{i}),\quad\text{ where }\tilde{y}^{i}=\frac{\partial\tilde{x}^{i}}{\partial x^{j}}y^{j},

    under the coordinate change xix~ix^{i}\mapsto\tilde{x}^{i} of MM.

We emphasize that the second property describes a large family of curves, including geodesics in Riemannian surfaces, geodesics in Finsler spaces, and magnetic geodesics, for more details see [3, Chapter 4], [20, Chapter 3], [5], and [6]. For such a curve, we can freely shift the parameter but rescaling it may give us a different one, since the trajectory depends on the energy level. This implies the first property is necessary for us to have a geodesic-like family of curves.

We define the integral transform along curves in Γ\Gamma as

Iwf(γ)=w(γ(t),γ˙(t))f(γ(t))𝑑t,γΓ,I_{w}f(\gamma)=\int w(\gamma(t),\dot{\gamma}(t))f(\gamma(t))dt,\quad\gamma\in\Gamma,

for any distribution ff compactly supported in MM. This integral transform has been studied in [7]. It is shown that if Γ\Gamma is analytic and regular, i.e., if GG is analytic and NΓN^{*}\Gamma covers TMT^{*}M without conjugate points, then IwI_{w} with an analytic and nonvanishing weight function is injective. A stability result for the localized normal operator is established there. In [2], the integral transform of functions and 11-forms over such curves in an oriented Finsler surface is studied and the injectivity is proved if there are no conjugate points.

However, when conjugate points exist, in many cases the microlocal stability are lost. The phenomenon of cancellation of singularities due to the existence of conjugate points are fully studied in the case of synthetic aperture radar imaging in [23], the X-ray transforms over a family of geodesic-like curves in [22], the geodesic ray transforms in [13, 18, 12]. In particular, [18] shows that regardless of the type of the conjugate points, the geodesic ray transforms on Riemannian surfaces are always unstable and we have loss of all derivatives, which leads to the artifacts in the reconstruction near pairs of conjugate points. The same phenomenon occur in the inversion of broken ray transforms with conjugate points, see [25]. For more references about the integral transform over geodesics, magnetic geodesics, or a general family of curves, see [1, 14, 16, 19, 26].

This work is inspired by [18] and we prove an analog of cancellation of singularities for IwI_{w} in Theorem 3.1 and 3.3. We prove that the local problem is ill-posed if there are conjugate points, i.e., singularities conormal to smooth curves in Γ\Gamma cannot be recovered uniquely. The microlocal kernel is described in Theorem 3.3. In Section 2, we introduce λ\lambda-geodesics to describe this generic family of smooth curves and we define the conjugate points. In Section 3, we consider the recovery of ff from the integral transform IwI_{w} using the local data. Here by the local data, we mean the integral transform is only known in a small neighborhood of a fixed curve γ(t)Γ\gamma(t)\in\Gamma. The transform IwI_{w} is an Fourier integral operator (FIO) with its canonical relation described in Proposition 6. We microlocalize the problem and prove Theorem 3.1 and Theorem 3.3. In Section 4, we present several examples of such family of curves and show the existence of conjugate points. In Section 5, we illustrate the artifacts arising in the reconstruction by numerical experiments. In particular, we show for the family of smooth curves in Example 1, the singularities of ff can be recovered without artifacts if we consider the global data and have the prior knowledge that ff is a distribution with compact support.

Acknowledgments

The author would like to thank Plamen Stefanov for suggesting this problem and for numerous helpful discussions with him throughout this project, and to thank Gabriel Paternain and Gunther Uhlmann for helpful suggestions. Part of this research was supported by NSF Grant DMS-1600327, the Simons Travel Grants, and was performed while the author was visiting IPAM.

2. The λ\lambda-geodesics

2.1. The generator GG

Locally we can rewrite the second order ODE as

γ¨x,vi(t)=Gi(γx,v,γ˙x,v)γ˙x,vkγ˙x,vlΓkli,i=1,2,\ddot{\gamma}^{i}_{x,v}(t)=G^{i}(\gamma_{x,v},\dot{\gamma}_{x,v})-\dot{\gamma}_{x,v}^{k}\dot{\gamma}_{x,v}^{l}\Gamma_{kl}^{i},\quad i=1,2,

where Γkli\Gamma_{kl}^{i} is the Christoffel symbol. The generator GG induces a vector field 𝐆\mathbf{G} on TMTM given by

𝐆=yixi+(Gi(x,y)ykylΓkli)yi\mathbf{G}=y^{i}\frac{\partial}{\partial x^{i}}+(G^{i}(x,y)-y^{k}y^{l}\Gamma_{kl}^{i})\frac{\partial}{\partial y^{i}}

in a local chart {xi,yi}\{x^{i},y^{i}\} of TMTM. It preserves the form under coordinate changes by the following arguments and thus is invariantly defined, see also (30) in [7]. Indeed, consider a local coordinate chart {xi,yi}\{x^{i},y^{i}\} near a fixed point (x,y)TM(x,y)\in TM. Under the coordinate change xix~ix^{i}\mapsto\tilde{x}^{i} of MM, we have y~i=(x~i/xj)yj\tilde{y}^{i}=({\partial\tilde{x}^{i}}/{\partial x^{j}})y^{j}. The double tangent bundle T(x,y)TMT_{(x,y)}TM has a natural basis {/xi,/yi}\{{\partial}/{\partial{x}^{i}},{\partial}/{\partial{y}^{i}}\}, which transforms as

xi=x~jxix~j+y~jxiy~j=x~jxix~j+yk2x~jxixky~j,\displaystyle\frac{\partial}{\partial{x}^{i}}=\frac{\partial\tilde{x}^{j}}{\partial x^{i}}\frac{\partial}{\partial\tilde{x}^{j}}+\frac{\partial\tilde{y}^{j}}{\partial x^{i}}\frac{\partial}{\partial\tilde{y}^{j}}=\frac{\partial\tilde{x}^{j}}{\partial x^{i}}\frac{\partial}{\partial\tilde{x}^{j}}+y^{k}\frac{\partial^{2}\tilde{x}^{j}}{\partial x^{i}\partial x^{k}}\frac{\partial}{\partial\tilde{y}^{j}},
yi=x~jxiy~j.\displaystyle\frac{\partial}{\partial{y}^{i}}=\frac{\partial\tilde{x}^{j}}{\partial x^{i}}\frac{\partial}{\partial\tilde{y}^{j}}.

In addition, the Christoffel symbol transforms as

Γkli=xix~mx~nxkx~pxlΓ~npm+2x~mxkxlxix~m.{\Gamma}_{kl}^{i}=\frac{\partial x^{i}}{\partial\tilde{x}^{m}}\frac{\partial\tilde{x}^{n}}{\partial x^{k}}\frac{\partial\tilde{x}^{p}}{\partial x^{l}}\tilde{\Gamma}_{np}^{m}+\frac{\partial^{2}\tilde{x}^{m}}{\partial x^{k}\partial x^{l}}\frac{\partial x^{i}}{\partial\tilde{x}^{m}}.

We plug these transformation laws above in G to have

G =yixi+(Gi(x,y)ykylΓkli)yi\displaystyle=y^{i}\frac{\partial}{\partial x^{i}}+(G^{i}(x,y)-y^{k}y^{l}\Gamma_{kl}^{i})\frac{\partial}{\partial y^{i}}
=yix~jxix~j+(yiyk2x~jxixk+(Giykyl(xix~mx~nxkx~pxlΓ~npm+2x~mxkxlxix~m))x~jxi)y~j\displaystyle=y^{i}\frac{\partial\tilde{x}^{j}}{\partial x^{i}}\frac{\partial}{\partial\tilde{x}^{j}}+\big{(}y^{i}y^{k}\frac{\partial^{2}\tilde{x}^{j}}{\partial x^{i}\partial x^{k}}+(G^{i}-y^{k}y^{l}(\frac{\partial x^{i}}{\partial\tilde{x}^{m}}\frac{\partial\tilde{x}^{n}}{\partial x^{k}}\frac{\partial\tilde{x}^{p}}{\partial x^{l}}\tilde{\Gamma}_{np}^{m}+\frac{\partial^{2}\tilde{x}^{m}}{\partial x^{k}\partial x^{l}}\frac{\partial x^{i}}{\partial\tilde{x}^{m}}))\frac{\partial\tilde{x}^{j}}{\partial x^{i}}\big{)}\frac{\partial}{\partial\tilde{y}^{j}}
=y~jx~j+(G~jy~ky~lΓ~klj)y~j.\displaystyle=\tilde{y}^{j}\frac{\partial}{\partial\tilde{x}^{j}}+(\tilde{G}^{j}-\tilde{y}^{k}\tilde{y}^{l}\tilde{\Gamma}_{kl}^{j})\frac{\partial}{\partial\tilde{y}^{j}}.

2.2. Reparameterize γ\gamma

Since γ˙\dot{\gamma} never vanishes, we reparameterize these curves such that they have unit speeds w.r.t. the Riemannian metric gg. Note that this reparameterization will also change the generator G(γ,γ˙)G(\gamma,\dot{\gamma}) and the weight function in the integral transform. In the following, we abuse the notation and use γx,v(t)\gamma_{x,v}(t) to denote the smooth curve with arc length passing (x,v)SM(x,v)\in SM, where SMSM is the unit circle bundle. The family of smooth curves with this arc length parameterization solves a new ODE

Dtγ˙=λ(γ,γ˙)γ˙,D_{t}\dot{\gamma}=\lambda(\gamma,\dot{\gamma})\dot{\gamma}^{\perp},

and are referred to as the λ\lambda-geodesics in [17, Chapter 7], where λ(γ,γ˙)C(SM)\lambda(\gamma,\dot{\gamma})\in C^{\infty}(SM) and γ˙\dot{\gamma}^{\perp} is the vector normal to γ{\gamma}^{\perp} with the same length and the rotation of π/2\pi/2. When λ\lambda is a smooth function on MM, the corresponding λ\lambda-geodesic flow is called the magnetic flow, see [9, 5]. When λ\lambda is a smooth function on SMSM, it is called the thermostats, see [4].

2.3. Extend Γ\Gamma

Moreover, it is convenient to extend Γ\Gamma, such that for any V=rvTpM0V=rv\in T_{p}M\setminus 0, with r>0r>0 and vSpMv\in S_{p}M, there exists a unique curve γx,V(t)\gamma_{x,V}(t) belonging to Γ\Gamma. This can be done by defining γx,V(t)=γx,v(rt)\gamma_{x,V}(t)=\gamma_{x,v}(rt). We note that γx,V(t)\gamma_{x,V}(t) satisfies the following ODE

Dtγ˙x,V(t)=|V|gλ(γx,V(t),1|V|gγ˙x,V(t))γ˙x,V(t),D_{t}\dot{\gamma}_{x,V}(t)=|V|_{g}\lambda({\gamma}_{x,V}(t),\frac{1}{|V|_{g}}\dot{\gamma}_{x,V}(t))\dot{\gamma}^{\perp}_{x,V}(t),

which is equivalent to extend λ(x,v)\lambda(x,v) to C(TM0)C^{\infty}(TM\setminus 0).

As we mentioned before, for a curve satisfying the equation in Property (A2) in general, rescaling it might not give us the same one, i.e., γp,v\gamma_{p,v} and γp,rv\gamma_{p,rv} are different curves. However, with the reparameterization and the extension above, the rescaling will cause no problem.

2.4. Conjugate points.

We define the exponential map as expp(t,v)=γp,v(t)\exp_{p}(t,v)=\gamma_{p,v}(t) for (p,v)SM(p,v)\in SM, see [22]. This definition uses polar coordinates and is independent of a change of the parameterization for curves in Γ\Gamma. For vSpMv\in S_{p}M, tt\in\mathbb{R}, the map V=tvexpp(t,v)V=tv\mapsto\exp_{p}(t,v) may not be smooth near V=0TpMV=0\in T_{p}M, see [5, A.3] for more details and properties.

