Research article Special Issues

Convergence for global curve diffusion flows

  • In this note we establish exponentially fast smooth convergence for global curve diffusion flows, and discuss open problems relating embeddedness to global existence (Giga's conjecture) and the shape of Type I singularities (Chou's conjecture).

    Citation: Glen Wheeler. Convergence for global curve diffusion flows[J]. Mathematics in Engineering, 2022, 4(1): 1-13. doi: 10.3934/mine.2022001

    Related Papers:

    [1] Takeyuki Nagasawa, Kohei Nakamura . Asymptotic analysis for non-local curvature flows for plane curves with a general rotation number. Mathematics in Engineering, 2021, 3(6): 1-26. doi: 10.3934/mine.2021047
    [2] Serena Della Corte, Antonia Diana, Carlo Mantegazza . Global existence and stability for the modified Mullins–Sekerka and surface diffusion flow. Mathematics in Engineering, 2022, 4(6): 1-104. doi: 10.3934/mine.2022054
    [3] Anoumou Attiogbe, Mouhamed Moustapha Fall, El Hadji Abdoulaye Thiam . Nonlocal diffusion of smooth sets. Mathematics in Engineering, 2022, 4(2): 1-22. doi: 10.3934/mine.2022009
    [4] Gabriele Grillo, Giulia Meglioli, Fabio Punzo . Global existence for reaction-diffusion evolution equations driven by the p-Laplacian on manifolds. Mathematics in Engineering, 2023, 5(3): 1-38. doi: 10.3934/mine.2023070
    [5] Matteo Novaga, Marco Pozzetta . Connected surfaces with boundary minimizing the Willmore energy. Mathematics in Engineering, 2020, 2(3): 527-556. doi: 10.3934/mine.2020024
    [6] Antonio Vitolo . Singular elliptic equations with directional diffusion. Mathematics in Engineering, 2021, 3(3): 1-16. doi: 10.3934/mine.2021027
    [7] Zaffar Mehdi Dar, M. Arrutselvi, Chandru Muthusamy, Sundararajan Natarajan, Gianmarco Manzini . Virtual element approximations of the time-fractional nonlinear convection-diffusion equation on polygonal meshes. Mathematics in Engineering, 2025, 7(2): 96-129. doi: 10.3934/mine.2025005
    [8] Tatsuya Miura . Polar tangential angles and free elasticae. Mathematics in Engineering, 2021, 3(4): 1-12. doi: 10.3934/mine.2021034
    [9] Arthur. J. Vromans, Fons van de Ven, Adrian Muntean . Homogenization of a pseudo-parabolic system via a spatial-temporal decoupling: Upscaling and corrector estimates for perforated domains. Mathematics in Engineering, 2019, 1(3): 548-582. doi: 10.3934/mine.2019.3.548
    [10] Filippo Gazzola, Elsa M. Marchini . The moon lander optimal control problem revisited. Mathematics in Engineering, 2021, 3(5): 1-14. doi: 10.3934/mine.2021040
  • In this note we establish exponentially fast smooth convergence for global curve diffusion flows, and discuss open problems relating embeddedness to global existence (Giga's conjecture) and the shape of Type I singularities (Chou's conjecture).



    The curve diffusion flow is the one-parameter family of immersed curves γ:S1×[0,T)R2 with normal velocity equal to gradH1(L(γ)), that is

    tγ=kssν. (CD)

    Here, s denotes arclength, k is the scalar curvature, and ν is the normal vector.

    In [10], the author proves that if the isoperimetric ratio and L2-normalised oscillation of curvature are close to their value on any circle (with an explicit estimate on the constant) then the flow exists globally and converges to a circle. However the rate of convergence was not established.

    In this note we prove a general convergence principle that can be summarised as:

    T=γ converges exponentially fast to an ω-circle in the smooth topology.

    An ω-circle is an immersion γ:S1R2 with winding number of the unit normal around the origin equal to ω. Note that also ω=12πkds.

    Theorem 1.1. Suppose γ:S×[0,)R2 is a global curve diffusion flow with smooth initial data γ0 that has winding number ω. Then γ converges exponentially fast to a round ω-circle in the smooth topology, with explicit estimates (here mN)

    ksm22cme¯k40t,

    with

    D(t)D0e4¯k40tandKosc(t)c0e2¯k40t.

