, D (), 1 − 2α0 > 2 2 2 N 2 p (2 p − 1) / 0 ρ 0 + (1 − α0 ) 1 N 2D2 () 1 ρ0 n0 − 1 · , − . p> + > 2 6 1 + (1 − α0 ) N α0 κ0 ρ 0 N
(4.14)
Then, for κ > κ0 , p and t0 a time for which (4.11) holds , there exists a set S ⊂ {1, . . . , N } such that |S| ≥ n 0 and
min xi , x j >
i, j∈S
1 − [1 − 4 p(1 − p + D2 ()/κ)]1/2 , t ≥ t0 . (4.15) 2p
Proof For notational simplicity, we set βi (t0 ) =
xi (t0 ), xc (t0 ) , i = 1, . . . , N and β0 := 1 − α0 , ρ(t0 )
and suppress t0 dependence in βi (t0 ), i.e., βi = βi (t0 ), i = 1, . . . , N , and define index sets: S+ := {i |βi > β0 }, S0 := {i | − β0 ≤ βi ≤ β0 }, S− := {i |βi < −β0 }. Then, clearly we have S+ ∪ S0 ∪ S− = {1, . . . , N }. Next, we claim that the set S+ plays the role of S in the statement of Lemma 4.5, i.e., S+ satisfies the conditions (4.15). This will be verified in two steps. • Step A (S+ satisfies (4.15)1 ): By condition (4.11)1 and the relation ρ = xc , we have |S0 | 1 |S+ | |S− | ρ 0 ≤ ρ(t0 ) = + (−β0 ) + β0 βi < N N N N i (4.16) |S0 | |S+ | = (1 + β0 ) + 2β0 − β0 , N N where we used the relations:
123
S.-Y. Ha et al.
|S+ | |S0 | |S− | =1− − . N N N
|βi | ≤ 1,
(4.17)
On the other hand, we use the relation (4.11)2 and (4.17) to see 1−
N |S0 | 2D2 () 2D2 () 1 2 |S+ | |S− | ≤1− βi ≤ ≤ + + β02 κρ0 κρ(t0 ) N N N N i=1
|S0 | |S0 | = 1 − (1 − β02 ) = 1 − α0 . N N This yields N 2D2 () · , i.e., |S0 | ≤ |S0 | ≤ α0 κ0 ρ0
/
0 N 2D2 () · . α0 κ0 ρ0
(4.18)
Then, it follows from (4.16) and (4.18) that we have ρ 0 + β02 2β0 |S0 | 1 |S+ | ρ 0 + β0 − · −1· > ≥ N 1 + β0 1 + β0 N N 1 + β02 n0 − 1 > . N
/
N 2D2 () · α0 κ0 ρ0
0 (4.19)
Here in the first inequality, we used ρ 0 + β02 ρ 0 + β0 ≥ 1 + β0 1 + β02
⇐⇒
1 − β0 ≥ ρ 0 (1 − β0 );
2β0 <1 1 + β0
and in the last inequality of (4.19), we used the last inequality of (4.14). Hence |S+ | ≥ n 0 . • Step B (S+ satisfies (4.15)2 ): Let i, j ∈ S+ . Then, we use xi (t0 ), xc (t0 ) = ρ(t0 )βi > ρ(t0 )β0 , to get
2
xi (t0 ) − β0 xc (t0 ) | = 1+β 2 − 2 β0 xi (t0 ), xc (t0 ) < 1−β 2 = α0 . (4.20) 0 0
ρ(t0 ) ρ(t0 ) Similarly, we have
x j (t0 ) − β0 xc (t0 ) < √α0 ,
ρ(t0 )
(4.21)
Then, we use (4.20) and (4.21) to get
x (t ) − x j (t0 ) 2
xi (t0 ) − x j (t0 ) < 2√α0 , i.e., xi (t0 ), x j (t0 ) = 1 − i 0 > 1 − 2α0 . 2 Hence, we have =
1 1 − [1 − 4 p(1 − p + D2 ()/κ)]1/2 > , 2p 2p
where we used the third relation in (4.14).
min xi (t0 ), x j (t0 ) > 1 − 2α0 >
i, j∈S
123
1 2 nN0
On the Relaxation Dynamics of Lohe Oscillators...
4.2.2 Proof of Theorem 4.2 Suppose that N ≥ 3, κ > 0,
D() > 0, xi0 = 1, ρ 0 > 0.
We split its proof into two steps as follows. • Step A: We claim that for N ≥ 3 and ρ 0 > 0, there exists 0 < α0 < 1, n 0 ∈ N, κ0 > 0 satisfying (4.14). • Case A (N even): We choose m 0 ∈ Z such that 2(m 0 + 1) 2m 0 < ρ0 ≤ , N N and for such m 0 , we set n0 =
1m 2 0
2
N 1152 D2 () ρ 0 /8, m 0 ≤ 1, + + 1, κ0 = , α = 0 m0 2 7 (ρ 0 )3 m 0 ≥ 2. 4N ,
(4.22)
• Case B (N odd): We choose m 0 ∈ Z such that 2m 0 − 1 2m 0 + 1 ≤ ρ0 < , N N and for such m 0 , we set n0 =
1m 2 0
2
N +1 640 D2 () + , κ0 = , α0 = 2 3 (ρ 0 )3
1 4N , m0 8N ,
m 0 ≤ 1, m 0 ≥ 2.
(4.23)
Then, for both cases, the relation N2 < n 0 ≤ N is easily verified. The other conditions will be checked in Appendix A. • Step B: We use Lemma 4.2 to conclude the p-entrainment. We set . 1152 D2 () 0 κ(ε) := max , κ (ε, ρ ) , 1 7 (ρ 0 )3 (4.24) 1 T (ε) := t0 + T1 (ε) ≤ + T1 (ε). 2D2 ()ρ 0 Here κ1 and t0 are the coupling strength and time appearing in Lemma 4.5, and T1 is the time we should wait for sup max xi (t) − x j (t) ≤ ε.
t>T (ε) i, j∈S
Note that κ(ε) and T1 (ε) are functions of ε, ρ 0 , D2 (), and d. This is because the function D2 () 2 f (y) = 2κ − py + y − 1 − p + 2κ used in Lemma 4.2 via Lemma 4.1 depends only on p, d, ρ 0 and κ. In detail, for a proper value of κ(ε), we can set 1 + [1 − 4 p(1 − p + D2 ()/κ)]1/2 1 > 1 − cos ε. 2p 2 On the other hand, from Step A,
123
S.-Y. Ha et al.
min xi (t0 ), x j (t0 ) >
i, j∈S
1 − [1 − 4 p(1 − p + D2 ()/κ)]1/2 . 2p
Hence for κ(ε), min xi (t0 + T1 (ε)), x j (t0 + T1 (ε)) > 1 − cos ε,
i, j∈S
for a finite time T1 (ε) which may depend on κ(ε), ε, ρ 0 , D2 (), and d. Therefore, sup max xi (t) − x j (t) ≤ ε.
t>T (ε) i, j∈S
5 Gradient Flow Formulation of the Lohe Models In this section, we present gradient flow formulations of the Lohe matrix models for i = , and using this new gradient flow formulation, we show that all initial configurations tend to the complete entrainment state asymptotically.
5.1 A Gradient Flow on Riemannian Manifolds Recall that for a finite-dimensional gradient flow with an analytical potential on the Euclidean space Rd , it is well known that a uniformly bounded flow converges. Below, we extend this result to a Riemannian setting. First, we recall the Łojasiewicz inequality in a Riemannian framework where the proof of Łojasiewicz inequality in Rd can be generalized in a straightforward manner. Theorem 5.1 Let (M, g) be a Riemannian C ω −manifold with Riemannian metric g. Then, for an open subset U ⊂ M, a point p ∈ U , and a real analytic function f : U → R, there exist constants γ ∈ [ 21 , 1), C L and an open neighborhood V satisfying p ∈ V ⊂ U such that | f (z) − f ( p)|γ ≤ C L ∇ f (z), ∀z ∈ V .
