You are on page 1of 25

SNELL’S LAW REVISITED AND GENERALIZED

VIA FINSLER GEOMETRY

STEEN MARKVORSEN AND ENRIQUE PENDÁS-RECONDO

Abstract. We study the variational problem of finding the fastest path


between two points that belong to different anisotropic media, each with
a prescribed speed profile and a common interface. The optimal curves
arXiv:2207.13515v1 [math.DG] 27 Jul 2022

are Finsler geodesics that are refracted – broken – as they pass through
the interface, due to the discontinuity of their velocities. This “breaking”
must satisfy a specific condition in terms of the Finsler metrics defined
by the speed profiles, thus establishing the generalized Snell’s law. In the
same way, optimal paths bouncing off the interface – without crossing
into the second domain – provide the generalized law of reflection. The
classical Snell’s and reflection laws are recovered in this setting when
the velocities are isotropic. If one considers a wave that propagates in
all directions from a given ignition point, the trajectories that globally
minimize the traveltime generate the wavefront at each instant of time.
We study in detail the global properties of such wavefronts in the Eu-
clidean plane with anisotropic speed profiles. Like the individual rays,
they break when they encounter the discontinuity interface. But they
are also broken due to the formation of cut loci – stemming from the
self-intersection of the wavefronts – which typically appear when they
approach a high-speed profile domain from a low-speed profile.

Keywords: Snell’s law, Fermat’s principle, Huygens’ principle, Zermelo


navigation, Finsler metrics, anisotropic discontinuous media, rays with
least traveltime, wave propagation.

2020 Mathematics Subject Classification. 53B40, 53C60, 53C22, 53Z05, 78M30.


1
2 S. MARKVORSEN AND E. PENDÁS-RECONDO

Contents
1. Introduction 2
Outline of the paper 3
2. Initial concepts 3
2.1. Traveltime and Finsler metrics 3
2.2. Snell’s law; the general setting 5
3. Time-minimizing trajectories 6
3.1. Variational definitions and conventions 6
3.2. Critical points of the traveltime functional 8
3.3. Generalized Snell’s law 9
3.4. The two-dimensional case 9
4. Reflected trajectories 11
4.1. The law of reflection 11
4.2. The total reflection phenomenon 13
5. Global considerations in R2 13
5.1. Existence and uniqueness of solution 14
5.2. New reflected trajectories 15
5.3. Globally time-minimizing trajectories 17
5.4. Anisotropic example 18
5.5. Construction of the wavefront 20
5.6. The cut locus 22
Acknowledgments 24
References 24

1. Introduction
Snell’s law is – in physical terms – concerned with the refraction of light
which in space passes from a point q1 to another point q2 through an interface
between two (usually isotropic) media Q1 and Q2 with different refraction
indices n1 and n2 , say. In its simplest form the law expresses the break of
the light path at the interface in terms of the incoming angle, the outgoing
angle, and the ratio between the two refraction indices. The light is supposed
to have a constant speed c/n1 in all directions in Q1 and similarly c/n2 in
Q2 (where c denotes the speed of light in vacuum). For convenience of the
reader, this classical formula is presented and discussed in Examples 3.7 and
4.5 below – in the more general framework of the present paper.
Snell’s law is classically derived from Fermat’s principle, which states that
the path of the light is the one which takes the least time to go from q1 to
q2 . This principle will also be our starting point. It motivates directly
the variational approach to the task of finding optimal traveltimes – also
through the interface of anisotropic inhomogeneous domains which is the
main concern of the present paper. The anisotropy is here given by having
different prescribed speed profiles V1 and V2 in the two domains. These
speed profiles, that we now allow to depend on both the position and the
direction of the travel path, are considered in a given Riemannian manifold
(Q, g) which thus defines the background for the domains Q1 ⊂ Q, Q2 ⊂ Q,
their interface η ⊂ Q as well as the space of paths under consideration.
SNELL’S LAW REVISITED 3

We then show how the traveltime through the interface η in this general
situation is minimized by broken geodesics of two Finsler metrics – one
on each domain – which are determined by the ambient metric g and the
prescribed speed profiles Vi . And we show and illustrate how the “breaking”
at the interface must satisfy a specific condition on these Finsler metrics,
i.e. the generalized Snell’s law. We note that the classical Snell’s law (with
isotropic constant speed profiles c/ni in (Q, g)) of course also follows from
the general version – but then with simpler Finsler metrics which are just
the conformally modified background metrics ρ2i · g, where the conformal
factors are ρi = ni /c.
Once the full set of relevant (possibly broken) time-parametrized Finsler
geodesics issuing from one point q1 have been found, they give rise to a folia-
tion of the domain Q by the wavefronts, each consisting of those points which
have the same traveltime from q1 – assuming that the geodesic equations
have solutions for sufficiently large time. Some geodesics are not globally
time-minimizing, so they do not contribute to the wavefront. This means
that they have encountered a cut locus on their way from q1 . We discuss in
detail, and illustrate via a simple example (with constant speed profiles Vi
in R2 ), how such cut loci typically appear in situations where V2 is “larger”
than V1 . In this way – already in this example – the creation of cut loci
follows from the discontinuous speed profile in Q notwithstanding the fact
that there is neither curvature nor structural topology present in neither Q1
nor Q2 .
Outline of the paper. We begin in § 2 with the equivalence between the
time spent when following a curve with a prescribed anisotropic velocity and
the length of the curve computed with a certain Finsler metric. As alluded
to above, when the velocity changes discontinuously from one medium to
another, we have a situation that generalizes the classical Snell’s law. In § 3
we apply calculus of variations to find the critical points of the traveltime
functional, i.e., the refracted trajectories, thus arriving at the generalized
Snell’s law (9). The precise results are summarized in Theorem 3.4. Follow-
ing an analogous procedure, we derive in § 4 a similar result, Theorem 4.1,
for the reflected trajectories, including the generalized law of reflection (11).
Finally, in § 5 we reduce the setting to R2 with constant Finsler metrics
(i.e., Minkowski norms). This allows us to address some global issues, such
as the existence and uniqueness of solutions to Snell’s law and the law of re-
flection, which are formulated in Theorems 5.2 and 5.3, as well as the study
of the global minima of the traveltime functional. When considering rays
that are shot in all directions from a given point of ignition, the trajectories
that globally minimize the traveltime are the ones that generate the wave-
front at each given time. This phenomenon is described and illustrated in
Theorem 5.8. When a wavefront overlaps itself (at the intersection of two
rays) it forms the cut locus, which is studied and characterized for a specific
example in § 5.6.

2. Initial concepts
2.1. Traveltime and Finsler metrics. Let (Q, g) be a smooth Riemann-
ian manifold of dimension n ≥ 2, and consider also its tangent bundle
4 S. MARKVORSEN AND E. PENDÁS-RECONDO

T Q. The metric g will measure actual distances in the space Q, i.e., if


γ : [0, t0 ] → Q is a (piecewise smooth) curve with tangent vector γ 0 (t), then
the distance traveled – when following this trajectory – is
Z t0 p Z t0
Lg [γ] := 0 0
g(γ (t), γ (t))dt = kγ 0 (t)kg dt (1)
0 0

(splitting the integral accordingly at any break). This distance, this length,
is independent of the parametrization of γ.
Assume now that we travel along γ with a prescribed velocity. We will
allow this velocity to be anisotropic, i.e., it may depend on the point of the
space and also on the direction. This means that there is a function

V : T Q \ 0 → (0, ∞)

such that V (p, v) represents a fixed speed at p ∈ Q in the direction v ∈


Tp Q \ {0}. Since V (p, v) depends on the (oriented) direction of v but not
on its length, this function must be positively homogeneous of degree zero,
i.e., V (p, v) = V (p, λv) for all λ > 0. Then the time spent traveling along γ
with this prescribed velocity is

kγ 0 (t)kg
Z t0
T [γ] := 0
dt, (2)
0 V (γ(t), γ (t))

which is again independent of reparametrizations (as long as they preserve


the orientation of γ) thanks to the zero-homogeneity of V . The preservation
of orientation is necessary, since in general we will not assume that V (p, v) =
V (p, −v). Nevertheless, observe that if γ is parametrized in such a way
that its tangent vector always coincides with the velocity given by V , i.e.
kγ 0 (t)kg = V (γ(t), γ 0 (t)), then the parameter t directly measures the time,
as T [γ] = t0 . We say in this case that γ is time-parametrized.
Let us denote by F the function given by the integrand in (2), i.e.,

F: T Q −→ [0, ∞)
kvkg
(p, v) 7−→ F (p, v) := V (p,v) ,

with F (p, 0) := 0 for all p ∈ Q. Then F is a Finsler metric if and only if


V is smooth and the following oval (the indicatrix of F ), that it defines at
each p ∈ Q is a strongly convex hypersurface of Tp Q:1

Σp := {v ∈ Tp Q : kvkg = V (p, v)} = {v ∈ Tp Q : F (p, v) = 1}.