    The constants cm depend only on the initial data γ0.

    The quantities D and K osc  are the isoperimetric deficit and the L2-normalised oscillation of curvature:

    D=L24ωπA and Kosc=Lγ(k¯k)2ds

    where L and A are the length and signed enclosed area of γ, and k, ¯k are the curvature scalar and the average of k respectively. Section 2 contains the proof of Theorem 1.1.

    Remark 1.2. Note that no smallness condition is required.

    Remark 1.3. The regularity hypothesis that we impose (initial data of class C) is not optimal. There are in the literature well-posedness (and regularisation) results for curve and surface diffusion with initial data of class C1,α [7] as well as initial data in a Besov space [4]. It remains an interesting problem to determine the weakest possible conditions on the initial data that allow generation of a unique, regularising solution to the flow.

    Remark 1.4. Theorem 1.1 is stated for flows of immersed multiply-covered circles which makes it applicable to the case considered by Miura-Okabe [8]. However the curve diffusion flow starting from a multiply-covered circle is not expected to be stable in general – perturbations that unbalance the length of different leaves are should lead to finite-time singualrities.* Miura-Okabe use a rotational symmetry assumption so that only perturbations that preserve the balance of the different leaves are studied, and the power of this hypothesis can be seen in the isoperimetric inequality for immersed rotationally symmetric curves that they establish. That is, preservation of the smallness of the isoperimetric ratio (see Lemma 2.4) is proved in [8] for rotationally symmetric curves without assuming global existence. This is the decisive new ingredient that allows their result on the curve diffusion flow to go through.

    * Rigorously establishing this is an interesting open problem.

    Yoshikazu Giga famously conjectured around ten years ago that:

    Conjecture (Giga's Conjecture). Suppose γ:S×[0,T)R2 is a curve diffusion flow with smooth initial data γ0 that has the property:

    γ(,t)  is  an  embedding  for  each  t[0,T).

    Then T=.

    If this conjecture holds, our Theorem 1.1 implies: an embedded curve diffusion flow converges smoothly and exponentially fast to a round circle.

    Note that in the case of the curve diffusion flow in homotopy classes outside ω-circles, shrinkers are expected and one explicit example is known (the Lemniscate of Bernoulli, see [5]). Indeed, Chou [2] suggests that parabolic rescalings of Type I singularities (a definition also due to Chou) are self-similar solutions to the curve diffusion flow:

    Conjecture (Chou’s Conjecture). Suppose γ:S×[0,T)R2 is a curve diffusion flow with T< that satisfies the estimate

    k22(t)C(Tt)1/4, (1.1)

    for some CR, and t[0,T).

    Then a parabolic rescaling (we assume the centre of mass of γ is the origin)

    η(s,τ)=(Tt)14γ(s,t),τη=kηssνη+14η

    where τ=log(Tt) yields a self similar solution η to the curve diffusion flow, that is, η converges as t to a solution η of

    η,νη=4kηss. (1.2)

    Curve diffusion flows satisfying (1.1) are said to be Type I. As suggested by Chou [2], this conjecture can be approached by studying all solutions to the shrinker Eq (1.2), and classifying all such solutions is an interesting problem. Since area is constant along the curve diffusion flow, and shrinkers must eventually enclose zero area, all shrinkers have zero signed enclosed area. This means that all solutions to (1.2) are non-embedded (and have zero signed enclosed area).

    The maximal time of smooth existence for the curve diffusion flow with ω=0 is always finite. This is because along the flow the curvature scalar always has a zero, and so the Wirtinger inequality implies that the flow can not exist past time T=L4016π4. This is not a sharp estimate, as equality in the estimates used by the proof implies k=const, and there is no such curve that also has ω=0. A reasonable conjecture is that the maximal existence time for a curve diffusion flow with winding number zero and initial length L0 is bounded by that of the Lemniscate of Bernoulli starting with length L0.

    As far as we know the connection between non-convexity and maximal existence time was observed first by Chou [2]. It was later used in [10] to estimate the total waiting time before a curve diffusion flow with small L2-oscillation of curvature and isoperimetric ratio becomes uniformly convex. Theorem 1.1 here combined with the waiting time estimate [10, Proposition 1.5] yields an estimate for the total waiting time before any global curve diffusion flow becomes uniformly convex.