(5.1)
Proof For the Euclidean space M = Rd , the Łojasiewicz inequality (5.1) has been proved in [26]. Since the inequality (5.1) has a local nature, the proof for the Euclidean space can be trivially extended to the Riemannian manifold M. We now apply Theorem 5.1 to study the asymptotic behavior of gradient flows with analytical potentials on Riemannian manifolds. The following theorem is an extended version of the arguments in [13], which was used for the Kuramoto model in R N . Theorem 5.2 Let f : U → R be a real analytic function defined on an open subset U of a Riemannian manifold (M, g), and suppose the smooth path θ : [0, ∞) → U is an integral curve of −∇ f with starting point θ 0 ∈ U , i.e., dθ (t) d = θ (0) = θ 0 , θ∗ = −∇ f (θ (t)). dt dt (Here, the gradient vector −∇ f is obtained from the covariant vector −d f under the natural identification induced by the Riemannian metric g on M. The pushforward θ∗ is the
123
On the Relaxation Dynamics of Lohe Oscillators...
corresponding derivative as well.) If the image of θ is contained in a compact subset K of U , then
d
= 0. θ∗ ∃ lim θ (t) = θ∞ ∈ K and lim
t→∞ t→∞ dt
Proof Without loss of generality, we may assume the initial condition θ 0 is not an equilibrium. Then, θ (t) can never be an equilibrium point in finite time since f is analytic. Since the image of θ is contained in the compact metric space K , we may find a sequence {tn }∞ n=1 and a point θ∞ ∈ M such that t1 ≤ t2 ≤ · · · ≤ tn → ∞,
lim θ (tn ) = θ∞ .
n→∞
On the other hand, f (θ (t)) satisfies d f (θ (t)) = −∇ f (θ (t))2 , dt and since f is bounded on K , there exists a limit limt→∞ f (θ (t)) = f ∞ and we must have f (θ∞ ) = f ∞ by continuity of f . Then, we apply Theorem 5.1 at θ∞ to obtain constants γ ∈ [ 21 , 1), C L and an open neighborhood θ∞ ∈ V ⊂ U such that | f (z) − f ∞ |γ ≤ C L ∇ f (z), ∀z ∈ V . Now consider the metric d induced on M through the Riemannian metric, and choose r > 0 such that Br (θ∞ ) ⊂ V . Define the function h : [0, ∞) → [0, ∞) by h(t) = ( f (θ (t)) − f ∞ )1−γ . Clearly, as t → ∞, h ↓ 0. So for any ε ∈ (0, ε0 ) there exists some T0 ∈ {tn } such that |h(t) − h(T0 )| ≤
ε(1 − γ ) ε , ∀t ≥ T0 and θ (T0 ) − θ∞ < . 3C L 3
We now intend to show that θ (t) ∈ Bε (θ∞ ) for t ≥ T0 . The following set T = {t1 > T0 : θ (t) ∈ B (θ∞ ) for all t ∈ (T0 , t1 )}
is nonempty by continuity of θ (t). Also, for any t1 ∈ T , d dh(t) = (1 − γ )( f (θ (t)) − f ∞ )−γ f (θ (t)) = −(1 − γ )( f (θ (t)) − f ∞ )−γ ∇ f (θ (t))2 dt dt 1−γ ≤− ∇ f (θ (t)), t ∈ (T0 , t1 ). CL Hence
3
t1 T0
3
θ∗ d dτ =
dτ
t1
∇ f (θ (τ ))dτ ≤ −
T0
C L (h(t1 ) − h(T0 )) ε ≤ . 1−γ 3
(5.2)
Now, we set T1 = sup T . If T1 < ∞ then clearly T1 ∈ T . Then by inequality (5.2), 3 T1 d d(θ (T1 ), θ∞ ) ≤ d(θ (T1 ), θ (T0 )) + d(θ (T0 ), θ∞ ) ≤ θ∗ ( )dτ + θ (T0 ) − θ∞ dτ T0 2 ≤ . 3
123
S.-Y. Ha et al.
This contradicts the definition of T1 since we may choose an element of T greater than T1 by continuity of θ (t). Hence T1 = ∞, i.e., θ (t) ∈ B (θ∞ ), t ≥ T0 . By taking the limit t1 → ∞ in (5.2), we obtain 3 ∞ ε θ˙ (τ )dτ ≤ , 3 T0 which implies the existance of the limit lim θ (t) = θ∞ .
t→∞
On the other hand, by the continuity of ∇ f , the limit
d
= lim − ∇ f (θ (t)) = − ∇ f (θ∞ ), lim θ∗ t→∞
dt t→∞ must be 0 since the limit limt→∞ θ (t) exists.
Corollary 5.1 Let f : M → R be a real analytic function defined on a compact Riemannian manifold M, and suppose that the smooth path θ : [0, ∞) → M is an integral curve of −∇ f with starting point θ 0 ∈ M, i.e. d 0 θ (0) = θ , θ∗ = −∇ f (θ (t)). dt Then the limit lim θ (t) = θ∞ ∈ M
t→∞
exists and
d
= 0. θ∗ lim
t→∞
dt
Proof We set M = U = K in Theorem 5.2 to get the proof.
5.2 The Kuramoto Model is a Gradient Flow We briefly discuss the gradient flow formulation of the Kuramoto model as a motivation to the Riemannian manifolds: θ˙i = νi +
N κ sin(θ j − θi ), i = 1, . . . , N . N
(5.3)
j=1
Recall that the Lohe matrix model in one-dimension d = 1 is reduced to the Kuramoto model via the correspondence relation Ui = e−iθi ,
Hi = νi , i = 1, . . . , N .
Then, we define an order parameter (Rk , φk ) and potential Vk as follows:
123
(5.4)
On the Relaxation Dynamics of Lohe Oscillators...
Rk eiφk :=
N 1 iθ j e , := (θ1 , . . . , θ N ), N j=1
ν := (ν1 , . . . , ν N ), Vk () := −ν · −
(5.5)
κ cos(θ j − θi ). N i, j
We use (5.4) to rewrite the potential Vk in terms of R: Vk () = −ν · − κ N Rk2 .
(5.6)
Then it is easy to see that the Kuramoto flow (5.3) is a gradient flow: ˙ = −∇ Vk (). This fact is used in [13] to prove the emergence of synchronization in the Kuramoto model whose domain is R N . Hence it is natural to ask whether the Lohe matrix and sphere models can be written as a gradient flow or not. This will be the main focus of the next subsections.
5.3 Gradient Flow Formulations of the Lohe Models In this subsection, we present gradient flow formulations for generalized Lohe models on four Riemannian manifolds: M : Sd , U(d),
SU(d), R × SU(d).
In the following, we consider Lohe flows on the above Riemannian manifolds and present a gradient flow formulation.
5.3.1 A Lohe Flow on Sd Consider the Lohe sphere model for identical oscillators: x˙i = +
N N 1 κ (xk − xi , xk xi ) = + κP⊥ x k , i = 1, . . . , N . xi N N k=1
(5.7)
k=1
where is a skew-symmtric matrix and P⊥ xi is the orthogonal projection onto the tangent plane perpendicular to xi : P⊥ xi y = y − x i , yx i . Without loss of generality, we assume = 0. Now, we introduce a potential function V0 : V0 (X ) := −
N κ xi , xk . 2N
(5.8)
i,k=1
Note that the potential function V0 is analytic, and recall that the surface gradient is defined as the projection of the gradient onto the tangent plane. Proposition 5.1 The Lohe sphere model (5.7) with = 0 is a gradient flow: # # x˙i = −∇xi V0 # d , Txi S
123
S.-Y. Ha et al.
# # where ∇xi V0 #
Txi Sd
denotes the orthogonal projection of the gradient vector ∇xi V0 ∈ Rd+1
onto the tangent plane Txi Sd at xi , which is the induced gradient vector on the manifold Sd . Proof It follows from (5.8) that for each i = 1, . . . , N , ∇ x i V0 = −
N κ xk . N k=1
This yields # # ∇ x i V0 #
Txi Sd
# N κ ## =− xk # # N k=1
=− Txi Sd
N κ (xk − xk , xi xi ) . N k=1
As a direct application of Theorem 5.2 and Proposition 5.1, we obtain the following convergence result for the Lohe flow from any initial data. Corollary 5.2 Let X = X (t) be a Lohe flow whose dynamics is governed by the system (5.7) with = 0. Then, there exists an equilibrium X ∞ such that X ∞ = lim X (t). t→∞
Proof Since (Sd ) N is a compact manifold, the Lohe flow X is always bounded. Thus it follows from Theorem 5.2 and Proposition 5.1 that any Lohe flow converges. Next, we briefly argue that why we need to restrict i = 0. One natural question is the existence of potential functions when the natural frequencies are not identical. For simplicity, let κ = 0 so that the interaction terms disappear: x˙i = i xi , i = 1, . . . , N , where the i ’s are skew-symmetric matrices. In this case, a possible ansatz on the nontrivial potential function might be V (x) = −
N
x tj j x j ,
j=1
but then the derivatives of V are zero due to skew-symmetry of i : −
∂V = (it + i )xi = 0, i.e., V = constant. ∂ xi
On the other hand, for the Kuramoto model, the potential for nonidentical oscillators still exists on the covering space (R1 ) N of (S1 ) N as we can see in (5.5), which is not well-defined on (S1 ) N . For the Lohe sphere model, let us consider d > 1, since d = 1 is exactly the same as the Kuramoto model. The sphere Sd for d > 1 is a simply connected manifold, which has no proper covering space. Therefore, the potential function should be on (Sd ) N if it exists. However, it cannot exist on (Sd ) N since this model has a periodic solution. We can easily see this from the following example. Example 5.1 Let κ = 0 and
123
On the Relaxation Dynamics of Lohe Oscillators...
⎛ 0 ⎜0 ⎜ ⎜ 1 = ⎜ ... ⎜ ⎝0 1
0 0 .. . 0 0
··· ··· .. . ··· ···
0 0 .. . 0 0
⎞ −1 0 ⎟ ⎟ .. ⎟ , = 0 for any k = 1. k . ⎟ ⎟ ⎠ 0 0
Then the i ’s are skew-symmetric matrices. With initial data x10 = (1, 0, . . . , 0)t , it has a periodic solution x1 (t) = exp(1 t)x10 = (cos(t), 0, . . . , 0, sin(t))t , with xk (t) = xk0 for k = 1. Remark 5.1 The Lohe sphere model on Sd for nonidentical oscillators does not have a potential function on Sd . However, for d = 1, it has a potential on its covering space (R1 ) N , as same as the Kuramoto model.