By dvexpp(t,v)d_{v}\exp_{p}(t,v), we always mean the differential of expp(t,v)\exp_{p}(t,v) w.r.t vSpMv\in S_{p}M. This notation is different from the one dvexpp(v)d_{v}\exp_{p}(v) for vTpMv\in T_{p}M used in the case of geodesics. We say a point q=γp,v(t0)q=\gamma_{p,v}(t_{0}) is conjugate to p=γp,v(0)p=\gamma_{p,v}(0) if the differential dt,vexpp(t,v)d_{t,v}\exp_{p}(t,v) of the exponential map in polar coordinates has the rank less than the maximal one at (t0,v)(t_{0},v).

To further describe the conjugate points, we consider the flow ϕt:SMSM\phi_{t}:SM\rightarrow SM given by ϕt(x,v)=(γx,v(t),γ˙x,v(t)).\phi_{t}(x,v)=(\gamma_{x,v}(t),\dot{\gamma}_{x,v}(t)). For a fixed curve γp0,v0(t)\gamma_{p_{0},v_{0}}(t), we set

(p1,v1)=ϕt(p0,v0),(p2,v2)=ϕt+s(p0,v0),(p_{1},v_{1})=\phi_{t}(p_{0},v_{0}),\quad(p_{2},v_{2})=\phi_{t+s}(p_{0},v_{0}),

where (pj,vj)SM(p_{j},v_{j})\in SM, for j=0,1,2j=0,1,2. Note that ϕt:SMSM\phi_{t}:SM\rightarrow SM is a local diffeomorphism, see [17, Chapter 7]. Its pushforward dϕt:T(p,v)SMTϕt(p,v)SM\mathop{}\!\mathrm{d}\phi_{t}:T_{(p,v)}SM\rightarrow T_{\phi_{t}(p,v)}SM is invertible. Let π\pi be the natural projection from SMSM to MM, which induces a map dπ\mathop{}\!\mathrm{d}\pi from T(p,v)SMT_{(p,v)}SM to TpMT_{p}M. We have the following proposition that relates the conjugate points with the pushforward of the flow ϕt\phi_{t}.

Proposition 1.

A point p2=γp0,v0(t+s)p_{2}=\gamma_{p_{0},v_{0}}(t+s) is conjugate to p1=γp0,v0(t)p_{1}=\gamma_{p_{0},v_{0}}(t) if and only if there exists ζ0T(p0,v0)SM\zeta_{0}\in T_{(p_{0},v_{0})}SM and cc\in\mathbb{R} satisfying

(1) dπ(dϕt)(ζ0)=0,dπ(dϕt+s)(ζ0)+cγ˙p0,v0(t+s)=0.\mathop{}\!\mathrm{d}\pi(\mathop{}\!\mathrm{d}\phi_{t})(\zeta_{0})=0,\qquad\mathop{}\!\mathrm{d}\pi(\mathop{}\!\mathrm{d}\phi_{t+s})(\zeta_{0})+c\dot{\gamma}_{p_{0},v_{0}}(t+s)=0.
Proof.

Let ζjT(pj,vj)SM\zeta_{j}\in T_{(p_{j},v_{j})}SM for j=0,1,2j=0,1,2 such that

ζ2=dϕs(ζ1)=dϕt+s(ζ0),\zeta_{2}=\mathop{}\!\mathrm{d}\phi_{s}(\zeta_{1})=\mathop{}\!\mathrm{d}\phi_{t+s}(\zeta_{0}),

which implies ζ1=dϕt(ζ0)\zeta_{1}=\mathop{}\!\mathrm{d}\phi_{t}(\zeta_{0}) by the group property of ϕt\phi_{t}. Let {pji,vji}\{p_{j}^{i},v_{j}^{i}\} be the local charts near (pj,vj)(p_{j},v_{j}) and they induce a natural local basis {/pji,/vji}\{\partial/\partial p_{j}^{i},\partial/\partial v_{j}^{i}\} in T(pj,vj)SMT_{(p_{j},v_{j})}SM, for j=0,1,2j=0,1,2. We write ζj=(αj,βj)\zeta_{j}=(\alpha_{j},\beta_{j}) in this local coordinates, where αj\alpha_{j} can be identified as an element in TpjMT_{p_{j}}M and βj\beta_{j} in TvjSpjMT_{v_{j}}S_{p_{j}}M. Note that dπ(ζj)=αj\mathop{}\!\mathrm{d}\pi(\zeta_{j})=\alpha_{j} for j=0,1,2j=0,1,2. Consider the representation of dϕt\mathop{}\!\mathrm{d}\phi_{t} in these local coordinates near (pj,vj)(p_{j},v_{j}) given by the Jacobian matrix

J(pj,vj,t)=[dpexpp(t,v)dvexpp(t,v)dpexp˙p(t,v)dvexp˙p(t,v)](pj,vj,t).J_{(p_{j},v_{j},t)}=\begin{bmatrix}\begin{array}[]{cc}\mathop{}\!\mathrm{d}_{p}\exp_{p}(t,v)&\mathop{}\!\mathrm{d}_{v}\exp_{p}(t,v)\\ \mathop{}\!\mathrm{d}_{p}{\dot{\exp}_{p}(t,v)}&\mathop{}\!\mathrm{d}_{v}{\dot{\exp}_{p}(t,v)}\end{array}\end{bmatrix}_{(p_{j},v_{j},t)}.

More explicitly, we have

(2) [α2β2]=J(p1,v1,s)[α1β1]=J(p0,v0,t+s)[α0β0],\begin{bmatrix}\begin{array}[]{c}\alpha_{2}\\ \beta_{2}\end{array}\end{bmatrix}=J_{(p_{1},v_{1},s)}\begin{bmatrix}\begin{array}[]{c}\alpha_{1}\\ \beta_{1}\end{array}\end{bmatrix}=J_{(p_{0},v_{0},t+s)}\begin{bmatrix}\begin{array}[]{c}\alpha_{0}\\ \beta_{0}\end{array}\end{bmatrix},

and

(3) [α1β1]=J(p0,v0,t)[α0β0].\begin{bmatrix}\begin{array}[]{c}\alpha_{1}\\ \beta_{1}\end{array}\end{bmatrix}=J_{(p_{0},v_{0},t)}\begin{bmatrix}\begin{array}[]{c}\alpha_{0}\\ \beta_{0}\end{array}\end{bmatrix}.

Recall that p2p_{2} is conjugate to p1p_{1} if only if there exists cc\in\mathbb{R} and β1Tv1Sp1M\beta_{1}\in T_{v_{1}}S_{p_{1}}M such that

ds,vexpp1(s,v1)(c,β1)=0dvexpp1(s,v1)(β1)=cγ˙p1,v1(s),\displaystyle d_{s,v}\exp_{p_{1}}(s,v_{1})(c,\beta_{1})=0\Rightarrow d_{v}\exp_{p_{1}}(s,v_{1})(\beta_{1})=-c\dot{\gamma}_{p_{1},v_{1}}(s),

On the one hand, if such c,β1c,\beta_{1} exist, then we pick ζ1=(0,β1)\zeta_{1}=(0,\beta_{1}) and let ζ0=J(p0,v0,t)1ζ1\zeta_{0}=J_{(p_{0},v_{0},t)}^{-1}\zeta_{1}. In this case, by equation (2) and (3) we have

0=α1=dπ(dϕt)(ζ0)=dpexpp0(t,v0)(α0)+dvexpp0(t,v0)(β0),\displaystyle 0=\alpha_{1}=\mathop{}\!\mathrm{d}\pi(\mathop{}\!\mathrm{d}\phi_{t})(\zeta_{0})=d_{p}\exp_{p_{0}}(t,v_{0})(\alpha_{0})+d_{v}\exp_{p_{0}}(t,v_{0})(\beta_{0}),
α2=dπ(dϕt+s)(ζ0)=dpexpp1(s,v1)(α1)+dvexpp1(s,v1)(β1)=cγ˙p1,v1(s),\displaystyle\alpha_{2}=\mathop{}\!\mathrm{d}\pi(\mathop{}\!\mathrm{d}\phi_{t+s})(\zeta_{0})=d_{p}\exp_{p_{1}}(s,v_{1})(\alpha_{1})+d_{v}\exp_{p_{1}}(s,v_{1})(\beta_{1})=-c\dot{\gamma}_{p_{1},v_{1}}(s),

which proves (1). On the other hand, if there is ζ0\zeta_{0} such that (1) is true, then we take (α1,β1)=J(p0,v0,t)(ζ0)(\alpha_{1},\beta_{1})=J_{(p_{0},v_{0},t)}(\zeta_{0}). By (1) we must have α1=0\alpha_{1}=0 and then dvexpp1(s,v1)(β1)=cγ˙p1,v1(s)d_{v}\exp_{p_{1}}(s,v_{1})(\beta_{1})=-c\dot{\gamma}_{p_{1},v_{1}}(s), which implies that p2p_{2} is conjugate to p1p_{1}. ∎

In the following, recall some results in [17, Chapter 4, 7]. We consider a moving frame {F(x,v),H(x,v),V(x,v)}\{F(x,v),H(x,v),V(x,v)\} on SMSM, where F(x,v)F(x,v) is the infinitesimal generator of ϕt\phi_{t}, H(x,v)H(x,v) is the horizontal vector field, and V(x,v)V(x,v) is the vertical vector field. For a detailed definition of HH and VV, see [17, Section 4.1]. One can show that

(4) dπ(F(x,v))=v,dπ(H(x,v))=v,dπ(V(x,v))=0.\displaystyle\mathop{}\!\mathrm{d}\pi(F(x,v))=v,\quad\mathop{}\!\mathrm{d}\pi(H(x,v))=v^{\perp},\quad\mathop{}\!\mathrm{d}\pi(V(x,v))=0.

The λ\lambda-geodesic vector field FF is related to the geodesic vector field XX by

F=X+λV.F=X+\lambda V.

One can show the commutators between them are given by

[H,V]=X=FλV,\displaystyle[H,V]=X=F-\lambda V,
[V,F]=H+V(λ)V,\displaystyle[V,F]=H+V(\lambda)V,
[F,H]=(KH(λ)+λ2)VλF,\displaystyle[F,H]=(K-H(\lambda)+\lambda^{2})V-\lambda F,

where KK is the Gaussian curvature.

Claim 1.

For any constant cc, we can find ζ0T(p0,v0)SM\zeta_{0}^{\prime}\in T_{(p_{0},v_{0})}SM such that

dπ(dϕt)(ζ0)=cγ˙p0,v0(t).\mathop{}\!\mathrm{d}\pi(\mathop{}\!\mathrm{d}\phi_{t})(\zeta_{0}^{\prime})=c\dot{\gamma}_{p_{0},v_{0}}(t).
Proof.

By [17, Section 7.1], we consider the infinitesimal generator F(x,v)F(x,v) of the flow ϕt\phi_{t} and define

Y(t)=c(dϕt)(F(ϕt(x,v))).Y(t)=c(\mathop{}\!\mathrm{d}\phi_{-t})(F(\phi_{t}(x,v))).

Differentiating both sides with respect to tt implies

Y˙(t)=c[F(x,v),F(x,v)]=0.\dot{Y}(t)=c[F(x,v),F(x,v)]=0.

Thus Y(t)Y(t) is a constant vector field on SMSM. By denoting it as Y=ζ0Y=\zeta_{0}^{\prime}, we have dϕt(ζ0)=cF(ϕt(x,v))d\phi_{t}(\zeta_{0}^{\prime})=cF(\phi_{t}(x,v)). With dπ(F(x,v))=v\mathop{}\!\mathrm{d}\pi(F(x,v))=v, applying dπ\mathop{}\!\mathrm{d}\pi to dϕt(ζ0)\mathop{}\!\mathrm{d}\phi_{t}(\zeta_{0}^{\prime}) claims what we need. ∎

2.5. Jacobi Fields

Inspired by Proposition 1, we say a vector field J(p0,v0)(t)J_{(p_{0},v_{0})}(t) is a Jacobi field along γp0,v0\gamma_{p_{0},v_{0}} if it can be written as

(5) J(p0,v0)(t)=dπ(dϕt)(ζ)+(c1+c2t)γ˙p0,v0(t),\displaystyle J_{(p_{0},v_{0})}(t)=\mathop{}\!\mathrm{d}\pi(\mathop{}\!\mathrm{d}\phi_{t})(\zeta)+(c_{1}+c_{2}t)\dot{\gamma}_{p_{0},v_{0}}(t),

for some ζT(p0,v0)SM\zeta\in T_{(p_{0},v_{0})}SM and c1,c2c_{1},c_{2}\in\mathbb{R}. We emphasize that the representation of a Jacobi field in form of (5) is not unique, which can be seen from Claim 1 and Proposition 5. Moreover, this definition is different from the usual one that the Jacobi fields are defined as the variation field along γp0,v0\gamma_{p_{0},v_{0}}, for example see [2]. However, they are essentially the same and the definition we choose here is more straightforward for the purpose of this work. To understand the last term tγ˙p0,v0(t)t\dot{\gamma}_{p_{0},v_{0}}(t), recall we extend Γ\Gamma in Section 2.3, such that for any V=rvTpM0V=rv\in T_{p}M\setminus 0 with r>0r>0 and vSpMv\in S_{p}M, there exists a unique curve γx,V(t)\gamma_{x,V}(t) belonging to Γ\Gamma.

