return

Celebratio Mathematica

Cathleen Morawetz

Morawetz inequalities

by Terence Tao

Morawetz worked to change the way we think about partial differential equations.
Photo courtesy of James Hamilton.

Cast a stone in­to a still lake. There is a large splash, and waves be­gin ra­di­at­ing out from the splash point on the sur­face of the wa­ter. But, as time passes, the amp­litude of the waves de­cays to zero.

This type of be­ha­vi­or is com­mon in phys­ic­al waves, and also in the par­tial dif­fer­en­tial equa­tions used in math­em­at­ics to mod­el these waves. Let us be­gin with the clas­sic­al wave equa­tion \begin{equation} \label{one} -\partial_{tt} u + \Delta u = 0, \end{equation} where \( u \colon \mathbb{R} \times \mathbb{R}^3 \rightarrow \mathbb{R} \) is a func­tion of both time \( t \in \mathbb{R} \) and space \( x \in \mathbb{R}^3 \), which is a simple mod­el for the amp­litude of a wave propagat­ing at unit speed in three-di­men­sion­al space; here \begin{equation*} \Delta = \frac{\partial^2}{\partial x^2_1} + \frac{\partial^2}{\partial x^2_2} + \frac{\partial^2}{\partial x^2_3} \end{equation*} de­notes the spa­tial Lapla­cian. One can veri­fy that one has the fam­ily of ex­pli­cit solu­tions \begin{equation} u(t, x) = \frac{F(t + |x|) - F(t - |x|)}{|x|} \end{equation} to \eqref{one} for any smooth, com­pactly sup­por­ted func­tion \( F \colon \mathbb{R} \rightarrow \mathbb{R} \), where \begin{equation*} |x| = \sqrt{x^2_1 + x^2_2 + x^2_3} \end{equation*} de­notes the Eu­c­lidean mag­nitude of a po­s­i­tion \( x \in \mathbb{R}^3 \). The dis­pers­ive nature of this equa­tion can be seen in the ob­ser­va­tion that the amp­litude \begin{equation*} \sup_{x\in\mathbb{R}^3} |u(t, x)| \end{equation*} of such solu­tions de­cays to zero as \( t \rightarrow \pm \infty \), whilst oth­er quant­it­ies such as the en­ergy \begin{equation*} \int_{\mathbb{R}^3} \frac{1}{2} |\partial_t u(t, x)|^2 + \frac{1}{2} |\nabla u(t, x)|^2\, dx \end{equation*} stay con­stant in time (and in par­tic­u­lar do not de­cay to zero).

The wave equa­tion can be viewed as a spe­cial case of the more gen­er­al lin­ear Klein–Gor­don equa­tion \begin{equation} \label{three} -\partial_{tt} u + \Delta u = m^2 u, \end{equation} where \( m \geq 0 \) is a con­stant. Even more im­port­ant is the lin­ear Schrödinger equa­tion, which we will nor­mal­ize here as \begin{equation} \label{four} i\partial_t u + \frac{1}{2} \Delta u = 0, \end{equation} where the un­known field \( u \colon \mathbb{R} \times \mathbb{R}^3 \rightarrow \mathbb{C} \) is now com­plex-val­ued. There are also non­lin­ear vari­ants of these equa­tions, such as the non­lin­ear Klein–Gor­don equa­tion \begin{equation} \label{five} - \partial_{tt} u + \Delta u = m^2 u + \lambda |u|^{p-1} u, \end{equation} and the non­lin­ear Schrödinger equa­tion \begin{equation} \label{six} i\partial_t u + \frac{1}{2}\Delta u = \lambda |u|^{p-1} u, \end{equation} where \( \lambda = \pm 1 \) and \( p > 1 \) are spe­cified para­met­ers. There are count­less oth­er fur­ther vari­ations (both lin­ear and non­lin­ear) of these dis­pers­ive equa­tions, such as Ein­stein’s equa­tions of gen­er­al re­lativ­ity, or the Korteweg–de Vries equa­tions for shal­low wa­ter waves.

Figure 1. Dispersion is illustrated in this numerical simulation of the Klein–Gordon equation implemented by Brian Leu, Albert Liu, and Parth Sheth using XSEDE when they were undergrads at U Michigan in 2013.

An im­port­ant way to cap­ture dis­per­sion math­em­at­ic­ally is through the es­tab­lish­ment of dis­pers­ive in­equal­it­ies that as­sert, roughly speak­ing, that if a solu­tion \( u \) to one of these equa­tions is suf­fi­ciently loc­al­ized in space at an ini­tial time, \( t = 0 \), then it will de­cay as \( t \rightarrow\infty \). (If a solu­tion \( u \) is not loc­al­ized enough in space ini­tially, it does not need to de­cay; con­sider for in­stance the trav­el­ing wave solu­tion \begin{equation*} u(t, x) = F(t - x_1) \end{equation*} to the wave equa­tion \eqref{one}.) This de­cay has to be meas­ured in suit­able func­tion space norms, such as the \( L^{\infty}_x (\mathbb{R}^3) \) norm.

One can rep­res­ent any solu­tion \( u \) to the lin­ear Schrödinger equa­tion ex­pli­citly in terms of the ini­tial data \( u(0) \) by the for­mula \begin{equation*} u (t,x)=\frac{1}{(2\pi it)^{3/2}}\int_{\mathbb{R}^3} e^{-i|x-y|^2/2t} u (0,y)\,dy \end{equation*} for all \( t \neq 0 \) and \( x \in \mathbb{R}^3 \), where the quant­ity \( (2\pi it)^{3/2} \) is defined us­ing a suit­able branch cut. From the tri­angle in­equal­ity, this im­me­di­ately gives the dis­pers­ive in­equal­ity \begin{equation} \|u(t)\|_{L^{\infty}_x(\mathbb{R}^3 )} \leq \frac{1}{(2\pi |t|)^{3/2}}\|u(0)\|_{L_x^1(\mathbb{R}^3)}. \end{equation} If the solu­tion is ini­tially spa­tially loc­al­ized in the sense that the \( L^1 \) norm \begin{equation*} \|u(0)\|_{L^{1}_x(\mathbb{R}^3 )} \end{equation*} is fi­nite, then the solu­tion \( u(t) \) de­cays uni­formly to zero as \( t \rightarrow \pm \infty \). A sim­il­ar (but slightly more com­plic­ated) dis­pers­ive in­equal­ity can also be ob­tained for solu­tions to the lin­ear Klein–Gor­don equa­tion \eqref{three}.

On the oth­er hand, solu­tions to the lin­ear Schrödinger equa­tion \eqref{four} sat­is­fy the point­wise mass con­ser­va­tion law \begin{equation} \partial_t |u|^2 = \sum^3_{j=1} \partial_{x_j} \mathrm{Im} (\bar{u}\,\partial_{x_j} u). \end{equation} From this, one can eas­ily de­rive con­ser­va­tion of the spa­tial \( L^2 \) norm of the solu­tion: \begin{equation*} \|u(t)\|_{L^2_x (\mathbb{R}^3)} = \|u(0)\|_{L^2_x (\mathbb{R}^3)} . \end{equation*} In par­tic­u­lar, the \( L^2 \) norm of the solu­tion will stay con­stant in time, rather than de­cay to zero.

Cathleen Synge Morawetz in 1964.
Photo courtesy of New York University.

To re­con­cile this fact with the dis­pers­ive es­tim­ate, we ob­serve that solu­tions to dis­pers­ive equa­tions such as the lin­ear Schrödinger equa­tion spread out in space as time goes to in­fin­ity (much as the ripples on a pond do), al­low­ing the \( L^{\infty} \) norm of such a solu­tion to go to zero even while the \( L^2 \) norm stays bounded away from zero. As men­tioned earli­er, this ef­fect can also be seen for the wave equa­tion \eqref{one}.

The above ana­lys­is of the lin­ear Schrödinger equa­tion re­lied cru­cially on hav­ing an ex­pli­cit fun­da­ment­al solu­tion at hand. What hap­pens if one works with non­lin­ear (and not com­pletely in­teg­rable) equa­tions, such as \eqref{five} or \eqref{six}, in which no ex­pli­cit and tract­able for­mula for the solu­tion is avail­able? For lin­ear equa­tions (such as the wave or Schrödinger equa­tion out­side of an obstacle, or in the pres­ence of po­ten­tials or mag­net­ic fields) one can still hope to use meth­ods from spec­tral the­ory to un­der­stand the long-time be­ha­vi­or (as is done for in­stance in the fam­ous RAGE the­or­em of Ruelle (1969), Am­reinGeorges­cu (1973), and Enss (1977)). However, such meth­ods are ab­sent for non­lin­ear equa­tions such as \eqref{five} or \eqref{six}, par­tic­u­larly when deal­ing with solu­tions that are too large for per­turb­at­ive the­ory to be of much use.

Re­call that in 1961, Mor­awetz proved the de­cay of solu­tions to the clas­sic­al wave equa­tion in the pres­ence of a star-shaped obstacle. Mor­awetz used the “Friedrichs abc meth­od,” in which one mul­ti­plied both sides of a PDE such as \eqref{five} or \eqref{six} by a mul­ti­pli­er \begin{equation*} a\,\partial_t u + b \cdot \nabla u + cu \end{equation*} for well-chosen func­tions \( a \), \( b \), \( c \), in­teg­rated over a space-time do­main, and re­arran­ging us­ing in­teg­ra­tion by parts and omit­ting some terms of def­in­ite sign, ob­tained a use­ful in­teg­ral in­equal­ity. The key dis­cov­ery of Mor­awetz (a ver­sion of which first ap­peared in work of Lud­wig) was that this meth­od was par­tic­u­larly fruit­ful when the mul­ti­pli­er was equal to the ra­di­al de­riv­at­ive \begin{equation*} \frac{x \cdot \nabla u}{|x|} \end{equation*} of the solu­tion (in some cases one also adds a lower-or­der term \( u/|x| \)).

In 1968, Mor­awetz ap­plied this tech­nique to study solu­tions \( u \) to the non­lin­ear Klein–Gor­don equa­tion \eqref{five}, as­sum­ing one is in the non­focus­ing case with \( \lambda = m =+1 \). (In the fo­cus­ing case \( \lambda = -1 \), the equa­tion \eqref{five} ad­mits “soliton” solu­tions that are sta­tion­ary in time and thus do not dis­perse.) By us­ing a mul­ti­pli­er of the above form, Mor­awetz ob­tained an in­equal­ity of the form \begin{equation} \label{nine} \int_{\mathbb{R}} \int_{\mathbb{R}^3} U(t, x)\, dx\, dt \leq C E(u(0)), \end{equation} where \begin{equation*} U(t, x) =\frac{ |u(t, x)|^2 + |u(t, x)|^{p+1}}{|x|}. \end{equation*} Here the con­stant \( C \) de­pends only on the ex­po­nent \( p \) and \( E(u(0)) \) is the en­ergy: \begin{equation*} E(u(0)) = \int_{\mathbb{R}^3} \frac{1}{2} |\nabla u(0, x)|^2 + \frac{1}{2} |\partial_t u(0, x)|^2 +\frac{1}{p+1} |u(0, x)|^{p+1}\, dx. \end{equation*} This type of es­tim­ate is now known as a Mor­awetz in­equal­ity. The key point here is that the left-hand side of the Mor­awetz in­equal­ity in \eqref{nine} con­tains an in­teg­ra­tion over the en­tire time do­main \( \mathbb{R} \) (as op­posed to a time in­teg­ral over a bounded in­ter­val). It im­me­di­ately rules out soliton-type solu­tions that move at bounded speed (as this would make the left-hand side of \eqref{nine} in­fin­ite). It forces some time-av­er­aged de­cay of the solu­tion near the spa­tial ori­gin \( x = 0 \). For in­stance, it is im­me­di­ate from \eqref{nine} that \begin{equation*} \frac{1}{T}\int_0^T \int_K |u(t, x)|^2\, dx\, dt \rightarrow 0 \end{equation*} as \( T \rightarrow\infty \) for any com­pact spa­tial re­gion \( K \subset \mathbb{R}^3 \).

Cathleen Morawetz and Walter Strauss in 2008.
Photo courtesy of Walter Strauss.

Once one has some sort of de­cay es­tim­ate for a dis­pers­ive equa­tion, it is of­ten pos­sible to “boot­strap” the es­tim­ate to ob­tain ad­di­tion­al de­cay es­tim­ates. For in­stance one might use the de­cay es­tim­ate one already has to bound the right-hand side of a non­lin­ear PDE such as \eqref{five} or \eqref{six}, and then solve the as­so­ci­ated (in­homo­gen­eous) lin­ear PDE to ob­tain a new de­cay es­tim­ate for the solu­tion.

An early res­ult of this type was de­veloped by Mor­awetz and Strauss in 1975. They showed that for any fi­nite en­ergy solu­tion \( u \) to the non­lin­ear Klein–Gor­don equa­tion \eqref{five} with \( \lambda = m = +1 \) and \( p = 3 \), the solu­tion de­cays like a solu­tion to the lin­ear Klein–Gor­don equa­tion \eqref{three}. More pre­cisely, there ex­ist fi­nite solu­tions \( u_+ , u_{-} \) to \eqref{three} such that \( u(t) - u_+ (t) \) (resp. \( u(t) - u_{-} (t) \)) goes to zero in the en­ergy norm as \( t\rightarrow +\infty \) (resp. \( t \rightarrow -\infty \)). This can be de­veloped fur­ther in­to a sat­is­fact­ory scat­ter­ing the­ory for such equa­tions, which among oth­er things gives a con­tinu­ous scat­ter­ing map from \( u_- \) to \( u_+ \) or vice versa. See Fig­ure 2.

In the dec­ades since Mor­awetz’s pi­on­eer­ing work, many ad­di­tion­al Mor­awetz in­equal­it­ies have been de­veloped. For in­stance, in 1978, Lin and Strauss de­veloped Mor­awetz in­equal­it­ies for the non­lin­ear Schrödinger equa­tion, and Mor­awetz her­self dis­covered fur­ther such es­tim­ates for the wave equa­tion out­side of an obstacle. In more re­cent years, “in­ter­ac­tion Mor­awetz in­equal­it­ies” were in­tro­duced, which could con­trol cor­rel­a­tion quant­it­ies such as \begin{equation} \label{ten} \int_{\mathbb{R}} \int_{\mathbb{R}^3} \int_{\mathbb{R}^3} \frac{|u(t,x)|^2|u(t,y)|^p}{|x-y|}\,dx\,dy\,dt \end{equation} for solu­tions \( u \) to the non­lin­ear Schrödinger equa­tion \eqref{six}.

Figure 2. Here we see \( u_- \) on the left in blue and \( u_+ \) on the right in red, with \( u \) in purple approximating \( u_- \) as \( t \rightarrow -\infty \) and approximating \( u_+ \) as \( t \rightarrow \infty \).
Figure by Penelope Chang.

One way to view Mor­awetz in­equal­it­ies is as an as­ser­tion of mono­ton­icity of the ra­di­al mo­mentum, which takes the form \begin{equation*} \int_{\mathbb{R}^3}(\partial_t u)\biggl(\frac{x}{|x|}\cdot \nabla u\biggr)\, dx \end{equation*} for wave or Klein–Gor­don equa­tions, and \begin{equation*} \int_{\mathbb{R}^3}\operatorname{Im}\biggl(\bar{u}\frac{x}{|x|}\cdot \nabla u\biggr)\, dx \end{equation*} for Schrödinger equa­tions. In­form­ally, this quant­ity is ex­pec­ted to be pos­it­ive when waves propag­ate away from the ori­gin, and neg­at­ive when they propag­ate to­wards the ori­gin. The in­tu­ition is that while waves can some­times propag­ate to­wards the ori­gin, even­tu­ally they will move past the ori­gin and be­gin ra­di­at­ing away from the ori­gin. However, in the ab­sence of fo­cus­ing mech­an­isms (such as a neg­at­ive sign \( \lambda = -1 \) in the non­lin­ear­ity), the re­verse phe­nomen­on of out­ward net ra­di­al mo­mentum be­ing con­ver­ted to in­ward net ra­di­al mo­mentum can­not oc­cur. Thus the ra­di­al mo­mentum is al­ways ex­pec­ted to be in­creas­ing in time.

On the oth­er hand, un­der hy­po­theses such as fi­nite en­ergy, this ra­di­al mo­mentum should be bounded. So by the fun­da­ment­al the­or­em of cal­cu­lus, the time de­riv­at­ive of the ra­di­al mo­mentum should have a bounded in­teg­ral in time. In­tu­it­ively, one ex­pects this time de­riv­at­ive to be large when the solu­tion has a strong pres­ence near the ori­gin, but not when the solu­tion is far away from the ori­gin. Far from the ori­gin the ra­di­al vec­tor field \begin{equation*} \frac{x}{|x|}\cdot \nabla \end{equation*} be­haves like a con­stant, and the ra­di­al mo­mentum ap­proaches a fixed co­ordin­ate of the total mo­mentum. This ex­plains why Mor­awetz in­equal­it­ies tend to in­volve factors such as \( 1/|x| \) that loc­al­ize the es­tim­ate to near the ori­gin.

The Mor­awetz in­equal­it­ies are in­dis­pens­ible as an in­gredi­ent in con­trolling the long-time be­ha­vi­or of solu­tions to a wide ar­ray of dis­pers­ive de­fo­cus­ing equa­tions, in­clud­ing a num­ber of en­ergy-crit­ic­al or mass-crit­ic­al equa­tions in which the ana­lys­is is par­tic­u­larly del­ic­ate and in­ter­est­ing; see for in­stance the texts [e1], [e3], [e4] for de­tailed cov­er­age of these top­ics. They have also been suc­cess­fully ap­plied to many equa­tions in gen­er­al re­lativ­ity (such as Ein­stein’s equa­tions for grav­it­a­tion­al fields), for in­stance to ana­lyze the asymp­tot­ic be­ha­viour around a black hole. The fun­da­ment­al tools that Mor­awetz has in­tro­duced to the field of dis­pers­ive equa­tions will cer­tainly un­der­lie fu­ture pro­gress in this field for dec­ades to come.