5.3.2 A Lohe Flow on U(d) In this part, we consider the Lohe matrix model on the unitary Hamiltonians Hi = H : ⎧ † † κ N ⎪ ˙ † ⎪ j=1 U j Ui − Ui U j , ⎨ Ui Ui = −iH + 2N N † † ˙ i = −iHUi + κ U , U U U − U U U j i i i j=1 i j 2N ⎪ ⎪ ⎩ Ui (0) = Ui0 ,
group U(d) with identical or equivalently t > 0,
(5.9)
i = 1, . . . , N .
Note that we may assume H = 0 in (5.9) without loss of generality, if necessary by considering a change of variable Ui → eiH t Ui , so that N κ U j Ui† Ui − Ui U †j Ui , t > 0. U˙ i = 2N
(5.10)
j=1
On the other hand, motivated by the Kuramoto model in Sect. 5.2, we introduce an order parameter R and a potential V1 for (5.9) with Hi = 0: R 2 :=
N κ 1 N and V1 {Ui }i=1 := − N R 2 . tr Ui U †j 2 N 2
(5.11)
i, j=1
Here, the Riemannian metric of U(d) is induced by the natural inclusion U(d) → Md,d (C). For the one-dimensional case, the relations (5.11) are exactly the same as (5.5) and (5.6) via the relation (5.4). Proposition 5.2 The Lohe model (5.9) with Hi = 0 is a gradient flow with analytical potential
V1 in (5.11):
# ∂ V1 ## U˙ i = − ∂Ui #TU
i
U(d)
, i = 1, . . . , N , t > 0. 2
Proof The function V1 has an obvious polynomial extension to all of Md,d (C) N = R2d N 2 viewed as a real analytic manifold. Since each variable Ui is in Md,d (C) = R2d , the partial
123
S.-Y. Ha et al.
derivatives of a matrix can be calculated by the partial derivatives of each real and imaginary 2 kl component of Ui on R2d . Let u ikl = aikl + ) ibi be the (k, l)-element of matrix Ui*. First, we use tr(Ui U †j ) = dk,l=1 u ikl u¯ klj = dk,l=1 (aikl a klj + bikl bklj ) + i(a klj bikl − aikl bklj ) to see V1 = −
d N N * κ κ ) kl kl (ai a j + bikl bklj ) + i(a klj bikl − aikl bklj ) tr(Ui U †j ) = − 2N 2N i, j=1
=−
κ 2N
N
i, j=1 k,l=1
d
) * aikl a klj + bikl bklj ,
i, j=1 k,l=1
where we cancel the imaginary term by symmetry of the indices i, j. This yields ∂ V1 ∂aikl
=−
N κ kl aj , N j=1
∂ V1 ∂bikl
=−
N κ kl bj , N j=1
2
and thus reverting back to the coordinates of Md,d (C) N = R2d N , we have ⎛ ⎞
# d d N N ∂ V1 ∂ V1 ## ∂ V1 2κ κ ⎝− = + i kl E kl = a klj − i bklj ⎠ E kl kl ∂Ui #TU Md,d (C) N N ∂a ∂b i i k,l=1 k,l=1 j=1 j=1 i =−
d d N N & κ % kl κ kl kl uj E a j + ibklj E kl = − N N j=1 k,l=1
=−
κ N
N
j=1 k,l=1
Uj,
j=1
where E kl denotes the d × d matrix whose (k, l)-coordinate is 1 and the other coordinates are 0. Now, we should project it on the tangent space TUi U(d) of U(d) at Ui . The orthogonal projection π : TId Md,d (C) → u(d) is given by A → 21 (A − A† ) since u(d) = TId U(d) = {X ∈ Md,d (C) | X + X † = 0} is the set of skew Hermitian matricies (See Chapter 4 in [34]). Since TUi U(d) is the right translate u(d)Ui of u(d), we can see that the orthogonal projection πUi : TUi Md,d (C) → TUi U(d) is given by AUi → π(A)Ui = 21 (A − A† )Ui for an element AUi ∈ TUi Md,d (C). Hence we may calculate
# # # ∂ V1 ## ∂ V1 ## ∂ V1 ## † = πUi U Ui =π ∂Ui #TU U(d) ∂Ui #TU Md,d (C) ∂Ui #TU Md,d (C) i i i i ⎞ ⎛ N N & κ κ % = π ⎝− U j Ui† ⎠ Ui = − U j Ui† − Ui U †j Ui . N 2N j=1
j=1
Therefore, we have −
# ∂ V1 ## ∂Ui #TU
i
U(d)
which yields the desired estimate.
123
Ui† =
N κ U j Ui† − Ui U †j , 2N j=1
On the Relaxation Dynamics of Lohe Oscillators...
As a direct application of Theorem 5.2 and Proposition 5.2, we have the convergence of the flow e−H t Ui as t → ∞. Corollary 5.3 Suppose that the Hamiltonians Hi are the same: Hi = H , i = 1, . . . , N , and let Ui = Ui (t) be a solution to the Cauchy problem (5.9). Then, the quantities e−iH t Ui converge for any initial configuration {Ui0 }. Proof Upon change of variables Ui → eiH t Ui , we may assume H = 0. Then we are done by Theorem 5.2 and Proposition 5.2. Remark 5.2 The above corollary implies that for the Lohe model (2.1) with positive coupling strength and zero natural frequencies, lim U˙i (t) = 0 for all i.
t→∞
This can be proved by more elementary means similar to those of Lemma 3.4, without employing the profound result of Łojasiewicz’s inequality and its implications on real analytic flows. However, this does not imply that the relative phases Ui U †j should converge, because the set of pairs of Ui U †j , (i, j = 1, . . . , N ) of equilibrium points is not in general discrete. This in turn highlights the significance of Łojasiewicz’s inequality in proving the convergence of Ui . Proof We first prove the existence of ρ ∞ where ρ is defined by the mean value, ρ = Uc , Uc =
N 1 Ui , N i=1
where we use the trace norm A =
tr(A A† ),
A ∈ Md,d (C).
Note that N κ κ U j Ui† − Ui U †j = (Uc Ui† − Ui Uc† ), U˙ i Ui† = 2N 2 j=1
so we may write κ U˙ i = (Uc − Ui Uc† Ui ). 2 From the definition of the norm, κ2 tr(Uc − Ui Uc† Ui )(Uc† − Ui† Uc Ui† ) 4 κ2 = tr(Uc Uc† − Ui Uc† Ui Uc† − Uc Ui† Uc Ui† + Ui Uc† Ui Ui† Uc Ui† ) 4 κ2 = tr(Uc Uc† − Ui Uc† Ui Uc† − Ui† Uc Ui† Uc + Uc† Uc ) 4 κ2 = Re tr(Uc Uc† − Ui Uc† Ui Uc† ). 2 On the other hand, U˙ i 2 =
123
S.-Y. Ha et al. N κ (Uc − Ui Uc† Ui ). U˙ c = 2N i=1
Thus, κ 2 ˙ 2 d 2 d (Uc Uc† − Ui Uc† Ui Uc† ) = Ui , ρ = Uc 2 = 2Re tr(U˙ c Uc† ) = Re tr dt dt N κN i=1 i=1 √ hence ρ is non-decreasing and bounded above by d, which implies the existence of ρ ∞ . Note that
κ
κ κ√
d(1 + d), U˙ i = (Ui − Uc Ui† Ui ) ≤ (Ui + Uc Ui† Ui ) ≤ 2 2 2 and thus that
N N
1
1 ˙ κ√
Ui ≤ d(1 + d). U˙ c =
U˙ i ≤
N
N 2 N
i=1
N
i=1
Now, according to the previous calculations, d ˙ 2 d κ2 Ui = Re tr(Uc Uc† − Ui Uc† Ui Uc† ) dt dt 2 κ2 = Re tr(U˙ c Uc† +Uc U˙ c† − U˙ i Uc† Ui Uc† −Ui U˙ c† Ui Uc† − Ui Uc† U˙ i Uc† −Ui Uc† Ui U˙ c† ) 2 = κ 2 Re tr(U˙ c Uc† − U˙ i Uc† Ui Uc† − Ui U˙ c† Ui Uc† ), and hence we conclude that # # #d # # U˙ i 2 # ≤ κ 2 (U˙ c U † + U˙ i U † Ui U † + Ui U˙ † Ui U † ) c c c c c # dt # ≤ Therefore,
κ3 d(1 + d)(1 + 2d). 2
d 2ρ is bounded, and by Barbalat’s lemma we obtain the conclusion of the remark. dt 2
As we discussed on Sd , we can consider the existence of potential functions for nonidentical oscillators. In U(d), a possible potential function for κ = 0 is V (U ) = i
N
tr(U †j H j U j ).
j=1
However, this is a constant function on U(d) since all U j are unitary. We can rigorously prove the nonexistence of potential functions on U(d) by showing periodic solutions. Example 5.2 For the case of κ = 0 and H1 = h 1 Id with a nonzero real scalar h 1 , the oscillator U1 follows the equation U˙1 = −ih 1 U1 . Therefore, we have a periodic solution U1 (t) = exp(−ih 1 t)U10 . Since U(d) is not a simply connected manifold, we have the following potential function on the covering space R × SU(d) similar to the Kuramoto model. Remark 5.3 The universal covering space of U(d) is R × SU(d). A Lohe flow on U(d) with nonidentical oscillators can have a potential function if and only if the natural frequencies Hi are all scalars. We will see this and analyze a Lohe flow on R × SU(d) in a later subsection.
123
On the Relaxation Dynamics of Lohe Oscillators...
5.3.3 A Lohe Flow on SU(d) Consider a Cauchy problem for the generalized Lohe model on SU(d), which was introduced in [20]: * ) † † † † κ N 1 U˙ i Ui† = −iHi + 2N j=1 U j Ui − Ui U j − d tr U j Ui − Ui U j Id , t > 0, (5.12) Ui (0) = Ui0 , i = 1, . . . , N , where Hi ∈ isu(d)(recall that su(d) consists of skew-hermitian matrices with zero trace). First, we show that system (5.12) preserves the structure of SU(d). Lemma 5.1 Special unitarity is preserved along the dynamics (5.12), i.e., Ui0 ∈ SU(d)
⇒ Ui (t) ∈ SU(d), t ≥ 0.
Proof Since the R.H.S. of (5.12)1 is anti-Hermitian, it is clear to see the unitarity of Ui . Thus, we only need to check whether detUi = 1 or not. Note that U˙ i Ui† has trace zero, and hence Ui−1 U˙ i = Ui−1 (U˙ i Ui† )Ui has trace zero. We use the Jacobi’s determinant formula (See Proposition 15.21 and Problem 15.7 in [34] for detail) to see that d detUi = detUi · tr(Ui−1 U˙ i ) = 0, dt so that detUi = 1 for t ≥ 0.
Proposition 5.3 Suppose that the constant hamiltonians Hi are the same, i.e., Hi = H , i = 1, . . . , N , and let Ui = Ui (t) be a solution to the Cauchy problem (5.12). Then, the Lohe model (5.12) can be rewritten as a gradient flow with analytical potential V1 in (5.11): # ∂ V1 ## ˙ , i = 1, . . . , N , t > 0. Ui = − ∂Ui #TU U(d) i
Here, the Riemannian metric on SU(d) is given by the inclusion SU(d) → Md,d (C). Proof The proof is almost the same with that of Proposition (5.2). Again the function V1 2 has an obvious polynomial extension to all of Md,d (C) N = R2d N viewed as a real analytic manifold, and has gradient # N κ ∂ V1 ## = − Uj. (5.13) ∂Ui #TU Md,d (C) N i
j=1
projection π : Now we want to project it on the tangent%space TUi SU(d). The orthogonal & TId Md,d (C) → su(d) is given by A → 21 A − A† − d1 tr(A − A† )Id where su(d) = {X ∈ Md,d (C) | X + X † = 0, tr(X ) = 0}. Note that the derivative of (detX = 1) at Id is (tr(X ) = 0) because of Jacobi’s determinant formula. (See Chapter 4 in [34] for details) The projection π comes from the composition of the projection from Md,d (C) onto u(d) and the projection from u(d) onto su(d). Since the projection π1 : u(d) → su(d) is given by the projection into the traceless matrices, A → A − d1 tr(A)Id , the projection π is obtained from the composition Md,d (C) → u(d) → su(d), 1 1 1 † † † A → (A − A ) → A − A − tr(A − A )Id . 2 2 d Therefore, the orthogonal projection of (5.13) on TUi U(d) is given by AUi → π(A)Ui ,
123
S.-Y. Ha et al.
# ∂ V1 ## − ∂Ui #TU
i
U(d)
Ui†
N * κ 1 ) † † † † U j Ui − Ui U j − tr U j Ui − Ui U j Id . = 2N d j=1
Then, as a direct application of Proposition 5.3, we have the following corollary. Corollary 5.4 Suppose that Hamiltonians Hi are the same: Hi = H , i = 1, . . . , N , and let Ui = Ui (t) be a solution to the Cauchy problem (5.10) with H = 0. Then, the quantities e−i H t Ui converge for any initial configuration {Ui0 }. Remark 5.4 A Lohe flow on SU(d) with nonidentical oscillators does not have a potential function since SU(d) is simply connected and there is a periodic solution as in Example 5.2.
5.3.4 A Lohe Flow on R × SU(d) Consider the universal covering space of U(d): R × SU(d) → U(d), (θ, V ) → eiθ V . If we use this representation as a parametrization Ui = eiθi Vi of each particle described under the Lohe model, {θi , Vi } behaves as N * κ ) iθ˙i Id + V˙i Vi† = −iHi + V j Vi† ei(θ j −θi ) − Vi V j† ei(θi −θ j ) . 2N j=1
Note that Vi ∈ SU(d), so
V˙i Vi† ∈ su(d), hence tr(V˙i Vi† ) = 0.
We may take the trace on both sides and use the above relation in order to split the derivatives for the R × SU(d) model: N * 1 iκ ) θ˙i = − trHi − tr V j Vi† ei(θ j −θi ) − Vi V j† ei(θi −θ j ) , d 2d N j=1
N i κ ) V j Vi† ei(θ j −θi ) − Vi V j† ei(θi −θ j ) V˙i Vi† = −iHi + trHi Id + d 2N j=1 * * ) 1 − tr V j Vi† ei(θ j −θi ) − Vi V j† ei(θi −θ j ) Id . d As a special case, we choose Hi = −νi Id , νi ∈ R to rewrite (5.14) as follows. N * iκ ) tr V j Vi† ei(θ j −θi ) − Vi V j† ei(θi −θ j ) , θ˙i = νi − 2d N
(5.14)
j=1
κ V˙i = 2N
N )
V j Vi† ei(θ j −θi ) − Vi V j† ei(θi −θ j )
j=1
* 1 − tr V j Vi† ei(θ j −θi ) − Vi V j† ei(θi −θ j ) Id Vi . d We set a real analytic potential on (R × SU(d)) N :
123
(5.15)
On the Relaxation Dynamics of Lohe Oscillators...
N V2 {θi , Vi }i=1
= −d
N i=1
νi θi −
N N κ κ i(θi −θ j ) dνi θi − tr e Vi V j† , N R2 = − 2 2N i=1
(5.16)
i, j=1
where R is the order parameter defined by the relation (5.11). Proposition 5.4 The Lohe model (5.15) is a gradient flow with analytical potential V2 in (5.16): # # ∂ V2 ## # θ˙i = −∇θi V2 # , V˙i = − , TR ∂ Vi #TV SU(d) i
where the Riemannian metric of the SU(d) components of (R × SU(d)) N are given by the inclusion SU (d) → Md,d (C), and those of the R components of (R × SU (d)) N are given as d times of the standard Riemannian metric of R. Proof Note that the function V2 given in (5.16) is defined on (R × Md,d (C)) N viewed as a real analytic manifold, and has gradient ⎞ ⎛ # N * ∂ V2 ## 1⎝ iκ ) − = tr V j Vi† ei(θ j −θi ) − Vi V j† ei(θi −θ j ) ⎠ , νi d − ∂θi #T R d 2N j=1
= νi −
iκ 2d N
N
* ) tr V j Vi† ei(θ j −θi ) − Vi V j† ei(θi −θ j ) ,
j=1
# N ∂ V2 ## κ − = V j ei(θ j −θi ) , ∂ Vi #T Md,d (C) N j=1
where we consider R N × Md,d (C) N with its ordinary metric times d on its R N component and with its ordinary metric on its Md,d (C) N component. We perform an orthogonal projection onto su(d) # to obtain ∂ V2 ## − Vi† ∂V # i TV SU (d) i
=
N * κ 1 ) V j Vi† ei(θ j −θi ) − Vi V j† ei(θi −θ j ) − tr V j Vi† ei(θ j −θi ) − Vi V j† ei(θi −θ j ) Id . 2N d j=1
Therefore, we can conclude that the Lohe flow (5.16) is a gradient flow with analytic potential V2 . As a direct corollary of Theorem 5.2 and Proposition 5.4, we have the following a priori convergence result. Corollary 5.5 Let (θi , Vi ) be a solution to system (5.15) such that the phase variables θi are uniformly bounded. Then, the quantities Ui = eiθi Vi converge.
6 Conclusion In this paper, we have studied the relaxation dynamics of the ensemble of Lohe sphere and matrix oscillators from generic initial configurations. For an ensemble of Lohe oscillators
123
S.-Y. Ha et al.
under the same Hamiltonians, we show that exponential position synchronization occurs asymptotically under the a priori assumption that complete position synchronization occurs. Moreover, we also showed that there are only two possible asymptotic states when the initial positions are all different and have a positive order parameter. This kind of conditions reminds us of the corresponding condition for the Kuramoto model in [19]. For distinct constant Hamiltonians, the relaxation dynamics are rather complicated. In this heterogeneous ensemble, we show that if the coupling strength is sufficiently large, then the heterogeneous ensemble tends to the practical NN−1 -entrained state asymptotically. However, in the intermediate coupling strength regime, we show that practical partial entrainment states will emerge from an initial configuration with positive order parameter. For the convergence result of the Lohe matrix oscillators, we employ a gradient flow approach. For the Kuramoto model, the gradient flow structure of the Kuramoto model plays a key role in the resolution of the complete synchronization in [19]. However, it is not known whether the Lohe matrix models admit gradient flow formulations with analytical potentials in a general setting. Fortunately, our finding in this paper says that at least for Lohe matrix models with the same Hamiltonian, the Lohe matrix flows can admit a gradient flow formulation so that the Lohe matrix flow on compact group manifolds such as U(d) and SU(d) always converge from any initial configuration. Of course, we cannot resolve the gradient flow formulation of the Lohe matrix model with heterogeneous Hamiltonians. We confirmed that it is no longer a gradient flow according to Remark 5.1, 5.3, and 5.4. Hence the asymptotic behaviors of nonidentical oscillators are still open problems for both sphere and matrix model. However, similar behaviors as in the Kuramoto model are observed through numerical simulations. We leave this issue for a future work. Acknowledgements The work of S.-Y. Ha was supported by the Samsung Science and Technology Foundation under Project Number SSTF-BA1401-03. The work of D. Ko was supported by the TJ Park fellowship from POSCO. The work of S. W. Ryoo was supported by the Student Directed Education program of Seoul National University.
Appendix A: Verification of Conditions (4.14) In this section, we show that the choices explicitly verify that the choices (4.22) and (4.23) satisfy the conditions (4.14), that is the conditions
2p (2 p − 1)2
D2 () < κ0 ,
1 − 2α0 >
1 , 2p
1 ρ0 + , 2 6 / 0 1 N 2D2 () n0 − 1 ρ 0 + (1 − α0 ) − . > · 1 + (1 − α0 ) N α0 κ0 ρ 0 N p>
A.1 N is even In this case, recall that we choose
123
(A.1) (A.2) (A.3) (A.4)
On the Relaxation Dynamics of Lohe Oscillators...
1m 2 N 2(m 0 + 1) 2m 0 0 < ρ0 ≤ , n0 = + + 1, N N 2 2 1152 D2 () ρ 0 /8, m 0 ≤ 1, , α0 = m 0 κ0 = 0 3 7 (ρ ) m 0 ≥ 2. 4N , m0 ≈ 21 + Here, n 0 was roughly chosen so that p = nN0 ≈ 21 + 2N would hold in most cases. On the other hand, (A.2) suggests that
α0 <
ρ0 4
(A.5)
and hence that (A.3)
1 1 1 1 1 1 1 p m0 − ≈ − ≈ − (1 − ( p − 1)) = = . 2 4p 2 4 1 + ( p − 1) 2 4 4 4N
+(1−α0 ) > n 0N−1 , hence we can finally choose κ0 large When m 0 ≥ 2, this happens to give ρ1+(1−α 0) enough so as to make (A.1) and (A.4) true. On the other hand, when m 0 ≤ 1, we need to choose a smaller α0 , because then n 0 = N2 + 1 and so 0
ρ 0 + (1 − α0 ) n0 − 1 1 > = 1 + (1 − α0 ) N 2
⇔
α0 < 2ρ 0 ,
m0 1 ≤ 4·2 = 18 . The while ρ0 could be very small. The denominator 8 was chosen because 4N funny coefficient 1152 7 in the choice for κ0 will be explained in the demonstration of (A.4) when m 0 ≥ 2; the condition (A.1) is weaker than (A.4).
• Case A (m 0 ≤ 1): We verify the relations (A.1)–(A.4) one by one. Note that by our choice, we have n0 =
N 4 ρ0 + 1, ρ 0 ≤ , α0 = . 2 N 8
• (Verification of (A.1)) We have p=
1 1 n0 = + , N 2 N
so 1 + (2/N ) N2 2p = = (2 p − 1)2 (2/N )2 16
8 1 8 1152 4+ ≤ 0 2 (4 + 4) = 0 2 ≤ , N (ρ ) (ρ ) 7(ρ 0 )3
hence 2p (2 p − 1)2
1152 D2 () = κ0 . 7 (ρ 0 )3
D2 () <
• (Verification of (A.2)) We have (1 − 2α0 )
ρ0 2 1 2 2n 0 1 = (1 − )(1 + ) ≥ (1 − )(1 + ) > 1, i.e., 1 − 2α0 > , N 4 N N N 2p
where we use the fact that N ≥ 3. • (Verification of (A.3)) We use (A.5) and
4 m0 5 2
= 0 to find
123
S.-Y. Ha et al.
n0 1 1 1 ρ0 1 ρ0 = + ≥ + > + . N 2 N 2 4 2 6 • (Verification of (A.4)) By the choice of α0 and κ0 in (A.5), we have 0 / N 2D2 () N N 2D2 () 7(ρ 0 )2 7N ρ 0 28 · = · 0< = 0. · = ≤ < 1, i.e., α0 κ0 ρ 0 ρ 0 /8 576 72 72 α0 κ0 ρ 0 p=
Hence, we have 1 ρ 0 + (1 − α0 ) − 1 + (1 − α0 ) N
/
N 2D2 () · α0 κ0 ρ 0
0 =
1 n0 − 1 1 + 7ρ 0 /8 −0> = . 2 − ρ 0 /8 2 N
• Case B (m 0 ≥ 2): Similar to Case A, we check the relations (A.1)–(A.4) one by one. • (Verification of (A.1)): Since 1m 2 N m0 N 0 + n0 = +1> + , 2 2 2 2 we have m0 2n 0 −1> . (A.6) 2p − 1 = N N This yields 1 N N 1 2p + 1 < = + 1 (2 p − 1)2 2p − 1 2p − 1 m0 m0 N N 2(m 0 + 1) 2(m 0 + 1) = +1 · · 2(m 0 + 1) m0 2(m 0 + 1) m0 1 5 1152 1 4 ≤ 0 ·4 ≤ ·4+1 ≤ 0 , ρ ρ0 ρ ρ0 7(ρ 0 )3 Hence 2p (2 p − 1)2
D2 () <
1152 D2 () = κ0 . 7 (ρ 0 )3
m0 and (A.6) to obtain • (Verification of (A.2)) We use α0 = 4N m 0 m0 m0 m0 2n 0 (1 − 2α0 ) p = (1 − 2α0 ) > 1− 1+ =1+ 1− > 1, N 2N N 2N N i.e., we have
1 − 2α0 >
1 . 2p
• (Verification of (A.3)) We again use the relation (A.6) to get p=
1 m0 1 2(m 0 + 1) m0 1 1 ρ0 2 n0 > + = + · ≥ + ρ0 · = + . N 2 2N 2 N 4(m 0 + 1) 2 4·3 2 6
• (Verification of (A.4)) By the choice of α0 =
m0 4N
and κ0 , we have
7N 2 (ρ 0 )2 4N 2 7(ρ 0 )2 7 · 4(m 0 + 1)2 N 2D2 () = · = · ≤ 0 α0 κ0 ρ m0 576 144m 0 144m 0 7m 0 1 2 = . 1+ 36 m0
123
(A.7)
On the Relaxation Dynamics of Lohe Oscillators...
On the other hand, we have 2m 0 m0 + 1 − 4N ρ 0 + (1 − α0 ) n 0 − 1 − > N m0 − 1 + (1 − α0 ) N 1 + 1 − 4N m 0 (7N + m 0 ) . = 2N (8N − m 0 )
m0 1 + 2N 2
=
4N + 7m 0 N + m0 − 8N − m 0 2N
(A.8) Thus, we use m 0 ≥ 2, (A.7) and (A.8) to find / 0 ρ 0 + (1 − α0 ) n 0 − 1 1 N 2D2 () · − − 1 + (1 − α0 ) N N α0 K0ρ0 (m 0 + 1)2 m 0 (7N + m 0 ) −7 2N (8N − m 0 ) 36N m 0 14N (5m 20 − 8m 0 − 4) + 25m 30 + 14m 20 + 7m 0 = 36N m 0 (8N − m 0 ) 14N (m 0 − 2)(5m 0 + 2) + 25m 30 + 14m 20 + 7m 0 = > 0. 36N m 0 (8N − m 0 ) >
i.e. we have ρ 0 + (1 − α0 ) 1 − 1 + (1 − α0 ) N
/
N 2D2 () · α0 K0ρ0
0 >
(A.9)
n0 − 1 . N
The role of the coefficient 1152 7 appears in the last line of (A.9). For any coefficient ζ , the last line of (A.9) would appear as replacing 1152 7 N a(m 0 ) + b(m 0 ) , N m 0 (8N − m 0 ) where a(m 0 ) is a quadratic in m 0 and b(m 0 ) is a cubic in m 0 , depending on ζ . If a(2) < 0, then we may choose N sufficiently large so that this rational function takes a negative value. Hence we must have a(2) ≥ 0, which forces us to choose ζ ≥ 1152 7 . This choice of ζ happens to make the rational function positive for all valid values of m 0 and N .