    We complete the proof in a sequence of lemmata. The hypotheses of Theorem 1.1 are not restated, but assumed throughout.

    Lemma 2.1. There exists a subsequence of times tj such that ks22(tj)0.

    Proof. Note that

    L=ks22.

    Integration and the monotonicity of length implies the result.

    As far as we are aware, Lemma 2.1 was first observed by Chou [2].

    Now we observe that the signed enclosed area of the curve under the flow can not be non-positive.

    Lemma 2.2. The signed enclosed area A of γ is strictly positive.

    Proof. Our convention is that the area of an ω-circle is positive (and equal to ωπr2, where r is its radius). If the area is not strictly positive, then the flow can exist for at most a finite amount of time. Since area is constant under the flow, if it is non-positive at any time, it must be so for all time. Then, since A(t)<0 implies that k(st,t)=0 for some point st, we use the Wirtinger and Poincaré inequalities to conclude

    L(t)π2L2k224π4L3

    which implies that the flow can exist at most for a finite time (in fact TL016π4). This is a contradiction.

    Now set

    D=L24ωπA.

    Recall that D is the isoperimetric defect. Note that D0 with equality if and only if γ is an ω-circle. Since length is monotone decreasing and signed enclosed area is constant in time, we immediately have:

    Lemma 2.3. The isoperimetric defect satisfies

    D(t)0.

    By combining Lemma 2.1, Lemma 2.2 and Lemma 2.3 we find the following.

    Corollary 2.4. Let ε>0 be arbitrary. There exists a tj such that for all t>tj,

    I(t):=L24ωA1+ε.

    Proof. The estimate

    DL22πK12osc,Kosc=Lγ(k¯k)2ds (2.1)

    (note that ¯k=2ωπ/L is the average of k) together with the Poincaré inequality implies

    DL72232ωπ32ks2. (2.2)

    The estimate (2.1) follows using an elementary argument (see the Appendix).

    Note that length is uniformly bounded below by the measure-theoretic isoperimetric inequality

    L2(t)4πA+

    where A+ is the area of γ changed so that the orientation of every loop is positive (no segments of negative area). Clearly we always have A+A and so (using the constancy of area along the flow)

    L2(t)4πA(0)

    follows. This is not a sharp estimate for ω1 but it is enough to give a uniform lower bound on evolving area for any winding number. Note that the shrinking Lemniscare of Bernoulli has A(0)=0 and T<. Lemma 2.2 implies that A(0)>0 and so in particular length is uniformly bounded from below by a universal constant.

    Therefore the estimate (2.2) implies that D(tj)0, which implies I(tj)1. Now take j sufficiently large that I(tj)1+ε, where ε>0 is as given by hypothesis. Then Lemma 2.3 implies that I(t)1+ε for all t>0, as required.

    Lemma 2.5. There are uniform bounds on all derivatives of curvature.

    Proof. We recall the estimate (2.2) and

    KoscL34ω2π2ks22, (2.3)

    which follows from the Poincaré inequality. Each of the estimates (2.1), (2.3) are uniform (from the uniform boundedness of length).

    These estimates imply that the isoperimetric defect and K osc  converge to zero at times tj as j. Therefore for j sufficiently large, the hypotheses of [10, Proposition 3.7] are satisfied at time tj. Given the eventual smallness of the isoperimetric ratio (Corollary 2.4), the argument in [10] (see also [8]) applies to yield uniform estimates on K osc . This then implies bounds on all derivatives of curvature by the evolutionary interpolation inequalities of Dziuk-Kuwert-Schätzle [3].

    The argument is by now standard in the literature, see for instance [1,Theorem 6.4] for a recent example.

    Lemma 2.6. The isoperimetric defect decays exponentially fast:

    D(t)D0e4¯k40t.

    Proof. Since

    D=2Lγk2sds,

    the estimate (2.1) implies

    DD64ω4π4L4D64ω4π4L40.

    Therefore

    (logD)4¯k40,

    where ¯k0 is the initial average of the curvature scalar. Then

    D(t)D0e4¯k40t,

    as required.

    Remark 2.7. One may also attempt the same strategy with the isoperimetric ratio in place of the isoperimetric defect. However, the resultant sharp estimate looks like

    (I1)¯k4.

    This results in linear decay of the isoperimetric ratio – not exponential.