Note that the length of γ computed with F coincides with the traveltime
along γ with the prescribed velocity V :
Z t0
LF [γ] := F (γ(t), γ 0 (t))dt = T [γ]. (3)
0

1Using any Euclidean metric in T Q, this means that the second fundamental form of
p
Σp with respect to the inner normal vector is positive definite or, equivalently, that Σp
has positive sectional curvature everywhere (see [16, Prop. 2.3]).
SNELL’S LAW REVISITED 5

2.2. Snell’s law; the general setting. In the classical Snell’s law set-
ting, a light ray travels between two different regions in R2 separated by a
straight line. Let us generalize this framework by having two open subsets
Q1 , Q2 ⊂ Q such that their closures Q1 , Q2 are smooth manifolds with the
same smooth boundary η := ∂Q1 = ∂Q2 and Q1 ∪ Q2 = Q. Assume also
that we have two different velocity functions – speed profiles – V1 , V2 defined
on T Q1 and T Q2 , respectively, and let F1 , F2 be their corresponding Finsler
metrics, so that (Q1 , F1 ) and (Q2 , F2 ) are smooth Finsler manifolds (and
(Q, g) is still a Riemannian one).
Then, if γ : [0, t0 ] → Q is a curve going from γ(0) = q1 ∈ Q1 to γ(t0 ) =
q2 ∈ Q2 and crossing η once at γ(τ ), for some τ ∈ (0, t0 ), we can still
compute the traveltime by
Z τ Z t0
T [γ] = F1 (γ(t), γ 0 (t))dt + F2 (γ(t), γ 0 (t))dt, (4)
0 τ
Rτ Rh R t0 Rt
where, formally, 0 := limh→τ − 0 and τ := limh→τ + h0 . To avoid inte-
grability and differentiability issues, we will further assume that F1 and F2
can be smoothly extended to η, so that the integrals always exist and we
can also compute F1 (γ(τ ), γ 0 (τ − )), F2 (γ(τ ), γ 0 (τ + )) and their derivatives.
Definition 2.1. Fixing an interval [0, t0 ], some τ ∈ (0, t0 ) and two points
q1 ∈ Q1 , q2 ∈ Q2 , let N be the set of all piecewise smooth curves γ : [0, t0 ] →
Q from γ(0) = q1 to γ(t0 ) = q2 and crossing η once at γ(τ ). We will call
TN : N → R defined by (4) the traveltime functional on N .
Note that any curve from q1 to q2 (piecewise smooth and crossing η once)
can be reparametrized so that it belongs to N , without affecting its travel-
time. Therefore, the choice of t0 and τ is arbitrary and non-restrictive (as
well as the choice of 0 as the initial time).
Among all the possible paths in N , light (or any wave, in general) will
choose the one that minimizes the traveltime fuctional, according to the
classical Fermat’s principle. In a smooth Finsler manifold (Q, F ), it is well
known that this trajectory (if it exists) must be a geodesic of F (see e.g. [2,
§ 5]). In our setting, we will have an F1 -geodesic from q1 to η (the incident
trajectory) and an F2 -geodesic from η to q2 (the refracted trajectory), with
a break at the point where the curve meets η. The goal is to find the change
in the direction at the break point, i.e., the generalized Snell’s law.
Remark 2.2. The functions V1 and V2 can represent not only the propa-
gation velocities of a wave, but also the maximum possible velocities of an
object or a person in two different media. In the former case, the problem is
related to Fermat’s principle (see e.g. [3, 4, 11]) and Huygens’ principle (see
e.g. [7, 13, 14, 19]), with real world applications in seismic theory [1, 20],
sound waves [12] or wildfire modeling [6, 10, 15, 18], among others. In the
latter, it becomes a general version of Zermelo’s navigation problem, which
seeks the fastest trajectory between two points for a moving object in a
medium which, in turn, may also be moving with respect to the observer
(see e.g. [5, 17]). Thus, real world applications of this generalized Snell’s law
are multiple and widely varied. Some examples are naturally encountered
when studying for example:
6 S. MARKVORSEN AND E. PENDÁS-RECONDO

• The trajectories of light rays between two anisotropic media.


• The firefront of a wildfire that spreads over two different crop or fuel
fields.
• The propagation of seismic waves when passing through layers of
different anisotropic inhomogeneous materials.
• The fastest trajectory between a point on the ground and a point in
the sea for a person who first has to run and then swim, taking the
water current into account.
• The fastest trajectory for a travel first by boat and then by zeppeling,
taking into account both the water current and the wind.

3. Time-minimizing trajectories
3.1. Variational definitions and conventions. In order to find time-
minimizing trajectories, one looks for the critical points of TN . Let γ ∈ N
be our candidate curve, with γ(0) = q1 ∈ Q1 , γ(t0 ) = q2 ∈ Q2 and γ(τ ) ∈ η.
Recall from Definition 2.1 that the choice of t0 and τ is arbitrary, so we can
further assume, without loss of generality, that γ is time-parametrized, i.e.,
F1 (γ(t), γ 0 (t)) = 1, ∀t ∈ [0, τ ),
0 (5)
F2 (γ(t), γ (t)) = 1, ∀t ∈ (τ, t0 ],
(so F1 (γ(τ ), γ 0 (τ − )) = 1 and F2 (γ(τ ), γ 0 (τ + )) = 1). The curve γ will be
the base curve in a family of curves defined by a variation, in the following
sense.
Definition 3.1. An admissible variation H of γ is a continuous map
H : (−ε, ε) × [0, t0 ] −→ Q
(ω, t) 7−→ H(ω, t) := γω (t)
such that:
• H(0, t) = γ0 (t) = γ(t) for all t ∈ [0, t0 ].
• H(ω, ·) = γω ∈ N for all ω ∈ (−ε, ε). In particular, γω (0) = q1 ,
γω (t0 ) = q2 and γω (τ ) ∈ η.
• For each t ∈ [0, t0 ], there exists a positive εt ≤ ε such that H(·, t) is
smooth on (−εt , εt ).
The last condition ensures that the tangent vector of the curve ω 7→ γω (t)
at ω = 0 is always well-defined. Let us denote by W (t) ∈ Tγ(t) Q this tangent
vector. Then the curve in T Q given by
[0, t0 ] −→ TQ
t 7−→ (γ(t), W (t))
is called the variational field of γ associated with H.
Finally, we say that γ is a critical point of TN if, for every admissible
variation H, we have that

d
TN [γω ] = 0.
dω ω=0
Remark 3.2. Note that the only time-parametrized curve in the variation
is the base curve γ = γ0 . In particular, the traveltime for γ is t0 and τ is the
time at which γ meets η. However, the parametrization of any other curve
SNELL’S LAW REVISITED 7

γω is not arbitrary: it must satisfy the conditions γω (0) = q1 , γω (t0 ) = q2


and γω (τ ) ∈ η. So, we cannot assume that γω is time-parametrized for
ω 6= 0, which means that the traveltime must be computed through (4) and
τ does not have to coincide with the actual time at which γω intersects η
for ω 6= 0.

Notation 3.3. From now on we will work in coordinates with the following
usual conventions. Let q = (q 1 , . . . , q n ) be a coordinate system on Q and
consider also the coordinate system (q, q̇) on T Q, being q̇ = (q̇ 1 , . . . , q̇ n )
the natural coordinates on the fiber induced by q. This means that every
(p, v) ∈ T Q will be written is these coordinates as (p1 , . . . , pn , v 1 , . . . , v n ) ∈
R2n , where pi = q i (p) and v i = q̇ i (v) = v(q i ), i.e., v = v i ∂q∂ i |p (using
Einstein’s summation convention). Then:
• Any function on T Q such as the Finsler metrics Fj can be written
in coordinates as Fj ≡ Fj (q, q̇) and it makes sense to compute the
following derivatives for j = 1, 2:
 
∂Fj ∂Fj ∂Fj
(p, v) := (p, v), . . . , n (p, v) ∈ Rn ,
∂q ∂q 1 ∂q
 
∂Fj ∂Fj ∂Fj
(p, v) := (p, v), . . . , n (p, v) ∈ Rn ,
∂ q̇ ∂ q̇ 1 ∂ q̇

• A curve γ : [0, t0 ] → Q will be written in these coordinates as γ(t) ≡


q(t) = (q 1 (t), . . . , q n (t)), where q i (t) := q i (γ(t)), and its tangent
vector as γ 0 (t) ≡ q̇(t) = (q̇ 1 (t), . . . , q̇ n (t)), where

dq i
q̇ i (t) := q̇ i (γ 0 (t)) = (t), i = 1, . . . , n.
dt
This way, we can also write for j = 1, 2:

Fj (t) := Fj (q(t), q̇(t)),


∂Fj ∂Fj
(t) := (q(t), q̇(t)),
∂q ∂q
∂Fj ∂Fj
(t) := (q(t), q̇(t)),
∂ q̇ ∂ q̇

• If we consider an admissible variation H of γ, then we write in


coordinates H(ω, t) ≡ qω (t), being q̇ω (t) the tangent vector of the
curve t 7→ qω (t). Also, the variational field is t 7→ (q(t), W (t)), with
W (t) = W i (t) ∂q∂ i |q(t) ≡ (W 1 (t), . . . , W n (t)) and

∂qωi (t)

i
W (t) = , i = 1, . . . , n.
∂ω ω=0

Finally, observe that even if q is not a global coordinate system on Q, the


compactness of [0, t0 ] allows one to find a finite number of coordinate charts
{(Ui , qi )}k+l
i=1 and a partition of the interval 0 = s1 < . . . < sk = τ < . . . <
8 S. MARKVORSEN AND E. PENDÁS-RECONDO

sk+l = t0 such that γ([si , si+1 ]) ⊂ Ui . So, we can compute


Z τ Z t0
T [γ] = F1 (γ(t), γ 0 (t))dt + F2 (γ(t), γ 0 (t))dt =
0 τ
k−1 Z
X si+1 k+l−1
X Z si+1
= F1 (qi (t), q̇i (t))dt + F2 (qi (t), q̇i (t))dt
i=1 si i=k si

and therefore, without loss of generality, we can assume that q is a global


coordinate system on Q (and so is (q, q̇) on T Q).
3.2. Critical points of the traveltime functional. Let us now derive
the conditions for the time-parametrized γ ∈ N to be a critical point of TN .
If such a curve exists, then

d
0= TN [γω ] =
dω ω=0
Z τ Z t0 (6)
d d
= F1 (qω (t), q̇ω (t))dt + F2 (qω (t), q̇ω (t))dt
dω ω=0 0 dω ω=0 τ
for every admissible variation of γ. The first integral is
Z τ

I1 = F1 (qω (t), q̇ω (t)) dt =
0 ∂ω ω=0
Z τ 
∂F1 i ∂F1 d i
= (t)W (t) + (t) W (t) dt =
0 ∂q i ∂ q̇ i dt
Z τ    
∂F1 d ∂F1 i d ∂F1 i
= (t) − (t) W (t) + (t)W (t) dt =
0 ∂q i dt ∂ q̇ i dt ∂ q̇ i
Z τ 
∂F1 − i ∂F1 d ∂F1
= i (τ )W (τ ) + (t) − (t) W i (t)dt.
∂ q̇ 0 ∂q i dt ∂ q̇ i
Analogously, the second integral is
Z t0  
∂F2 + i ∂F2 d ∂F2
I2 = − i (τ )W (τ ) + i
(t) − i
(t) W i (t)dt.
∂ q̇ τ ∂q dt ∂ q̇
Note that in both cases we have used that W (0) = W (t0 ) = 0, since the
variation has fixed endpoints. Then (6) reduces to
  Z τ 
∂F1 − ∂F2 + i ∂F1 d ∂F1
0= (τ ) − (τ ) W (τ ) + (t) − (t) W i (t)dt
∂ q̇ i ∂ q̇ i 0 ∂q i dt ∂ q̇ i
Z t0  
∂F2 d ∂F2
+ (t) − (t) W i (t)dt
τ ∂q i dt ∂ q̇ i
and this must hold for every admissible variation H(ω, t) ≡ qω (t), which
implies that γ must satisfy:
∂F1 d ∂F1
(t) − (t) = 0, ∀t ∈ [0, τ ),
∂q i dt ∂ q̇ i
(7)
∂F2 d ∂F2
(t) − (t) = 0, ∀t ∈ (τ, t0 ],
∂q i dt ∂ q̇ i
and the following additional condition at the break point γ(τ ):
 
∂F1 − ∂F2 +
(τ ) − (τ ) W i (τ ) = 0, ∀H. (8)
∂ q̇ i ∂ q̇ i
SNELL’S LAW REVISITED 9

On the one hand, the two equations in (7) are the well-known Euler-
Lagrange equations, equivalent to the (pre-)geodesic equations for the met-
rics F1 , F2 . Along with the time-parametrization conditions in (5), this tells
us that γ is an F1 -unit speed geodesic from γ(0) = q1 ∈ Q1 to γ(τ ) ∈ η
and an F2 -unit speed geodesic from there to γ(t0 ) = q2 ∈ Q2 . On the other
hand, (8) is the condition at the break point γ(τ ) that provides the desired
Snell’s law, as we will see next. Note that if γ satisfies these conditions, it is
specifically a local minimum of TN , as we can easily increase the traveltime.

3.3. Generalized Snell’s law. Let us now examine the equation (8). First,
note that for every admissible variation H(ω, t) ≡ qω (t), we have that qω (t) ∈
η for all ω ∈ (−ε, ε), where τ is fixed (recall Definition 3.1). Therefore:

∂qω (τ )
= W (τ ) ∈ Tq(τ ) η
∂ω ω=0
and this restriction must hold for all H, so (8) can be rewritten as
 
∂F1 − ∂F2 +
(τ ) − (τ ) ui = 0, ∀u ∈ Tq(τ ) η, (9)
∂ q̇ i ∂ q̇ i
where ui = q̇ i (u). This is the condition at the break point that governs
the change in the trajectory’s direction, and thus we arrive at the general
version of Snell’s law:
Theorem 3.4. Let γ : [0, t0 ] → Q be a time-parametrized curve in N . Then
γ is a critical point (specifically, a minimum) of TN if and only if it is an
F1 -unit speed geodesic from γ(0) ∈ Q1 to γ(τ ) ∈ η, an F2 -unit speed geodesic
from γ(τ ) ∈ η to γ(t0 ) ∈ Q2 , and it satisfies the generalized Snell’s law given
by (9) at the break point γ(τ ) ∈ η. In this case we say that γ is a refracted
trajectory.
Remark 3.5. The formulation we have carried out here is typical in classical
mechanics, where instead of a Finsler metric F one has a Lagrangian defined
as L = K − P , being K and P the total kinetic and potential energy of a
system of particles,
R respectively, and instead of minimizing the traveltime
functional
R T = F one looks for the critical points of the action functional
S = L. In this setting, (9) states that the total linear momentum of the
system must be conserved in the tangent directions to η (see e.g. [8]).
Snell’s law is essentially an equation system involving q(τ ), q(τ − ) and
q(τ + ). Typically, q(τ ), q(τ − ) are known from the initial conditions and
the geodesic equations in the first region Q1 , so one looks for q(τ + ). Since
dim(η) = dim(Tq(τ ) η) = n − 1, (9) provides exactly n − 1 independent
equations. With the additional condition F2 (q(τ ), q(τ + )) = 1, we have n
equations to find the n-component vector q(τ + ). Alternatively, one can just
find the direction of q(τ + ) from (9), which is usually more convenient in
dimension 2, as we will see next.

3.4. The two-dimensional case. If dim(Q) = n = 2, then the direction of


the incident and refracted trajectory can be written in terms of the following
angles.
10 S. MARKVORSEN AND E. PENDÁS-RECONDO

Definition 3.6. Let n = 2 and γ ∈ N . We define the angles of incidence


and refraction θ1 , θ2 ∈ (−π, π] as the (oriented) directions of q̇(τ − ), q̇(τ + ),
respectively, measured with respect to the q̇ 1 -axis. Namely,
q̇ 2 (τ − ) q̇ 2 (τ + )
tan θ1 = , tan θ2 = ,
q̇ 1 (τ − ) q̇ 1 (τ + )
with the restriction that vθ1 := (cos θ1 , sin θ1 ) and vθ2 := (cos θ2 , sin θ2 ) must
point to the Q2 -side of Tq(τ ) η.
This way, for a fixed break point q(τ ) ∈ η, Snell’s law is transformed
into an equation between the angles of incidence and refraction. Specifi-
cally, it can always be written explicitly in terms of tan θ1 and tan θ2 . In-
deed, since F1 and F2 are positively homogeneous of degree one, Fj (λv) =
|λ|Fj (sgn(λ)v) for all v ∈ Tq(τ ) η, λ ∈ R and, as a consequence,
∂Fj ∂Fj
(λv) = (sgn(λ)v), j = 1, 2,
∂ q̇ ∂ q̇
so, in particular,
q̇ 2
  