A.2 N is odd Recall that we choose m 0 to satisfy 2m 0 − 1 2m 0 + 1 ≤ ρ0 < . N N Then, for such m 0 , we set n0 =
1m 2 0
2
N +1 640 D2 () + , α0 = , κ0 = 2 3 (ρ 0 )3
(A.10)
1 4N , m0 8N ,
m 0 ≤ 1, m 0 ≥ 2.
(A.11)
We again claim that the choices (A.10)–(A.11) satisfy (A.1)–(A.4). The rationale for choosing n 0 , and α0 when m 0 ≥ 2, is similar to A.1. However, when m 0 ≤ 1, we have −1 n 0 = N 2+1 and n 0N−1 = N2N , which is less than 21 , so we can be a little more lenient with our choice of α0 so as to make
ρ 0 +(1−α0 ) 1+(1−α0 )
>
n 0 −1 N
even when ρ 0 ≈ 0. Again, the coefficient
123
S.-Y. Ha et al.
in the choice of κ0 is necessitated by (A.4) when m 0 ≥ 2; the condition (A.1) is weaker than (A.4). 640 3
• Case C (m 0 ≤ 1): Note that the relations (A.11) yield 1 N +1 3 , n0 = and ρ 0 < . 4N 2 N • (Verification of (A.1)) We use the relations (A.12) to obtain α0 =
(A.12)
2 nN0 2p 9(N + 1) N 2 640 1 = = N (N + 1) = . · < 12 · 0 2 < & % 2 0 )3 n0 (2 p − 1)2 N 9 (ρ ) 3(ρ 2N −1 Thus, we have 640 D2 () 2p D2 () < = κ0 . (2 p − 1)2 3 (ρ 0 )3 • (Verification of (A.2)) We again use (A.12) to get (1 − 2α0 )2 p = (1 − 2α0 )
2n 0 1 N +1 1 = (1 − )· =1+ N 2N N 2N
1−
1 N
> 1.
• (Verification of (A.3)) We use (A.12) to get p=
1 1 1 ρ0 n0 = + > + . N 2 2N 2 6
• (Verification of (A.4)) Note that 0<
3(ρ 0 )2 N 2D2 () 3N 2 (ρ 0 )2 27 · = 4N 2 · = < < 1, i.e., 0 α0 κ0 ρ 320 80 80 0 / N 2D2 () · = 0. α0 κ0 ρ 0
This yields 1 ρ 0 + (1 − α0 ) − 1 + (1 − α0 ) N ≥
1− 2−
1 4N 1 4N
−0−
/
0 N 2D2 () n0 − 1 · − α0 κ0 ρ 0 N
N −1 4N − 1 N −1 7N − 1 = − = > 0, 2N 8N − 1 2N 2N (8N − 1)
i.e., 1 ρ 0 + (1 − α0 ) − 1 + (1 − α0 ) N
/
N 2D2 () · α0 κ0 ρ 0
0 >
n0 − 1 > 0. N
• Case D (m 0 ≥ 2): By definition of n 0 in (A.11), we have 1m 2 N + 1 m0 N +1 m0 − 1 N 0 + > −1+ = + , n0 = 2 2 2 2 2 2 1m 2 N + 1 m0 N +1 0 + n0 = ≤ + . 2 2 2 2 These yield 2p =
123
m0 − 1 m0 + 1 2n 0 > + 1 and 2 p < + 1. N N N
(A.13)
(A.14)
On the Relaxation Dynamics of Lohe Oscillators...
• (Verification of (A.1)) It can easily be seen that x →
2x (2x−1)2
=
1 2x−1
1 · ( 2x−1 + 1) is
decreasing for x > 21 . Thus we can use (A.14)1 to see We use (A.10) and (A.14) to see m 0 −1 2m 0 + 1 1 2 N (m 0 − 1 + N ) N (m 0 − 1 + N ) 2p N +1 < < < · · (m 0 − 1)2 (m 0 − 1)2 N ρ0 (2 p − 1)2 ( m 0N−1 )2 >1
2N 2 2(m 0 − 1) + 3 1 2 · ≤ · (m 0 − 1)2 N ρ0 2 2(m 0 − 1) + 3 1 =2· · 0 2 m0 − 1 (ρ ) 1 640 ≤ 2 · 52 · 0 2 < , (ρ ) 3(ρ 0 )3 where in the third inequality, we used (A.10) to find the relation: 2m 0 − 1 ≤ ρ0 ≤ 1 N
⇒
m 0 − 1 < 2m 0 − 1 ≤ N .
Thus, we have 2p (2 p − 1)
2
D2 () <
• (Verification of (A.2)) We use α0 =
m0 8N
640 D2 () = κ0 . 3 (ρ 0 )3
and (A.14) to obtain
m0 m0 − 1 3m 0 − 1 m 0 (m 0 − 1) 2n 0 > (1− )(1+ ) = 1+ − N 4N N 4N 4N 2 m 0 (m 0 − 1) m0 − 1 m0 m0 − 2− > 1, ≥1+ =1+ 2N 4N 2 4N N
(1 − 2α0 )2 p = (1−2α0 )
where in the last inequality, we also used the following relation from (A.14): 2 ≥ 2p >
m0 − 1 +1 N
⇒
N > m 0 − 1.
Hence, we have 1 − 2α0 > • (Verification of (A.3)) We use
m0 2 !
≥
m 0 −1 2
n0 ≥
1 . 2p
and (A.11) to find
m0 + N . 2
This yields p=
2 1 m0 1 2m 0 + 1 m0 1 1 ρ0 n0 ≥ + = + · > + ρ0 · > + . N 2 2N 2 N 2(2m 0 + 1) 2 2·5 2 6
• (Verification of (A.4)) We use α0 =
m0 8N
and κ0 =
640 D2 () 3 (ρ 0 )3
to obtain
N 2D2 () 8N 2 3(ρ 0 )2 3(2m 0 + 1)2 3N 2 (ρ 0 )2 · = · < . = α0 κ0 ρ 0 m0 320 40m 0 40m 0
(A.15)
On the other hand, we have
123
S.-Y. Ha et al. 2m 0 −1 N
m0 + 1 − 8N m0 − 1 + N − m0 1 + 1 − 8N 2N 8N + 15m 0 − 8 m 0 − 1 + N = − 16N − m 0 2N m 0 (15N + m 0 − 1) . = 2N (16N − m 0 )
ρ 0 + (1 − α0 ) n 0 − 1 − ≥ 1 + (1 − α0 ) N
(A.16)
Thus, it follows from (A.15), (A.16) and m 0 ≥ 2 that we have 0 / ρ0 + (1 − α0 ) n 0 − 1 1 N 2D2 () − − · 1 + (1 − α0 ) N N α0 κ0 ρ 0 m 0 (15N + m 0 − 1) 3(2m 0 + 1)2 − 2N (16N − m 0 ) 40N m 0 2 12N (9m 0 − 16m 0 − 4) + 32m 30 − 8m 20 + 3m 0 = 40m 0 N (16N − m 0 ) 12N (m 0 − 2)(9m 0 + 2) + 8m 20 (4m 0 − 1) + 3m 0 = 40m 0 N (16N − m 0 ) > 0. >
(A.17)
This yields ρ0 + (1 − α0 ) 1 − 1 + (1 − α0 ) N
/
N 2D2 () · α0 K 0 ρ0
0 >
n0 − 1 . N
The role of the coefficient 640 3 appears in the penultimate term in (A.17), and the reason for this choice is the same as Case B.
Appendix B: Linear Instability of the Bipolar State In this section, we discuss the linear stability of the bipolar state for the Lohe model with identical oscillators: x˙i = xi +
N κ xi , xk xk − xi , i = 1, . . . , N . N xi , xi
(B.1)
k=1
By replacing the variables xi by e−t xi , we may assume = 0, so that x˙i =
N κ xi , xk xi , i = 1, . . . , N . xk − N xi , xi
(B.2)
k=1
By the O(d)-symmetry of the above model, i.e., by replacing yi by Ayi for some A ∈ O(d), and by permuting the indices, we may assume the bipolar state (y1 , . . . , y N ) in question consists of the following: (0, . . . , 0, 1), 1 ≤ i ≤ r, yi = (0, . . . , 0, −1), r < i ≤ N ,
123
On the Relaxation Dynamics of Lohe Oscillators...
where r = 0, . . . , N describes the number of oscillators on the north pole. We consider a perturbation {xi } of {yi }: ⎧ 6 ⎨(x 1 , . . . , x d , 1 − (x 1 )2 − . . . − (x d )2 ), 1 ≤ i ≤ r , i i i i 6 xi = ⎩(x 1 , . . . , x d , − 1 − (x 1 )2 . . . − (x d )2 ), r < i ≤ N , i
i
i
i
√ where |xia | " 1/ d for i = 1, . . . , N , a = 0, . . . , d − 1. Discarding quadratic terms, we can see that 1 + O(x2 ), 1 ≤ i, j ≤ r or r < i, j ≤ N , xi , x j = −1 + O(x2 ), 1 ≤ i ≤ r < j ≤ N or 1 ≤ j ≤ r < i ≤ N , N d−1 a 2 where x2 := i=1 a=0 (x i ) . Hence, if we project (B.2) onto the first d variables, we obtain the system of equations
x˙ia =
N κ a κ(2r − N ) a xk − xi + O(x3 ), i = 1, . . . , r , a = 0, . . . , d − 1. (B.3) N N k=1
x˙ia =
N κ a κ(N − 2r ) a xk − xi + O(x3 ), i = r + 1, . . . , N , a = 0, . . . , d − 1. N N k=1
(B.4) Note that we do not need the (d + 1)th component, since it is determined by the first d components. Linearizing (B.3), (B.4) yields z˙ ia =
N κ a κ(2r − N ) a z i , i = 1, . . . , r , a = 0, . . . , d − 1. zk − N N
(B.5)
N κ a κ(N − 2r ) a z i , i = r + 1, . . . , N , a = 0, . . . , d − 1. zk − N N
(B.6)
k=1
z˙ ia =
k=1
Then our result on linear stability is as follows: Theorem B.1 Let (B.5), (B.6) be the linear approximation to (2.7) at the bipolar state (y1 , . . . , y N ). (1) When r = 0 or r = N , i.e., near the completely synchronized state, the linear system (B.5), (B.6) is diagonalizable with eigenvalue 0 with multiplicity d, and −1 with multiplicity d(N − 1). Hence the linear stability is indeterminate. (2) When 1 ≤ r ≤ N − 1, i.e., near a bipolar state that is not the completely synchronized state, the linear system (B.5), (B.6) is diagonalizable with eigenvalue 0 with multiplicity d, 1 with multiplicity d, 2r N−N with multiplicity d(N − r − 1), N −2r with multiplicity N d(r − 1); hence it is linearly unstable. Remark B.1 In both cases, 0 is an eigenvalue with multiplicity at least d. This follows from the fact that (2.7) possesses an O(d)-symmetry, and from the fact that the orbit of the corresponding action of O(d) on the bipolar states in (S d ) N has dimension d. Proof This system is just the linear ODE z = κz,
123
S.-Y. Ha et al.
where zT = (z 10 , . . . , z 0N , . . . , z 1d−1 , . . . , z d−1 N ), and is the block matrix = diag(0 , . . . , 0 ). d copies
Here, 0 is defined as 1 2r − N N − 2r A− IN − 2 B N N N where A is the N × N -matrix all of whose entries are 1, and B is the N × N -matrix in block form I 0 B= r 0 0 0 :=
where Ir is the r ×r unit matrix, and the 0’s represent zero matrices of the appropriate dimension. According to Corollary C.1 in the next section, is diagonalizable, with eigenvalues ⎧ ⎪ if r = 0, N , ⎨0 with multiplicity d, − 1 with multiplicity d(N − 1) 0 with multiplicity d, 1 with multiplicity d, ⎪ ⎩ 2r −N N −2r N with multiplicity d(N − r − 1), N with multiplicity d(r − 1) if 1 ≤r ≤ N −1.
Appendix C: Calculation of Some Eigenvalues Lemma C.1 Let k be a field of characteristic zero. Let N ∈ N , m ∈ Z such that 0 ≤ r ≤ N . Let A be the N × N -matrix all of whose entries are 1 ∈ k, and let B be the N × N -matrix in the block form I 0 B= r 0 0 where Ir is the r ×r unit matrix over k, and the 0’s represent zero matrices of the appropriate dimension. Then for any μ ∈ k, the matrix A − μB has characteristic polynomial ⎧ N −1 (t − N ) ⎪ if r = 0, ⎨t φ(t) = t N −r −1 (t + μ)r −1 (t 2 + (μ − N )t + μ(r − N )) if 1 ≤ r ≤ N − 1, (C.1) ⎪ ⎩ if r = N , (t + μ) N −1 (t + (μ − N )) over k. Proof • Case A (r = 0): Then B = 0, so we are searching for the characteristic polynomial of A. If N = 1, then it is immediate that the characteristic polynomial is φ(t) = t − 1, as asserted above. So we assume N ≥ 2. Then it is easy to verify that the relation A2 − N A = 0, holds,as well as the relations A = 0,
123
A − N I N = 0.
On the Relaxation Dynamics of Lohe Oscillators...
Hence the minimal polynomial m(t) of A is m(t) = t 2 − N t = t(t − N ), and since the minimal polynomial and the characteristic polynomial must share the same irreducible factors, we conclude that φ(t) must be of the form φ(t) = t N −a (t − N )a , where a ∈ Z is an integer 1 ≤ a ≤ N − 1. This allows us to calculate, using the binomial theorem, (coefficient of t N −1 in φ(t)) = −a N , but we can also directly calculate (coefficient of t N −1 in φ(t)) = −tr(A) = −N . Hence we must have a = 1, i.e., φ(t) = t N −1 (t − N ), as asserted. • Case B (r = N ): Then B = I N , the N × N identity matrix, so φ(t) = det(t I N − (A − μB)) = det((t + μ)I N − A) = (t + μ) N −1 (t + μ − N ), by using the result of Case A. This agrees with our assertion (C.1). • Case C (1 ≤ r ≤ N − 1): Without loss of generality, we may consider μ to be a variable over k, i.e., that the field of rational functions k(μ) is a transcendental extension of k, and consider A − μB as a matrix over k(μ). Then we must verify the relation (C.1) over the polynomial ring k(μ)[t]. First, it is easy to verify the relations A2 = N A,
B 2 = B,
AB A = r A.
(C.2)
Using these relations, we proceed to verify the following by direct expansion: (A − μB)2 = N A + μ2 B − μAB − μB A, (A − μB)3 = (N 2 − μr )A − μ3 B + (μ2 − μN )AB + (μ2 − μN )B A + μ2 B AB, (A − μB)4 = (N 3 − 2N μr + r μ2 )A + μ4 B + (−μ3 + μ2 N − μN 2 + μ2 r )AB + (−μ3 + μ2 N − μN 2 + μ2 r )B A + (−2μ3 + μ2 N )B AB. This permits us to verify the relation (A − μB)4 + (2μ − N )(A − μB)3 + (μ2 − 2μN + μr )(A − μB)2 − μ2 (N − r )(A − μB) = 0. (The rationale for the preceding computation is that, the relations (C.2) force the powers of A − μB to be linear combinations of the five terms A, B, AB, B A, BAB only, and hence that five powers of A − μB suffice to give a linear relation. Luckily, we needed only four.) Hence the minimal polynomial m(t) of A − μB over k(μ) is a divisor of t 4 + (2μ − N )t 3 + (μ2 − 2μN + μr )t 2 − μ2 (N − r )t = t(t + μ)(t 2 + (μ − N )t + μ(r − N )). But t 2 + (μ − N )t + μ(r − N ) is irreducible over k(μ) if and only if it factors into linear factors, that is to say, if its discriminant (as a quadratic polynomial in t) (μ − N )2 − 4μ(r − N ) = μ2 − 2(2r − N )μ + N 2
123
S.-Y. Ha et al.
is a square in k(μ), which holds if and only if its discriminant (as a quadratic polynomial in μ) 4(2r − N )2 − 4N 2 = 16r (r − N ) = 0 is zero in k. Since k has characteristic zero, this is equivalent to r = 0 or r = N in Z. These cases were already excluded. Hence t 2 + (μ − N )t + μ(r − N ) is irreducible over k(μ). Since m(t) and φ(t) share the same irreducible factors, we conclude that φ(t) = t N −a−2b (t + μ)a (t 2 + (μ − N )t + μ(r − N ))b , where a, b are nonnegative integers such that a + 2b ≤ N . As in Case A, we use the binomial theorem to calculate (coefficient of t N −1 in φ(t)) = aμ + b(μ − N ), and compare this with (coefficient of t N −1 in φ(t)) = −tr(A) = r μ − N . Thus aμ + b(μ − N ) = r μ − N holds in k(μ), and since k has characteristic zero, we must have a = r − 1, b = 1 in Z, i.e., φ(t) = t N −r −1 (t + μ)r −1 (t 2 + (μ − N )t + μ(r − N )),
as asserted.
Remark C.1 We will only be interested in the case μ = 4r − 2N . The reason for introducing μ as a variable is that, in the case 1 ≤ r ≤ N − 1, the calculation of the characteristic polynomial becomes easier once we consider μ as a transcendental element over k. Corollary C.1 The matrix 1 2r − N N − 2r A− IN − 2 B N N N over R is diagonalizable, with eigenvalues ⎧ ⎪ if r = 0, N , ⎨0 with multiplicity 1, − 1 with multiplicity N − 1 0 with multiplicity 1, 1 with multiplicity 1, ⎪ ⎩ 2r −N N −2r N with multiplicity N − r − 1, N with multiplicity r − 1 if 1 ≤ r ≤ N − 1. 0 :=
Proof Diagonalizability follows from the symmetry of 0 and the spectral theorem for real symmetric matrices. Let μ = 2(2r − N ). Then we can see that t 2 + (μ − N )t + μ(r − N ) = t 2 + (4r − 3N )t + 2(2r − N )(r − N ) = (t + 2r − N )(t + 2r − 2N ), so (C.1) becomes
123
On the Relaxation Dynamics of Lohe Oscillators...
⎧ N −1 (t − N ) ⎪ if r = 0, ⎨t φ(t) = t N −r −1 (t + 4r − 2N )r −1 (t + 2r − N )(t + 2r − 2N ) if 1 ≤r ≤ N −1, (C.3) ⎪ ⎩ if r = N , (t + 2N ) N −1 (t + N ) (we used μ = 2N when r = N ). Thus, the desired characteristic polynomial is 1 2r − N N − 2r det t I N − A− IN − 2 B N N N 1 = N det [N t I N − (A − (N − 2r )I N − 2(2r − N )B)] N 1 = N det [(N t + N − 2r )I N − (A − μB)] N 1 = N φ(N t + N − 2r ) N ⎧ 1 (N t + N − 2 · 0) N −1 (N t + N − 2 · 0 − N ) if r = 0, ⎪ ⎪ NN ⎪ ⎨ 1 (N t + N − 2r ) N −r −1 ((N t + N − 2r ) + (4r − 2N ))r −1 = NN ⎪ ·((N t + N − 2r ) + 2r − N )((N t + N − 2r ) + (2r − 2N )) if 1 ≤ r ≤ N − 1, ⎪ ⎪ ⎩ 1 (N t + N − 2 · N + 2N ) N −1 ((N t + N − 2 · N ) + N ) if r = N , NN ⎧ N −1 ⎪ if r = 0, ⎨t(t + 1) 2r −N r −1 N −r −1 = t(t − 1)(t + N −2r ((t + N )) if 1 ≤ r ≤ N − 1, N ) ⎪ ⎩ if r = N . t(t + 1) N −1 This completes the proof. (The diagonalizability of 0 , which is equivalent to the diagonalizability of A − μB, can also be shown as follows. If r = 0 or N , we use the diagonalizability of A. If 1 ≤ r ≤ N − 1, we use the fact that the minimal polynomial of A − μB divides t(t + μ)(t 2 + (μ − N )t + μ(r − N )) = t(t + 4r − 2N )(t + 2r − N )(t + 2r − 2N ), which consists of distinct linear factors, and hence the minimal polynomial of A −μB must factor into distinct linear factors.)
References 1. Acebron, J.A., Bonilla, L.L., Pérez Vicente, C.J.P., Ritort, F., Spigler, R.: The Kuramoto model: A simple paradigm for synchronization phenomena. Rev. Mod. Phys. 77, 137–185 (2005) 2. Aeyels, D., Rogge, J.: Stability of phase locking and existence of frequency in networks of globally coupled oscillators. Prog. Theor. Phys. 112, 921–941 (2004) 3. Benedetto, D., Caglioti, E., Montemagno, U.: On the complete phase synchronization for the Kuramoto model in the mean-field limit. Commun. Math. Sci. 13, 1775–1786 (2015) 4. Buck, J., Buck, E.: Biology of synchronous flashing of fireflies. Nature 211, 562 (1966) 5. Chi, D., Choi, S.-H., Ha, S.-Y.: Emergent behaviors of a holonomic particle system on a sphere. J. Math. Phys. 55, 052703 (2014) 6. Choi, S.-H., Ha, S.-Y.: Complete entrainment of Lohe oscillators under attractive and repulsive couplings. SIAM J. App. Dyn. 13, 1417–1441 (2013) 7. Choi, S.-H., Ha, S.-Y.: Quantum synchronization of the Schödinger–Lohe model. J. Phys. A 47, 355104 (2014) 8. Choi, S.-H., Ha, S.-Y.: Emergent behaviors of quantum Lohe oscillators with all-to-all couplings. J. Nonlinear Sci. 25, 1257–1283 (2015) 9. Choi, S.-H., Ha, S.-Y.: Large-time dynamics of the asymptotic Lohe model with a small-time delay. J. Phys. A 48, 425101 (2015)
123
S.-Y. Ha et al. 10. Choi, S.-H., Ha, S.-Y.: Time-delayed interactions and synchronization of identical Lohe oscillators. Q. Appl. Math. 74, 297–319 (2016) 11. Choi, Y., Ha, S.-Y., Jung, S., Kim, Y.: Asymptotic formation and orbital stability of phase-locked states for the Kuramoto model. Physica D 241, 735–754 (2012) 12. Chopra, N., Spong, M.W.: On exponential synchronization of Kuramoto oscillators. IEEE Trans. Autom. Control 54, 353–357 (2009) 13. Dong, J.-G., Xue, X.: Synchronization analysis of Kuramoto oscillators. Commun. Math. Sci. 11, 465–480 (2013) 14. Dörfler, F., Bullo, F.: On the critical coupling for Kuramoto oscillators. SIAM J. Appl. Dyn. Syst. 10, 1070–1099 (2011) 15. Dörfler, F., Bullo, F.: Exploring synchronization in complex oscillator networks. In: IEEE 51st Annual Conference on Decision and Control (CDC), pp. 7157–7170 (2012) 16. Dörfler, F., Bullo, F.: Synchronization in complex networks of phase oscillators: a survey. Automatica 50, 1539–1564 (2014) 17. Ha, S.-Y., Ryoo, S.: On the emergence and orbital stability of phase-locked states for the Lohe model. J. Stat. Phys. 163, 411–439 (2016) 18. Ha, S.-Y., Li, Z., Xue, X.: Formation of phase-locked states in a population of locally interacting Kuramoto oscillators. J. Differ. Equ. 255, 3053–3070 (2013) 19. Ha, S.-Y., Kim, H.W., Ryoo, S.W.: Emergence of phase-locked states for the Kuramoto model in a large coupling regime. Commun. Math. Sci. 14, 1073–1091 (2016) 20. Ha, S.-Y., Ko, D., Ryoo, S.W.: Emergent dynamics of a generalized lohe model on some class of Lie groups. J. Stat. Phys. 168, 171–207 (2017) 21. Jadbabaie, A., Motee, N., Barahona M.: On the stability of the Kuramoto model of coupled nonlinear oscillators. In: Proceedings of the American Control Conference, pp. 4296–4301 (2004) 22. Kuramoto, Y.: International Symposium on Mathematical Problems in Theoretical Physics. Lecture Notes Theoretical Physics, vol. 30, p. 420 (1975) 23. Kuramoto, Y.: Chemical Oscillations, Waves and Turbulence. Springer, Berlin (1984) 24. Lohe, M.A.: Non-Abelian Kuramoto model and synchronization. J. Phys. A 42, 395101–395126 (2009) 25. Lohe, M.A.: Quantum synchronization over quantum networks. J. Phys. A 43, 465301 (2010) 26. Lojasiewicz, A.: Une propriété topologique des sous-ensembles analytiques réels, pp. 87–89. Les Équations aux Dérivées Partielles, Centre National de la Recherche Scientifique, Paris (1963) 27. Mirollo, R., Strogatz, S.H.: Stability of incoherence in a population of coupled oscillators. J. Stat. Phys. 63, 613–635 (1991) 28. Mirollo, R., Strogatz, S.H.: The spectrum of the locked state for the Kuramoto model of coupled oscillators. Physica D 205, 249–266 (2005) 29. Mirollo, R., Strogatz, S.H.: The spectrum of the partially locked state for the Kuramoto model. J. Nonlinear Sci. 17, 309–347 (2007) 30. Olfati-Saber, R.: Swarms on sphere: a programmable swarm with synchronous behaviors like oscillator networks. In: Proceedings of the 45th IEEE conference on Decision and Control, pp. 5060–5066 (2006) 31. Peskin, C.S.: Mathematical aspects of heart physiology. Courant Institute of Mathematical Sciences, New York (1975) 32. Pikovsky, A., Rosenblum, M., Kurths, J.: Synchronization: A Universal Concept in Nonlinear Sciences. Cambridge University Press, Cambridge (2001) 33. Strogatz, S.H.: From Kuramoto to Crawford: exploring the onset of synchronization in populations of coupled oscillators. Physica D 143, 1–20 (2000) 34. Tu, L.W.: An Introduction to Manifolds. Springer, New York (2011) 35. van Hemmen, J.L., Wreszinski, W.F.: Lyapunov function for the Kuramoto model of nonlinearly coupled oscillators. J. Stat. Phys. 72, 145–166 (1993) 36. Verwoerd, M., Mason, O.: Global phase-locking in finite populations of phase-coupled oscillators. SIAM J. Appl. Dyn. Syst. 7, 134–160 (2008) 37. Verwoerd, M., Mason, O.: On computing the critical coupling coefficient for the Kuramoto model on a complete bipartite graph. SIAM J. Appl. Dyn. Syst. 8, 417–453 (2009) 38. Winfree, A.T.: Biological rhythms and the behavior of populations of coupled oscillators. J. Theor. Biol. 16, 15–42 (1967) 39. Winfree, A.T.: The Geometry of Biological Time. Springer, New York (1980)
123