    Lemma 2.8. All derivatives of curvature decay exponentially fast, with the explicit estimate

    ksm22cme¯k40t.

    Here ksm=msk.

    Proof. Since the previous lemma establishes uniform bounds for all derivatives of curvature, a Fourier series argument (taught to us by Ben Andrews many years ago and used in for example [1]; we give an explanation in the appendix).

    The author recently learned that similar Fourier series arguments (and inequalities that imply the below) also appear in [9].

    KosccDL

    where c depends on the estimates for k and ks. Now the isoperimetric inequality (and constancy of area) implies that L is uniformly bounded from below, so the above implies that K osc  satisfies

    Koscc0e2¯k40t.

    Integration by parts and the uniform estimates imply that for any m we have

    ksm22cme¯k40t,

    as required.

    The exponential decay implies convergence of the position vector γ to an ω-circle with a standard argument (for example this was used by Huisken [6] in his seminal work on mean curvature flow). Briefly, this is because we then have estimates on γ by simply integrating the evolution equation in time. We may convert from arclength derivatives to arbitrary ones in a manner analogous to [3,proof of Theorem 3.1]. Note that this integration in time can not be done with only linear decay estimates.

    The author thanks his colleagues for discussions on the curve diffusion flow, especially relating to its convergence properties. In particular he would like to thank Tatsuya Miura for extensive discussion and encouraging the dissemination of this note. Without the encouragement of his peers this note would not have been written. This work was completed under the financial support of the Okinawa Institute of Science and Technology, hosted by James McCoy. He is grateful for their support and hospitality. The author would additionally like to thank the anonymous referee for their comments and suggestions that led to an improvement in this article.

    We would like to express our great gratitiude to Ben Andrews, who taught us the technique we are about to use (and many, many other things). In particular Ben demonstrated how to prove a great variety of inequalities for curves, and how they can be applied to many different kinds of curvature flow. This occurred while the author was his guest at the Mathematical Sciences Center at Tsinghua University in Beijing, 2011. These results should be thought of as due to Ben Andrews and not the author.

    The author declares no conflict of interest.

    Both estimates used in this article are contained in the following statement.

    Lemma A.1. Let γ:SR2 be a smooth immersed curve. Then

    2πL4D2KosccD12L. (A.1)

    where c is such that

    L3ks22+L5/2k36K12oscc.

    Note that when the upper estimate is used, curvature and its derivative(s) are uniformly bounded, so the hypothesis involving c is satisfied.

    Let us prove first the lower estimate. To begin, we translate the curve γ to ˜γ so that

    ˜γds=0

    and the estimate

    |˜γ|L2π

    holds. The lower estimate in (6.1) is invariant under translation (this is true of both sides), so if we can prove the estimate for ˜γ it will be true for γ also.

    Recall that

    A=12˜γ,νds.

    Then integration by parts using ˜γ,˜γss=1+k˜γ,ν implies

    D=L+2ωπ˜γ,νds=kL˜γ,ν+2ωπ˜γ,νds=(kL+2ωπ)˜γ,νds.

    Since ¯k=1Lkds=2ωπL, we have ¯kk=1L(kL+2ωπ). Therefore

    D|kL+2ωπ|L2πdsL22πk¯k1.

    The lower estimate in (A.1) follows now from Hölder's inequality.

    Remark A.2. The estimate

    |γ,ν1Lγ,νds|L

    is false in general. Take for example a circle with unit radius centred at (P,0) in the plane. Then,

    sup|γ,ν1Lγ,νds|(P+1)+2π2π=P+2.

    Since P is arbitrary, this quantity is unbounded. This estimate is used in [9], although it is not a major issue (translation invariance as used here can also be used there).

    As in [1,Proof of Theorem 6.1] we work in the complex plane. Let us identify γ(s)=(x(s),y(s)) with sx(s)+iy(s). The idea is to use a Fourier series decomposition for γ to reduce the question of proving geometric inequalities to properties of infinite series of integers. Here we have γ smooth, but in order to express γ as a convergent Fourier series we only require γW1,2. Our manipulations of the series below require significantly more regularity; for instance, the highest order one needs γW4,2.

    We write

    γ(s)=pZˆγ(p)cp(s)

    where the coefficients cp are given by

    cp(s)=1Le2iωπps/L,

    with projection ˆγ defined via

    ˆγ(p)=γγ(s)¯cp(s)ds.