Proposition 2.

A point p2=γp0,v0(t2)p_{2}=\gamma_{p_{0},v_{0}}(t_{2}) is conjugate to p1=γp0,v0(t1)p_{1}=\gamma_{p_{0},v_{0}}(t_{1}) if and only if there exists a nonvanishing Jacobi field J(t)J(t) along γp0,v0\gamma_{p_{0},v_{0}} such that J(t1)=J(t2)=0J(t_{1})=J(t_{2})=0.

Proof.

If p2=γp0,v0(t2)p_{2}=\gamma_{p_{0},v_{0}}(t_{2}) is conjugate to p1=γp0,v0(t1)p_{1}=\gamma_{p_{0},v_{0}}(t_{1}), then by Proposition 1, there exists ζ0\zeta_{0} and cc such that

dπ(dϕt1)(ζ0)=0,dπ(ϕt2)(ζ0)+cγ˙p0,v0(t2)=0.\mathop{}\!\mathrm{d}\pi(\mathop{}\!\mathrm{d}\phi_{t_{1}})(\zeta_{0})=0,\qquad\mathop{}\!\mathrm{d}\pi(\phi_{t_{2}})(\zeta_{0})+c\dot{\gamma}_{p_{0},v_{0}}(t_{2})=0.

Let J(p0,v0)(t)=dπ(dϕt)(ζ0)+ctt1t2t1γ˙p0,v0(t)J_{(p_{0},v_{0})}(t)=\mathop{}\!\mathrm{d}\pi(\mathop{}\!\mathrm{d}\phi_{t})(\zeta_{0})+c\frac{t-t_{1}}{t_{2}-t_{1}}\dot{\gamma}_{p_{0},v_{0}}(t). Then we have J(t1)=J(t2)=0J(t_{1})=J(t_{2})=0.

Conversely, suppose there is a nonzero Jacobi field with J(p0,v0)(t1)=J(p0,v0)(t2)=0J_{(p_{0},v_{0})}(t_{1})=J_{(p_{0},v_{0})}(t_{2})=0. More explicitly, we have

(6) dπ(dϕt1)(ζ0)+λ1γ˙p0,v0(t1)=0,dπ(dϕt2)(ζ0)+λ2γ˙p0,v0(t2)=0,\displaystyle\begin{split}&\mathop{}\!\mathrm{d}\pi(\mathop{}\!\mathrm{d}\phi_{t_{1}})(\zeta_{0})+\lambda_{1}\dot{\gamma}_{p_{0},v_{0}}(t_{1})=0,\\ &\mathop{}\!\mathrm{d}\pi(\mathop{}\!\mathrm{d}\phi_{t_{2}})(\zeta_{0})+\lambda_{2}\dot{\gamma}_{p_{0},v_{0}}(t_{2})=0,\end{split}

where λ1=c1+c2t1,λ2=c1+c2t2\lambda_{1}=c_{1}+c_{2}t_{1},\lambda_{2}=c_{1}+c_{2}t_{2}. It suffices to find ζ0\zeta_{0}^{\prime} such that dπ(dϕt1)(ζ0)=λ1γ˙p0,v0(t)\mathop{}\!\mathrm{d}\pi(\mathop{}\!\mathrm{d}\phi_{t_{1}})(\zeta_{0}^{\prime})=\lambda_{1}\dot{\gamma}_{p_{0},v_{0}}(t). Such ζ0\zeta_{0}^{\prime} can be found by Claim 1. In this way, the term with γ˙p0,v0(t1)\dot{\gamma}_{p_{0},v_{0}}(t_{1}) in the first equation can replaced by considering ζ~0=ζ0+ζ0\tilde{\zeta}_{0}=\zeta_{0}+\zeta_{0}^{\prime}. By Proposition 1 we have p2p_{2} is conjugate to p1p_{1}. ∎

The same conclusion is proved in [2, Theorem 5.3] based on the definition of the Jacobi fields as the variational fields.

2.6. The Jacobi equation.

In this subsection, we show the following proposition in the case of λ\lambda-geodesic flow, analogous to [17, Proposition 4.14 ] in the case of the geodesic flow.

Proposition 3 ([2, Lemma 4.6]).

For any ζT(p,v)SM\zeta\in T_{(p,v)}SM, if we write dϕt(ξ)\mathop{}\!\mathrm{d}\phi_{t}(\xi)

dϕt(ζ)=x(t)F(ϕt(x,v))+y(t)H(ϕt(x,v))+z(t)V(ϕt(x,v)),\mathop{}\!\mathrm{d}\phi_{t}(\zeta)=x(t)F(\phi_{t}(x,v))+y(t)H(\phi_{t}(x,v))+z(t)V(\phi_{t}(x,v)),

in the moving frame {F,H,V}\{F,H,V\}, then the smooth functions x,y,zx,y,z should satisfy

x˙=λy,\displaystyle\dot{x}=\lambda y,
y˙=z,\displaystyle\dot{y}=z,
z˙=(KH(λ)+λ2)y+V(λ)z.\displaystyle\dot{z}=-({K-H(\lambda)+\lambda^{2}})y+V(\lambda)z.

Combining the last two equations above, one can see that yy should satisfy a second-order ODE. Notice γ˙(t),γ˙(t)\dot{\gamma}(t)^{\perp},\dot{\gamma}(t) give us a moving frame along γp0,v0(t)\gamma_{p_{0},v_{0}}(t). This proves the following proposition analogous to [2, Lemma 5.2].

Proposition 4.

If J(t)J(t) is a Jacobi field along γp0,v0\gamma_{p_{0},v_{0}} expressed as

J(t)=x(t)γ˙p0,v0(t)+y(t)γ˙p0,v0(t),J(t)=x(t)\dot{\gamma}_{p_{0},v_{0}}(t)+y(t)\dot{\gamma}_{p_{0},v_{0}}^{\perp}(t),

then x(t),y(t)x(t),y(t) satisfy the Jacobi equations

(7) x˙=λy+c,\displaystyle\dot{x}=\lambda y+c,
(8) y¨V(λ)y˙+(KH(λ)+λ2)y=0,\displaystyle\ddot{y}-V(\lambda)\dot{y}+(K-H(\lambda)+\lambda^{2})y=0,

where cc is a constant.

Remark 1.

If λ=0\lambda=0, we have the case of geodesics. Compared to [2, Lemma 5.2], we have an extra constant cc in (7), since by extending Γ\Gamma we include tγ˙(t)t\dot{\gamma}(t) in our definition of the Jacobi field.

Proof.

Recall J(t)J(t) has the representation

J(p0,v0)(t)=dπ(dϕt)(ζ)+(c1+c2t)γ˙p0,v0(t).J_{(p_{0},v_{0})}(t)=\mathop{}\!\mathrm{d}\pi(\mathop{}\!\mathrm{d}\phi_{t})(\zeta)+(c_{1}+c_{2}t)\dot{\gamma}_{p_{0},v_{0}}(t).

Applying dπ\mathop{}\!\mathrm{d}\pi to Proposition 3, we have the desired result by (4). ∎

3. Microlocal analysis of the local problem

3.1. Parameterization of \mathcal{M}

Since the group action of \mathbb{R} on MM by ϕt\phi_{t} is free and proper, the family of curves form a smooth manifold =SM/ϕt\mathcal{M}=SM/\phi_{t} of dimension 2×22=22\times 2-2=2. We parameterize \mathcal{M} below near a fixed curve γΓ\gamma\in\Gamma, in the same way for the case of the geodesics in [18, Section 3.1]. This parameterization gives us a local charts of the manifold of curves.

For this purpose, for fixed γ\gamma we choose a hypersurface HH such that γ\gamma hits HH transversally at p0=γ(0)p_{0}=\gamma(0) and v0=γ˙(0)v_{0}=\dot{\gamma}(0). Let \mathcal{H} consist of all (p,v)SM(p,v)\in SM in a small neighborhood of (p0,v0)(p_{0},v_{0}), such that pHp\in H and vv is transversal to HH. Then we can parameterize all curves near fixed γ\gamma by the intersection point pp and the direction vv, i.e., by an element (p,v)(p,v)\in\mathcal{H}. Suppose HH is locally given by x1=0x^{1}=0 and has coordinate yy. We can write each element in \mathcal{H} using (y,η)(y,\eta) instead, where η\eta is the parameterization of vS(0,y)Mv\in S_{(0,y)}M. In the following, we use the new notation γ(y,η,t)\gamma(y,\eta,t) to denote the smooth curve starting from p=(0,y)p=(0,y) in the direction vv. In some cases, we write γ(t),γy(t),γη(t)\gamma(t),\frac{\partial\gamma}{\partial y}(t),\frac{\partial\gamma}{\partial\eta}(t) and omit the variables y,ηy,\eta for simplification, if there is no confusion caused.

Proposition 5.

With this parameterization, the Jacobi field J(t)J(t) along γ(y,η,t)\gamma(y,\eta,t) can be simplified as

J(t)=κ1γy(t)+κ2γηz(t)+c1γ˙(t)+c2tγ˙(t),J(t)=\kappa_{1}\frac{\partial\gamma}{\partial y}(t)+\kappa_{2}\frac{\partial\gamma}{\partial\eta}z(t)+c_{1}\dot{\gamma}(t)+c_{2}t\dot{\gamma}(t),

where κ1,κ2,c1,c2\kappa_{1},\kappa_{2},c_{1},c_{2}\in\mathbb{R}.

Proof.

It is shown in Claim 1 that for any cc\in\mathbb{R}, there exists ζT(p,v)SM\zeta\in T_{(p,v)}SM such that

S(t)=dπ(dϕt)(ζ)+cγ˙(y,η,t)0,S(t)=\mathop{}\!\mathrm{d}\pi(\mathop{}\!\mathrm{d}\phi_{t})(\zeta)+c\dot{\gamma}(y,\eta,t)\equiv 0,

where in local coordinates p=(0,y)p=(0,y) and we assume c0c\neq 0. On the other hand, in the local coordinates (x1,y,η)(x^{1},y,\eta) of SMSM, the pushforward dϕt\mathop{}\!\mathrm{d}\phi_{t} has a representation and we can write

dπ(dϕt)(ζ)=κ1γy+κ2γx1+κ3γη.\mathop{}\!\mathrm{d}\pi(\mathop{}\!\mathrm{d}\phi_{t})(\zeta)=\kappa_{1}\frac{\partial\gamma}{\partial y}+\kappa_{2}\frac{\partial\gamma}{\partial x^{1}}+\kappa_{3}\frac{\partial\gamma}{\partial\eta}.

Then it suffices to show that κ20\kappa_{2}\neq 0 and therefore γx\frac{\partial\gamma}{\partial x} can be written as a linear combination of γy,γη,γ˙\frac{\partial\gamma}{\partial y},\frac{\partial\gamma}{\partial\eta},\dot{\gamma}. Indeed, we have the following initial conditions

(9) γη(0)=0,(γy(0),γ˙(0))g0,\frac{\partial\gamma}{\partial\eta}(0)=0,\quad(\frac{\partial\gamma}{\partial y}(0),\dot{\gamma}^{\perp}(0))_{g}\neq 0,

where the first one comes from γ(y,η,0)p\gamma(y,\eta,0)\equiv p and the second one is from the assumption that γ(y,η,t)\gamma(y,\eta,t) hit HH transversally at t=0t=0. Now we compute

0(S(0),γ˙(0))=κ1(γy(0),γ˙(0))g+κ2(γx1(0),γ˙(0))g.0\equiv(S(0),\dot{\gamma}^{\perp}(0))=\kappa_{1}(\frac{\partial\gamma}{\partial y}(0),\dot{\gamma}^{\perp}(0))_{g}+\kappa_{2}(\frac{\partial\gamma}{\partial x^{1}}(0),\dot{\gamma}^{\perp}(0))_{g}.

