Math 597F, Notes 8: unstable Orr-Sommerfeld solutions for stable profiles

We now turn to the delicate case: Orr-Sommerfeld solutions for stable profiles to Rayleigh. The results reported here are in a joint work with E. Grenier and Y. Guo, directing some tedious details of the proof to our paper. We consider the Orr-Sommerfeld problem:

\displaystyle Ray_\alpha(\phi) = \epsilon \Delta_\alpha ^2 \phi,

with zero boundary conditions on {\phi} and {\phi'}, in which {Ray_\alpha(\cdot)} denotes the Rayleigh operator and {\Delta_\alpha = \partial_z^2 - \alpha^2}.

We assume that the boundary layer {U} is stable to the Rayleigh operator, that is, there is no unstable eigenvalue with {\Im c_0 >0} to the Rayleigh problem (equivalently, the spectrum of Euler lies precisely on the imaginary axis). In term of the Wronskian determinant, this asserts that {E(\alpha,c) \not =0} for all {c} with {\Im c>0}. The dispersion relation studied in the last section shows that there is no eigenvalue to the Orr-Sommerfeld problem, with {\Im c(\epsilon)} being bounded away from zero in the limit {\epsilon \rightarrow 0}. Hence, in search for a possible unstable point spectrum of Navier-Stokes, we will construct solutions to Orr-Sommerfeld equations in the regime where all parameters {\alpha, c} and {\epsilon} are sufficiently small. In particular, {\alpha \log \Im c \ll 1}, in order to take the inverse of the {Ray_\alpha} operator as studied in the last section. Throughout the notes, {z_c} denotes a complex number so that {U(z_c) = c}, for each fixed complex number {c}.

Clearly, {\epsilon \Delta_\alpha^2 Ray_\alpha^{-1}} is not a good iteration operator, since by a view of Rayleigh solutions, {Ray_\alpha^{-1}(f)} typically has a singularity of the form {(z-z_c)\log (z-z_c)} and therefore {\epsilon \Delta_\alpha^2 Ray_\alpha^{-1}} has a singularity of order {(z-z_c)^{-3}}. To deal with the singularity, we need to examine the leading operator in Orr-Sommerfeld equations near the singular point {z = z_c}. Indeed, introducing the blow-up variable: {Z = (z-z_c)/\delta} and search for the ansatz solution {\phi = \phi_\mathrm{cr}(Z)}, which is in the literature often referred as the critical layer. It follows that {\phi_\mathrm{cr}} approximately solves

\displaystyle \partial_Z^4 \phi_\mathrm{cr} \approx Z\partial_Z^2 \phi_\mathrm{cr}

with the critical layer size {\delta \approx \epsilon^{1/3}}. That is, within the critical layer, {\partial_z^2 \phi_\mathrm{cr}} solves the classical Airy equation. In the light of the toy model problem in the previous section, we introduce the following iteration operator:

\displaystyle Iter: = \underbrace{Airy^{-1}}_{\mbox{critical layer}} \circ \quad \underbrace{\epsilon \Delta^2_\alpha}_{\mbox{error}} \quad \circ \quad \underbrace{Ray_\alpha^{-1}}_{\mbox{inviscid}}.

It suffices to show that the Iter operator is contractive in a suitable function space. This would take care o the singularity. (in the proof, we in fact need a slightly more complicated iteration operator to take care of a possible linear growth in {z} of the Rayleigh solutions at infinity). Here, the Airy operator is defined by

\displaystyle Airy(\phi) := \epsilon \partial_z^4 \phi - (U - c + 2 \epsilon \alpha^2) \partial_z^2 \phi . \ \ \ \ \ (1)

Several issues to overcome. We need to appropriately define {Airy^{-1}} and {Ray_\alpha^{-1}}, and most of all, to study the smoothing effect of Airy operator near the critical layer. The inverse {Ray_\alpha^{-1}} is studied at length in the previous lecture. Similarly, the inverse {Airy^{-1}} will be constructed through its Green kernel.

8.1. Green kernel

We recall the basic construction of the Green kernel of a linear, variable-coefficient, second-order ODE operator:

\displaystyle L \phi : = a(z) \partial_z^2\phi + b(z)\partial_z\phi + c(z) \phi = f(z),\qquad z\ge 0,

with {a(z)} never vanishing. For simplicity, we assume that there are two independent solutions {\phi_1, \phi_2} to the homogenous equation {L\phi =0}, with {\phi_1} being bounded and {\phi_2} unbounded at infinity {z = \infty}. We search for a particular (bounded) solution to the non-homogenous equation in the integral form:

\displaystyle \phi (z) = \int_0^\infty G(x,z) f(x) \; dx,\ \ \ \ \ (2)

in which {G(x,z)} is called the Green kernel of {L}, which by the equation satisfies {L G(x,z) = \delta_x(z)}, where {\delta_x(z)} denotes the delta function centered at {z=x}. Comparing the singularity in the equation for {G(x,z)}, one asserts that {G(x,z)} must be continuous and {\partial_zG(x,z)} has a jump across {z=x}; precisely,

\displaystyle [G(x,z)]_{\vert_{z=x}} : = \lim_{z \rightarrow x^+} G(x,z) - \lim_{z\rightarrow x^-} G(x,z) = 0, \qquad [a(z)\partial_z G(x,z)]_{\vert_{z=x}} = 1.

We now construct a Green kernel {G(x,z)} (which is non unique, unless we impose a boundary condition). As {z \not =x}, {G(x,z)} solves the homogenous equation and so can be constructed as a linear combination of the other two independent solutions {\phi_1,\phi_2}. This yields

\displaystyle G(x,z) = \left \{ \begin{aligned} C_1(x) \phi_1(z) , \qquad & z>x\\ - C_2(x) \phi_2(z), \qquad & z<x. \end{aligned} \right.

The jump conditions across {z=x} yield the linear equations:

\displaystyle \begin{pmatrix} \phi_1 & \phi_2 \\ \phi'_1 & \phi'_2 \end{pmatrix} \begin{pmatrix} C_1\\C_2\end{pmatrix}(x) = \begin{pmatrix} 0 \\ 1/a(x)\end{pmatrix} ,

which gives {C_1(x)} and {C_2(x)}. Hence, the Green kernel {G(x,z)} can be defined as

\displaystyle G(x,z) = \frac{1}{a(x)W[\phi_1, \phi_2](x)} \left \{ \begin{aligned} \phi_2(x) \phi_1(z) , \qquad & z>x\\ \phi_1(x) \phi_2(z), \qquad & z<x \end{aligned} \right. \ \ \ \ \ (3)

in which {W[\phi_1, \phi_2]} denotes the Wronskian determinant, which can be computed through the Abel’s identity: {W(x) = W(0) \mbox{exp}(-\int_0^x b(z)/a(z)\; dz)}. In particular, when {b\equiv 0}, the Wronskian determinant is constant. Note that the Green kernel is not unique, as we did not take care of the boundary conditions. Certainly, this construction yields a particular solution in the integral form (2).

8.2. Airy operator

We now study the Airy operator, defined as in (1). Note that {\partial_z^2\phi} does not exactly solve the classical Airy equation: {\partial_z^2 u - z u =0}. We shall make a change of variables and unknowns in order to go back to the classical Airy equation. This change is very classical in physical literature, and called the Langer’s transformation: {(z,\phi) \mapsto (\eta,\Phi)}, with {\eta = \eta(z)} defined by

\displaystyle \eta (z) = \Big[ \frac 32 \int_{z_c}^z \Big( \frac{U-c}{U'_c}\Big)^{1/2} \; dz \Big]^{2/3} \ \ \ \ \ (4)

and {\Phi = \Phi(\eta)} defined by the relation

\displaystyle \partial_z^2 \phi (z) = \dot z ^{1/2} \Phi(\eta), \ \ \ \ \ (5)

in which {\dot z = \frac{d z( \eta)}{d \eta} } and {z = z(\eta)} is the inverse of the map {\eta = \eta(z)}. By a view of the definition (4), we note that {(U-c)\dot z^2 = U'_c \eta}, with {U'_c = U'(z_c)}. The following lemma links the Airy operator (1) with the classical Airy equation.

Lemma 1 Let {(z,\phi) \mapsto (\eta, \Phi)} be the Langer’s transformation defined as above. The function {\Phi(\eta)} solves the classical Airy equation:

\displaystyle \epsilon \partial^2_\eta \Phi - U_c' \eta \Phi = f(\eta)\ \ \ \ \ (6)

if and only if the function {\phi = \phi(z)} solves

\displaystyle Airy ( \phi) = \dot z ^{-3/2} f(\eta(z))+ \epsilon [ \partial_z^2 \dot z^{1/2} \dot z^{-1/2} - 2\alpha^2 ]\partial_z^2 \phi (z) .\ \ \ \ \ (7)

Proof: The lemma follows from direct calculations. \Box

8.2.1. The classical Airy. Thanks to the Langer’s transformation, we first solve the classical Airy equation (6) for {\Phi}. Let us denote

\displaystyle \delta = \Bigl( { \epsilon\over U_c'} \Bigr)^{1/3} = e^{-i \pi / 6} (\alpha R U_c')^{-1/3}

to be the critical layer size, and introduce the notation {Z = \delta^{-1} \eta }. Then {\Psi(Z) = \Phi(\eta)} solves the truly classical Airy: {\Psi'' - Z \Psi =U_c' \delta f(\delta Z)}. We use the following classical lemma:

Lemma 2 The classical Airy equation {\Psi'' - z \Psi =0} has two independent solutions {Ai(z)} and {Ci(z)} so that the Wronskian determinant of {Ai} and {Ci} equals to one. In addition, {Ai(e^{i \pi /6} x)} and {Ci(e^{i \pi /6} x)} converge to {0} as {x\rightarrow \pm \infty} ({x} being real), respectively. Furthermore, there hold asymptotic bounds:

\displaystyle \Bigl| Ai(k, e^{i \pi / 6} x) \Bigr| \le {C| x |^{-k/2-1/4} } e^{-\sqrt{2 | x|} x / 3}, \qquad k\in \mathbb{Z}, \quad x\in \mathbb{R},

and

\displaystyle \Bigl| Ci(k, e^{i \pi / 6} x) \Bigr| \le {C| x |^{-k/2-1/4} } e^{\sqrt{2 | x|} x / 3}, \qquad k\in \mathbb{Z},\quad x\in \mathbb{R},

in which {Ai(0,z) = Ai(z)}, {Ai(k,z) = \partial_z^{-k} Ai(z)} for {k\le 0}, and {Ai(k,z)} is the {k^{th}} primitives of {Ai(z)} for {k\ge 0} and is defined by the inductive path integrals

\displaystyle Ai(k, z ) = \int_\infty^z Ai(k-1, w) \; dw

so that the integration path is contained in the sector with {|\arg(z)| < \pi/3}. The Airy functions {Ci(k,z)} for {k\not =0} are defined similarly.

Thus, we can now introduce the Green kernel of the classical Airy equation (6):

\displaystyle G_{a}(X,Z) = \delta \epsilon^{-1} \left\{ \begin{array}{rrr} Ai(X)Ci(Z), \qquad &\mbox{if}\qquad \xi > \eta,\\ Ai (Z) Ci(X) , \qquad &\mbox{if}\qquad \xi < \eta, \end{array}\right.

in which {X = \delta^{-1} \xi, Z = \delta^{-1}\eta}. By definition, we have

\displaystyle \epsilon\partial_\eta^2 G_{a}(X,Z) - U_c' \eta G_{a}(X,Z) = \delta_\xi(\eta). \ \ \ \ \ (8)

It follows directly that

Lemma 3 There hold pointwise estimates:

\displaystyle \begin{aligned} |\partial_Z^\ell \partial_X^k G_\mathrm{a}(X,Z) | &\le C |\delta|^{-2}(1+|Z|)^{(k+\ell-1)/2} e^ {- {\sqrt{2} \over 3} \sqrt{|Z|}|X-Z| } \end{aligned}

for all {k, \ell\ge 0}.

8.2.2. An approximate Green kernel for Airy operator. Let us take {\xi = \eta(x)} and {\eta = \eta(z)} where {\eta(\cdot)} is the Langer’s transformation and denote {\dot x = 1/\eta'(x)} and {\dot z = 1/ \eta'(z)}. By a view of (5), we define the function {G(x,z)} so that

\displaystyle \partial_z^2G(x,z) =\dot x ^{3/2} \dot z^{1/2} G_{a}(\delta^{-1}\eta(x),\delta^{-1}\eta(z)), \ \ \ \ \ (9)

in which the factor {\dot x ^{3/2}} was added simply to normalize the jump of {G(x,z)}. It then follows from Lemma 1 together with {\delta_{\eta(x)} (\eta(z)) = \delta_x(z)} that

\displaystyle Airy( G(x,z) )= \delta_x(z) + \epsilon [ \partial_z^2 \dot z^{1/2} \dot z^{-1/2} - 2\alpha^2]\partial_z^2G(x,z) . \ \ \ \ \ (10)

That is, {G(x,z)} is indeed an approximate Green function of the Airy operator, defined as in (1), up to a small error term of order {\epsilon \partial_z^2 G = \mathcal{O}(\delta)}. It remains to solve (9) for {G(x,z)}, retaining the jump conditions on {G(x,z)} across {x=z}. In view of primitive Airy functions, let us denote

\displaystyle \widetilde Ci(1,z) =\delta^{-1} \int_0^z \dot y^{1/2} Ci(\delta^{-1}\eta(y))\; dy, \qquad \widetilde Ci(2,z) = \delta^{-1}\int_0^z \widetilde Ci(1,y)\; dy

and

\displaystyle \widetilde Ai(1,z) = \delta^{-1}\int_\infty^z \dot y^{1/2} Ai(\delta^{-1}\eta(y))\; dy, \qquad \widetilde Ai(2,z) =\delta^{-1} \int_\infty^z \widetilde Ai(1,y)\; dy.

Thus, together with our convention that the Green function {G(x,z)} should vanish as {z} goes to {+\infty} for each fixed {x}, we are led to introduce

\displaystyle G(x,z) = i \delta^3 \pi \epsilon^{-1} \dot x^{3/2} \left\{ \begin{aligned} \Big[ Ai(\delta^{-1}\eta(x)) \widetilde Ci(2,z) + \delta^{-1} a_1 (x) (z-x) + a_2(x) \Big] , &\quad \mbox{if }x>z,\\ Ci(\delta^{-1}\eta(x)) \widetilde Ai(2,z) , &\quad \mbox{if } x<z, \end{aligned} \right.

in which {a_1(x), a_2(x)} are chosen so that the jump conditions (see below) hold. Clearly, by definition, {G(x,z)} solves (9), and hence (10). Here the jump conditions on the Green function read:

\displaystyle \begin{aligned} ~[G(x,z)]_{\vert_{x=z}} = [\partial_z G(x,z)]_{\vert_{x=z}} = [\partial_z^2 G(x,z)]_{\vert_{x=z}} =0 \end{aligned}\ \ \ \ \ (11)

and

\displaystyle \begin{aligned} ~[\epsilon \partial_z^3G(x,z)]_{\vert_{x=z}} = 1. \end{aligned}\ \ \ \ \ (12)

We note that from (9) and the jump conditions on {G_a(X,Z)} across {X=Z}, the above jump conditions of {\partial_z^2 G} and {\partial_z^3 G} follow easily. In order for the jump conditions on {G(x,z)} and {\partial_zG(x,z)}, we take

\displaystyle \begin{aligned} a_1 (x) &= Ci(\delta^{-1}\eta(x)) \widetilde Ai(1,x) - Ai(\delta^{-1}\eta(x)) \widetilde Ci(1,x) ,\\ a_2(x) &= Ci(\delta^{-1}\eta(x))\widetilde Ai(2,x) - Ai(\delta^{-1}\eta(x)) \widetilde Ci(2,x) . \end{aligned} \ \ \ \ \ (13)

This yields an approximate Green kernel for the Airy operator. Direct, but much tedious, calculations give sufficient point-wise estimates on the Green kernel, yielding the following convolution estimates; see this paper for details of the proof.

Lemma 4 Let {G(x,z)} be the approximate Green kernel constructed as above. There holds

\displaystyle \Big\|\int_{0}^{\infty} \partial^k_z G(x,\cdot ) f(x) dx \Big \|_{{\eta'}}\le \frac {C|\delta|^{-k}}{\eta - \eta'}\|f\|_{\eta} , \qquad k\ge 0,

for arbitrary {0\le \eta'<\eta}, and for {f \in X_\eta}, denoting the function space that consists of decaying functions at an exponential rate of {e^{-\eta z}}.

8.2.3. Inverse for Airy operator. We study the inhomogeneous Airy equation: {Airy(\phi) = f}, for some source {f(z)}. Thanks to the above construction of an approximate Green kernel, we introduce an approximate inverse of {Airy} operator:

\displaystyle AirySolver(f) := \int_{0}^{\infty} G(x,z) f(x) dx . \ \ \ \ \ (14)

Then, there holds

\displaystyle Airy(AirySolver(f)) = f + AiryErr(f) \ \ \ \ \ (15)

where the error operator {AiryErr(\cdot)} is introduced due to the approximation of the Green kernel and thus is defined as

\displaystyle AiryErr(f) : = \epsilon\int_{0}^{\infty} [ \partial_z^2 \dot z^{1/2} \dot z^{-1/2} - 2\alpha^2]\partial_z^2G(x,z) f(x) dx .

From the convolution estimates, Lemma 4, there hold

\displaystyle \| AirySolver(f)\|_{{\eta'}} \le C_\eta \| f\|_{\eta} , \qquad \| AiryErr(f) \|_{\eta} \le C|\delta| \|f \|_{\eta},

for all {f\in X_\eta} and {\eta' <\eta}. Recalling that {\delta \rightarrow 0}, and so the error is indeed small of order {\mathcal{O}(\delta)}. It is worth noting that unlike {AirySolver(\cdot)}, the {AiryErr(\cdot)} maps from {X_\eta} into itself. The reason is that the slight loss of an exponential decay in {AirySolver(\cdot)} is due precisely to the linear growth in {z} in the approximate Green function {G(x,z)}, whereas this linear growth vanishes in the calculation for {\partial_z^2 G(x,z)}. We recover the same rate of decay for {AiryErr(f)} as that of {f}.

We now construct an inverse for Airy operator by iteration. Let us start with a fixed {f \in X_\eta}. Define

\displaystyle \begin{aligned} \phi_n &= - AirySolver(E_{n-1}) \\ E_n &= - AiryErr(E_{n-1}) \end{aligned} \ \ \ \ \ (16)

for all {n \ge 1}, with {E_0 = f}. It follows by induction that

\displaystyle Airy (S_n) = f + E_n,\qquad S_n = \sum_{k=1}^n \phi_k ,

for all {n\ge 1}. By induction, { \| E_n\|_\eta \le C \delta \|E_{n-1}\|_\eta \le (C\delta)^n \| f\|_\eta} and hence {E_n \rightarrow 0} in {X_\eta} as {n \rightarrow \infty}, for sufficiently small {\delta}. In addition, { \| \phi_n \|_{\eta'} \le C \| E_{n-1}\|_\eta \le C (C\delta)^{n-1} }, which shows that {\phi_n} converges to zero in {X_{\eta'}} for arbitrary fixed {\eta' <\eta} as {n \rightarrow \infty}. Furthermore the series

\displaystyle S_n \rightarrow S_\infty

in {X_{\eta'}} as {n \rightarrow \infty}, for some {S_\infty \in X_{\eta'}}. We then denote {AirySolver_\infty(f) = S_\infty}, for each {f \in X_\eta}. In addition, we have { Airy (S_\infty) = f,} that is, {AirySolver_\infty(f) } is the exact inverse for the Airy operator. Summarizing, we have proved the following theorem.

Lemma 5 Let {\eta'<\eta} be positive numbers. Assume that {\delta} is sufficiently small. There exists an exact solver {AirySolver_\infty(\cdot)}, which is a bounded operator from {X_\eta} to {X_{\eta'}}, so that

\displaystyle Airy(AirySolver_\infty (f)) = f.

8.3. Singularities and contraction of Iter operator

In this section, we study the smoothing effect of the modified Airy function. Precisely, let us consider the Airy equation with a singular source:

\displaystyle Airy(\phi) = \epsilon \partial_z^4 f(z)\ \ \ \ \ (17)

in which {f \in Y_{4,\eta}}, that is {f(z)} and its derivatives decay exponentially at infinity and behaves as {(z-z_c)\log(z-z_c)} near the critical layer {z=z_c}. The singular source {\epsilon \partial_z^4f} arises as an error of the inviscid solution when solving the full viscous problem. The key for the contraction of the iteration operator lies in the following lemma:

Lemma 6 Assume that {z_c,\delta \lesssim \alpha}. Let {AirySolver_\infty(\cdot)} be the exact Airy solver of the {Airy(\cdot)} operator constructed as in Proposition 5 and let {f\in Y_{4,\eta}}. There holds the estimate:

\displaystyle \begin{aligned} \Big\| AirySolver_\infty( \epsilon \partial_x^4 f ) \Big\|_{X_{2,\eta'}} \le C_\eta\|f\|_{Y_{4,\eta}} \delta(1+|\log \delta|) (1+|z_c/\delta| ) \end{aligned}\ \ \ \ \ (18)

for arbitrary {\eta' < \eta}.

Proof: The rough idea is that the convolution can be computed as

\displaystyle G \star \epsilon \partial_z^4 f = - \epsilon \partial_z^3 G \star \partial_z f ,

in which {\epsilon \partial_z^3 G} is bounded and is localized near the critical layer of the size of order {\delta}. This indicates the bound by {\delta \log \delta} as stated in the estimate (18). The factor {1 + |z_c/\delta|} is precisely due to the linear growth in {z} in the Green kernel {G(x,z)}. We refer to the paper, Section 5, for details of the proof. \Box

Having provided the estimates on {Ray_\alpha^{-1} } and {Airy^{-1}}, and the convolution estimates, we can obtain the contraction of the {Iter} operator, defined by

\displaystyle Iter: = \underbrace{Airy^{-1}}_{\mbox{critical layer}} \circ \quad \underbrace{\epsilon \Delta^2_\alpha}_{\mbox{error}} \quad \circ \quad \underbrace{Ray_\alpha^{-1}}_{\mbox{inviscid}}.

Precisely, we can prove the following lemma:

Lemma 7 For {g \in X_{2,\eta}}, the {Iter(\cdot)} operator is a well-defined map from {X_{2,\eta}} to {X_{2,\eta}}. Furthermore, there holds

\displaystyle \| Iter(g)\|_{X_{2,\eta}} \le C \delta(1+|\log \delta|) (1+|z_c/\delta| ) \|g\|_{X_{2,\eta}},\ \ \ \ \ (19)

for some universal constant {C}.

8.4. Slow modes

In this paragraph we explicitly compute the boundary contribution of the first terms in the expansion of the slow Orr-Sommerfeld modes, which are obtained from the Rayleigh solutions:

\displaystyle \begin{aligned} \phi_s(z;c) &= \phi_{Ray}(z;c) + Airy^{-1} (\epsilon \Delta_\alpha \phi_{Ray})(z;c) + \cdots \end{aligned}\ \ \ \ \ (20)

in which the second term is obtained by the interation via the Iter operator, plus higher order terms. We recall that the Rayleigh solution, again obtained via a perturbative analysis, is of the form:

\displaystyle \phi_{Ray} (z;c)= e^{-\alpha z} (U-c + \mathcal{O}(\alpha)).

It is crucial to note that the possible {z\log z} singularity in the Rayleigh solution arises only at the order of {\alpha}. That is, we apply the Airy smoothing operator, Lemma 6, precisely to the {\mathcal{O}(\alpha)} term, yielding

\displaystyle \| Airy^{-1} (\epsilon \Delta_\alpha \phi_{Ray})\|_\eta \le C \epsilon + C\alpha \delta(1+|\log \delta|) (1+|z_c/\delta| ).

This yields at once the following lemma:

Lemma 8 Let {\phi_s} be the slow mode constructed above. For small {z_c, \alpha, \delta}, such that {\delta \lesssim \alpha} and {z_c\approx \alpha}, there hold

\displaystyle \begin{aligned} \frac{\phi_s(0;c)}{\partial_z\phi_s(0;c)} &= \frac{1}{U'_0}\Big[ U_0 - c + \alpha \frac{(U_+-U_0)^2}{U'_0} + \mathcal{O}(\alpha^2\log \alpha) \Big] . \end{aligned} \ \ \ \ \ (21)

8.5. Fast modes

Similarly, the fast modes are constructed as a perturbation from the second primitive Airy solutions:

\displaystyle \phi_{f,0}(z) : = \gamma_0 Ai(2,\delta^{-1}\eta(z)) , \qquad \gamma_0 := Ai(2,\delta^{-1}\eta(0))^{-1}.

Here, {\gamma_0} is to normalize the possible blow-up value of {Ai(2,\cdot)} on the boundary {z=0}, since {\delta^{-1} \eta(0) \approx e^{i 7\pi /6} |z_c/\delta|} could be arbitrarily large. By construction, there holds the following expansion of the fast mode {\phi_f} on the boundary {z=0}:

\displaystyle \phi_f(0) = \phi_{f,0}(0) + \mathcal{O}(\delta), \qquad \phi'_f(0) = \phi'_{f,0}(0) + \mathcal{O}(1).

By definition, we have { \phi_{f,0}(0) = 1} and

\displaystyle \phi'_{f,0}(0) = \delta^{-1} {Ai(1,\delta^{-1} \eta(0)) \over Ai(2,\delta^{-1} \eta(0)) } .

In the study of the linear dispersion relation, we are interested in the ratio {\phi_f/\phi'_f}. From the above estimates on {\phi_f(0)} and {\phi'_{f}(0)}, it follows at once that

\displaystyle {\phi_{f}(0) \over \phi'_{f}(0)} = \frac{\delta C_{Ai}(\delta^{-1} \eta(0))}{1+\mathcal{O}( \delta) C_{Ai}(\delta^{-1} \eta(0)) } ( 1 + \mathcal{O}(\delta)), \qquad C_{Ai} (Y):= {Ai(2,Y) \over Ai(1,Y) } .

As will be calculated below, {\delta C_{Ai}(\delta^{-1} \eta(0)) \approx \delta (1+|\eta(0)/\delta|)^{-1/2} \ll 1}. Hence, the above ratio is estimated by

\displaystyle {\phi_{f}(0) \over \phi'_{f}(0)} =\delta C_{Ai}(\delta^{-1} \eta(0)) ( 1 + \mathcal{O}(\delta)) . \ \ \ \ \ (22)

Here, we recall that {\delta = e^{-i \pi / 6} (\alpha R U_c')^{-1/3}}, and {\eta(0) = - z_c + \mathcal{O}(z_c^2)}. Therefore, we are interested in the ratio {C_{Ai}(Y)} for complex {Y = - e^{i \pi /6}y}, for {y} being in a small neighborhood of { \mathbb{R}^+}. Without loss of generality, in what follows, we consider {y \in \mathbb{R}^+}. Directly from the asymptotic behavior of the Airy functions, we obtain the following lemma:

Lemma 9 Let {C_{Ai}(\cdot)} be defined as above. Then, {C_{Ai}(\cdot)} is uniformly bounded on the ray {Y = e^{7i\pi/6} y} for {y \in \mathbb{R}^+}. In addition, there holds

\displaystyle C_{Ai}(- e^{i \pi /6} y) = - e^{ 5i \pi / 12} y^{-1/2} (1+\mathcal{O}(y^{-3/2}))

for all large {y\in \mathbb{R}^+}. At {y = 0}, we have { C_{Ai} (0) = - 3^{1/3} \Gamma(4/3).}

This yields at once the following estimate on the ratio (22):

Lemma 10 As long as {z_c/\delta} is sufficiently large, there holds

\displaystyle {\phi_{f}(0) \over \phi'_{f}(0)} = - e^{\pi i/4} |\delta| |z_c/\delta|^{-1/2} (1+\mathcal{O}(|z_c/\delta|^{-3/2})) \ \ \ \ \ (23)

In particular, the imaginary part of {\phi_f / \phi'_f} becomes negative when {z_c/\delta} is large (the ratio has a positive imaginary part when {z_c/\delta} is small).

8.6. Linear dispersion relation

Any linear combination of the slow and fast modes is an exact solution to the Orr-Sommerfeld equation. The zero boundary conditions yield the dispersion relation:

\displaystyle \frac{\phi_{s}(0; \alpha,\epsilon,c)}{ \phi'_{s}(0; \alpha,\epsilon,c)} = \frac{\phi_{f}(0; \alpha,\epsilon,c)}{\phi'_{f}(0; \alpha,\epsilon,c) }. \ \ \ \ \ (24)

We shall show that for some ranges of {(\alpha,\epsilon)}, the dispersion relation yields the existence of unstable eigenvalues {c}. By Lemmas 8 and 10, the linear dispersion relation (24) simply becomes

\displaystyle \begin{aligned} \Big[ U_0 - c + \frac{\alpha (U_+-U_0)^2 }{U'_0} + \mathcal{O}(\alpha^2\log \alpha ) \Big] = - e^{\pi i/4} |\delta| |z_c/\delta|^{-1/2} (1+\mathcal{O}(|z_c/\delta|^{-3/2})). \end{aligned}\ \ \ \ \ (25)

We are interested in the region where {z_c/\delta} is large. The dispersion relation yields

\displaystyle | U_0 - c| \le C \alpha + C \delta (1+|z_c/\delta|)^{-1/2}.\ \ \ \ \ (26)

Hence as {\alpha, \epsilon, \delta \rightarrow 0}, the eigenvalue {c} converges to {U_0} and so {z_c \approx \alpha}, from the equation {U(z_c) = c}. The existence of a unique {c = c(\alpha,\epsilon)} near {c_0 = U_0} so that the linear dispersion (25) holds, when {\alpha, \epsilon} are sufficiently small, follows easily from the Implicit Function Theorem.

8.6.1 Lower stability branch: {\alpha_\mathrm{low} \approx R^{-1/4}}

Let us consider the case {\alpha = A R^{-1/4}}, for some constant {A}. We recall that {\delta \approx (\alpha R)^{-1/3} = A^{-1/3} R^{-1/4}}. That is, {\alpha \approx \delta} for fixed constant {A}. In addition, since {z_c\approx \alpha}, we have

\displaystyle z_c/\delta \quad \approx\quad A^{4/3}.\ \ \ \ \ (27)

Thus, we are in the case that the critical layer goes up to the boundary with {z_c/\delta} staying bounded in the limit {\alpha,\epsilon \rightarrow 0}. We obtain the following lemma.

Lemma 11 Let {\alpha = A R^{-1/4}}. For {R} sufficiently large, there exists a critical constant {A_{c}} so that the eigenvalue {c = c(\alpha,\epsilon)} has its imaginary part changing from negative (stability) to positive (instability) as {A} increases past {A = A_c}. In particular,

\displaystyle \Im c \quad\approx \quad A^{-1} R^{-1/4}.

Proof: The imaginary part of the dispersion relation (25) yields

\displaystyle (-1 + \mathcal{O}(\alpha)) \Im c + \mathcal{O}(\alpha^2 \log \alpha) = \mathcal{O}(\delta (1+|z_c/\delta|)^{-1/2}).\ \ \ \ \ (28)

which gives {\Im c = \mathcal{O}(\delta (1+|z_c/\delta|)^{-1/2})} and so {\Im c \approx A^{-1} R^{-1/4}}. Next, also from Lemma 10, the right-hand side is positive when {z_c/\delta} is small, and becomes negative when {z_c/\delta \rightarrow \infty}. Consequently, together with (27), there must be a critical number {A_c} so that for all {A > A_c}, the right-hand side is positive, yielding the lemma as claimed. \Box

8.6.2. Intermediate zone: {R^{-1/4} \ll \alpha \ll R^{-1/6}}

Let us now turn to the intermediate case when

\displaystyle \alpha = A R^{-\beta}

with {1/10 < \beta < 1/4}. In this case {\delta \approx \alpha^{-1/3} R^{-1/3} \approx A^{-1/3} R^{\beta/3 - 1/3}} and hence {\delta \ll \alpha}. That is, the critical layer is away from the boundary: {\delta \ll z_c}. We have the following lemma.

Lemma 12 Let {\alpha = A R^{-\beta}} with {1/6<\beta<1/4}. For arbitrary fixed positive {A}, the eigenvalue {c = c(\alpha,\epsilon)} always has positive imaginary part (instability) with

\displaystyle \Im c \quad \approx\quad A^{-1 }R^{\beta-1/2}.

Proof: As mentioned above, {z_c/\delta} is unbounded in this case. Since {z_c \approx \alpha}, we indeed have

\displaystyle z_c/\delta \quad \approx\quad A^{4/3} R^{(1-4\beta)/3} \rightarrow \infty,

as {R \rightarrow \infty} since {\beta <1/4}. By Lemma 10, the right hand side of the dispersion relation reads

\displaystyle \Im\Big( {\phi_{f}(0) \over \phi'_{f}(0)} \Big) = \mathcal{O}(\delta (1+|z_c/\delta|)^{-1/2}) \approx A^{-1} R^{\beta -1/2}, \ \ \ \ \ (29)

which is positive, since {z_c/\delta \rightarrow \infty}. It is crucial to note that in this case { \alpha^2 \log \alpha \approx R^{-2\beta} \log R,} which can be neglected in the dispersion relation (28) as compared to the size of the imaginary part of {\phi_f/\phi'_f}. This yields the lemma. \Box

8.6.3 Upper stability branch: {\alpha_\mathrm{up} \approx R^{-1/6}}

The upper branch of marginal stability is more delicate to handle. This is the case when the term of order {\alpha^2 \log\alpha} on the left hand side of the dispersion relation has the same order as the right-hand side, and hence it is not clear whether {\Im c} remains positive. We expect that the Rayleigh solutions will dominate when {\alpha \gg \alpha_{\mathrm{up}}}, and the imaginary part of {c} will change from positive (instability) to negative (stability).

Print

Leave a Reply