∂Fj 1 2 ∂Fj 1 ∂Fj
(q̇ , q̇ ) = i sgn(q̇ ) 1, 1 = (sgn(q̇ 1 )(1, tan θj )),
∂ q̇ i ∂ q̇ q̇ ∂ q̇ i
for i, j = 1, 2, where the references to q(τ ) and τ ± have been omitted.
Therefore, in general one can solve (9) for tan θ2 and then obtain θ2 ∈
(−π, π] by imposing the restriction that vθ2 must point to the Q2 -side of
Tq(τ ) η.
Example 3.7. Let us recover the classical Snell’s law in the Euclidean
plane. In this usual setting, Q = R2 , (q, q̇) = (x, y, ẋ, ẏ) and the background
Riemannian metric g = dx2 + dy 2 is the natural one in R2 . Let Q1 =
{(x, y) ∈ R2 : x < 0} and Q2 = {(x, y) ∈ R2 : x > 0}, so η is the vertical
line x = 0. In this situation, assume that the propagation of light is isotropic
and its velocity depends on the refractive index of the medium. Assume also
that Q1 and Q2 are homogeneous, i.e., their respective refractive indices
n1 , n2 are constant. Then:
c n1 p 2
V1 (x, y, ẋ, ẏ) = ⇒ F1 (x, y, ẋ, ẏ) = ẋ + ẏ 2 ,
n1 c
c n2 p 2
V2 (x, y, ẋ, ẏ) = ⇒ F2 (x, y, ẋ, ẏ) = ẋ + ẏ 2 ,
n2 c
where c is the speed of light in vacuum. The angles of incidence and refrac-
tion are given by
ẏ(τ − ) ẏ(τ + )
tan θ1 = , tan θ2 = ,
ẋ(τ − ) ẋ(τ + )
so, omitting the reference to τ ± we get:
∂Fj nj ẏ nj ẏ nj tan θj
= p = p =± p =
∂ ẏ c ẋ2 + ẏ 2 c |ẋ| 1 + ẏ 2 /ẋ2 c 1 + tan2 θj
nj nj
=± tan θj | cos θj | = sin θj , j = 1, 2,
c c
SNELL’S LAW REVISITED 11

where the ± vanishes because sgn(ẋ(τ − )) = sgn(cos θ1 ) and sgn(ẋ(τ + )) =


sgn(cos θ2 ). This way, since in this particular example {(0, 1)} is a basis of
T(0,y(τ )) η at any break point (0, y(τ )) ∈ η, (9) yields the classical Snell’s law:
n1 sin θ1 = n2 sin θ2 , (10)
with the restriction that sgn(cos θ1 ) = sgn(cos θ2 ) > 0, so that both vθ1 and
vθ2 point to Q2 .

4. Reflected trajectories
4.1. The law of reflection. So far, given a curve γ departing from Q1 and
intersecting η, we have been looking for the refracted trajectory, i.e., the
one that crosses η and continues into Q2 . It is interesting also to seek the
reflected trajectory, which will be the one that minimizes the traveltime in
a slightly different set of curves than N in Definition 2.1.
Specifically, fixing [0, t0 ], τ ∈ (0, t0 ) and two points q1 , q2 ∈ Q1 , we define

N as the set of all piecewise smooth curves γ : [0, t0 ] → Q from γ(0) = q1
to γ(t0 ) = q2 such that γ remains entirely in Q1 , touching η once at γ(τ ).
If one now looks for the critical points of TN ∗ : N ∗ → R in a similar way
as in the previous section, one finds, on the one hand, the Euler-Lagrange
equations (compare with (7)):
∂F1 d ∂F1
i
(t) − (t) = 0, ∀t ∈ [0, τ ) ∪ (τ, t0 ],
∂q dt ∂q i
which now state that the candidate curve γ(t) ≡ q(t) must be an F1 -geodesic
from q1 to η and also from η to q2 , and on the other hand, instead of Snell’s
law (9) (i.e., the refraction law), we get the following reflection law:
 
∂F1 − ∂F1 +
(τ ) − (τ ) ui = 0, ∀u ∈ Tq(τ ) η. (11)
∂ q̇ i ∂ q̇ i
Therefore, analogously to Theorem 3.4, we have the following result.
Theorem 4.1. Let γ : [0, t0 ] → Q be a time-parametrized curve in N ∗ .
Then γ is a critical point (specifically, a minimum) of TN ∗ if and only if it
is an F1 -unit speed geodesic from γ(0) ∈ Q1 to γ(τ ) ∈ η and from γ(τ ) ∈ η
to γ(t0 ) ∈ Q1 , and it satisfies the law of reflection given by (11) at the break
point γ(τ ) ∈ η. In this case we say that γ is a reflected trajectory.
Moreover, the discussion developed in § 3.4 becomes valid here as well,
i.e., in the two-dimensional case the law of reflection can be written in terms
of the following angles.
Definition 4.2. Let n = 2 and γ ∈ N ∗ . We define the angles of incidence
and reflection θ1 , θ3 ∈ (−π, π] as the (oriented) directions of q̇(τ − ), q̇(τ + ),
respectively, measured with respect to the q̇ 1 -axis. Namely,
q̇ 2 (τ − ) q̇ 2 (τ + )
tan θ1 = , tan θ3 = ,
q̇ 1 (τ − ) q̇ 1 (τ + )
with the restriction that vθ1 := (cos θ1 , sin θ1 ) must point to the Q2 -side and
vθ3 := (cos θ3 , sin θ3 ) to the Q1 -side of Tq(τ ) η.
12 S. MARKVORSEN AND E. PENDÁS-RECONDO

Figure 1. In the two-dimensional case, Snell’s law or the law


of reflection can be written in terms of the angle of incidence
θ1 and the angle of refraction θ2 or reflection θ3 , respectively.
When looking for the refracted trajectory γ ∈ N (resp. the
reflected trajectory γ ∈ N ∗ ), q̇(τ + ) must point to the Q2 -side
(resp. the Q1 -side) of Tq(τ ) η.

Note the difference between the angles of refraction and reflection (com-
pare with Definition 4.2): now q̇(τ + ) must point to the Q1 -side of Tq(τ ) η
because otherwise, γ would cross to Q2 and thus, γ ∈/ N ∗ (see Figure 1).
Example 4.3. In the same setting as in Example 3.7, let us recover the
classical reflection law. The angles of incidence and reflection are
ẏ(τ − ) ẏ(τ + )
tan θ1 = , tan θ3 = ,
ẋ(τ − ) ẋ(τ + )
with sgn(ẋ(τ − )) = sgn(cos θ1 ) and sgn(ẋ(τ + )) = sgn(cos θ2 ), so
∂F1 − n1 ∂F1 + n1
(τ ) = sin θ1 , (τ ) = sin θ3 .
∂ ẏ c ∂ ẏ c
Thus, (11) directly yields
sin θ1 = sin θ3 , (12)
with the restriction that sgn(cos θ1 ) > 0 and sgn(cos θ3 ) < 0, so that vθ1
points to Q2 and vθ3 to Q1 . For example, if θ1 ∈ [0, π2 ), then θ3 = π − θ1 . If
θ1 ∈ (− π2 , 0), then θ3 = −π − θ1 .2

2Another usual way to define the angles of incidence, refraction and reflection is to
restrict their values to [0, π2 ) and measure them with respect to the normal direction to
T(x(τ ),y(τ )) η. In this case, the classical Snell’s law would still be (10) but without any
additional restriction, and the classical reflection law would read θ1 = θ3 .
SNELL’S LAW REVISITED 13

4.2. The total reflection phenomenon. In general, the existence of solu-


tions to Snell’s law (9) is not guaranteed. Namely, given an incident velocity
q̇(τ − ), there might not be a solution q̇(τ + ) pointing to the Q2 -side of Tq(τ ) η.
In this case, there is no refraction and the incoming trajectory is totally
reflected without being able to cross into Q2 . This is known as the total
reflection phenomenon. The limit between the existence and non-existence
of solution is given by the situation where q̇(τ + ) is tangent to Tq(τ ) η, which
motivates the following definition.
Definition 4.4. We say that an incident velocity q̇(τ − ) (or an angle of
incidence θ1 in the two-dimensional case) is critical if there exists a solution
q̇(τ + ) to Snell’s law (9) which is tangent to Tq(τ ) η.
The trajectory γ associated with the critical velocity can still be consid-
ered as an actual refracted trajectory as long as immediately after touching
η, it passes to Q2 or remains on η with the F2 metric, i.e. if there exists
δ > 0 such that γ(τ + t) ∈ (Q2 , F2 ) for all t ∈ (0, δ). We say in this case that
γ is a critical trajectory. In fact, critical trajectories along η are relevant in
situations such as the one we will describe in the next section.
Example 4.5. In the classical setting of Example 3.7, vθ2 is tangent to η
when θ2 = ± π2 , so the critical angles are given by
 
n2
θc± = ± arcsin .
n1
Note that these angles only exist if n1 > n2 , which means that the velocity
V2 on Q2 is greater than V1 on Q1 . If, for example, n1 = 1 and n2 = 21 , then
θc± = ± π6 . In the cases θ1 > π6 or θ1 < π6 , there does not exist an angle θ2
satisfying the classical Snell’s law (10) and pointing to Q2 . In consequence,
there is no time-minimizing path from any q1 ∈ Q1 to any q2 ∈ Q2 which
meets η with θ1 > π6 or θ1 < π6 , so there is no refraction and the incident
trajectory is totally reflected with an angle θ3 satisfying the classical law of
reflection (12). In other terms, suppose q1 has Euclidean distance d to η,
then the Euclidean size of the “window” through which a time-minimizing
trajectory can go through η is √2d·n 2
2
2
.
n1 −n2

5. Global considerations in R2
In order to reduce the casuistry we can find in the most general case
and fix some ideas, we will reduce our study in this section to the classical
setting (see Example 3.7), although maintaining the Finslerian nature of the
metrics. Namely, Q = R2 , (q, q̇) = (x, y, ẋ, ẏ), g = dx2 + dy 2 , Q1 = {(x, y) ∈
R2 : x < 0}, Q2 = {(x, y) ∈ R2 : x > 0} and F1 , F2 are Finsler metrics
independent of the position (i.e., Minkowski norms):
p
ẋ2 + ẏ 2
Fj (x, y, ẋ, ẏ) = Fj (ẋ, ẏ) = , j = 1, 2.
Vj (ẋ, ẏ)
In particular, this means that the geodesics of F1 , F2 are straight lines. Note
also that the angles of incidence, refraction and reflection are restricted to
the values θ1 ∈ (− π2 , π2 ), θ2 ∈ [− π2 , π2 ], θ3 ∈ (−π, − π2 ) ∪ ( π2 , π], and there are
at most two critical angles θc± , with θc+ > θc− : θc+ corresponds to θ2 = π2
14 S. MARKVORSEN AND E. PENDÁS-RECONDO

and θc− to θ2 = − π2 (the uniqueness of θc± when they exist is proven below
in Lemma 5.1).

5.1. Existence and uniqueness of solution. In this setting, Snell’s law


(9) reduces to
∂F1 − ∂F2 +
(τ ) = (τ ).
∂ ẏ ∂ ẏ
∂Fj
Now, being F1 and F2 Minkowski norms, we have that ∂y = 0 and thus,
from the Euler-Lagrange equations (7):
d ∂F1
(t) = 0, ∀t ∈ [0, τ ),
dt ∂ ẏ
d ∂F2
(t) = 0, ∀t ∈ (τ, t0 ].
dt ∂ ẏ
which means, together with Snell’s law, that there is a conserved quantity
along the entire refracted curve γ ∈ N .3 This constant is called the raypath
parameter (see [3]). It only depends on the initial direction, due to the
∂F
zero-homogeneity of ∂ ẏj . So, Snell’s law becomes
P1 (θ1 ) = P2 (θ2 ), (13)
with
 
∂Fj sin θ ∂ 1
Pj (θ) := (θ) = + cos θ , j = 1, 2, (14)
∂ ẏ Vj (θ) ∂θ Vj (θ)
2
where we have used that ∂∂ẏ = cosẋ θ ∂θ

. Note that Pj (θ) is smooth, thanks
to the smoothness and positiveness of Vj (θ).
Lemma 5.1. The functions Pj : [− π2 , π2 ] → R are strictly increasing.
Also, the critical angles θc± exist (and are necessarily unique) if and only
if V1 (± π2 ) < V2 (± π2 ), respectively.
Proof. One can easily verify that
" #
dVj 2 d2 Vj
 
dPj cos θ 2
= Vj + 2 − Vj 2 ,
dθ Vj3 dθ dθ
where the term in brackets is exactly the second fundamental form of the
indicatrix Σj of Fj (given by the curve (π, π] 3 θ 7→ Vj (cos θ, sin θ)) with
respect to the natural Euclidean metric in R2 and the inner normal vector.
dP
The strong convexity of Σj and the positiveness of Vj ensure that dθj > 0
for all θ ∈ (− π2 , π2 ), so we conclude that Pj (θ) is strictly increasing in θ ∈
[− π2 , π2 ]. In particular, this means that the maximum of Pj (θ) is Pj ( π2 ) =
1 π 1 ± π π
Vj ( π ) and the minimum Pj (− 2 ) = − Vj (− π ) . So θc ∈ (− 2 , 2 ) exists if and
2 2
only if P1 (θc± ) = P2 (± π2 ), i.e. if and only if 1
V1 (± π2 ) > 1
V2 (± π2 ) (and in this
case θc± is unique due to the fact that P1 (θ) is strictly increasing). 
3This is related to Noether’s theorem: the translational symmetry of the setting along
the y-axis implies the existence of a constant of motion.
SNELL’S LAW REVISITED 15

Theorem 5.2. Given any angle of incidence θ1 ∈ (θc− , θc+ ) (replace θc±
with ± π2 if the critical angles do not exist), there exists a unique angle of
refraction θ2 ∈ (− π2 , π2 ) satisfying Snell’s law. Equivalently, given any q1 ∈
Q1 and q2 ∈ Q2 , there exists a unique refracted trajectory γ ∈ N joining
both points.

Proof. Given θ1 ∈ (θc− , θc+ ) (or θ1 ∈ (− π2 , π2 ) if the critical angles do not


exist), from Lemma 5.1 we know that P2 (− π2 ) < P1 (θ1 ) < P2 ( π2 ), so there
exists θ2 ∈ (− π2 , π2 ) satisfying (13) and since P2 (θ) is strictly increasing, this
angle is necessarily unique. Note that the converse is also true, i.e., given
any θ2 ∈ (− π2 , π2 ) we can find θ1 ∈ (θc− , θc+ ) (or θ1 ∈ (− π2 , π2 )) satisfying
Snell’s law. In R2 with straight lines as geodesics, this ensures that all the
refracted trajectories cover Q2 entirely, independently of the initial point
q1 ∈ Q1 . 

Of course, one can follow the same steps and derive an analogous result
for the existence and uniqueness of solutions to the law of reflection, which
reads P1 (θ1 ) = P1 (θ3 ) in terms of the raypath parameter.

Theorem 5.3. Given any angle of incidence θ1 ∈ (− π2 , π2 ), there exists


a unique angle of reflection θ3 ∈ (−π, − π2 ) ∪ ( π2 , π] satisfying the law of
reflection. Equivalently, given any q1 , q2 ∈ Q1 , there exists a unique refracted
trajectory γ ∈ N ∗ joining both points.

5.2. New reflected trajectories. The existence and uniqueness of re-


fracted and reflected trajectories ensure that, once the initial and final points
are fixed, there is only one critical point of TN and TN ∗ , which therefore
must be a global minimum. However, recall that curves in N and N ∗ have
the restriction that they must intersect η only once. One might wonder if
more contact points with η could result in a curve that spends even less time
than the usual refracted or reflected trajectory. Namely, we can define new
sets N̂ and N̂ ∗ in the same way as N and N ∗ , respectively, but with the re-
striction that the (piecewise smooth) curves must meet η at least once (but
possibly at more points), and look for the critical points of the traveltime
functional on these sets.
Let γ be a curve from q1 to q2 that meets η for the first time at τ1 and
leaves it for the last time at τ2 . In order to minimize the traveltime, the curve
from γ(τ1 ) to γ(τ2 ) must be the segment of the straight line η joining these
points (since it is a geodesic in the present R2 setting under consideration).
Now, if γ ∈ N̂ , by the triangle inequality (see e.g. [2, Thm. 1.2.2]) any
two-segment trajectory from q1 ∈ Q1 to γ(τ ) ∈ η (τ ∈ [τ1 , τ2 ]) and from
there to q2 ∈ Q2 spends less time than γ. So the critical points of TN̂
must always belong to N , i.e., they are the usual two-segment refracted
trajectories crossing η once. However, if γ ∈ N ∗ , the travel along η with
the metric F2 can be an advantage. A real world example is when moving
between two points in the sea next to the shore: it can be faster to swim first
to the shore, then run along the beach line edge and finally swim again to the
final point. Obviously, the traveltime on η has to be computed through F2
(otherwise, the usual reflected trajectory and hence also the direct straight
16 S. MARKVORSEN AND E. PENDÁS-RECONDO

line trajectory in Q1 is always faster):


Z τ1 Z τ2 Z t0
0 0
T [γ] = F1 (γ (t))dt + F2 (γ (t))dt + F1 (γ 0 (t))dt.
0 τ1 τ2
Taking this into account, we can now define a suitable variation and apply
the calculus of variations in the same way as in § 3.2. This gives the following
conditions for a time-parametrized curve γ to be a critical point of TN̂ ∗ :
• γ is composed of three straight lines: an F1 -unit segment γ1 from
q1 ∈ Q1 to γ(τ1 ) ∈ η, an F2 -unit segment γ2 ⊂ η between γ(τ1 ) and
γ(τ2 ), and a final F1 -unit segment γ3 from γ(τ2 ) ∈ η to q2 ∈ Q1 (see
Figure 2).
• At the break points, γ(t) ≡ (x(t), y(t)) (for t = τ1 and t = τ2 ,
respectively) must satisfy:
∂F1 − ∂F2 +
(τ ) = (τ ),
∂ ẏ 1 ∂ ẏ 1
∂F2 − ∂F1 +
(τ ) = (τ ).
∂ ẏ 2 ∂ ẏ 2
Equivalently, in terms of the (constant) directions θ1 , θ2 , θ3 ∈ (−π, π]
of each segment and the raypath parameter (14):
P1 (θ1 ) = P2 (θ2 ) = P1 (θ3 ). (15)
In particular, this means that θ1 must be critical (so that θ2 = ± π2 )
and θ3 must be exactly the angle of reflection we would obtain at τ1 ,
being γ1 the incident trajectory (see Figure 2).
If it exists, this three-segment trajectory is always faster than the usual
reflection. However, its existence is not guaranteed: it depends on whether
the critical angle exists (which in turn depends on whether the velocity
V2 on η is greater than V1 on Q1 ; recall Lemma 5.1) and, if it does, on
the initial and final points. So, fixing two points q1 , q2 ∈ Q1 , the global
minimum of traveltime between q1 and q2 among curves in TN̂ ∗ is the three-
segment trajectory satisfying the above conditions, if it exists, or the usual
reflected trajectory, if it does not. This global comparison should, of course,
be benchmarked against the traveltime for the direct straight line between
q1 and q2 . This will be done in § 5.3 below.
Remark 5.4. At first glance, these new reflected trajectories might only
seem relevant in situations when you can willingly change the directions
midway at points of η. Obviously, a light ray refracted at the critical angle
will continue along η indefinitely, without returning to Q1 . When computing
the whole propagation of a wave though, it is crucial to take into account all
the possible trajectories. Indeed, Huygens’ principle states that each point
of the wavefront can be regarded as an independent source that shoots wave
trajectories in all directions. Among all of them, those that globally mini-
mize the traveltime are the ones that keep providing points in the wavefront
at later instants of times. So, a curve such as γ in Figure 2 can indeed
represent a realistic wave trajectory that contributes to the formation of the
wavefront (see § 5.5).
SNELL’S LAW REVISITED 17

Figure 2. There are situations where moving along γ is


faster than following the usual reflected trajectory ϕ or even
the straight line between two fixed points q1 , q2 ∈ Q1 . This
can only happen if γ2 is traveled with velocity V2 ( π2 ) along
η and if θ1 and θ3 are, respectively, the critical and reflec-
tion angles of γ1 . In particular, γ3 must be parallel to the
reflection ρ of γ1 .

5.3. Globally time-minimizing trajectories. Taking a step further, we


can now remove the restriction that the curves must touch η.

Definition 5.5. Fixing any two points q1 , q2 ∈ Q, let M be the set of all
the piecewise smooth curves from q1 to q2 . We say that γ ∈ M globally
minimizes the traveltime if T [γ] ≤ T [ϕ] for all ϕ ∈ M.

If q1 ∈ Q1 , q2 ∈ Q2 , then it is clear that M = N̂ , so the refracted tra-


jectories always globally minimize the traveltime. This includes the critical
trajectories along η, as stated in the following result.

Proposition 5.6. If γ : [0, t0 ] → Q is a critical trajectory (associated with


one of the critical angles θc± ), then γ globally minimizes the traveltime be-
tween γ(0) ∈ Q1 and γ(t0 ) ∈ η.

Proof. Given γ(t0 ) = (0, y0 ) ∈ η, suppose there exists another curve ξ1 ∈ M


that arrives earlier at this point from γ(0) ∈ Q1 . Consider now (δ, y0 ) ∈ Q2
with δ > 0 and let ϕδ be the unique refracted trajectory from γ(0) to (δ, y0 ),
which globally minimizes the traveltime. The smoothness of the raypath
parameter given by (14) ensures that

lim T [ϕδ ] = T [γ] > T [ξ1 ].


δ→0
18 S. MARKVORSEN AND E. PENDÁS-RECONDO

So, for a sufficiently small δ > 0, the straight line ξ2 from (0, y0 ) to (δ, y0 )
satisfies that T [ϕδ ] > T [ξ1 ] + T [ξ2 ], which means that we have built a path
in M faster than ϕδ , arriving at a contradiction. 
If q1 , q2 ∈ Q1 though, the global minimum of TN̂ ∗ competes against the
straight line in Q1 between q1 and q2 . It is clear that the straight line is
always faster than the usual reflection, due to the triangle inequality for
the metric F1 in Q1 , but in some situations the three-segment reflected
trajectory can globally minimize the traveltime between q1 and q2 , thus
creating a point on the cut locus from q1 – we elaborate further on this issue
in § 5.6 below. We first illustrate the setting for these possibilities next with
an example.
5.4. Anisotropic example. Let us consider the following velocity func-
tions in R2 :
aj (1 − ε2j )
Vj (θ) = , j = 1, 2,
1 − εj cos(θ − φj )
whose indicatrices are ellipses of eccentricity εj , respectively, centered at
one of their foci, with semi-major axis aj oriented in the direction φj (with
respect to the ẋ-axis) and such that φj = 0 points to their other respective
focus point. Pecisely, Vj (θ) is the (Euclidean) length of the vector from the
centered focus to the ellipse in the direction θ, and this ellipse becomes the
indicatrix of the Finsler metric4
p
ẋ2 + ẏ 2
    

Fj (ẋ, ẏ) = 1 − ε j cos arctan − φ j , j = 1, 2.
aj (1 − ε2j ) ẋ
One can easily verify that
∂Fj sin θ − εj sin φj
(θ) = , j = 1, 2,
∂ ẏ aj (1 − ε2j )
so (15) becomes
∂F1 ∂F2 sin θk − ε1 sin φ1 sin θ2 − ε2 sin φ2
(θk ) = (θ2 ) ⇒ 2 = ,
∂ ẏ ∂ ẏ a1 (1 − ε1 ) a2 (1 − ε22 )
for k = 1, 3. In particular, in order to find the three-segment reflected
trajectory, θ1 must be the (critical) angle such that θ2 = ± π2 , i.e.,
a1 (1 − ε21 )
 
θ1 = arcsin (±1 − ε2 sin φ2 ) + ε1 sin φ1 , (16)
a2 (1 − ε22 )
and θ3 must satisfy the same equation but pointing to Q1 .5
Now, suppose we want to go from q1 = (−x0 , −y0 ) to q2 = (−x0 , y0 ), for
some x0 , y0 > 0, and let us set a1 = a2 = 1, ε1 = ε2 = 21 , φ1 = 0 and
φ2 = π2 . Then (16) becomes θ1 = arcsin ±1 − 12 , which does not have


solution for the − sign (corresponding to θ2 = − π2 ), but it does for the +


sign (corresponding to θ2 = π2 ). Namely, there is only one critical angle
θ1 = θc+ = arcsin 12 = π6 and thus, θ3 = π − θ1 = 5π
6 . We can now compute
4This type of Finsler metrics are used to model the spread of wildfires (see e.g. [9, 15]).
5Even though the velocities are anisotropic, the law of reflection in this example is the
same as in the case with Euclidean metrics (see Example 4.3). However, this is not true
in general.
SNELL’S LAW REVISITED 19

the intersection points with η to see if the three-segment reflection exists.


Since we can completely determine γ1 and γ3 (we know q1 , q2 and θ1 , θ3 ), it
is easy to show that
   
x0 x0
γ1 ∩ η = 0, −y0 + √ , γ3 ∩ η = 0, y0 − √ .
3 3
We conclude (see Figure 3):
x0
• If y0 > √ 3
, then the three-segment curve exists and it is the global
minimum of TN̂ ∗ .
x0
• If y0 = √ 3
, then γ2 reduces to the point (0, 0) and γ3 becomes the
reflection of γ1 corresponding to the critical angle.
x0
• If y0 < √ 3
, then you cannot reach q2 through a three-segment curve
with the given angles θ1 , θ2 , θ3 . In this case, the global minimum
of TN̂ ∗ is the usual reflected trajectory, with an angle of incidence
smaller than the critical angle.
Finally, we can find the trajectory that globally minimizes the traveltime.
This can be done simply by comparing the traveltimes of the straight line
x0
ξ and the three-segment reflection γ in the case y0 > √ 3
(in any other case
the straight line is always the fastest trajectory). Using the endpoints and
direction of each segment, they can be parametrized as follows:
ξ(t) = (−x0 , −y0 ) + t(0, 2), t ∈ [0, y0 ],

 
x0
γ1 (t) = (−x0 , −y0 ) + t( 3, 1), t ∈ 0, √ ,
3
   
x0 x0
γ2 (t) = 0, −y0 + √ + t(0, 2), t ∈ 0, y0 − √ ,
3 3

   
x0 x0
γ3 (t) = 0, y0 − √ + t(− 3, 1), t ∈ 0, √
3 3
(note that this is not the time-parametrization). Therefore:
Z y0
4
T [ξ] = F1 (0, 2)dt = y0 ,
0 3
x x x
Z √0 √ Z y0 − √0 Z √0 √
3 3 3
T [γ] = F1 ( 3, 1)dt + F2 (0, 2)dt + F1 (− 3, 1)dt =
0
0 0