It follows that κ20\kappa_{2}\neq 0, otherwise one has κ2=κ1=0\kappa_{2}=\kappa_{1}=0, which leads to

κ3γη+cγ˙(y,η,t)=0.\kappa_{3}\frac{\partial\gamma}{\partial\eta}+c\dot{\gamma}(y,\eta,t)=0.

This contradicts with the initial condition in (9) and that γ˙(y,η,t)\dot{\gamma}(y,\eta,t) is nonvanishing. ∎

3.2. IwI_{w} as an FIO

The same analysis for the geodesic transform in [18, Section 3.1] works for IwI_{w}. We briefly state these arguments for IwI_{w} in this subsection. For more details see [18, Section 3.1].

We consider the curve-point relation

Z={(γ(y,η,t),x)×M:xγ},Z=\{(\gamma(y,\eta,t),x)\in\mathcal{M}\times M:\ x\in\gamma\},

from the double fibration point of view of [8, 11]. Note that ZZ is a smooth manifold with the conormal bundle

NZ={(γ,γ^,x,ξ)T(×M)0:(γ^,ξ) vanishes on T(γ,x)Z0}.N^{*}Z=\{(\gamma,\hat{\gamma},x,\xi)\in T^{*}(\mathcal{M}\times M)\setminus 0:\ (\hat{\gamma},\xi)\text{ vanishes on }T_{(\gamma,x)}Z_{0}\}.

The integral transform along curves in Γ\Gamma defined by

Iwf(γ)=w(γ(t),γ˙(t))f(γ(t))dt,f(M),γΓ.I_{w}f(\gamma)=\int w(\gamma(t),\dot{\gamma}(t))f(\gamma(t))\mathop{}\!\mathrm{d}t,\quad f\in\mathcal{E}^{\prime}(M),\ \gamma\in\Gamma.

has a delta-type Schwartz kernel wδZw\delta_{Z}, which is a conormal distribution supported in NZN^{*}Z. Therefore, it is an FIO of order 1/2-{1}/{2} and associated with the canonical relation C=NZC=N^{*}Z^{\prime}, the twisted version of NZN^{*}Z given by

C={(γ,γ^,x,ξ):{(γ,γ^,x,ξ)NZ}.C=\{(\gamma,\hat{\gamma},x,-\xi):\ \{(\gamma,\hat{\gamma},x,\xi)\in N^{*}Z\}.

By the local parameterization of \mathcal{M}, we can show (y,η,y^,η^,x,ξ)C(y,\eta,\hat{y},\hat{\eta},x,\xi)\in C if and only if there exists tt with x=γ(y,η,t)x=\gamma(y,\eta,t) and

(10) ξiγ˙i=0,ξiγiy=y^,ξiγiη=η^,\displaystyle\xi_{i}\dot{\gamma}^{i}=0,\quad\xi_{i}\frac{\partial\gamma^{i}}{\partial y}=\hat{y},\quad\xi_{i}\frac{\partial\gamma^{i}}{\partial\eta}=\hat{\eta},

where the Einstein summation convention is used.

3.3. The description of CC

Inspired by [18, Section 3.2], in this subsection, we write CC in a more explicit form and describe its properties. Let

e1(t)=γ˙(t),e2(t)=γ˙(t)e_{1}(t)=\dot{\gamma}(t)^{\perp},\quad e_{2}(t)=\dot{\gamma}(t)

be a moving frame. We have the corresponding dual basis {e1(t),e2(t)}\{e^{1}(t),e^{2}(t)\}. If we regard ξ\xi as a function of tt, then the first condition in (10) means

ξ(t)=h(t)e1(t),\xi(t)=h(t)e^{1}(t),

for some nonvanishing function h(t)h(t). Suppose we have the following expansion w.r.t the frame

(11) γy=a1(t)e1(t)+a2(t)e2(t),γη=b1(t)e1(t)+b2(t)e2(t).\displaystyle\frac{\partial\gamma}{\partial y}=a_{1}(t)e_{1}(t)+a_{2}(t)e_{2}(t),\quad\frac{\partial\gamma}{\partial\eta}=b_{1}(t)e_{1}(t)+b_{2}(t)e_{2}(t).

The second and third conditions in (10) imply

h(t)a1(t)=y^,h(t)b1(t)=η^.\displaystyle h(t)a_{1}(t)=\hat{y},\quad h(t)b_{1}(t)=\hat{\eta}.

Therefore, the canonical relation has the local representation

C={(y,η,μa1(t,y,η)y^,μb1(t,y,η)η^,γ(t,y,η),μγ˙(t,y,η)):(y,η),μ0},\displaystyle C=\{(y,\eta,\underbrace{\mu a_{1}(t,y,\eta)}_{\hat{y}},\underbrace{\mu b_{1}(t,y,\eta)}_{\hat{\eta}},\gamma(t,y,\eta),\mu\dot{\gamma}^{\perp}(t,y,\eta)):\ (y,\eta)\in\mathcal{H},\ \mu\neq 0\},

where

(12) a1(t,y,η)=(e1(t),γ(t,y,η)y)g,b1(t,y,η)=(e1(t),γ(t,y,η)η)ga_{1}(t,y,\eta)=(e_{1}(t),\frac{\partial\gamma(t,y,\eta)}{\partial y})_{g},\quad b_{1}(t,y,\eta)=(e_{1}(t),\frac{\partial\gamma(t,y,\eta)}{\partial\eta})_{g}

are projections along γ˙(t,y,η)\dot{\gamma}^{\perp}(t,y,\eta). The canonical relation CC can be parameterized by (y,η,t,μ)(y,\eta,t,\mu) and therefore we have dimC=4\dim C=4.

Proposition 6.

Let πM:CTM0\pi_{M}:C\rightarrow T^{*}M\setminus 0 and π:CT0\pi_{\mathcal{M}}:C\rightarrow T^{*}\mathcal{M}\setminus 0 be the natural projections. Then

  • (a)

    πM:(y,η,y^,η^,x,ξ)(x,ξ)\pi_{M}:(y,\eta,\hat{y},\hat{\eta},x,\xi)\mapsto(x,\xi) is a diffeomorphism,

  • (b)

    π:(y,η,y^,η^,x,ξ)(y,η,y^,η^)\pi_{\mathcal{M}}:(y,\eta,\hat{y},\hat{\eta},x,\xi)\mapsto(y,\eta,\hat{y},\hat{\eta}) is a local diffeomorphism.

Proof.

We prove (a) first. For each (x,ξ)TM0(x,\xi)\in T^{*}M\setminus 0, with the property (A1), there is a unique curve γ\gamma passing xx at time tt and conormal to ξ\xi. Suppose γ\gamma hits HH transversally at yy in the direction parameterized by η\eta, at t=0t=0. Then (y,η)(y,\eta) can be derived from the flow ϕt(x,v)\phi_{-t}(x,v), composed with the restriction on HH, where vv is the unit vector corresponds to ξ\xi. Thus, in a small neighborhood (y,η)(y,\eta) depends on (x,ξ)(x,\xi) smoothly and (y^,η^)=(μa1(t,y,η),μb1(t,y,η))(\hat{y},\hat{\eta})=(\mu a_{1}(t,y,\eta),\mu b_{1}(t,y,\eta)) with μ=|ξ|\mu=|\xi| also depends on (x,ξ)(x,\xi) in a smooth way, if we are away from the zero section. This implies one can regard (x,ξ)(x,\xi) as a parameterization of CC and πM\pi_{M} is a diffeomorphism.

For (b), in local parameterization we have π:(y,η,t,μ)(y,η,μa1,μb1)\pi_{\mathcal{M}}:(y,\eta,t,\mu)\mapsto(y,\eta,\mu a_{1},\mu b_{1}). We compute dπ\mathop{}\!\mathrm{d}\pi_{\mathcal{M}} to have

detdπ=μ(a˙1b1b˙1a1)μW(t),\det\mathop{}\!\mathrm{d}\pi_{\mathcal{M}}=\mu(\dot{a}_{1}b_{1}-\dot{b}_{1}a_{1})\equiv\mu W(t),

where W(t)W(t) is the Wronskian and it satisfies W˙(t)=V(λ)W(t)\dot{W}(t)=V(\lambda)W(t). Indeed, recall a1,b1a_{1},b_{1} are the projections of Jacobi fields onto γ˙(t)\dot{\gamma}^{\perp}(t) and therefore both of they should satisfy the equation in (8). Solving the equation of the Wronskian, we have W(t)=W(0)eV(λ)dtW(t)=W(0)e^{\int V(\lambda)\mathop{}\!\mathrm{d}t}, which indicates W(t)W(t) is zero at any tt, if it vanishes at some point. When W(t)W(t) vanishes, we have a1(t)=cb1(t)a_{1}(t)=cb_{1}(t) for some constant cc. However, since a1(0)0a_{1}(0)\neq 0 and b1(0)=0b_{1}(0)=0 by the initial condition (9), it is impossible and therefore we have W(t)0W(t)\neq 0. ∎

We say the integral transform IwI_{w} satisfies the Bolker condition, see [10], if the projection π\pi_{\mathcal{M}} is an injective immersion.

Theorem 3.1.

The map 𝒞(x,ξ)ππM1(x,ξ)\mathcal{C}(x,\xi)\equiv\pi_{\mathcal{M}}\circ\pi_{M}^{-1}(x,\xi) is a local diffeomorphism from TM0T^{*}M\setminus 0 to T0T^{*}\mathcal{M}\setminus 0. Moreover, we have 𝒞(p1,ξ1)=𝒞(p2,ξ2)\mathcal{C}(p_{1},\xi^{1})=\mathcal{C}(p_{2},\xi^{2}) if and only if there is a curve γ(t,y,η)\gamma(t,y,\eta) joining p1p_{1} at t1t_{1} and p2p_{2} at t2t_{2}, with t2>t1t_{2}>t_{1}, such that

  • (a)

    p2p_{2} is conjugate to p1p_{1}.

  • (b)

    ξ1=μa1(t2)γ˙(t1)\xi^{1}=\mu a_{1}(t_{2})\dot{\gamma}^{\perp}(t_{1}) and ξ2=μa1(t1)γ˙(t2)\xi^{2}=\mu a_{1}(t_{1})\dot{\gamma}^{\perp}(t_{2}) for some nonzero μ\mu, where a1(t)a_{1}(t) is defined in (12).

In particular, if there are no conjugate points, then the Bolker condition is satisfied and 𝒞\mathcal{C} is a diffeomorphism.

Proof.

If there exist different (p1,ξ1)(p_{1},\xi^{1}) at t1t_{1} and (p2,ξ2)(p_{2},\xi^{2}) at t2t_{2} such that 𝒞(p1,ξ1)=𝒞(p2,ξ2)=(y,η,y^,η^)\mathcal{C}(p_{1},\xi^{1})=\mathcal{C}(p_{2},\xi^{2})=(y,\eta,\hat{y},\hat{\eta}), we must have y^b1(t)η^a1(t)=0\hat{y}b_{1}(t)-\hat{\eta}a_{1}(t)=0 is true for both t=t1,t2t=t_{1},t_{2}. Consider the vector field

c0(y,η,t)=y^γηη^γη=(y^b1(t)η^a1(t))e1(t)+(y^b2(t)η^a2(t))e2(t).c_{0}(y,\eta,t)=\hat{y}\frac{\partial\gamma}{\partial\eta}-\hat{\eta}\frac{\partial\gamma}{\partial\eta}=(\hat{y}b_{1}(t)-\hat{\eta}a_{1}(t))e_{1}(t)+(\hat{y}b_{2}(t)-\hat{\eta}a_{2}(t))e_{2}(t).

It satisfies

c0(y,η,t1)=(y^b2(t1)η^a2(t1))e2(t1)μ1γ˙(t1).\displaystyle c_{0}(y,\eta,t_{1})=(\hat{y}b_{2}(t_{1})-\hat{\eta}a_{2}(t_{1}))e_{2}(t_{1})\equiv\mu_{1}\dot{\gamma}(t_{1}).
c0(y,η,t2)=(y^b2(t2)η^a2(t2))e2(t2)μ2γ˙(t2).\displaystyle c_{0}(y,\eta,t_{2})=(\hat{y}b_{2}(t_{2})-\hat{\eta}a_{2}(t_{2}))e_{2}(t_{2})\equiv\mu_{2}\dot{\gamma}(t_{2}).

It follows that we can define the Jacobi field

c(y,η,t)=y^γηη^γη(μ1tt2t1t2+μ2tt1t2t1)γ˙(t),c(y,\eta,t)=\hat{y}\frac{\partial\gamma}{\partial\eta}-\hat{\eta}\frac{\partial\gamma}{\partial\eta}-(\mu_{1}\frac{t-t_{2}}{t_{1}-t_{2}}+\mu_{2}\frac{t-t_{1}}{t_{2}-t_{1}})\dot{\gamma}(t),

which satisfies c(y,η,t1)=c(y,η,t2)=0c(y,\eta,t_{1})=c(y,\eta,t_{2})=0. This implies p2p_{2} is conjugate to p1p_{1}. Suppose

ξ1=μ1γ˙(t1),ξ2=μ2γ˙(t2)\xi^{1}=\mu_{1}\dot{\gamma}^{\perp}(t_{1}),\quad\xi^{2}=\mu_{2}\dot{\gamma}^{\perp}(t_{2})

By (10), they satisfy

y^=μ1a1(t1)=μ2a1(t2).\hat{y}=\mu_{1}a_{1}(t_{1})=\mu_{2}a_{1}(t_{2}).

This indicates we can find nonzero μ\mu such that μ1=μa1(t2)\mu_{1}=\mu a_{1}(t_{2}) and μ2=μa1(t1)\mu_{2}=\mu a_{1}(t_{1}).

Conversely, if p2p_{2} is conjugate to p1p_{1}, then there is a nonzero Jacobi field J(t)J(t) so that J(t1)=J(t2)=0J(t_{1})=J(t_{2})=0. More precisely, we have

κ1γy+κ2γη+c1γ˙(t)+c2tγ˙(t)=0,for t=t1,t2.\kappa_{1}\frac{\partial\gamma}{\partial y}+\kappa_{2}\frac{\partial\gamma}{\partial\eta}+c_{1}\dot{\gamma}(t)+c_{2}t\dot{\gamma}(t)=0,\quad\text{for }t=t_{1},t_{2}.

The projection on γ˙(t)\dot{\gamma}^{\perp}(t) shows

(13) {κ1a1(t1)+κ2b1(t1)=0κ1a1(t2)+κ2b1(t2)=0[a1(t1)b1(t1)a1(t2)b1(t2)][κ1κ2]=0.\begin{cases}\kappa_{1}a_{1}(t_{1})+\kappa_{2}b_{1}(t_{1})=0\\ \kappa_{1}a_{1}(t_{2})+\kappa_{2}b_{1}(t_{2})=0\end{cases}\iff\left[\begin{array}[]{cc}a_{1}(t_{1})&b_{1}(t_{1})\\ a_{1}(t_{2})&b_{1}(t_{2})\end{array}\right]\left[\begin{array}[]{cc}\kappa_{1}\\ \kappa_{2}\end{array}\right]=0.

Notice κ1,κ2\kappa_{1},\kappa_{2} cannot be both zero, otherwise we have the trivial Jocabi field tangent to γ\gamma. This implies the matrix in (13) is singular. Now since ξ1=μa1(t2)γ˙(t1)\xi^{1}=\mu a_{1}(t_{2})\dot{\gamma}^{\perp}(t_{1}) and ξ2=μa1(t1)γ˙(t2)\xi^{2}=\mu a_{1}(t_{1})\dot{\gamma}^{\perp}(t_{2}) for some nonzero μ\mu, by (10) one have

y^=μa1(t2)a1(t1)=μa1(t1)a1(t2),η^=μa1(t2)b1(t1)=μa1(t1)b1(t2),\hat{y}=\mu a_{1}(t_{2})a_{1}(t_{1})=\mu a_{1}(t_{1})a_{1}(t_{2}),\quad\hat{\eta}=\mu a_{1}(t_{2})b_{1}(t_{1})=\mu a_{1}(t_{1})b_{1}(t_{2}),

where the last equality is from the zero determinant of the matrix in (13). This proves 𝒞(p1,ξ1)=𝒞(p2,ξ2)\mathcal{C}(p_{1},\xi^{1})=\mathcal{C}(p_{2},\xi^{2}). ∎

For (p1,ξ1)(p_{1},\xi^{1}) and (p2,ξ2)(p_{2},\xi^{2}) satisfying Theorem 3.1, we call them the conjugate covectors.

Theorem 3.2 ([7, Proposition 2]).

If there are no conjugate points, the normal operator N=IwIwN=I_{w}^{*}I_{w} is a Ψ\PsiDO of order 1-1 with principal symbol

σp(N)(x,ξ)=2π|ξ|(|w(x,ξ)|2+|w(x,ξ)|2),\sigma_{p}(N)(x,\xi)=\frac{2\pi}{|\xi|}(|w(x,\xi_{\perp})|^{2}+|w(x,-\xi_{\perp})|^{2}),

where ξ\xi_{\perp} is the vector conormal to the covector ξ\xi with the same length and the rotation of π/2\pi/2.

Proof.

This theorem is proved in [7, Section 4]. By the proof of [7, Proposition 2], the normal operator is a Ψ\PsiDO with the principal symbol

σp(N)(x,ξ)=2πS1A(x,0,ω)δ(ξ(ω))dω,\sigma_{p}(N)(x,\xi)=2\pi\int_{S^{1}}A(x,0,\omega)\delta(\xi(\omega))\mathop{}\!\mathrm{d}\omega,

where δ\delta is the Dirac delta function and in local coordinates we have

A(x,0,ω)=J1(x,0,ω)|w(x,ω)|2.A(x,0,\omega)=J^{-1}(x,0,\omega)|w(x,\omega)|^{2}.

Here we write

expx(t,v)x=tm(t,v;x)\exp_{x}(t,v)-x=tm(t,v;x)

and introduce the change of variables (r,w)×S1(r,w)\in\mathbb{R}\times S^{1} by

r=t|m(t,v;x)|,ω=m(t,v;x)/|m(t,v;x)|,r=t|m(t,v;x)|,\quad\omega=m(t,v;x)/|m(t,v;x)|,

with the Jacobian

J(x,t,v)det(r,ω)(t,v).J(x,t,v)\equiv\det\frac{\partial(r,\omega)}{\partial(t,v)}.

For more details, see equations (32) - (34) in [7]. Note that

m(0,v;x)=γ˙x,v(0)=vm(0,v;x)=\dot{\gamma}_{x,v}(0)=v

since we use arc length parameterization. Thus, we have that J(x,0,v)=1J(x,0,v)=1 and the principal symbol of NN is given as claimed. See also [22, Theorem 5.2] and its proof. ∎

3.4. The local problem

In this following, we present the cancellation of singularities arising in the local inverse problem for the integral transform IwI_{w}, if there are conjugate covectors. This is the analog to the case of geodesic ray transforms in [18].

Consider a fixed curve γΓ\gamma\in\Gamma with conjugate covectors (x1,ξ1)(x_{1},\xi^{1}) at t1t_{1} and (x2,ξ2)(x_{2},\xi^{2}) at t2t_{2}. Let VjV^{j} be small conic neighborhoods of (xj,ξj)(x_{j},\xi^{j}), with base UjU_{j} as a small neighborhood of xjx_{j}, for j=1,2j=1,2. With 𝒞\mathcal{C} being a local diffeomorphism, it maps a small conic neighborhood of (xj,ξj)(x_{j},\xi^{j}) to one of 𝒞(xj,ξj)\mathcal{C}(x_{j},\xi^{j}), for j=1,2j=1,2. Notice VjV^{j} should have two disjoint components, for more details see [18]. By shrinking those neighborhoods a bit, one can assume that 𝒞(V1)=𝒞(V1)𝒱\mathcal{C}(V^{1})=\mathcal{C}(V^{1})\equiv\mathcal{V}. We define the restriction 𝒞j𝒞|Vj\mathcal{C}_{j}\equiv\mathcal{C}|_{V^{j}} for j=1,2j=1,2. Note that 𝒞j\mathcal{C}_{j} are diffeomorphisms. It follows that

𝒞12𝒞11𝒞2:V2V1,𝒞21𝒞21𝒞1:V1V2\mathcal{C}_{12}\equiv\mathcal{C}_{1}^{-1}\mathcal{C}_{2}:V^{2}\rightarrow V^{1},\quad\quad\mathcal{C}_{21}\equiv\mathcal{C}_{2}^{-1}\mathcal{C}_{1}:V^{1}\rightarrow V^{2}

are also diffeomorphisms. Let IjI_{j} be IwI_{w} restricted to distributions with wave front sets supported in VjV^{j}, for j=1,2j=1,2. Then IjI_{j} are FIOs with canonical relations Cj{C}_{j}, where CjC_{j} are restriction of the canonical relation CC to Vj×𝒱V^{j}\times\mathcal{V}. When the weights are nonvanishing, the restriction IjI_{j} are elliptic FIOs, and therefore we can define

(14) F21=I21I1,F12=I11I2.\displaystyle F_{21}=I_{2}^{-1}I_{1},\quad F_{12}=I_{1}^{-1}I_{2}.

Note that they are FIOs with canonical relations C21C1C_{2}^{-1}\circ C_{1} and C11C2C_{1}^{-1}\circ C_{2} respectively. We can show the following result by the same arguments in [18].

Theorem 3.3.

Suppose w(x1,ξ1),w(x2,ξ2)0w(x_{1},\xi^{1}_{\perp}),w(x_{2},\xi^{2}_{\perp})\neq 0. Let fj(Uj)f_{j}\in\mathcal{E}^{\prime}(U_{j}) with WF(fj)Vj\text{$\operatorname{WF}$}(f_{j})\subset V^{j}, for j=1,2j=1,2. Then the local data

Iw(f1+f2)Hs(𝒱){{I_{w}}}(f_{1}+f_{2})\in H^{s}(\mathcal{V})

if and only if

f1+F12f2Hs1/2(V1)F21f1+f2Hs1/2(V2),f_{1}+F_{12}f_{2}\in H^{s-1/2}(V^{1})\Leftrightarrow F_{21}f_{1}+f_{2}\in H^{s-1/2}(V^{2}),

where F12F_{12} and F21F_{21} are elliptic FIOs defined in (14).

This theorem indicates that given a distribution f1f_{1} singular in V1V^{1}, there exists a distribution f2=F21f1f_{2}=-F_{21}f_{1} singular in V2V^{2} such that the transform is smooth. In other words, if we suppose f=f1f=f_{1}, the singularities of ff cannot be resolved from the singularities of the transform IwfI_{w}f. Indeed, the singularities of ff can only be recovered up to an error in the microlocal kernel, i.e., an error in form of (IdF21)h1(\text{Id}-F_{21})h_{1} with some h1h_{1} singular in V1V^{1}, since Iw((IdF21)h1)I_{w}((\text{Id}-F_{21})h_{1}) is always smooth. For a more detailed description, see [12].

3.5. Artifacts

In this subsection, we describe the artifacts arising in the reconstruction from the local data, when there are conjugate points. For convenience, we assume the weight w=1w=1 and use the notation II instead of IwI_{w} in the following.

First, we consider the backprojection II{{I}}^{*}{{I}} to reconstruct ff, in the presence of conjugate covectors. Suppose γ\gamma is the λ\lambda-geodesic in Theorem 3.1 with conjugate covectors (x,ξ)(x,\xi) in V1V^{1} and (y,η)(y,\eta) in V2V^{2}. Let f=f1+f2f=f_{1}+f_{2} with fjf_{j} singular in VjV^{j}, j=1,2j=1,2. In a small neighborhood of γ\gamma, we have

IIf=I1I1f1+I1I2f2+I2I1f1+I2I2f2.{{I}}^{*}{{I}}f=I_{1}^{*}I_{1}f_{1}+I_{1}^{*}I_{2}f_{2}+I_{2}^{*}I_{1}f_{1}+I_{2}^{*}I_{2}f_{2}.