    Curvature arises by differentiating γ, which is the same as differentiating the Fourier decomposition. The basic relation is:

    Lemma A.3. For q2 we have

    pZpq|ˆγ(p)|2=i(2iωπ/L)qkQq1ds=iq1Lq(2ωπ)qkQq1ds

    where Qq=(q1sγ)¯sγ. For the q=1 case we have

    pZp|ˆγ(p)|2=iL2ωπQ1ds.

    Proof. We provide a sketch with the essential steps. First, observe that for q2 we have

    Qqds=ikQq1ds. (A.2)

    This is because Qq=ikQq1+sQq1 (this equality can be proved by a straightforward induction argument). But now we use the series decomposition with orthonormality of cp to conclude

    Qqds=(q1sγ)¯sγds=(p(2iωπ/L)q1pq1cp(s)ˆγ(p))¯(p(2iωπ/L)pcp(s)ˆγ(p))ds=(2iωπ/L)qpZpq|ˆγ(p)|2.

    In the q=1 case this finishes the proof; for q2 we combine this with (A.2) to finish the proof.

    Lemma A.3 allows us to simply express the isoperimetric defect and the L2 oscillation of curvature in terms of infinite series. To see this, we consider the cases of q=1,2,3,4,5,6 in Lemma A.3. For each of computation we make the notation

    T=γ,sγ,andN=γ,ν.

    Note that sT=1+kN and sN=kT. We calculate

    (q=1)pZp|ˆγ(p)|2=iL2ωπQ1ds=iL2ωπ(x+iy)(xsiys)ds=iL2ωπT+iNds=iL2ωπiNds=LAωπ
    (q=2)pZp2|ˆγ(p)|2=i3L2(2ωπ)2kQ1ds=iL2(2ωπ)2kT+ikNds=L3(2ωπ)2
    (q=3)pZp3|ˆγ(p)|2=i4L3(2ωπ)3kQ2ds=L3(2ωπ)3ik2Q1+ksQ1ds=L3(2ωπ)3(ik2+ks)T+i(ik2+ks)Nds=L3(2ωπ)3kds=L3(2ωπ)2

    In fact, Q2=(sγ)(¯sγ)=1. We continue to calculate:

    (q=4)pZp4|ˆγ(p)|2=i5L4(2ωπ)4kQ3ds=iL4(2ωπ)4(ik2+ks)Q2ds=iL4(2ωπ)4ik2ds=L4(2ωπ)4k2ds
    (q=5)pZp5|ˆγ(p)|2=i6L5(2ωπ)5kQ4ds=L5(2ωπ)5(ik2+ks)Q3ds=L5(2ωπ)5(ik2+ks)(ik)ds=L5(2ωπ)5k3ds.
    (q=6)pZp6|ˆγ(p)|2=i7L6(2ωπ)6kQ5ds=iL6(2ωπ)6(ik2+ks)(k2+iks)ds=iL6(2ωπ)6(ik4k2ks2k2ks+ikkss)ds=L6(2ωπ)6(k4+k2s)ds.

    Although we won't need it here, one can continue to calculate more of these identities for greater values of q. They can be used to establish higher order inequalities - we include an example (inequality (A.3)). Here are the next two:

    (q=7)pZp7|ˆγ(p)|2=i8L7(2ωπ)7kQ6ds=L7(2ωπ)7(ik2+ks)(ik33kks+ikss)ds=L7(2ωπ)7k56ik3ks4k2kss3kk2s+ikksssds=L7(2ωπ)7k5+5kk2sds.
    (q=8)pZp8|ˆγ(p)|2=i9L8(2ωπ)8kQ7ds=iL8(2ωπ)8(ik2+ks)(k46ik2ks4kkss3k2s+iksss)ds=iL8(2ωπ)8ik6+10k4ks15ik2k2s10ik3kss10kkskss5k2ksss+ikks4ds=L8(2ωπ)8k6+15k2k2s+10k3kss+10ikkskss+5ik2ksss+k2ssds=L8(2ωπ)8k615k2k2s+k2ssds.

    This last identity can be used trivially to deduce the estimate

    15k2k2sdsk6+k2ssds. (A.3)

    This estimate does not seem to follow from using more common methods.