4 x0
= √ (4 − 3) + 1 ,
3 3
and we conclude (see Figure 3):
x0

• If y0 = √ 3
(4 − 3) + 1, then γ and ξ reach q2 at the same time and
both globally minimize the traveltime. In this case q2 belongs to the
cut locus of q1 –√see § 5.6 below.
x0
• If y0 > √ 3
(4 − 3) + 1, only γ globally minimizes the traveltime.
x

• If y0 < √03 (4 − 3) + 1, only ξ globally minimizes the traveltime.
20 S. MARKVORSEN AND E. PENDÁS-RECONDO

Figure 3. In this example, the velocity at each point is given


by a focus-centered ellipse with orientation φ1 = 0 in Q1 and
φ2 = π2 in Q2 , which favors the travel along η. Fixing x0 = 1
and always considering the travel from q1 = (−1, −y0 ) to its
symmetrical q2 = (−1, y0 ), there are three different regions
for y0 : if 0 < y0 < √13 , the fastest trajectory via η is the usual
reflection passing through (0, 0), with an angle of incidence
smaller than the critical one; if √13 < y0 < √43 , then the three-
segment reflection becomes faster than the usual reflected
trajectory, but still in these first two cases ξ is the fastest
path; when y0 > √43 , the three-segment reflection becomes
even faster than ξ. The limit cases are depicted explicitly: ϕ
when y0 = √13 and γ when y0 = √43 .

5.5. Construction of the wavefront. Fixing a time t0 > 0, consider all


the time-parametrized trajectories γ that globally minimize the traveltime
from q1 ∈ Q1 to γ(t0 ) (and thus, to any γ(t) with t ∈ (0, t0 ]). They provide
the outermost points that can be reached by a wave spreading from q1 during
the time lapse t0 , i.e., each one of these trajectories provide a point in the
wavefront.
Throughout this section we have seen that only three types of curves
can globally minimize the traveltime in the classical Finslerian setting: the
straight lines in Q1 , the refracted trajectories (crossing η once or remaining
on it) and the three-segment reflected trajectories. Separately, they provide
the standard, refracted and reflected wavefronts, respectively, which compete
against each other to generate the actual wavefront.
Let τη and τ ± be the times when the standard wavefront and the critical
trajectories reach η, respectively (τη coincides with the F1 -distance from q1
SNELL’S LAW REVISITED 21

to η). The refracted wavefront only comes into existence from τη , and the
reflected ones from τ ± .
Lemma 5.7. Fixing any q1 ∈ Q1 and a time t0 > τ ± :
(i) The standard wavefront Γ̃t0 at t = t0 is given by the indicatrix of F1
scaled up by t0 , which is equivalent to
Γ̃t0 = {p ∈ R2 : F1 (p − q1 ) = t0 }, (17)
and Γ̃t0 intersects η at two points.
(ii) There exists one reflected wavefront Γ̂± t0 at t = t0 for each critical angle
θc± . If ϕ± is the usual reflection at θc± and γ ± is the corresponding crit-
ical trajectory along η (both time-parametrized), then Γ̂± t0 is the straight
± ±
line joining γ (t0 ) and ϕ (t0 ).
(iii) The intersection between Γ̃t0 and each Γ̂± t0 is a unique point, provided
±
that Γ̂t0 exists.
Proof. For the claim (i), if we consider F1 -unit speed straight lines ξ, note
that they travel a distance t0 kξ 0 (t)kg with kξ 0 (t)kg constant and F1 (ξ 0 (t)) =
1, so t0 Σ1 provides the distance traveled by ξ in all directions, being Σ1
the indicatrix of F1 . Centered at q1 , t0 Σ1 gives the points (17). Also, as
t0 > τ ± ≥ τη , the strong convexity of Σ1 ensures that Γ̃t0 intersects η at
exactly two points.

We will prove (ii) and (iii) for Γ̂+ t0 , being the proof for Γ̂t0 completely
analogous. For (ii), let ρ be any (time-parametrized) three-segment reflected
trajectory with angle of incidence θ1 = θc+ . Assume ρ reaches η at time
τ1 > 0 (the same as ϕ+ and γ + ) and leaves it at τ2 , with τ1 < τ2 < t0 .
Splitting the curves in their different segments, note that ρ1 = ϕ+ +
1 = γ1 , ρ2
is part of γ2+ and ρ3 is parallel to ϕ+ 2 . Thus:
kρ03 (t)kg (t0 − τ2 ) kρ03 (t)kg k(ϕ+ 0
2 ) (t)kg k(ϕ+ 0
2 ) (t)kg (t0 − τ1 )
= = = ,
k(γ2+ )0 (t)kg (t0 − τ2 ) k(γ2+ )0 (t)kg k(γ2+ )0 (t)kg k(γ2+ )0 (t)kg (t0 − τ1 )
which means, by Thales’ theorem, that ρ(t0 ) is in the segment Γ̂+ t0 joining
γ (t0 ) and ϕ (t0 ). This also ensures that given any point p ∈ Γ̂+
+ +
t0 , there
+
exists a three-segment reflected trajectory from q1 to p. So Γ̂t0 is the reflected
wavefront.
For the claim (iii), note first that Γ̃t0 ∩ Γ̂+ t0 has at most two points, given
+
the strong convexity of Γ̃t0 . Now, γ always globally minimizes the trav-
eltime by Proposition 5.6, but ϕ+ never does. In particular, this means
that
F1 (ϕ+ (t0 ) − q1 ) < t0 < F1 (γ + (t0 ) − q1 ),
so there must be at least one point p ∈ Γ̂+ t0 where F1 (p − q1 ) = t0 and thus,
+
p ∈ Γ̃t0 ∩ Γ̂t0 . If there was another intersection point, it would mean that
ϕ+ globally minimizes the traveltime, which is a contradiction. 
Note that Γ̃t0 and Γ̂± t0 can be explicitly parametrized as (smooth) curves
in R2 as follows:
Γ̃t0 (θ) = t0 V1 (θ)(cos θ, sin θ) + q1 , θ ∈ (−π, π],
(18)
Γ̂± ± ± ±
t0 (s) = γ (t0 ) + s(ϕ (t0 ) − γ (t0 )), s ∈ [0, 1],
22 S. MARKVORSEN AND E. PENDÁS-RECONDO

and Lemma 5.7 ensures there are exactly two angles θη± such that Γ̃(θη± ) ∈ η
and, if θc± exists, we can find s± ± ± ±
0 and θ0 such that Γ̃(θ0 ) = Γ̂(s0 ). Note
± ±
that θη± , s0 and θ0 depend on t0 .

Theorem 5.8. Assume for simplicity that θc+ exists but θc− does not (anal-
ogous results are obtained in any other case). Then the actual wavefront Γt0
from q1 ∈ Q1 at t = t0 is a piecewise smooth closed curve composed of the
following parts:
• If t0 ∈ [0, τη ], then Γt0 is the whole standard wavefront at t = t0 (see
Figure 4(a)).
• If t0 ∈ (τη , τ + ], then Γt0 = Γ1 ∪ Γ2 , where

Γ1 = {Γ̃t0 (θ) ∈ R2 : θ ∈ (−π, θη− ] ∪ [θη+ , π]}


and Γ2 is the whole refracted wavefront at t = t0 (see Figure 4(b)).
• If t0 > τ + , then Γt0 = Γ1 ∪ Γ2 ∪ Γ3 , where
Γ1 := {Γ̃t0 (θ) ∈ R2 : θ ∈ (−π, θη− ] ∪ [θ0+ , π]},
Γ3 := {Γ̂+ 2 +
t0 (s) ∈ R : s ∈ [0, s0 ]}

and Γ2 is the whole refracted wavefront at t = t0 (see Figures 4(c)


and 4(d)).
Proof. Note first that Γ1 and Γ3 are always smooth curves by (18) and so
is Γ2 by Theorem 5.2 and the smoothness of the raypath parameter (14).
Now, if t0 ∈ [0, τη ], only the standard wavefront exists. If t0 ∈ (τη , τ + ], the
reflected wavefront still does not exist, so the part of the standard wavefront
Γ1 in Q1 globally minimizes the traveltime, and so does the whole refracted
wavefront Γ2 (recall the discussion in § 5.3). Also, the endpoints of Γ1
and Γ2 coincide, which ensures that Γt0 is a piecewise smooth closed curve.
Finally, if t0 > τ + , it is clear from § 5.3 and Lemma 5.7 that each part
Γ1 , Γ2 , Γ3 globally minimizes the traveltime. In this case, the endpoints of
Γ1 are Γ2 (θη− ) and Γ3 (0), and Γ2 (θ0+ ) = Γ3 (s+
0 ). So Γt0 is still a piecewise
smooth closed curve. 

5.6. The cut locus. Fixing any q ∈ Q, let us define the following curves:
σ ± : (τ ± , ∞) −→ R2
t 7−→ σ (t) = Γ̃t ∩ Γ̂±
±
t ,

where Γ̃t , Γ̂±


t are the standard and reflected wavefronts from q at time t and
τ ± is the time at which the critical trajectory γ ± reaches η. Provided that
the critical angles exist, note that each σ ± is well defined, as τ ± is the time
from which the reflected wavefront begins to exist and Γ̃t ∩ Γ̂± t is just one
point by Lemma 5.7 (see Figures 4(c) and 4(d)).
By definition, each point of these curves is reached at the same time by
two different globally minimizing trajectories from q, constituting what is
known as the cut locus of q:
Cq := Im(σ + ) ∪ Im(σ − ).
SNELL’S LAW REVISITED 23

(a) t0 < τη (b) t0 = τ +

(c) t0 > τ + (d) t0 > τ +

Figure 4. Evolution of the wavefront over time. The cut


locus in (c) and (d) given by the intersection of Γ1 and Γ3 for
t ∈ (τ + , t0 ] develops into Q1 from the point of first encounter
of the expanding ellipse with η in the critical direction in (b).
The Finsler metrics are the same as in § 5.4.

Proposition 5.9. Fix q ∈ Q and let Mq be the set of all piecewise smooth
curves γ : [0, ∞) → Q (and thus, future inextendible) with γ(0) = q. If we
define the map tq : Mq → (0, ∞] by
tq (γ) := sup{t ∈ R : γ|[0,t] globally minimizes the traveltime},
then the cut locus of q is
Cq = {γ(tq (γ)) ∈ R2 : γ ∈ Mq , 0 < tq (γ) < ∞}. (19)

Proof. Let p = Γ̃t0 ∩ Γ̂± ±


t0 ∈ Cq for any t0 > τ . There exist two (time-
parametrized) curves ξ, γ ∈ Mq such that ξ(t0 ) = γ(t0 ) = p and ξ[0,t0 ] , γ[0,t0 ]
24 S. MARKVORSEN AND E. PENDÁS-RECONDO

globally minimize the traveltime. Suppose γ|[0,t0 +δ] is still globally time-
minimizing for some δ > 0. Then the concatenation of ξ[0,t0 ] with γ|[t0 ,t0 +δ]
also globally minimizes the traveltime. But if we concatenate ξ[0,t0 −δ] with
the straight line joining ξ(t0 − δ) and γ(t0 + δ) we reduce the traveltime by
the triangle inequality (note that γ and ξ cannot be parallel), which is a
contradiction. So t0 = tq (γ) and p is in the set (19).
Conversely, let p = γ(tq (γ)) for some γ ∈ Mq with 0 < tq (γ) < ∞.
Since γ|[0,tq (γ)] globally minimizes the traveltime, γ is either a refracted
trajectory, a straight line or a three-segment reflected trajectory. Now, if
it was a refracted trajectory, then tq (γ) = ∞, which is a contradiction, so
assume without loss of generality that γ(tq (γ)) ∈ Γ̂+ tq (γ) . For each δ > 0, the
global time-minimizing curve from q to γ(tq (γ) + δ) must be the straight
line joining both points (it cannot be a tree-segment reflected trajectory
because its last segment would not intersect γ) and in the limit δ → 0,
we find the (time-parametrized) straight line ξ from q to γ(tq (γ)) satisfying
ξ(tq (γ)) ∈ Γ̃tq (γ) . So p = γ(tq (γ)) ∈ Γ̃tq (γ) ∩ Γ̂+
tq (γ) ∈ Cq . 

Essentially, any wavefront trajectory crossing one of the curves σ ± loses


the property of being globally time-minimizing, no longer providing points
in the wavefront afterwards. In practice, σ ± represent the merging regions
of two wavefronts coming from different directions, becoming extremely
relevant in some situations. When dealing with wildfires, for instance,
these curves are danger zones for the firefighters, as they can easily become
trapped by the fire and the convergence of several fire trajectories leads to
the corresponding increase in the heat intensity (see also the discussion in
[15]).

Acknowledgments
This work is the result of a research stay of EPR at DTU Compute, par-
tially supported by The International Doctoral School of the University of
Murcia (EIDUM). EPR was also partially supported by MICINN/FEDER
project with reference PGC2018-097046-B-I00 and by Contratos Predoctor-
ales FPU-Universidad de Murcia and Ministerio de Universidades, Spain.

References
[1] P.L. Antonelli, A. Bóna and M.A. Slawiński. Seismic rays as Finsler geodesics.
Nonlinear Anal. RWA 4 (5), 711–722 (2003).
[2] D. Bao, S.-S. Chern and Z. Shen. An Introduction to Riemann-Finsler geometry.
In: Graduate Texts in Mathematics, vol. 200, Springer-Verlag, New York, 2000.
[3] A. Bóna and M.A. Slawiński. Raypaths as parametric curves in anisotropic,
nonuniform media: differential-geometry approach Nonlinear Analysis 51, 983–994
(2002).
[4] E. Caponio, M.A. Javaloyes and M. Sánchez. On the interplay between
Lorentzian causality and Finsler metrics of Randers type. Rev. Mat. Iberoam. 27
(3), 919–952 (2011).
[5] E. Caponio, M.A. Javaloyes and M. Sánchez. Wind Finslerian structures: from
Zermelo’s navigation to the causality of spacetimes. ArXiv e-prints, arXiv:1407.5494
[math.DG] (2014). To appear in Memoirs of AMS.
[6] H.R. Dehkordi. Applications of Randers geodesics for wildfire spread modelling.
Applied Mathematical Modelling 106, 45–59 (2022).
SNELL’S LAW REVISITED 25

[7] H.R. Dehkordi and A. Saa. Huygens’ envelope principle in Finsler spaces and
analogue gravity. Classical Quantum Gravity 36 (8), 085008 (2019).
[8] R.C. Fetecau, J.E Marsden, M. Ortiz and M. West. Nonsmooth Lagrangian
Mechanics and Variational Collision Integrators. SIAM J. Applied Dynamical Systems
2 (3), 381–416 (2019).
[9] M.A. Finney. FARSITE: Fire Area Simulator-model development and evaluation.
USDA Forest Service, Res. Pap. RMRS-RP-4, Rocky Mountain Research Station,
Ogden, UT, 1998 (revised 2004).
[10] G. W. Gibbons. A Spacetime Geometry picture of Forest Fire Spreading and of
Quantum Navigation. ArXiv e-prints, arXiv:1708.02777 [gr-qc] (2017).
[11] G.W. Gibbons, C.A.R. Herdeiro, C.M. Warnick and M.C. Werner. Station-
ary metrics and optical Zermelo-Randers-Finsler geometry. Phys. Rev. D 79, 044022
(2009).
[12] G.W. Gibbons and C.M. Warnick. The geometry of sound rays in a wind. Con-
temp. Phys. 52 (3), 197–209 (2011).
[13] M.A. Javaloyes and E. Pendás-Recondo. Lightlike hypersurfaces and time-
minimizing geodesics in cone structures. ArXiv e-prints, arXiv:2109.07969 [math.DG]
(2021). To appear in: A.L. Albujer et al. (eds.), Developments in Lorentzian Geome-
try, Springer Proceedings in Mathematics & Statistics 389.
[14] M.A. Javaloyes, E. Pendás-Recondo and M. Sánchez. Applications of cone
structures to the anisotropic rheonomic Huygens’ principle. Nonlinear Analysis 209,
112337 (2021).
[15] M.A. Javaloyes, E. Pendás-Recondo and M. Sánchez. A general model for wild-
fire propagation with wind and slope. ArXiv e-prints, arXiv:2110.03364 [math.DG]
(2021).
[16] M.A. Javaloyes and M. Sánchez. On the definition and examples of Finsler met-
rics. Ann. Sc. Norm. Super. Pisa Cl. Sci. 13 (3), 813–858 (2014).
[17] M.A. Javaloyes and M. Sánchez. On the definition and examples of cones and
Finsler spacetimes. RACSAM 114, 30 (2020).
[18] S. Markvorsen. A Finsler geodesic spray paradigm for wildfire spread modelling.
Nonlinear Anal. RWA 28, 208–228 (2016).
[19] S. Markvorsen. Geodesic sprays and frozen metrics in rheonomic Lagrange mani-
folds. ArXiv e-prints, arXiv:1708.07350 [math.DG] (2017).
[20] T. Yajima and H. Nagahama. Finsler geometry of seismic ray path in anisotropic
media. Proc. R. Soc. A 469, 1763–1777 (2009).

DTU Compute, Mathematics,


Technical University of Denmark,
DK-2800 Kgs. Lyngby, Denmark.
Email address: stema@dtu.dk

Departamento de Matemáticas,
Universidad de Murcia,
Campus de Espinardo,
30100 Espinardo, Murcia, Spain.
Email address: e.pendasrecondo@um.es

You might also like