Recall I1I_{1} and I2I_{2} are defined microlocally and are elliptic FIOs of order 1-1 with canonical relations C1{C}_{1} and C2{C}_{2}, which are diffeomorphisms. Then by Theorem 3.2, I1I1I_{1}^{*}I_{1} and I2I2I_{2}^{*}I_{2} are elliptic Ψ\PsiDOs of order 1-1 with principal symbol 4π/|ξ|{4\pi}/{|\xi|}. For I1I2I_{1}^{*}I_{2} and I2I1I_{2}^{*}I_{1}, by [15, Theorem 25.2.2] and the transversal composition calculus, they are FIOs of order 1-1 associated with canonical relations C11C2C^{-1}_{1}\circ C_{2} and C21C1C^{-1}_{2}\circ C_{1} respectively.

Let Δg-\Delta_{g} be the Laplacian operator in MM and its square root Δg\sqrt{-\Delta_{g}} is a Ψ\PsiDO of order 11 with principal symbol |ξ|g|\xi|_{g}. Let Λ=Δg/(4π)\Lambda=\sqrt{-\Delta_{g}}/{(4\pi)}. Then module lower order operators, one has

ΛI1I1Id,ΛI1I1Id,ΛI2I1F21,ΛI1I2F12.\Lambda I_{1}^{*}I_{1}\equiv\text{Id},\quad\Lambda I_{1}^{*}I_{1}\equiv\text{Id},\quad\Lambda I_{2}^{*}I_{1}\equiv F_{21},\quad\Lambda I_{1}^{*}I_{2}\equiv F_{12}.

It follows that

(15) ΛIIf=f1+F21f1+f2+F12f2\displaystyle\Lambda I^{*}If=f_{1}+F_{21}f_{1}+f_{2}+F_{12}f_{2}

up to lower order terms. This implies that we recover the singularities of f1+f2f_{1}+f_{2} together with F12f1+F21f2F_{12}f_{1}+F_{21}f_{2} from the backprojection. The later are artifacts. If we write a distribution singular in V1V2V^{1}\cup V^{2} as a vector-valued function with the first component equal to its restriction to V1V^{1} and the second component equal to its restriction to V2V^{2}, then

ΛIIf=[IdF12F21Id][f1f2]M[f1f2]\Lambda{{I}}^{*}{{I}}f=\begin{bmatrix}\begin{array}[]{ll}\text{Id}&F_{12}\\ F_{21}&\text{Id}\end{array}\end{bmatrix}\begin{bmatrix}\begin{array}[]{l}f_{1}\\ f_{2}\end{array}\end{bmatrix}\equiv M\begin{bmatrix}\begin{array}[]{l}f_{1}\\ f_{2}\end{array}\end{bmatrix}

up to lower-order terms. Especially when f2=0f_{2}=0, from the filtered backprojection ΛIIf\Lambda{{I}}^{*}{{I}}f, we recover f1+F21f1f_{1}+F_{21}f_{1} and therefore the artifacts equal to F21f1F_{21}f_{1} arises in the reconstruction.

Next, we consider the numerical reconstruction by using the Landweber iteration as in [12]. For more details of the method, see [24]. We follow the same argument in [18, 12], see also [25]. Let χ\chi be a smooth cutoff in 2\mathbb{R}^{2} with χ=1\chi=1 in a small neighborhood of MM to avoid dealing with non-local operators. We set =ΛχII\mathcal{L}=\Lambda\chi I^{*}I and compute

=IIχ14π(Δg)χII=[Id+F21F21F21+F12F12+F21Id+F12F12].\mathcal{L}^{*}\mathcal{L}=I^{*}I\chi\frac{1}{4\pi}({-\Delta_{g}})\chi I^{*}I=\begin{bmatrix}\begin{array}[]{ll}\text{Id}+F^{*}_{21}F_{21}&F^{*}_{21}+F_{12}\\ F_{12}^{*}+F_{21}&\text{Id}+F_{12}^{*}F_{12}\end{array}\end{bmatrix}.

Let gg be the local data and it is assumed be in the range of I{{I}}. Now we use the Landweber iteration to solve the equation If=g{{I}}f=g. We write

(16) (Id(Idγ))f=γΛχIg.(\text{Id}-(\text{Id}-\gamma\mathcal{L}^{*}\mathcal{L}))f=\gamma\mathcal{L}^{*}\Lambda\chi I^{*}g.

Then with a small enough and suitable γ>0\gamma>0, it can be solved by the Neumann series and we have the truncated scheme

f(N)=k=0N(Idγ)kγΛχIg.f^{(N)}=\sum_{k=0}^{N}(\text{Id}-\gamma\mathcal{L}^{*}\mathcal{L})^{k}\gamma\mathcal{L}^{*}\Lambda\chi I^{*}g.

This series converge to the minimal norm solution to f=ΛχIg\mathcal{L}f=\Lambda\chi I^{*}g. Suppose the original function is f=f1+f2f=f_{1}+f_{2} with f2=0f_{2}=0. The analysis in [12, Section 3.2.3] shows that

Landweber solution=[f1(Id+F21F21)1f1]+[F21(Id+F21F21)1f1],\displaystyle\text{Landweber solution}=[f_{1}-(\text{Id}+F_{21}^{*}F_{21})^{-1}f_{1}]+[F_{21}(\text{Id}+F^{*}_{21}F_{21})^{-1}f_{1}],

where the first square brackets are terms microlocally supported in V1V^{1} and the second term in V2V^{2}. The artifacts arising in the reconstruction is

fLandweber solution=[(Id+F21F21)1f1][F21(Id+F21F21)1f1].\displaystyle f-\text{Landweber solution}=[(\text{Id}+F_{21}^{*}F_{21})^{-1}f_{1}]-[F_{21}(\text{Id}+F^{*}_{21}F_{21})^{-1}f_{1}].

We emphasize that the artifacts above arises in the reconstruction from the local data. If we consider the recovery from the global data, i.e., with the knowledge of the integral transform over all curves in Γ\Gamma, then the singularities of ff might be recoverable. This is because the singularities can be probed by more than one smooth curves in Γ\Gamma. In some cases, the recovery of certain singularity depends on a discrete dynamical system, i.e., a sequence of conjugate covectors, inside MM, see (19). If this sequence goes out of MM, then we can resolve the corresponding singularity as is discussed in [23, 25]. For more details, see Proposition 7.

4. Examples of λ\lambda-geodesics with conjugate points

In this section, we present several examples of the family of curves that we study. These curves are different from geodesics and we show the conjugate points exist.

Example 1.

The first example comes from [22]. Let Γ1\Gamma_{1} consist all the unit circles in 2\mathbb{R}^{2} with a fixed orientation. These circles are actually the magnetic geodesics w.r.t the Euclidean metric and a constant non-zero magnetic field by [5]. Suppose they have a fixed orientation and we parameterize a unit circle through point p=(p1,p2)p=(p^{1},p^{2}) in the direction of v=(cosθ,sinθ)v=(\cos\theta,\sin\theta) by

γp,θ(t)=(p1,p2)+(cos(t+θ),sin(t+θ))(cosθ,sinθ).\gamma_{p,\theta}(t)=(p^{1},p^{2})+(\cos(t+\theta),\sin(t+\theta))-(\cos\theta,\sin\theta).

We have

γ˙p,θ(t)=(sin(t+θ),cos(t+θ)),γ¨p,θ(t)=(cos(t+θ),sin(t+θ)),\dot{\gamma}_{p,\theta}(t)=(-\sin(t+\theta),\cos(t+\theta)),\quad\ddot{\gamma}_{p,\theta}(t)=-(\cos(t+\theta),\sin(t+\theta)),

and therefore

γ¨p,θ(t)=(0110)γ˙p,θ(t).\ddot{\gamma}_{p,\theta}(t)=\begin{pmatrix}0&-1\\ 1&0\end{pmatrix}\dot{\gamma}_{p,\theta}(t).

One can check that the properties (A1) and (A2) are satisfied. As is shown in [22], since

(17) detds,vexpp(t,v)=detγp,θ(t)(t,θ)=sint,\displaystyle\det\mathop{}\!\mathrm{d}_{s,v}\exp_{p}(t,v)=\det\frac{\partial\gamma_{p,\theta}(t)}{\partial(t,\theta)}=\sin t,

for each point pp, it has conjugate points corresponding to t=πt=\pi and any θ\theta\in\mathbb{R}. Thus, the conjugate locus of pp is

Σ(p)={γp,θ(π),θ}={y:|yp|=2},\Sigma(p)=\{\gamma_{p,\theta}(\pi),\ \theta\in\mathbb{R}\}=\{y:\ |y-p|=2\},

which is the circle centered at pp with radius equal to 22.

Now let (p1,ξ1)TM(p_{1},\xi^{1})\in T^{*}M and suppose ff is singular near (p1,ξ1)(p_{1},\xi^{1}). We would like to find out the conjugate covector (p2,ξ2)(p_{2},\xi^{2}) of (p1,ξ1)(p_{1},\xi^{1}). To parameterize Γ1\Gamma_{1}, we assume p1p_{1} is near the origin and ξ1\xi^{1} is in a conic neighborhood of ξ0=(1,0)\xi^{0}=(1,0). Consider the line H={(y,0),y}H=\{(y,0),y\in\mathbb{R}\}. By the notations used in Section 3, a unit circle passing the point (y,0)2(y,0)\in\mathbb{R}^{2} in the direction of v=(cosη,sinη)v=(\cos\eta,\sin\eta) is now parameterized by

γ(t,y,η)=(y,0)+(cos(t+η),sin(t+η))(cosη,sinη).\gamma(t,y,\eta)=(y,0)+(\cos(t+\eta),\sin(t+\eta))-(\cos\eta,\sin\eta).

Suppose p1=(y1,0)p_{1}=(y_{1},0) and ξ1=(cosη1,sinη1).\xi^{1}=(\cos\eta_{1},\sin\eta_{1}). By Theorem 3.1, (p2,ξ2)(p_{2},\xi^{2}) and (p1,ξ1)(p_{1},\xi^{1}) are conjugate covectors when p2p_{2} is conjugate to p1p_{1} and

ξ1=μa1(t2)γ˙(t1),ξ2=μa1(t1)γ˙(t2)\xi^{1}=\mu a_{1}(t_{2})\dot{\gamma}^{\perp}(t_{1}),\quad\xi^{2}=\mu a_{1}(t_{1})\dot{\gamma}^{\perp}(t_{2})

for some nonzero μ\mu. Then by equation (17), we write

p2=γ(π,y1,η1),or p2=γ(π,y1,η1)p_{2}=\gamma(\pi,y_{1},\eta_{1}),\quad\text{or }p_{2}=\gamma(\pi,y_{1},-\eta_{1})

with t1=0,t2=πt_{1}=0,t_{2}=\pi, which depends on the different sign of μ\mu. We denote the first case by (p2(1),ξ(1)2)(p_{2}^{(1)},\xi^{2}_{(1)}) and the second case by (p2(2),ξ(2)2)(p_{2}^{(2)},\xi^{2}_{(2)}). We compute

p2(1)=(y1,0)2(cosη1,sinη1),p2(2)=(y1,0)+2(cosη1,sinη1).p^{(1)}_{2}=(y_{1},0)-2(\cos\eta_{1},\sin\eta_{1}),\quad p^{(2)}_{2}=(y_{1},0)+2(\cos\eta_{1},\sin\eta_{1}).

Since a1(t)=(γ˙(t),γ(t,y,η)y)a_{1}(t)=(\dot{\gamma}^{\perp}(t),\frac{\partial\gamma(t,y,\eta)}{\partial y}), we can find

ξ(1)2=ξ1,ξ(2)2=ξ1.\xi^{2}_{(1)}=\xi^{1},\quad\xi^{2}_{(2)}=\xi^{1}.

Therefore, there are two covectors that are conjugate to (p1,ξ1)(p_{1},\xi^{1}) given by

(18) (p2(1),ξ(1)2)=(p12(cosη1,sinη1),ξ1),(p2(2),ξ(2)2)=(p1+2(cosη1,sinη1),ξ1).(p_{2}^{(1)},\xi^{2}_{(1)})=(p_{1}-2(\cos\eta_{1},\sin\eta_{1}),\xi^{1}),\quad(p_{2}^{(2)},\xi^{2}_{(2)})=(p_{1}+2(\cos\eta_{1},\sin\eta_{1}),\xi^{1}).
Example 2.

In this example, let Γ2\Gamma_{2} consist all identical ellipses with a fixed orientation starting from different points

γp(t)=(p1,p2)+(acosta,bsint),\gamma_{p}(t)=(p^{1},p^{2})+(a\cos t-a,b\sin t),

where a,ba,b are constants. Different values of a,ba,b give us different Γ2\Gamma_{2}. We have

γ˙p(t)=(asint,bcost),γ¨p(t)=(acost,bsint).\dot{\gamma}_{p}(t)=(-a\sin t,b\cos t),\quad\ddot{\gamma}_{p}(t)=-(a\cos t,b\sin t).

One can verify that Γ\Gamma satisfies the properties (A1) and (A2). Indeed, first we have

γ¨p(t)=(0a/bb/a0)γ˙p(t).\ddot{\gamma}_{p}(t)=\begin{pmatrix}0&-{a}/{b}\\ {b}/{a}&0\end{pmatrix}\dot{\gamma}_{p}(t).

Then, for any point x=(x1,x2)x=(x_{1},x_{2}) and v=(cosθ,sinθ)v=(\cos\theta,\sin\theta), there is a curve γp\gamma_{p} passing xx in the direction of vv if we can find some p0,t0p_{0},t_{0} such that

asint0=μcosθ,bcost0=μsinθ,γp(t)=x,-a\sin t_{0}=\mu\cos\theta,\quad b\cos t_{0}=\mu\sin\theta,\quad{\gamma}_{p}(t)=x,

for some μ>0\mu>0. This is true since we can solve t0t_{0} from

cost0=μacosθ,sint0=μbsinθ,\cos t_{0}=\frac{-\mu}{a}\cos\theta,\quad\sin t_{0}=\frac{\mu}{b}\sin\theta,

where μ=abb2cos2θ+a2sin2θ\mu=\frac{ab}{\sqrt{b^{2}\cos^{2}\theta+a^{2}\sin^{2}\theta}} and then p0=x(acost0a,bsint0)p_{0}=x-(a\cos t_{0}-a,b\sin t_{0}). One can see that t0t_{0} is unique in the sense of modulo 2π2\pi and it depends on θ\theta in a smooth way. We denote it by t0(θ)t_{0}(\theta) and compute

dt0dθ=b2cos2θ+a2sin2θab.\displaystyle\frac{\mathop{}\!\mathrm{d}t_{0}}{\mathop{}\!\mathrm{d}\theta}=\frac{b^{2}\cos^{2}\theta+a^{2}\sin^{2}\theta}{ab}.

To parametrize Γ2\Gamma_{2}, we pick H=(0,x2)H={(0,x^{2})}. Notice for every point x=(0,x2)Hx=(0,x^{2})\in H and any direction v=(cosθ,sinθ)v=(\cos\theta,\sin\theta), there is a unique γp(t)\gamma_{p}(t) passing xx in the direction of vv. From the analysis above, we reparameterize this ellipse as

ρx2,θ(t)=(acos(t+t0)acost0,bsin(t+t0)bsint0+x2).\rho_{x^{2},\theta}(t)=(a\cos(t+t_{0})-a\cos t_{0},b\sin(t+t_{0})-b\sin t_{0}+x^{2}).

Thus, for any x=(0,x2)x=(0,x^{2}) we have

detds,vexpx(t,v)=detρx2,θ(t)(t,θ)=dt0dθabsint,\det\mathop{}\!\mathrm{d}_{s,v}\exp_{x}(t,v)=\det\frac{\partial\rho_{x^{2},\theta}(t)}{\partial(t,\theta)}=\frac{\mathop{}\!\mathrm{d}t_{0}}{\mathop{}\!\mathrm{d}\theta}ab\sin t,

and its conjugate points are corresponding to t=πt=\pi and any θ\theta. This gives us the conjugate locus Σ(x)={ρx2,θ(π),θ}={x2(acost0,bsint0),t0}\Sigma(x)=\{\rho_{x^{2},\theta}(\pi),\ \theta\in\mathbb{R}\}=\{x-2(a\cos t_{0},b\sin t_{0}),t_{0}\in\mathbb{R}\}, which is a larger ellipse centered at xx.

Example 3.

In this example, we choose the a family of smooth curves that are locally defined. For p=(p1,p2)p=(p^{1},p^{2}), define

γp(t)=eat(cos(tφ)cos(φ),sin(tφ)sin(φ))+p,\gamma_{p}(t)=e^{at}(\cos(t-\varphi)-\cos(-\varphi),\sin(t-\varphi)-\sin(-\varphi))+p,

where aa is a nonzero constant and φ\varphi satisfies

cosφ=aa2+1,sinφ=1a2+1.\cos\varphi=\frac{a}{\sqrt{a^{2}+1}},\quad\sin\varphi=\frac{1}{\sqrt{a^{2}+1}}.

We compute

γ˙p(t)=eata2+1(cost,sint),γ¨p(t)=eata2+1(acostsint,asint+cost),\dot{\gamma}_{p}(t)=e^{at}\sqrt{a^{2}+1}(\cos t,\sin t),\quad\ddot{\gamma}_{p}(t)=e^{at}\sqrt{a^{2}+1}(a\cos t-\sin t,a\sin t+\cos t),

and one can show that

γ¨p(t)=(a11a)γ˙p(t),\ddot{\gamma}_{p}(t)=\begin{pmatrix}a&-1\\ 1&a\end{pmatrix}\dot{\gamma}_{p}(t),

which implies the property (A2). To satisfy the property (A1) at least locally, we fix a curve

γ0(t)=eat(cos(tφ),sin(tφ)), for t(π4,7π4)\gamma_{0}(t)=e^{at}(\cos(t-\varphi),\sin(t-\varphi)),\text{ for }t\in(-\frac{\pi}{4},\frac{7\pi}{4})

which passes the point (0,0)(0,0) at t=0t=0 in the direction θ=0\theta=0. Now let

Γ3={γ0(t)+p, where t(π4,7π4)}.\Gamma_{3}=\{\gamma_{0}(t)+p,\text{ where }t\in(-\frac{\pi}{4},\frac{7\pi}{4})\}.

Then Γ3\Gamma_{3} is an open family of smooth curves. For each point x=(x1,x2)x=(x_{1},x_{2}) and direction v=(cosθ,sinθ)v=(\cos\theta,\sin\theta) with θπ4\theta\neq-\frac{\pi}{4}, there is a unique

t0=θ,p0=xeaθ(cos(θφ),sin(θφ)+(cosφ,sinφ),t_{0}=\theta,\quad p_{0}=x-e^{a\theta}(\cos(\theta-\varphi),\sin(\theta-\varphi)+(\cos\varphi,-\sin\varphi),

such that γp0(t0)\gamma_{p_{0}}(t_{0}) through xx in the direction of vv. We reparameterize Γ3\Gamma_{3} using H=(0,x2)H={(0,x^{2})} by writing these curves as

ρx2,θ(t)\displaystyle\rho_{x^{2},\theta}(t) =ea(t+θ)(cos(t+θφ),sin(t+θφ))\displaystyle=e^{a(t+\theta)}(\cos(t+\theta-\varphi),\sin(t+\theta-\varphi))
eaθ(cos(θφ),sin(θφ)))+(0,x2), for t(π4,7π4).\displaystyle-e^{a\theta}(\cos(\theta-\varphi),\sin(\theta-\varphi)))+(0,x^{2}),\text{ for }t\in(-\frac{\pi}{4},\frac{7\pi}{4}).

Notice that ρx2,θ(t)\rho_{x^{2},\theta}(t) passes (0,x2)(0,x^{2}) at t=0t=0 in the direction of θ\theta. We mention that Γ3\Gamma_{3} forms a neighborhood of ρ0,0(t)\rho_{0,0}(t) and we compute

detds,vexpx(t,v)=detρx2,θ(t)(t,θ)=e2a+θsint.\det\mathop{}\!\mathrm{d}_{s,v}\exp_{x}(t,v)=\det\frac{\partial\rho_{x^{2},\theta}(t)}{\partial(t,\theta)}=e^{2a+\theta}\sin t.

Thus, for the point x=(0,0)x=(0,0), there are conjugate points corresponding to t=±πt=\pm\pi and θ(π/4,7π/4)\theta\in(-{\pi}/{4},{7\pi}/{4}). The conjugate locus is

Σ(x)\displaystyle\Sigma(x) ={ρ0,θ(±π),θ}\displaystyle=\{\rho_{0,\theta}(\pm\pi),\ \theta\in\mathbb{R}\}
={(e±aπ+1)eaθ(cos(θφ),sin(θφ)), for θ(π4,7π4)}.\displaystyle=\{-(e^{\pm a\pi}+1)e^{a\theta}(\cos(\theta-\varphi),\sin(\theta-\varphi)),\text{ for }\theta\in(-\frac{\pi}{4},\frac{7\pi}{4})\}.

In Figure 1, we choose a=1/4a=-1/4 (this implies φ=arccos(1/17)\varphi=\arccos(1/\sqrt{17})) and we plot all ρ0,θ(t)\rho_{0,\theta}(t) for θ(π/4,π/4)\theta\in(-{\pi}/{4},{\pi}/{4}). The light green curve is part of the conjugate locus Σ(x)\Sigma(x).

Refer to caption
Figure 1. Light green curve: part of conjugate locus Σ(x)\Sigma(x).

5. Numerical experiments

This section aims to illustrate the artifacts arising in the reconstruction by numerical experiments. We consider the family of unit circles with a fixed orientation, i.e., the magnetic geodesics w.r.t. the Euclidean metric and a constant nonzero magnetic field. See Example 1 in Section 4 and [22] for more details.

More explicitly, let MM be a bounded domain without boundary in 2\mathbb{R}^{2}, for example, the open disk of radius R=3R=3 centered at the origin. Suppose ff is a smooth function supported in MM. We define

If(x1,x2)=|w|=1f((x1,x2)+w)𝑑lw=02πf(x1+cosα,x2+sinα)𝑑αIf(x_{1},x_{2})=\int_{|w|=1}f((x_{1},x_{2})+w)dl_{w}=\int_{0}^{2\pi}f(x_{1}+\cos\alpha,x_{2}+\sin\alpha)d\alpha

as the integral transform performed over unit circles with radius 11. These circles can be parameterized by their centers. Then the adjoint operator is

Ig(y1,y2)=|w|=1g((y1+y2)w)𝑑lw=02πg(y1cosα,y2sinα)𝑑αI^{*}g(y_{1},y_{2})=\int_{|w|=1}g((y_{1}+y_{2})-w)dl_{w}=\int_{0}^{2\pi}g(y_{1}-\cos\alpha,y_{2}-\sin\alpha)d\alpha

which coincides with II itself. Numerically we compute IfIf by the following steps.

  • (1)

    Discretization. We introduce 6N×6N6N\times 6N equispaced points in the square domain [3,3]×[3,3][-3,3]\times[-3,3] so the grid spacing is 1/N{1}/{N}. We discretize the input function ff over the grids.

  • (2)

    We compute the integral by the Trapezoidal method

    If(x1,x2)=k=1nf(zk1,wk1)+f(zk,wk)2Δα+O(1n2),If(x_{1},x_{2})=\sum_{k=1}^{n}\frac{f(z_{k-1},w_{k-1})+f(z_{k},w_{k})}{2}\Delta\alpha+O(\frac{1}{n^{2}}),

    where zk=x1+cos2πk/n,wk=x2+sin2πk/nz_{k}=x_{1}+\cos{{2\pi k}/{n}},w_{k}=x_{2}+\sin{{2\pi k}/{n}} and Δα=2πn\Delta\alpha=\frac{2\pi}{n}. When the input function is only given over grids, we can use the bilinear interpolation to approximate its values at {(zk,wk),k=1,,n}\{(z_{k},w_{k}),k=1,\ldots,n\} in Step (2) before the numerical integration.

Then we use the same method to compute II^{*}.

5.1. Backprojection

First, we consider the reconstruction of ff from the transform by the backprojection IIfI^{*}If. To check the numerical implementation of the backprojection, one can compare the numerical result with the analytical one given by the formula

IIf(x1,x2)=0202π44r2f(x1+rcosθ,x2+rsinθ)𝑑θdr.I^{*}If(x_{1},x_{2})=\int_{0}^{2}\int_{0}^{2\pi}\frac{4}{4-r^{2}}f(x_{1}+r\cos\theta,x_{2}+r\sin\theta)d\theta\mathop{}\!\mathrm{d}r.

in [22]. Let f1f_{1} be a truncated Gaussian concentrated near the origin, as an approximation of a delta function at the origin, and let f2=0f_{2}=0. We choose N=40N=40 and the relative error between the numerical result RNR_{N} and the analytical one

RNIIf2IIf2=0.0140\frac{\|{R_{N}-I^{*}If}\|_{2}}{\|I^{*}If\|_{2}}=0.0140

is relatively small. In this case, ff is discretized as a matrix of 240×240240\times 240 and we integrate it over unit circles. From Figure 2, we can see the artifacts appear exactly in the location of conjugate points. Indeed, by the analysis in Example 1, the conjugate locus of the origin is the circle centered at the origin of radius 22. With ff as an approximation of the delta function, the singularities of ff are located near the origin in all directions. The conjugate covectors of the singularities of the delta function are described by the equation (18), which are covectors conormal to the circle of radius 22. Then by (15), from the backprojection we recover both the singularities of f=f1f=f_{1} and the singularities of F12f1F_{12}f_{1}. Note here we can omit Λ\Lambda in (15) if we only consider the singularities, since Λ\Lambda is an elliptic Ψ\PsiDO.

Refer to caption
Refer to caption
Figure 2. The true ff and the reconstruction from the backprojection.

5.2. Landweber iteration

In the following, we choose f=f1f=f_{1} to be a modified Gaussian with singularities located both in certain space and in certain direction, that is, a coherent state, as is shown in Figure 3 (a). Note the singularities of ff are actually semi-classical and they are located near the origin. The same analysis works (see [21] for FIOs and semi-classical wave front sets) and here we use this coherent state to illustrate the artifacts in the analysis before. If we use the backprojection to reconstruct ff, the artifacts appear in the location of conjugate points, see Figure 3 (b). This is described by (18) and (15).

Next, we use the Landweber iteration to reconstruct ff. The analysis in Section 3.5 explains the artifacts in the reconstruction from the local data. However, if we use the global data and have the prior knowledge that ff is supported in a compact set, then we can recover the singularities of ff without artifacts. The following proposition is an analog to [25, Corrollary 3].

Proposition 7.

Suppose f(2)f\in\mathcal{E}^{\prime}(\mathbb{R}^{2}) and IfC()If\in C^{\infty}(\mathcal{M}). Then ff is smooth.

Proof.

Let (x0,ξ0)TM(x_{0},\xi^{0})\in T^{*}M and assume it is in the wave front set of ff. This singularity can be canceled by its conjugate covectors, if they exist. As in [25, Section 5], we define

(19) (x0,ξ0)={(xk,ξk),\displaystyle\mathcal{M}(x_{0},\xi^{0})=\{(x_{k},\xi^{k}), if it exists and is conjugate to (xk,ξk),\displaystyle\text{ if it exists and is conjugate to $(x_{k^{\prime}},\xi^{k^{\prime}})$},
where k=ksgnk,for k=±1,±2,}\displaystyle\text{where }k^{\prime}=k-\operatorname{sgn}k,\text{for }k=\pm 1,\pm 2,\ldots\}

as the set of all conjugate covectors related to (x0,ξ0)(x_{0},\xi^{0}). We can assume ξ0=(cosη0,sinη0)\xi^{0}=(\cos\eta_{0},\sin\eta_{0}) is in a conic neighborhood of the covector ξ=(1,0)\xi=(1,0). By (18), we have

(x0,ξ0)={(xk,ξ0):xk=x0+2k(cosη0,sinη0), for k=±1,±2,}.\displaystyle\mathcal{M}(x_{0},\xi^{0})=\{(x_{k},\xi^{0}):x_{k}=x_{0}+2k(\cos\eta_{0},\sin\eta_{0}),\text{ for }k=\pm 1,\pm 2,\ldots\}.

Let VkV^{k} be a small conic neighborhoods of (xk,ξk)(x0,ξ0)(x_{k},\xi^{k})\in\mathcal{M}(x_{0},\xi^{0}). Let fkf_{k} be ff microlocally restricted to VkV^{k} and IkI_{k} be II restricted to distributions singular in VkV^{k}. For each kk, near the curve γk\gamma_{k} where (xk1,ξk1)(x_{k-1},\xi^{k-1}) and (xk,ξk)(x_{k},\xi^{k}) are conjugate covectors, we have the equation of cancellation of singularities

(20) Ik1fk1+Ikfk=0modC,I_{k-1}f_{k-1}+I_{k}f_{k}=0\mod C^{\infty},

if we shrink VkV^{k} such that C(Vk)=Vk1C(V^{k})=V^{k-1}. Since ff has compact support, there exist k+>0k_{+}>0 and k<0k_{-}<0 such that ff is smooth near all (xk,ξk)(x_{k},\xi^{k}) with k<k+k<k_{+} or k<kk<k_{-}. Then we have

Ikfk=0modC, for any k<k+ or k<k.I_{k}f_{k}=0\mod C^{\infty},\text{ for any }k<k_{+}\text{ or }k<k_{-}.

By (20), it follows that Ikfk=0modCI_{k}f_{k}=0\mod C^{\infty} for all kk. Thus, we have ff is smooth. ∎

Refer to caption
(a) true ff
Refer to caption
(b) backprojection f(1)f^{(1)}
Refer to caption
(c) f(100)f^{(100)}
Figure 3. Reconstruction for a modified Gaussian function ff, where e=ff(100)2f2e=\frac{\|f-f^{(100)}\|_{2}}{\|f\|_{2}} is the relative error.

In the numerical experiment, suppose ff is compactly supported in the disk DD that is centered at the origin with radius 33. When we use Landweber iteration, after performing the backprojection operator in each step, we smoothly cut the function such that it is still supported in radius 33. More explicitly, we consider the operator φ=φ=φΛχII,\mathcal{L}_{\varphi}=\varphi\mathcal{L}=\varphi\Lambda\chi I^{*}I, where φ\varphi is a smooth cutoff function with φ=1\varphi=1 in DD and supported in a slightly larger disk DD^{\prime}. With ff compacted supported in DD, we have φ:L2(D)L2(D).\mathcal{L}_{\varphi}:L^{2}(D^{\prime})\rightarrow L^{2}(D^{\prime}). In this case, to solve the equation If=g{{I}}f=g, we write

(Id(Idγφφ))f=γφ(φΛχIg)(\text{Id}-(\text{Id}-\gamma\mathcal{L}_{\varphi}^{*}\mathcal{L}_{\varphi}))f=\gamma\mathcal{L}_{\varphi}^{*}(\varphi\Lambda\chi I^{*}g)

and use the truncated Neumann series

f(N)=k=0N(Idγφφ)kγφ(φΛχIg).f^{(N)}=\sum_{k=0}^{N}(\text{Id}-\gamma\mathcal{L}^{*}_{\varphi}\mathcal{L}_{\varphi})^{k}\gamma\mathcal{L}^{*}_{\varphi}(\varphi\Lambda\chi I^{*}g).

This series converges to the minimal norm solution to φf=φΛχIg\mathcal{L}_{\varphi}f=\varphi\Lambda\chi I^{*}g in L2(D)L^{2}(D^{\prime}). After 100 steps of iteration, we get a quite good reconstruction (with the relative error e=0.055e=0.055). This illustrates the result of Proposition 7.

References

  • [1] Gareth Ainsworth. The attenuated magnetic ray transform on surfaces. Inverse Problems and Imaging, 7, 2013.
  • [2] Yernat M. Assylbekov and Nurlan S. Dairbekov. The X-ray transform on a general family of curves on Finsler surfaces. The Journal of Geometric Analysis, 2017.
  • [3] Ioan Bucataru and Radu Miron. Finsler-Lagrange geometry: Applications to dynamical systems. Editura Academiei Romane Bucureşti, 2007.
  • [4] Nurlan S Dairbekov and Gabriel P Paternain. Entropy production in Gaussian thermostats. Communications in mathematical physics, 269(2):533–543, 2007.
  • [5] Nurlan S. Dairbekov, Gabriel P. Paternain, Plamen Stefanov, and Gunther Uhlmann. The boundary rigidity problem in the presence of a magnetic field. Adv. Math., 216(2):535–609, 2007.
  • [6] Jesse Douglas. The general geometry of paths. Annals of Mathematics, pages 143–168, 1927.
  • [7] Bela Frigyik, Plamen Stefanov, and Gunther Uhlmann. The X-ray transform for a generic family of curves and weights. Journal of Geometric Analysis, 18(1):89–108, 2007.
  • [8] I. M. Gelfand, M. I. Graev, and Z. Ya. Shapiro. Differential forms and integral geometry. Functional Analysis and Its Applications, 3(2):101–114, 1969.
  • [9] Viktor L Ginzburg. On closed trajectories of a charge in a magnetic field. an application of symplectic geometry. Contact and symplectic geometry (Cambridge, 1994), 8:131–148, 1996.
  • [10] Victor Guillemin and Shlomo Sternberg. Some problems in integral geometry and some related problems in micro-local analysis. American Journal of Mathematics, 101(4):915, 1979.
  • [11] Sigurdur Helgason. The Radon transform. Springer US, 2013.
  • [12] Sean Holman, François Monard, and Plamen Stefanov. The attenuated geodesic X-ray transform. Inverse Problems, 34(6):064003, 2018.
  • [13] Sean Holman and Gunther Uhlmann. On the microlocal analysis of the geodesic X-ray transform with conjugate points. volume 108, pages 459–494. Lehigh University, 2018.
  • [14] Andrew Homan and Hanming Zhou. Injectivity and stability for a generic class of generalized Radon transforms. The Journal of Geometric Analysis, 27(2):1515–1529, 2017.
  • [15] Lars Hörmander. The Analysis of Linear Partial Differential Operators IV. Springer Berlin Heidelberg, 2009.
  • [16] Peer Christian Kunstmann, Eric Todd Quinto, and Andreas Rieder. Seismic imaging with generalized radon transforms: stability of the Bolker condition. 2021.
  • [17] Will Merry and Gabriel Paternain. Inverse Problems in Geometry and Dynamics Lecture notes. incomplete version, 2011.
  • [18] François Monard, Plamen Stefanov, and Gunther Uhlmann. The geodesic ray transform on Riemannian surfaces with conjugate points. Comm. Math. Phys., 337(3):1491–1513, 2015.
  • [19] Gabriel Paternain, Miko Salo, and Gunther Uhlmann. Geometric inverse problems, with emphasis in two dimensions. Cambridge University Press, to appear.
  • [20] Zhongmin Shen. Differential geometry of spray and Finsler spaces. Springer Science & Business Media, 2013.
  • [21] Plamen Stefanov and Samy Tindel. Sampling linear inverse problems with noise. arXiv preprint arXiv:2011.13489, 2020.
  • [22] Plamen Stefanov and Gunther Uhlmann. The geodesic X-ray transform with fold caustics. Anal. PDE, 5(2):219–260, 2012.
  • [23] Plamen Stefanov and Gunther Uhlmann. Is a curved flight path in SAR better than a straight one? SIAM J. Appl. Math., 73(4):1596–1612, 2013.
  • [24] Plamen Stefanov and Yang Yang. Multiwave tomography with reflectors: Landweber’s iteration. Inverse Problems & Imaging, 11(2), 2017.
  • [25] Yang Zhang. Artifacts in the inversion of the broken ray transform in the plane. Inverse Problems & Imaging, 14(1):1–26, 2020.
  • [26] Hanming Zhou. The local magnetic ray transform of tensor fields. SIAM Journal on Mathematical Analysis, 50(2):1753–1778, 2018.