    Now we return to our aim of proving the upper estimate. First, we write key quantities in terms of infinite series.

    Lemma A.4. The equalities

    D=(2ωπ)2LpZp(p1)|ˆγ(p)|2

    and

    Kosc=(2ωπ)4L3pZp2(p21)|ˆγ(p)|2=(2ωπ)4L3pZp3(p1)|ˆγ(p)|2

    hold.

    Proof. We calculate

    D=L24ωπA=4ω2π2L(L3(2ωπ)2ALωπ).

    Using the expressions for q=2 and q=1, this shows

    D=4ω2π2L(pZ(p2p)|ˆγ(p)|2),

    as required. Now

    Kosc=L(k¯k)2ds=Lk2ds(2ωπ)2

    which, using the expressions for q=4 and q=2, gives

    Kosc=(2ωπ)4L3pZp4|ˆγ(p)|2(2ωπ)4L3pZp2|ˆγ(p)|2=(2ωπ)4L3pZ(p4p2)|ˆγ(p)|2.

    For the second term we could have used the equality for q=3 instead, which gives the last equality and finishes the proof.

    Finally, we use the Cauchy-Schwartz inequality in 2(Z) to conclude essentially the upper estimate.

    Lemma A.5. We have

    K2oscLD(ks22+L1/2k36K12osc).

    Proof. We use the Cauchy-Schwartz inequality to estimate

    Kosc=(2ωπ)4L3pZp3(p1)|ˆγ(p)|2(2ωπ)4L3(pZp(p1)|ˆγ(p)|2)12(pZp5(p1)|ˆγ(p)|2)12.

    Now Lemma A.4 implies

    Kosc(2ωπ)4L3D12L122ωπ(pZp5(p1)|ˆγ(p)|2)12.

    The expressions for q=5 and q=6 now imply

    K2osc(2ωπ)8L6DL(2ωπ)2(L6(2ωπ)6k4+k2s2ωπLk3ds)=LDk2s+k3(k¯k)dsLD(ks22+L1/2k36K12osc),

    as required.

    Now we finish the proof of the upper estimate. Since by hypothesis we have

    L3ks22+L5/2k36K12oscc,

    Lemma A.5 implies

    KosccD12L

    as required.



    [1] B. Andrews, J. McCoy, G. Wheeler, V. M. Wheeler, Closed ideal planar curves, Geom. Topol., 24 (2020), 1019-1049.
    [2] K. S. Chou, A blow-up criterion for the curve shortening flow by surface diffusion, Hokkaido Math. J., 32 (2003), 1-19.
    [3] G. Dziuk, E. Kuwert, R. Schätzle, Evolution of elastic curves in Rn: existence and computation, SIAM J. Math. Anal., 33 (2002), 1228-1245.
    [4] J. Escher, P. B. Mucha, The surface diffusion flow on rough phase spaces, Discrete Cont. Dyn. A, 26 (2010), 431-453.
    [5] M. Edwards, A. Gerhardt-Bourke, J. McCoy, G. Wheeler, V. M. Wheeler, The shrinking figure eight and other solitons for the curve diffusion flow, J. Elasticity, 119 (2015), 191-211.
    [6] G. Huisken, Flow by mean-curvature of convex surfaces into spheres, J. Differ. Geom., 20 (1984), 237-266.
    [7] J. LeCrone, Y. Z. Shao, G. Simonett, The surface diffusion and the Willmore flow for uniformly regular hypersurfaces, Discrete Cont. Dyn. S, 13 (2020), 3503-3524.
    [8] T. Miura, S. Okabe, On the isoperimetric inequality and surface diffusion flow for multiply winding curves, 2019, arXiv: 1909.08816.
    [9] T. Nagasawa, K. Nakamura, Interpolation inequalities between the deviation of curvature and the isoperimetric ratio with applications to geometric flows, Adv. Differential Equ., 24 (2019), 581-608.
    [10] G. Wheeler, On the curve diffusion flow of closed plane curves, Ann. Mat. Pur. Appl., 192 (2013), 931-950.
  • This article has been cited by:

    1. M.K. Cooper, G. Wheeler, V.-M. Wheeler, Theory and numerics for Chen's flow of curves, 2023, 362, 00220396, 1, 10.1016/j.jde.2023.02.064
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2410) PDF downloads(195) Cited by(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog