return

Celebratio Mathematica

Cathleen Morawetz

Cathleen Morawetz
and the scattering of acoustic waves

by Leslie Greengard and Tonatiuh Sánchez-Vizuet

Cath­leen Mor­awetz was a force at the Cour­ant In­sti­tute when one of us (L.G.) ar­rived as a postdoc­tor­al fel­low. It was the last year of her dir­ect­or­ship, but she made the time to wel­come all new­comers. Her gen­er­os­ity of spir­it was un­matched — she en­cour­aged young people in every dis­cip­line, and her hu­mour and en­thu­si­asm were in­fec­tious.

When she began to study the de­cay prop­er­ties of acous­tic waves after impinging on an obstacle, es­sen­tially no gen­er­al res­ults were avail­able. To un­der­stand the rel­ev­ant is­sues, let us be­gin with the for­mu­la­tion of the prob­lem in terms of the gov­ern­ing lin­ear, scal­ar wave equa­tion in \( \mathbb{R}^3 \), with a for­cing term which is turned on for a fi­nite time: \begin{equation} \label{eleven} u_{tt} (\mathbf{x}, t) = \Delta u(\mathbf{x}, t) + f(\mathbf{x}, t). \end{equation} Here, \( \Delta u \) is the Lapla­cian op­er­at­or act­ing on the scal­ar func­tion \( u(\mathbf{x}, t) \) and \( f(\mathbf{x}, t) \) is nonzero only in the fi­nite time in­ter­val \( 0 \leq t \leq T \). We as­sume that we have zero ini­tial (Cauchy) data at time \( t = 0 \): \[ u(\mathbf{x}, 0) = 0\quad\text{ and } \quad u_t (\mathbf{x}, 0) = 0. \] We also as­sume that \( f(\mathbf{x}, t) \) is a smooth, com­pactly sup­por­ted and square in­teg­rable func­tion in space-time, such as \begin{equation} \label{twelve} f(\mathbf{x},t) = W(\|\mathbf{x} - \mathbf{x}_0\|)\, W\, \biggl(\frac{2t-T}{T}\biggr), \end{equation} where \( \mathbf{x}_0 \in \mathbb{R}^3 \) and \( W(x) \) is a stand­ard \( C^{\infty} \) bump func­tion such as \[ W(x) = \begin{cases} e^{-\frac{1}{1-x^2}} & \text{for } |x| < 1;\\ 0 & \text{otherwise}. \end{cases} \] Then, it is well known that \begin{equation} \label{thirteen} u(\mathbf{x}, t) = \frac{1}{4\pi}\int_{B_{\mathbf{x}_0} (1)} \frac{f (\mathbf{x}^{\prime}, t - \|\mathbf{x} - \mathbf{x}^{\prime}\|)}{\|\mathbf{x}-\mathbf{x}^{\prime}\|}\,d\mathbf{x}^{\prime}, \end{equation} where \( B_{\mathbf{x}_0} (1) \) de­notes the unit ball centered at \( \mathbf{x}_0 \). From this for­mula it is clear that at any point \( \mathbf{x} \) in space, the solu­tion first be­comes nonzero at \( t = d_{\min} \) , where \( d_{\min} \) is the dis­tance from \( \mathbf{x} \) to the closest point in \( B_{\mathbf{x}_0} (1) \). It van­ishes identic­ally at \( \mathbf{x} \) as soon as \( t > T + d_{\max} \), where \( d_{\max} \) is the dis­tance from \( \mathbf{x} \) to the farthest point \( B_{\mathbf{x}_0} (1) \).

Figure 3. The evolution of an acoustic wave impinging upon a non-star-shaped object, \( \pi \), at times \( t = 1, 4, 8, 13, 19, 26 \) implemented using a high-order integral equation solver. A video of the simulation can be found below.
Editor’s note: The stills in the above figure and the video simulation from which they are taken are courtesy of Tonatiuh Sánchez-Vizuet, and are based on the numerical algorithms for time-domain integral equations developed in [e3], [e4], and [e6].

Sup­pose now that, rather than propagat­ing in free-space, the out­go­ing spher­ic­al wave­front em­an­at­ing from \( \mathbf{x}_0 \) hits an ob­ject as in Fig­ure 3. That is, we as­sume there is a “sound-soft,” smooth, bounded obstacle \( \Omega \) with bound­ary \( \partial\Omega \), which is at some dis­tance from \( B_{\mathbf{x}_0} (1) \), so that \[ \Omega \cap B_{\mathbf{x}_0} (1)= \emptyset .\] Then, us­ing the lan­guage of scat­ter­ing the­ory, the total acous­tic field is giv­en by \[ u(\mathbf{x}, t) + u^{\operatorname{scat}} (\mathbf{x}, t) ,\] where the scattered field sat­is­fies the ho­mo­gen­eous wave equa­tion \[ u^{\operatorname{scat}}_{tt} (\mathbf{x}, t) - \Delta u^{\operatorname{scat}} (\mathbf{x}, t) = 0 \] for \( t > 0 \), with ini­tial data \[ u^{\operatorname{scat}} (\mathbf{x}, 0) = 0 \] and \[ u^{\operatorname{scat}}_t (\mathbf{x}, 0) = 0 \] and Di­rich­let bound­ary con­di­tions \[ u^{\operatorname{scat}} (\mathbf{x}, t) = -u(\mathbf{x}, t) \] for \( x \in \partial\Omega \).

Let \( \mathbf{y} \) de­note some fixed point away from both the ball \( B_{\mathbf{x}_0} (1) \) and the obstacle \( \Omega \). The ques­tion is: can one prove that the scattered field de­cays at \( \mathbf{y} \), and if so, at what rate? Very little pro­gress had been made on this ques­tion un­til 1959, when Wil­cox pub­lished a short note show­ing that in the case of a spher­ic­al obstacle, an ex­act solu­tion could be ex­pressed in terms of spher­ic­al har­mon­ics. From this, Wil­cox was able to con­clude that the solu­tion de­cays ex­po­nen­tially fast. While an im­port­ant step, his res­ult yiel­ded no sug­ges­tion as to how to pro­ceed in the gen­er­al case.

Figure 4. The evolution of an acoustic wave being scattered off a star-shaped obstacle at the same sequence of times as in Figure 3. In the final panel, the scattered wave has almost completely left the simulation region, in contrast with the final panel of Figure 3.
Editor’s note: This sequence of images and the video from which they are taken are courtesy of Tonatiuh Sánchez-Vizuet (see the video link in the caption for Figure 3)

In 1961, Mor­awetz [1] made a crit­ic­al step for­ward. She showed that if the re­flect­ing obstacle is star-shaped, then the solu­tion to the wave equa­tion de­cays like \( t^{1/2} \). A re­gion \( \Omega \) is said to be star-shaped if there ex­ists a point \( \mathbf{p} \in \Omega \), such that for all \( \mathbf{x} \in \Omega \), the line seg­ment from \( \mathbf{p} \) to \( \mathbf{x} \) is con­tained in \( \Omega \). The ob­ject in Fig­ure 3, for ex­ample, is not star-shaped, while the ob­ject in Fig­ure 4 is. It is per­haps sur­pris­ing that for non-star-shaped obstacles, very little is un­der­stood to the present day.

The qual­it­at­ive dif­fer­ence in the be­ha­vi­or of waves re­flect­ing from obstacles that are not star-shaped and those that are is il­lus­trated in Fig­ures 3 and 4. In the first three pan­els of each fig­ure, as the in­com­ing wave hits the ob­ject, the scattered wave is clearly vis­ible, with en­ergy propagat­ing out­wards in all dir­ec­tions. In the next three pan­els, more of the en­ergy is car­ried away. In Fig­ure 3, some of the en­ergy re­mains be­hind for quite some time, and in the last pan­el a sig­ni­fic­ant amount of en­ergy has fo­cused in a small neigh­bor­hood. In Fig­ure 4, the en­ergy has propag­ated out­ward without sig­ni­fic­ant con­cen­tra­tion and ap­pears to de­cay much more rap­idly.

Re­mark.  The sim­u­la­tions in these fig­ures are ac­tu­ally for the two-di­men­sion­al wave equa­tion, with an in­com­ing plane wave of the form \[ u^{\operatorname{inc}} (\mathbf{x}, t) = \chi (s/\alpha) \sin^3 (s/\alpha), \quad s := \mathbf{x} \cdot \mathbf{d} - t. \] The unit vec­tor \( \mathbf{d} \) points in the dir­ec­tion in which the wave propag­ates, \( \chi(\,\cdot\,) \) is a smooth ap­prox­im­a­tion to the char­ac­ter­ist­ic func­tion of the in­ter­val \( [0, 2\pi] \), and \( \alpha \) is a scal­ing factor that has the ef­fect of shrink­ing (if \( \alpha < 1 \)) or dilat­ing (if \( \alpha > 1 \)) the wave pro­file. In two di­men­sions, waves do not de­cay ex­po­nen­tially fast, even in the ab­sence of a scat­ter­er, but the fo­cus­ing/trap­ping ef­fect caused by non-star-shaped obstacles is sim­il­ar.
Peter Lax with Cathleen Morawetz at the 2008 Conference on Nonlinear Phenomena in Mathematical Physics: Dedicated to Cathleen Synge Morawetz on her 85th Birthday.
Photo courtesy of the Fields Institute.

Mor­awetz’s writ­ing style was very much that of a storyteller. To get a sense of that, here is the be­gin­ning of the proof of the main the­or­em in her 1961 pa­per [1]:

The proof is based on en­ergy iden­tit­ies, i.e., quad­rat­ic in­teg­ral re­la­tions sat­is­fied by all solu­tions. This is one of the most power­ful tools for get­ting es­tim­ates for solu­tions of el­lipt­ic, hy­per­bol­ic or mixed equa­tions. The most fa­mil­i­ar iden­tity of this kind for the wave equa­tion is ob­tained by mul­tiply­ing \( u_{tt} = \Delta u \) by \( u_t \) and in­teg­rat­ing in the slab \( 0 \leq t \leq t_1 \); the res­ult­ing in­teg­ral iden­tity sat­is­fies the con­ser­va­tion of en­ergy. Here we use an­oth­er mul­ti­pli­er in the place of \( u_t \) in­tro­duced by Prot­ter for an­oth­er pur­pose. The sig­ni­fic­ance of us­ing al­tern­at­ive mul­ti­pli­ers has been fre­quently em­phas­ized by Friedrichs and is of­ten re­ferred to as Friedrichs’s \( a \), \( b \), \( c \)-meth­od. The mul­ti­pli­er here is \begin{equation} \label{fourteen} xu_x + yu_y + zu_z + tu_t + u \end{equation} and from the res­ult­ing iden­tity we con­clude that all the en­ergy is car­ried out­ward.

In truth, Mor­awetz was be­ing overly mod­est. It was her keen in­sight that al­lowed for the se­lec­tion of a mul­ti­pli­er which would yield the de­sired res­ult. The power and gen­er­al­ity of this ap­proach led to break­throughs in many wave propaga­tion prob­lems, with the state of the art col­lec­ted in Mor­awetz’s 1966 mono­graph “En­ergy iden­tit­ies for the wave equa­tion,” ori­gin­ally re­leased as a Cour­ant In­sti­tute tech­nic­al re­port.

A second ma­jor step for­ward in un­der­stand­ing the de­cay of waves scattered from star-shaped obstacles came in 1963, in joint work with Lax and Phil­lips. They showed that, in fact, such solu­tions de­cay ex­po­nen­tially (as they do for a sphere), not just as \( t^{-1/2} \). The proof re­lies on an ob­ser­va­tion of Lax and Phil­lips that there is a func­tion \( Z(t) \) which sat­is­fies the semig­roup prop­erty \begin{equation} \label{fifteen} Z(t + s) = Z(t)Z(s), \end{equation} and whose norm con­trols the de­cay of the solu­tion. In this con­text, Mor­awetz’s 1961 pa­per shows that for some time \( t = \tau \), \( |Z(\tau)| \) has de­cayed to less than one: \[ |Z(\tau)| < 1 = e^{-\alpha}\quad\text{ for some } \alpha > 0. \] That is enough to guar­an­tee ex­po­nen­tial de­cay! One simply writes \[ t = n\tau + t_1\quad\text{ where }t_1 < \tau, \] from which \[ |Z(t)| = |Z(t_1 )|\,|[Z(\tau)]^n | \leq |Z(t_1 )|e^{-\alpha n} = Ce^{-\alpha t/\tau}. \]

Nontrapping objects

Figure 5. In 1977, Morawetz, Ralston, and Strauss generalized the class of scatterers for which decay results could be proven. They showed, in three dimensions, that if the object \( \Omega \) does not trap rays, then the local energy of the wave must decay exponentially. Star-shaped objects are a subset of this much larger class. The path taken by a ray is depicted, reflecting from the surface each time according to geometrical optics.
Figure by Leslie Greengard.
One of the fea­tures of star-shaped ob­jects is that rays impinging on them can­not be trapped. A ray here is the path taken by an in­fin­itely thin beam of light which re­flects from the sur­face ac­cord­ing to geo­met­ric­al op­tics. For a com­plic­ated scat­ter­er, one can ima­gine that a ray could un­der­go suc­cess­ive bounces without es­cap­ing from the con­vo­lu­tions of the sur­face \( \partial\Omega \) in any fi­nite time in­ter­val (see Fig­ure 5).

This situ­ation was stud­ied by Mor­awetz, Ral­ston, and Strauss in their 1977 art­icle, where they proved a re­mark­able ex­ten­sion of Mor­awetz’s earli­er res­ults; if the ob­ject \( \Omega \) does not trap rays, then the scattered wave de­cays ex­po­nen­tially. The proof in­volves the in­tro­duc­tion of an es­cape func­tion (a gen­er­al­iz­a­tion of the geo­met­ric in­tu­ition of an “es­cape path of fi­nite length”) and a dif­fer­ent mul­ti­pli­er from that in Mor­awetz’s 1961 pa­per. The use of such Mor­awetz mul­ti­pli­ers is now ubi­quit­ous in the ana­lys­is of PDEs.

De­not­ing by \( S \) a sphere which con­tains the smooth scat­ter­er \( \Omega \), con­sid­er­ing \( \mathbf{x} \in S\backslash \Omega \), and let­ting \( \xi \) be a unit vec­tor in \( \mathbb{R}^3 \), \( p(\mathbf{x}, \xi) \) is said to be an es­cape func­tion if it is real-val­ued, \( C^{\infty} \), and, in­form­ally speak­ing, “strictly in­creas­ing along rays, \( \xi \) be­ing the ray dir­ec­tion at \( \mathbf{x} \).” Rays are said to be not trapped if the total path length in \( S\backslash \Omega \) is bounded and waves are said to be not trapped if the loc­al en­ergy in \( S\backslash \Omega \) de­cays to zero uni­formly. Without en­ter­ing in­to de­tails, Mor­awetz, Ral­ston, and Strauss showed (1) that if rays are not trapped, then there ex­ists an es­cape func­tion and (2) that if there ex­ists an es­cape func­tion, then waves are not trapped, from which the res­ult fol­lows.

Geometric optics and frequency domain analysis

In the study of lin­ear wave propaga­tion, much of our un­der­stand­ing comes from the fre­quency do­main — that is, ana­lyz­ing the Four­i­er trans­form of the wave equa­tion \eqref{eleven}: \begin{equation} \label{sixteen} - k^2 U(\mathbf{x}, k) - \Delta U(\mathbf{x}, k) = F(\mathbf{x}, k). \end{equation} De­pend­ing on the con­text, this is re­ferred to as the Helm­holtz or re­duced wave equa­tion. In 1968, Mor­awetz, to­geth­er with Don Lud­wig, began an in­vest­ig­a­tion of ex­ter­i­or scat­ter­ing from star-shaped sur­faces in the fre­quency do­main [2]. Two ma­jor res­ults were presen­ted there. First, they provided a key proof of the well-posed­ness of the scat­ter­ing prob­lem for sound-soft bound­ar­ies (ho­mo­gen­eous Di­rich­let bound­ary con­di­tions) with re­spect to the bound­ary data and for­cing term \( F(\mathbf{x}, k) \) in \eqref{sixteen}. They also in­tro­duced what are now called Mor­awetz iden­tit­ies for the Helm­holtz equa­tion. Second, they showed that the for­mu­las pro­duced by the the­ory of geo­met­ric­al op­tics are asymp­tot­ic to the ex­act solu­tion. The rel­ev­ant asymp­tot­ic re­gimes are il­lus­trated in Fig­ure 6.

Figure 6. The asymptotic regimes for geometrical optics. For a fixed point \( \mathbf{x}_0 \), the two tangent lines to the scatterer define the shadow boundary (dashed black lines), which separates the illuminated region from the shadow region. The penumbra is a neighborhood of the shadow boundary, formed by the union of all shadow boundaries of spherical waves with centers in a neighborhood of \( \mathbf{x}_0 \).
Figure by Leslie Greengard, adapted from [2]

Without en­ter­ing in­to tech­nic­al de­tails, geo­met­ric­al op­tics is based on ex­pand­ing the in­com­ing and scattered waves in terms of a series in in­verse powers of the wavenum­ber \( k \) about the point \( \mathbf{x}_0 \) (see Fig­ure 6). Lud­wig had earli­er pro­posed an ex­pan­sion for the pen­um­bra re­gion as well. Mor­awetz and Lud­wig showed that all of these ex­pan­sions are truly asymp­tot­ic: to the solu­tion in the il­lu­min­ated re­gion and pen­um­bra, and asymp­tot­ic­ally zero in the deep shad­ow.

Al­though Mor­awetz her­self did little nu­mer­ic­al com­pu­ta­tion, her ana­lyt­ic work (es­pe­cially on mul­ti­pli­ers) has played a ma­jor role in the design of nu­mer­ic­al meth­ods. We can­not do justice to the lit­er­at­ure here, but refer the read­er to three re­cent pa­pers: one on ei­gen­value com­pu­ta­tion, one on fre­quency do­main scat­ter­ing, and one on time-do­main in­teg­ral equa­tions [e1] [e2] [e5]. We have only been able to scratch the sur­face of her leg­acy in this note. Her con­tri­bu­tions are pro­found and deep, and have changed the way we think about par­tial dif­fer­en­tial equa­tions. She was a won­der­ful friend and col­league and is greatly missed.

Works

[1] C. S. Mor­awetz: “The de­cay of solu­tions of the ex­ter­i­or ini­tial-bound­ary value prob­lem for the wave equa­tion,” Comm. Pure Ap­pl. Math. 14 : 3 (August 1961), pp. 561–​568. MR 132908 Zbl 0101.​07701 article

[2] C. S. Mor­awetz and D. Lud­wig: “An in­equal­ity for the re­duced wave op­er­at­or and the jus­ti­fic­a­tion of geo­met­ric­al op­tics,” Comm. Pure Ap­pl. Math. 21 : 2 (March 1968), pp. 187–​203. MR 223136 Zbl 0157.​18701 article