The case in which the water has infinite depth is considered, the effect of surface tension as well.
Problem Prescription
Disturbance: pressure point, and finite depth: \(h\)
velocity potential \[ \begin{equation} \Phi=U x+\phi \end{equation} \] Laplace Equation \[ \begin{equation} \phi_{x x}+\phi_{y y}+\phi_{z z}=0, \end{equation} \] boundary condition \[ \begin{equation} \begin{aligned} &\phi_z=0 \quad \text { on } \quad z=-h\\ &P \delta(x) \delta(y)+\frac{1}{2} \rho\left(U^2+2 U \phi_x\right)+\rho g \eta-T\left(\eta_{x x}+\eta_{y y}\right)=\text { constant }=\frac{1}{2} \rho U^2 \quad \text { on } \quad z=0 \end{aligned} \end{equation} \] Kinematic condition \[ \begin{equation} \frac{\mathrm{d} \eta}{\mathrm{~d} t}=U \eta_x=\phi_z \end{equation} \] Using Fourier Transform, we have \[ \begin{equation} \eta(x, y, z)=\int_{-\infty}^{\infty} \int_{-\infty}^{\infty} \mathrm{e}^{\mathrm{i}(\alpha x+\beta y)} \frac{F(\alpha, \beta)}{G(\alpha, \beta)} \mathrm{d} \alpha \mathrm{~d} \beta, \end{equation} \] where \[ \begin{equation} \begin{gathered} F(\alpha, \beta)=\frac{P k \sinh \{k(h+z)\}}{4 \pi^2 \rho \cosh (k h)} ; \\ G(\alpha, \beta)=\alpha^2-k\left(1+\sigma k^2\right) \tanh (k h) ; \\ k=\sqrt{\alpha^2+\beta^2} \end{gathered} \end{equation} \] It is evident form inspection that \(\alpha =k_x,\;\beta=k_y\), and \(\sigma=T/\rho\).
Implementation steps
Rotating \(\alpha\) and \(\beta\) axes through an angle \(\theta\), where \(x=r\cos \theta,\;y=r\sin \theta\), for the particular direction \(\mathbf{r}=(x,y)\) along which we evaluate \(\eta\) \[ \begin{equation} \eta=\int_{-\infty}^{\infty} \mathrm{d} \bar{\beta} \int_{-\infty}^{\infty} \mathrm{e}^{\mathrm{i} \bar{\alpha} r} \frac{F(\bar{\alpha}, \bar{\beta})}{G(\bar{\alpha}, \bar{\beta})} \mathrm{d} \bar{\alpha}, \end{equation} \] where \(\alpha=\bar{\alpha} \cos \theta-\bar{\beta} \sin \theta ; \quad \beta=\bar{\alpha} \sin \theta+\bar{\beta} \cos \theta\).
--- note -------------------------------------------------
? why is there only \(\bar{\alpha}\)? how do we vanish \(\bar\beta\) in exponential term \(\mathrm{e}^{\mathrm{i} \bar{\alpha} r}\)
✏️A: \[ \alpha x+\beta y=(\bar{\alpha} \cos \theta-\bar{\beta} \sin \theta) r \cos \theta+(\bar{\alpha} \sin \theta+\bar{\beta} \cos \theta) r \sin \theta . \] collect the \(\bar{\alpha}\) and \(\bar{\beta}\) terms \[ \begin{aligned} \alpha x+\beta y & =\bar{\alpha} r\left(\cos ^2 \theta+\sin ^2 \theta\right)+\bar{\beta} r(-\sin \theta \cos \theta+\cos \theta \sin \theta) \\ & =\bar{\alpha} r+\bar{\beta} r(0) \\ & =\bar{\alpha} r . \end{aligned} \] Also, regarding the differential term, it follows that \[ \begin{aligned} & d \alpha=\cos \theta d \bar{\alpha}-\sin \theta d \bar{\beta} \\ & d \beta=\sin \theta d \bar{\alpha}+\cos \theta d \bar{\beta} \end{aligned} \] subsequently \[ \begin{aligned} d \alpha \wedge d \beta & =(\cos \theta d \bar{\alpha}-\sin \theta d \bar{\beta}) \wedge(\sin \theta d \bar{\alpha}+\cos \theta d \bar{\beta}) \\ & =-\sin ^2 \theta d \bar{\beta} \wedge d \bar{\alpha}+\cos ^2 \theta d \bar{\alpha} \wedge d \bar{\beta} \\ & =d \bar{\alpha} \wedge d \bar{\beta} \end{aligned} \]
The integrals depends on contributions from the singularities. Here it is \[ \alpha^2=k \tanh (k h) \] when surface tension is neglected, and \(h\) is infinite, \[ \begin{aligned} & \alpha^2=\sqrt{\alpha^2+\beta^2} \\ & \alpha^4-\alpha^2-\beta^2=0 \\ & \alpha^2=\frac{1 \pm \sqrt{1+4 \beta^2}}{2} \\ & \alpha^2 \geq 0 \Rightarrow \\ & \alpha^2=\frac{1+\sqrt{1+4 \beta^2}}{2} \\ & \alpha= \pm \frac{1}{\sqrt{2}}\left[1+\left(1+4 \beta^2\right)^{\frac{1}{2}}\right]^{\frac{1}{2}} \end{aligned} \]
Because of using Cauchy principal value of the integral, it leads to gravity waves appearing ahead of the pressure point, and some means must be found to give the correct physical solution. Therefore to solve this, we apply a radiation condition which allow the pressure increase exponentially with time like \(e^{\epsilon t}(\epsilon>0)\). Then \(G\) is replaced by \[ \begin{equation} G-\mathrm{i} 2 \alpha \epsilon-\epsilon^2 \end{equation} \] the poles are now at complex values of \(\bar{\alpha}\). We next shift the path of integration to the line \(\Im \bar{\alpha}=H>0\). This introduces a factor \(\mathrm{e}^{-H r}\)
--- note ---------------------------------------------
By shifting the integration contour to the line where \(\operatorname{Im} \bar{\alpha}=H\) \[ \bar{\alpha}=x+i H, \quad x \in(-\infty, \infty), \]
the differential becomes \(d \bar{\alpha}=d x\). Substituting this transformation into the integral yields:
\[ \int_{-\infty}^{\infty} e^{i \bar{\alpha} r} \frac{F(\bar{\alpha}, \bar{\beta})}{G(\bar{\alpha}, \bar{\beta})} d \bar{\alpha}=\int_{-\infty}^{\infty} e^{i(x+i H) r} \frac{F(x+i H, \bar{\beta})}{G(x+i H, \bar{\beta})} d x \]
The exponential term can be decomposed as: \[ e^{i(x+i H) r}=e^{i x r} e^{-H r} \]
This factorization reveals that \(e^{Hr}\) emerges as a constant multiplier independent of \(x\). Thus, the translated contour integral reduces to: \[ e^{-H r} \int_{-\infty}^{\infty} e^{i x r} \frac{F(x+i H, \bar{\beta})}{G(x+i H, \bar{\beta})} d x \]
This derivation clarifies the origin of the \(e^{Hr}\) term.
into the integral, making it negligible, but we must add on \(2 \pi i\) times the sum of the residues at poles with \(0<\Im \bar{\alpha}<H\).
Singularities
Three canonical categories of singularities: Removable, Pole, and Branch Points.
Singularities | Topological Property | Operational Identification | Expansion Structure |
---|---|---|---|
Removable | Defect removable by analytic continuation | \(\operatorname{Res}\left(f, z_0\right)=0\) | Regular Taylor expansion |
Pole of Order \(m\) | Localized divergence | \(\lim _{z \rightarrow z_0} \frac{1}{\left(z-z_0\right)^m f(z)} \neq 0 \text { (exists) }\) | Laurent series terminating at \(-\left(z-z_0\right)^{-n}\) |
Branch Singularity | Path-dependent function values | \(\oint_\gamma f(z) d z \neq 0\) for loops encircling \(z_0\) | Requires branch cut; no single-valued expansion |
A residue exists if and only if a function’s Laurent expansion about \(z_0\) contains a finite \((z−z_0)^{−1}\)term. This definition fails when:
- Removable singularities (all \(\left(z-z_0\right)^{-n}\) coefficients vanish),
- Essential singularities (non-terminating negative-power terms),
- No single-valued Laurent expansion exists (branch points).
Residue theorem:
Suppose \(f\) is holomorphic within the region enclosed by a simple closed curve \(C\), with the exception of isolated singularities at points \(\left\{a_k\right\}\). Under these conditions:
\[ \oint_C f(z) d z=2 \pi i \sum_k \operatorname{Res}\left(f, a_k\right) \]
explanation on singularities in complex analysis
Isolated Singularity
Definition: A point \(z_0\) is called an isolated singularity of \(f(z)\) if \(f(z)\) is analytic in a punctured neighborhood \(0 < |z - z_0| < R\) but not at \(z_0\).
Subclasses:
- Removable Singularity
- Pole
- Essential Singularity
Discriminant: Classified by the structure of negative-power terms in the Laurent series expansion:
- Removable Singularity: No negative-power terms (all coefficients \(a_{-n} = 0\)).
- Pole: Finite negative-power terms (highest negative power: \((z - z_0)^{-n}\), \(n \in \mathbb{N}^+\)).
- Essential Singularity: Infinitely many negative-power terms.
Removable Singularity
Definition: A singularity \(z_0\) is removable if there exists an analytic function \(g(z)\) such that \(f(z) = g(z)\) for all \(z \neq z_0\) in a neighborhood of \(z_0\).
Key Properties:
- \(\displaystyle \lim_{z \to z_0} f(z)\) exists and is finite.
- The singularity is eliminated by redefining \(f(z_0) = \lim_{z \to z_0} f(z)\). Example: \(f(z) = \frac{\sin z}{z}\) has a removable singularity at \(z = 0\).
Pole
Definition: A singularity \(z_0\) is a pole of order \(m\) if there exists \(m \in \mathbb{N}^+\) such that \((z - z_0)^m f(z)\) is analytic and non-zero at \(z_0\). Key Properties:
- \(\displaystyle \lim_{z \to z_0} |f(z)| = +\infty\).
- The Laurent expansion terminates at \((z - z_0)^{-m}\). Example: \(f(z) = \frac{1}{(z - 1)^3}\) has a third-order pole at \(z = 1\).
Essential Singularity
Definition: A singularity \(z_0\) is essential if it is isolated but neither removable nor a pole.
Key Properties (Weierstrass-Casorati Theorem):
- The image of any punctured neighborhood of \(z_0\) is dense in \(\mathbb{C} \cup \{\infty\}\).
- The Laurent series contains infinitely many negative-power terms. Example: \(f(z) = e^{1/z}\) has an essential singularity at \(z = 0\).
Branch Singularity
Definition: A point \(z_0\) is a branch singularity (or branch point) if analytic continuation of \(f(z)\) around a closed path encircling \(z_0\) results in a multivalued function. Key Properties:
- Non-isolated: Multivaluedness persists in any neighborhood of \(z_0\).
- Requires branch cut(s) to define single-valued branches.
Example: \(f(z) = \sqrt{z}\) has a branch point at \(z = 0\).
Distinctions and Connections
Core Differences
Singularity Type | Isolated? | Laurent Series Existence | Typical Behavior | Residue Defined? |
---|---|---|---|---|
Removable | ✔️ | Yes (no negative powers) | Finite limit | Yes (trivially zero) |
Pole | ✔️ | Yes (finite negative powers) | Diverges to infinity | Yes (non-zero) |
Essential | ✔️ | Yes (infinite negative powers) | Dense in \(\mathbb{C} \cup \{\infty\}\) | Yes (may be non-zero) |
Branch | ❌ | No (multivalued) | Multivaluedness under monodromy | No |
Connections
- Isolated Singularity Family: Removable singularities, poles, and essential singularities are subclasses of isolated singularities, distinguished by Laurent series structures.
- Uniqueness of Branch Singularities:
- Non-isolated nature precludes Laurent series analysis; requires Riemann surfaces for multivaluedness.
- Topological distinction: Branch singularities alter the global structure of the complex plane, while isolated singularities are local phenomena.
- Non-isolated nature precludes Laurent series analysis; requires Riemann surfaces for multivaluedness.
Comparative Examples
- Removable vs. Pole:
- \(f(z) = \frac{\sin z}{z}\) (removable): \(\lim_{z \to 0} f(z) = 1\).
- \(f(z) = \frac{1}{z^2}\) (pole): \(\lim_{z \to 0} |f(z)| = +\infty\).
- \(f(z) = \frac{\sin z}{z}\) (removable): \(\lim_{z \to 0} f(z) = 1\).
- Pole vs. Essential Singularity:
- \(f(z) = \frac{1}{z^3}\) (pole): Laurent series contains only \(z^{-3}\).
- \(f(z) = e^{1/z}\) (essential): Laurent series \(\sum_{n=0}^\infty \frac{z^{-n}}{n!}\).
- \(f(z) = \frac{1}{z^3}\) (pole): Laurent series contains only \(z^{-3}\).
- Essential vs. Branch Singularity:
- \(e^{1/z}\) (essential): Locally dense but single-valued.
- \(\sqrt{z}\) (branch): Monodromy \(\sqrt{z} \to -\sqrt{z}\) upon encircling \(z = 0\).
- \(e^{1/z}\) (essential): Locally dense but single-valued.
Applications
- Residue Theorem: Applies only to isolated singularities (poles and essential singularities); poles yield straightforward residue calculations.
- Integral Evaluation:
- Branch singularities require contour deformation around branch cuts (e.g., keyhole contours).
- Essential singularities may involve infinite series for residue computation.
- Branch singularities require contour deformation around branch cuts (e.g., keyhole contours).
- Geometric Function Theory:
- Branch singularities correspond to Riemann surface branching.
- Essential singularities reflect chaotic behavior in complex dynamics.
- Branch singularities correspond to Riemann surface branching.
Summary
- Isolated Singularities (removable, poles, essential) are classified by local Laurent series properties.
- Branch Singularities are global, multivalued phenomena requiring topological tools (branch cuts, Riemann surfaces).
- Essential singularities represent the "most pathological" isolated singularities, while branch singularities exist outside the isolated singularity framework.
--- note --------------------------------
Q: why is \(\sqrt{z}\) single-valued function?
A:
- Commencing with the definition of the complex square root
- For a complex number \(z = re^{i\theta}\) (expressed in exponential form, where \(r\) is the modulus and \(\theta\) the argument), its square root is defined as \(\sqrt{z} = \sqrt{r}e^{i\theta/2}\).
- The argument \(\theta\) of a complex number exhibits periodicity, i.e., \(\theta\) and \(\theta + 2k\pi\) (\(k \in \mathbb{Z}\)) represent equivalent angular orientations. For instance, when \(k = 1\), \(\theta + 2\pi\) corresponds to the same direction as \(\theta\).
- Analysis of the multivalued nature of the square root function
- Substituting into the square root expression yields \(\sqrt{z}=\sqrt{r} e^{i(\theta+2 k \pi) / 2}=\sqrt{r} e^{i \theta / 2} e^{i k \pi}\).
- For distinct integer values of \(k, e^{i k \pi}\) assumes two distinct cases:
- When \(k\) is even (e.g., \(k=0,2,4, \ldots\) ), \(e^{i k \pi}=1\), resulting in \(\sqrt{z}=\sqrt{r} e^{i \theta / 2}\).
- When \(k\) is odd (e.g., \(k=1,3,5, \ldots\) ), \(e^{i k \pi}=-1\), yielding \(\sqrt{z}=-\sqrt{r} e^{i \theta / 2}\).
- Thus, for any nonzero \(z\), the square root function bifurcates into two distinct values, demonstrating its multivalued character.
- Multivaluedness through the lens of analytic continuation
- Consider a circular path encircling the origin, starting from the positive real axis. As \(z\) traverses this path with \(\theta\) increasing from \(0\) to \(2\pi\), the argument of the square root evolves from \(0\) to \(\pi\).
- Upon completing a full rotation (\(\theta \rightarrow \theta + 2\pi\)), the square root function does not revert to its initial value but instead assumes the negative of its original value. This monodromy phenomenon underscores the multivaluedness inherent to the square root function.
- Branch cuts as a resolution to multivaluedness
- To enforce single-valuedness, a branch cut is introduced. For the square root function, a conventional choice is the ray extending from the origin along the negative real axis to infinity.
- This cut partitions the complex plane into two regions. By restricting the argument to a principal branch (e.g., \(-\pi < \theta \leq \pi\)), the square root becomes single-valued within each region. Crossing the branch cut induces a discontinuous transition between branches, further illustrating the multivalued nature.
surface displacement without surface tension
Evaluating the inner integral, we get \[ \begin{equation} \eta=\int_{-\infty}^{\infty}\left\{2 \pi \mathrm{i} \Sigma \mathrm{e}^{\mathrm{i} \bar{\alpha} r} \frac{F}{\partial G / \partial \bar{\alpha}}+o\left(r^{-N}\right)\right\} \mathrm{d} \bar{\beta}, \end{equation} \]
Notes
Q: how to get eqn (9)
A: we use Residue theorem, firstly, the pole is \[ \begin{aligned} & \quad \alpha=\bar{\alpha}_0+\frac{i \epsilon 2(\bar{\alpha} \cos \theta-\bar{\beta} \sin \theta)}{\partial G / \partial \bar{\alpha}} \\ & \oint e^{i \bar{\alpha} r} \frac{F}{G} d \bar{\alpha} \\ & =2 \pi i \operatorname{Res}\left[f, a_i\right]+0 \\ & =2 \pi i \lim _{\bar{\alpha} \rightarrow \bar{\alpha}_0}\left(\bar{\alpha}-\bar{\alpha}_0\right) e^{i \bar{\alpha} r} \frac{F}{G} \\ & =2 \pi i \lim _{\bar{\alpha} \rightarrow \bar{\alpha}_0}\left(\bar{\alpha}-\bar{\alpha}_0\right) e^{i \bar{\alpha} r} \frac{F\left(\bar{\alpha}_0\right)}{\left(\bar{\alpha}-\bar{\alpha}_0\right) \partial G / \partial \bar{\alpha}} \\ & =2 \pi i e^{i \bar{\alpha}_0 r} \frac{F\left(\bar{\alpha}_0\right)}{\partial G / \partial \bar{\alpha}} \end{aligned} \]
Q: how to get the \(o(r^-{N})\)
A: \[ \int_{-\infty}^{\infty} e^{i \bar{\alpha} r }\frac{F}{G} d \bar{\alpha}=\oint e^{i \bar{\alpha} r }\frac{F}{G} d \bar{\alpha}-\int_{\text {semicircle }} e^{i \bar{\alpha} r} \frac{F}{G} d \bar{\alpha} \] now we estimate the integral \[ \int_{\text {Semicircle }} e^{i \bar{a} r} \frac{F}{G} d \bar{\alpha} \] \(\frac{F}{G}\) grows at polynomially in \(|\bar{a}|\). and assuming \(\bar{\alpha}=\xi+i \lambda\) \[ \begin{aligned} \left|e^{i \bar{\alpha} r} \frac{F}{G}\right| &\leqslant\left|e^{-r \lambda} e^{i \xi r} \frac{F}{G}\right|\\ &\leqslant\left||a|^m e^{-r \lambda}\right| \end{aligned} \] as \(r \rightarrow \infty,|a|^m \leq C\) the remainder
\[ \begin{aligned} |a|^m e^{-r \lambda} & \leqslant C e^{-r \lambda} \\ & \leqslant C \;\mathscr{o}\left(r^{-N}\right) \end{aligned} \]
because an exponential decay is faster than any algebraic power.
The inner integral in approached by contour integration along the lines (this refers to the sectorial integration domain encompassing singularities), but the only other contribution is from a branch point at \(\bar{\alpha} =i\abs{\beta}\), and this term decays like \(e^{-\abs{\beta}r}\), which can be neglected.
--- note --------------------------------------------------
Q: branch point \(\bar{\alpha}=i\abs{\beta}\)
A: The square root introduces multivaluedness in \(k\), rendering it a two-valued function. The condition \(\bar{\alpha}^2+\) \(\bar{\beta}^2=0\), which yields \[ \bar{\alpha}= \pm i|\bar{\beta}| \]
identifies the branch points of the equation \(k=0\). Conventionally, these branch points are joined by a branch cut along the imaginary axis, inducing a sign change in \(k\) when crossing the cut.
we put \(s=\bar{\alpha}-i|\bar{\beta}|\), we may rewrite \(s=s e^{ \pm i \pi}\), and, \(s=0\) is the branch points, then \[ \begin{equation} \begin{aligned} \bar{\alpha}^2+\bar{\beta}^2&=(i \abs{\beta})^2+2 i b s e^{ \pm i \pi}+O\left(s^2\right)+\abs{\beta}^2\\ &=2 i b s+O\left(s^2\right) \end{aligned} \end{equation} \] Neglecting higher-order infinitesimal terms, it follows that \[ \begin{equation} k \equiv \sqrt{\bar{\alpha}^2+\bar{\beta}^2} \simeq \sqrt{2 b} s^{1 / 2} e^{\pm i \pi / 4} \end{equation} \] Increasing \(\arg s\) by \(2\pi\) flips the sign of \(e^{\pm i \pi / 4}\) (e.g. from \(-\pi\) to \(\pi\)), accounting for the sign difference across the branch point.
Next, we consider \(e^{i\bar{\alpha}r}\), traversing the branch cut, yields: \[ \begin{equation} e^{i \bar{\alpha} r}=e^{i\left(i \abs{\beta}+s e^{ \pm i \pi}\right) r}=e^{-\abs{\beta} r} e^{-s r} e^{ \pm 0^{+} r} \end{equation} \]
- shared prefactor \(e^{-\abs{\beta}r}\) which is the dominant exponential decay
- the phase difference between upper/lower sides is infinitesimal
The main contribution for large \(r\) comes from the points according to stationary phase method where \[ \begin{equation} \mathrm{d} \bar{\alpha} / \mathrm{d} \bar{\beta}=0 \quad(\mathrm{~d} \alpha / \mathrm{d} \beta=-\tan \theta=-y / x) \end{equation} \] At these points normal to the line \(G=0\) parallel to the \(\bar\alpha\) axis.
--- note --------------------------------------------------
- Q:❓ At these points normal to the line \(G=0\) parallel to the \(\bar\alpha\) axis.
A: Recall that the dispersion curve \[ \mathcal{C}: G(\bar{\alpha}, \bar{\beta})=0 \]
is a line in the two-dimensional \((\bar{\alpha}, \bar{\beta})\)-plane. Along that line we can regard \(\bar{\alpha}\) as a function of \(\bar{\beta}\) :
\[ (\bar{\alpha}, \bar{\beta})=(\bar{\alpha}(\bar{\beta}), \bar{\beta}), \quad \text { with } \quad \frac{d \bar{\alpha}}{d \bar{\beta}}=-\frac{G_{\bar{\beta}}}{G_{\bar{\alpha}}} . \] then normal to the line \(G=0\), then we consider tangent vector to the curve \[ \mathbf{t}=\left(\frac{d \bar{\alpha}}{d \bar{\beta}}, 1\right) \]
When \(d \bar{\alpha} / d \bar{\beta}=0\) this becomes \(\mathbf{t}=(0,1)\), i.e. it points straight up the \(\bar{\beta}\)-axis. A line whose tangent is vertical has its normal horizontal. In other words \[ \mathbf{n} \| \bar{\alpha} \text {-axis } \quad(\mathbf{n} \cdot \mathbf{t}=0) \]
- tangent vector
For a plane curve given implicitly by \(F(x,y)=0\), pick any regular point \((x_0,y_0)\)
where the gradient \(\nabla F(x_0,y_0)=(F_x,F_y)\neq(0,0)\). A tangent vector \(\mathbf{t}\) at that point must be orthogonal to the gradient, because the gradient points in the normal direction:
\[ \nabla F\cdot\mathbf{t}=0 \;\Longrightarrow\; F_x\,t_x+F_y\,t_y=0. \]
A convenient choice (unique up to a non‑zero scalar factor) is
\[ \mathbf{t}=(-F_y,\;F_x)\qquad\text{evaluated at the point }(x_0,y_0). \]
Equivalently, if \(F_y\neq 0\) we can write the slope form
\[ \frac{dy}{dx}=-\frac{F_x}{F_y}, \quad\text{so}\quad \mathbf{t}=(1,\,-F_x/F_y), \]
and if \(F_x\neq 0\) swap the roles. Any non‑zero scalar multiple of these vectors points along the tangent line to the curve at the chosen point.
Then rewrite \(\bar \alpha\) as Taylor series \[ \begin{equation} \bar{\alpha}=\bar{\alpha}_m+\frac{1}{2} \kappa\left(\bar{\beta}-\bar{\beta}_m\right)^2, \end{equation} \] where \(\kappa=\left(\mathrm{d}^2 \bar{\alpha} / \mathrm{d} \bar{\beta}^2\right)_m\) is the curvature of the line \(G=0\) (assumed non-zero here) and we get \[ \begin{equation} \eta \sim 2 \pi \mathrm{i} \sum_m \frac{F}{\partial G / \partial \bar{\alpha}} \exp \left(\mathrm{i} \bar{\alpha}_m r+\frac{1}{4} \pi \mathrm{i} \operatorname{sgn} \kappa\right) \sqrt{ \frac{2 \pi}{|\kappa| r}}+O\left(r^{-1}\right) \end{equation} \]
--- note -------------------------------------------------
Q: Where does the \(O(r^{-1})\) error term originate?
A: Outside the small neighbourhoods of the stationary points \(\varphi^{\prime}(\bar{\beta})\) is bounded away from zero. On that set we integrate by parts once: (path: non-stationary) \[ \int_{\text {non-stationary }} a e^{i r \varphi} d \bar{\beta}=\frac{1}{i r}\left[\frac{a}{\varphi^{\prime}} e^{i r \varphi}\right]_{\text {ends }}-\frac{1}{i r} \int e^{i r \varphi} \frac{d}{d \bar{\beta}}\left(\frac{a}{\varphi^{\prime}}\right) d \bar{\beta} . \]
- The bracketed end terms vanish because \(a(\bar{\beta})\) falls off quicker than any power as \(|\bar{\beta}| \rightarrow \infty\).
- The remaining integral is uniformly bounded, so the whole piece is \(O\left(r^{-1}\right)\).
at points \(m\), \(\partial G / \partial \bar{\alpha}= \pm|\nabla G|\).
proof:
we start from the derivatives \[ \frac{\partial(\bar{\alpha}, \bar{\beta})}{\partial(\alpha, \beta)}=\left(\begin{array}{cc} \cos \theta & -\sin \theta \\ \sin \theta & \cos \theta \end{array}\right)=R \] then \[ \begin{aligned} \nabla G & =\partial_{(\alpha, \beta)} G=\partial_{(\bar{\alpha}, \beta)} G \partial_{(\alpha, \beta)}(\bar{\alpha}, \bar{\beta}) \\ & =R \partial_{(\bar{\alpha}, \bar{\beta})} G \\ & =\binom{\partial_\bar{\alpha} G \cos \theta-\partial_{\bar{\beta}} G \sin \theta}{\partial_\bar{\alpha} G \sin \theta+\partial_{\bar{\beta}} G \cos \theta} \end{aligned} \] since \(\frac{d \bar{\alpha}}{d \bar{\beta}}=-\frac{\partial_{\bar\beta} G}{\partial_{\bar\alpha} G}=0\) then
\[ \partial_\bar\beta G=0 \]
therefore
\[ \begin{aligned} \nabla G & =\left(\partial_{\overline{\alpha}} G \cos \theta, \partial_{\overline{\alpha}} G \sin \theta\right) \\ & |\nabla G|=\left|\partial_{\overline{\alpha}} G\right| \\ \Rightarrow & \partial_\bar{\alpha} G= \pm|\nabla G| \end{aligned} \] Now we get the radiation condition \[ \frac{\partial G}{\partial \bar{\alpha}}=\frac{1}{\cos \theta} \frac{\partial G}{\partial \alpha}=\frac{1}{\sin \theta} \frac{\partial G}{\partial \beta} \] Subsequently, at the stationary points \(m\) \[ \kappa=-\frac{\partial^2 G}{\partial \bar{\beta}^2}/\frac{\partial G}{\partial \bar{\alpha}} \] at points \(m\). When the depth is infinite this is
\[ \alpha^2 \cos \theta /\left(2 \alpha^2-1\right)>0 \]
Notes
Proof:
\[ \frac{\bar{\alpha} \cos \theta-\bar{\beta} \sin \theta}{\partial G / \partial \bar{\alpha}}=\frac{\alpha \cos \theta}{\partial G / \partial \alpha}>0 \] A: we have \[ \alpha=\bar{\alpha} \cos \theta-\bar{\beta} \sin \theta \] and \[ \begin{aligned} \partial_\alpha G & =\frac{\partial \bar{\beta}}{\partial \alpha} \frac{\partial G}{\partial \bar{\beta}}+\frac{\partial \bar{\alpha}}{\partial \alpha} \frac{\partial G}{\partial \bar{\alpha}} \\ & =\sin \theta \underbrace{\partial_\bar{\beta} G}_0+\cos \theta \partial_\bar{\alpha} G \\ & =\cos \theta \partial_\bar{\alpha} G \end{aligned} \] Therefore \[ \frac{\bar{\alpha} \cos \theta-\bar{\beta} \sin \theta}{\partial G / \partial \bar{\alpha}}=\frac{\alpha \cos \theta}{\partial G / \partial \alpha} \] Proof: \[ \frac{\partial G}{\partial \bar{a}}=\frac{1}{\sin \theta} \frac{\partial G}{\partial \beta} \] A: since \[ \frac{d \alpha}{d \beta} =-\frac{\partial_\beta G}{\partial_\alpha G}=-\tan \theta \]
\(\Rightarrow\)
\[ \partial_\alpha G =\partial_\beta G / \tan \theta \] It follows that
\[ \begin{aligned} \partial \bar{\alpha} G & =\partial_{\alpha} G / \cos \theta \\ & =\partial_\beta G / \sin \theta \end{aligned} \] Proof:
\[ \alpha^2 \cos \theta /\left(2 \alpha^2-1\right)>0 \] When depth is infinite \[ G=\alpha^2-k, k=\sqrt{\alpha^2+\beta^2}, \] and we have \(G=0\), then we get \(k=\alpha^2\), and \[ \partial_\alpha G=2\alpha-\frac{\alpha}{k}=2\alpha-\frac{1}{\alpha} \] Thus, \[ \begin{aligned} & \alpha \cos \theta /\left(2 \alpha-\frac{1}{\alpha}\right) \\ & =\alpha^2 \cos \theta /\left(2 \alpha^2-1\right) \end{aligned} \]
From the foregoing analysis, \[ G(\alpha, \beta)=\alpha^2-k=0, \quad k=\sqrt{\alpha^2+\beta^2} \]
Set \(k=\alpha^2\) in the definition of \(k\) :
\[ k^2=\alpha^4=\alpha^2+\beta^2 \Longrightarrow \beta^2=\alpha^2\left(\alpha^2-1\right) . \] therefore \(\alpha^2-1 \geqslant 0\) and \(\cos \theta >0\).
Because of the symmetry of the curves \(G=0\) about the axes, it is clear that the points \(\left(\alpha_m, \beta_m\right)\) and \(\left(-\alpha_m,-\beta_m\right)\) have normals in the same direction.
Notes
Why \(G\) is symmetric
For the dispersion functions that appear in surface-wave theory \[ G(\alpha, \beta)=\alpha^2-k\left(1+\sigma k^2\right) \tanh k h, \quad k=\sqrt{\alpha^2+\beta^2} \]
\(G\) depends on \(\alpha\) and \(\beta\) only through the even combinations \(\alpha^2\) and \(\beta^2\).
Hence
\[ G(\alpha, \beta)=G(-\alpha, \beta)=G(\alpha,-\beta)=G(-\alpha,-\beta), \]
so the curve \(G=0\) is symmetric with respect to both coordinate axes and to the origin.
How the gradient transforms under the half-turn \[ (\alpha, \beta) \mapsto(-\alpha,-\beta) \]
Because \(G\) is even in each variable, its first derivatives are odd:
\[ G_\alpha(-\alpha,-\beta)=-G_\alpha(\alpha, \beta), \quad G_\beta(-\alpha,-\beta)=-G_\beta(\alpha, \beta) . \]
Thus
\[ \nabla G(-\alpha,-\beta)=-\nabla G(\alpha, \beta) . \]
The equation previously tells us that the gradients at the two points are anti-parallel; geometrically they lie on the same line through the two points. The normal line is determined only up to sign, so an overall factor \(-1\) makes no difference:
normal at \(\left(-\alpha_m,-\beta_m\right) \|\) normal at \(\left(\alpha_m, \beta_m\right)\).
Therefore, we can write \[ \begin{equation} \eta \sim-\frac{2^{\frac{5}{2}} \pi^{\frac{3}{2}}}{r^{\frac{1}{2}}} \sum_{m+} \frac{F \cos \theta}{\partial G / \partial \alpha(\sqrt{|\kappa|)}} \sin \left(\alpha x+\beta y+\frac{1}{4} \pi \operatorname{sgn} \kappa\right)+O\left(\frac{1}{r}\right), \end{equation} \]
where \(\kappa=-\frac{\partial^2 G}{\partial \bar{\beta}^2} / \frac{\partial G}{\partial \bar{\alpha}}\)
Notes
From previous analysis \[ \bar{\alpha}=\alpha \cos \theta+\beta \sin \theta \]
\(\Rightarrow\)
\[ \begin{aligned} r \bar{\alpha} & =a r \cos \theta+\beta r \sin \theta \\ & =a x+\beta y \end{aligned} \]
It follows that \[ \eta\;\sim\;2\pi i \sum_{m}\, \frac{F}{\partial_{\bar\alpha}G}\; \exp\!\bigl\{i(\alpha x+\beta y)+\tfrac{\pi i}{4}\operatorname{sgn}\kappa\bigr\}\; \sqrt{\frac{2\pi}{|\kappa|\,r}} \;+\;O(r^{-1}) \]
Q: Show \[ \eta\;\sim\; -\frac{2^{5/2}\pi^{3/2}}{r^{1/2}} \sum_{m+} \frac{F\cos\theta}{\displaystyle (\partial_\alpha G)\sqrt{|\kappa|}}\; \sin\!\Bigl(\alpha x+\beta y+\tfrac{\pi}{4}\operatorname{sgn}\kappa\Bigr) \;+\;O(r^{-1}). \]
A: Replace \(\partial_{\bar\alpha}G\) by \(\partial_\alpha G\), at every stationary point \(m\) we showed earlier that
\[ \partial_{\bar\alpha}G =\frac{1}{\cos\theta}\,\partial_\alpha G \]
Substituting this into the equation above; all denominators pick up a factor \(\cos\theta\):
\[ 2\pi i\;\frac{F}{\partial_{\bar\alpha}G} =2\pi i\; \frac{F\cos\theta}{\partial_\alpha G} \]
Because the level curve \(G=0\) is even in both \(\alpha\) and \(\beta\), each positive-\(\alpha\) stationary point \((\alpha_m,\beta_m)\) has a partner \((-\alpha_m,-\beta_m)\). Denoting the phase \(\Phi=\alpha x+\beta y\).
- Even factors: \(F\), \(|\kappa|\), \(\cos\theta\) are unchanged.
- Odd factor: \(\partial_\alpha G\) is odd, so \((\partial_\alpha G)_{-\alpha}=-\,(\partial_\alpha G)_{+\alpha}\).
- Second derivative: \(\kappa=-\partial_{\bar\beta}^2G/\partial_{\bar\alpha}G\) changes sign, hence \(\operatorname{sgn}\kappa\) changes sign.
Here, we show that \(\partial_{\bar\beta}^2G/\partial_{\bar\alpha}G\) changes sign
The two stationary points that come from the symmetry of the dispersion curve are \[ m^{+}:(\bar\alpha,\bar\beta)=(\bar\alpha_m,\bar\beta_m), \quad m^{-}:(\bar\alpha,\bar\beta)=(-\bar\alpha_m,-\bar\beta_m). \]
Because \(G\bigl(\bar\alpha,\bar\beta\bigr)\) depends only on \(\bar\alpha^{2},\bar\beta^{2}\) it is even in each of its arguments:
\[ G(-\bar\alpha,-\bar\beta)=G(\bar\alpha,\bar\beta). \]
Along \(G(\bar\alpha,\bar\beta)=0\) we have, by implicit differentiation,
\[ \frac{d\bar\alpha}{d\bar\beta} =-\frac{G_{\bar\beta}}{G_{\bar\alpha}} . \]
At a stationary point of the outer integral
\[ G_{\bar\beta}=0\qquad\Longrightarrow\qquad \frac{d\bar\alpha}{d\bar\beta}=0 . \]
Differentiate the preceding relation once more and then use \(G_{\bar\beta}=0\):
\[ \kappa\;=\; \frac{d^{2}\bar\alpha}{d\bar\beta^{2}} =-\frac{G_{\bar\beta\bar\beta}}{G_{\bar\alpha}}\; \tag{N1} \]
The numerator \(G_{\bar\beta\bar\beta}\) contains two \(\bar\beta\)-derivatives and is therefore even under \((\bar\alpha,\bar\beta)\mapsto(-\bar\alpha,-\bar\beta)\). The denominator \(G_{\bar\alpha}\) contains one \(\bar\alpha\)-derivative and is odd. Hence at the partner point
\[ G_{\bar\beta\bar\beta}\bigl(-\bar\alpha_m,-\bar\beta_m\bigr) = \;G_{\bar\beta\bar\beta}\bigl(\bar\alpha_m,\bar\beta_m\bigr),\\[6pt] G_{\bar\alpha }\bigl(-\bar\alpha_m,-\bar\beta_m\bigr) =-\,G_{\bar\alpha }\bigl(\bar\alpha_m,\bar\beta_m\bigr). \]
Insert these into equation \((N1)\):
\[ \kappa(m^{-}) =-\frac{G_{\bar\beta\bar\beta}}{-\,G_{\bar\alpha}} =+\frac{G_{\bar\beta\bar\beta}}{ G_{\bar\alpha}} =-\kappa(m^{+}). \]
The second derivative—and therefore its sign—reverses when one moves from the stationary point \((\bar\alpha_m,\bar\beta_m)\) to its symmetric partner \((-\bar\alpha_m,-\bar\beta_m)\):
\[ \kappa\bigl(-\bar\alpha_m,-\bar\beta_m\bigr) =-\kappa\bigl(\bar\alpha_m,\bar\beta_m\bigr)\; \]
Because \(\operatorname{sgn}\kappa\) flips, the phase shift \(\tfrac{\pi}{4}\operatorname{sgn}\kappa\) in the exponential (or sine) terms also changes sign, exactly as required in the derivation of the final far-field formula.
Q.E.D
Hence the two contributions for a given pair are
\[ \begin{aligned} \eta_{(+)} &=A\; e^{\,i(\Phi+\frac{\pi}{4}\operatorname{sgn}\kappa)},\\[2pt] \eta_{(-)} &= -A\; e^{-i(\Phi+\frac{\pi}{4}\operatorname{sgn}\kappa)}, \end{aligned} \qquad A\;=\;2\pi i\; \frac{F\cos\theta}{\partial_\alpha G}\, \sqrt{\frac{2\pi}{|\kappa|\,r}} . \]
Adding the pair, \[ \eta_{(+)}+\eta_{(-)} =2\pi i\frac{F\cos\theta}{\partial_\alpha G} \sqrt{\frac{2\pi}{|\kappa|\,r}}\; \bigl[e^{i\Theta}-e^{-i\Theta}\bigr], \qquad \Theta:=\Phi+\tfrac{\pi}{4}\operatorname{sgn}\kappa \]
\[ e^{i\Theta}-e^{-i\Theta}=2i\sin\Theta \;\;\Longrightarrow\;\; \eta_{(+)}+\eta_{(-)} =-\,4\pi \frac{F\cos\theta}{\partial_\alpha G} \sqrt{\frac{2\pi}{|\kappa|\,r}}\; \sin\Theta . \tag{N2} \]
Collect constants and sum only over \(m+\) Equation \((N2)\) already contains both points of each symmetric pair, so the remaining sum can be restricted to the "positive-\(\alpha\)" members, traditionally labelled \(m+\). Gather the numerical factors:
\[ 4\pi\sqrt{2\pi} =4\sqrt{2}\,\pi^{3/2}=2^{5/2}\pi^{3/2}. \]
Insert this into \((N2)\) and write \(\Theta=\alpha x+\beta y+\pi \operatorname{sgn}\kappa/4\):
\[ \eta \;\sim\; -\,\frac{2^{5/2}\pi^{3/2}}{r^{1/2}} \sum_{m+} \frac{F\cos\theta}{(\partial_\alpha G)\sqrt{|\kappa|}}\; \sin\!\Bigl(\alpha x+\beta y+\tfrac{\pi}{4}\operatorname{sgn}\kappa\Bigr) +O\!\bigl(r^{-1}\bigr), \] Q.E.D
we shall begin with the case of water of infinite depth, for this case \(G=0\), then we get
\[ \begin{equation} \begin{gathered} \alpha_m=\frac{1}{2 \sqrt{2}|\tan \theta|}\left\{4 \tan ^2 \theta+1 \pm \sqrt{\left(1-8 \tan ^2 \theta\right) }\right\}^{\frac{1}{2}} \\ \beta_m=-\alpha_m\left(2 \alpha_m^2-1\right) \tan \theta \end{gathered} \end{equation} \]
Notes
As previously demonstrated, the dispersion curve \(G=0\), the stationary–phase condition \(G_\beta=\tan\theta\,G_\alpha\), and the reality/radiation requirement (25) \(\alpha^2\ge1\).
For \(h\to\infty\) and \(\sigma=0\) \[ G(\alpha,\beta)=\alpha^{2}-k=0,\qquad k=\sqrt{\alpha^{2}+\beta^{2}} \]
gives
\[ k=\alpha^{2},\qquad \beta^{2}=k^{2}-\alpha^{2}=\alpha^{2}\bigl(\alpha^{2}-1\bigr). \tag{N3} \]
According to the derivatives of \(G\) \[ \begin{aligned} G_\alpha &=2\alpha-\frac{\alpha}{k} =\;\frac{2\alpha^{2}-1}{\alpha},\\ G_\beta &=-\frac{\beta}{k} =\;-\frac{\beta}{\alpha^{2}} \end{aligned}\tag{N4} \]
By applying stationary phase condition
\[ G_\beta=\tan\theta\,G_\alpha \quad\Longrightarrow\quad -\frac{\beta}{\alpha^{2}} =\tan\theta\;\frac{2\alpha^{2}-1}{\alpha}. \]
Multiply by \(\alpha^{2}\):
\[ \beta=-\alpha\bigl(2\alpha^{2}-1\bigr)\tan\theta, \tag{N5} \]
Insert \((N5)\) into \(\beta^{2}=\alpha^{2}(\alpha^{2}-1)\):
\[ \alpha^{2}\bigl(2\alpha^{2}-1\bigr)^{2}\tan^{2}\theta \;=\;\alpha^{2}(\alpha^{2}-1). \]
Divide by \(\alpha^{2}\;(\alpha\neq0)\)
\[ (2\alpha^{2}-1)^{2}\tan^{2}\theta=\alpha^{2}-1. \]
Therefore we obtain
\[ x=\alpha^{2} =\frac{4\tan^{2}\theta+1\;\pm\;\sqrt{\,1-8\tan^{2}\theta}}{\;8\tan^{2}\theta}. \]
Hence the pair
\[ \bigl(\alpha_m,\beta_m\bigr) =\left(\; \frac{1}{2\sqrt{2}|\tan\theta|} \sqrt{\,4\tan^{2}\!\theta+1\pm\sqrt{1-8\tan^{2}\!\theta}}\;,\; -\alpha_m\bigl(2\alpha_m^{2}-1\bigr)\tan\theta \right) \]
Consequently, we derive,
\[ \begin{equation} \eta \sim-\frac{P}{\rho} \sqrt{\frac{2}{\pi r}}\sum_{m+} \mathrm{e}^{\alpha^2 z}\left\{\frac{\alpha^7 \cos \theta}{\left(2 \alpha^2-1\right)\left|1-6 \alpha^2 \sin ^2 \theta\right|}\right\}^{\frac{1}{2}} \sin \left(\alpha x+\beta y+\frac{1}{4} \pi \operatorname{sgn} \kappa\right)+O\left(\frac{1}{r}\right). \end{equation} \]
Notes
Based on the preceding derivation
symbol | definition | deep-water value \(h=\infty, \sigma=0\) |
---|---|---|
\(F\) | \(F=\frac{P k \sinh k z}{4 \pi^2 \rho}\) (because \(\cosh k h \rightarrow \frac{1}{2} \mathrm{e}^{k h}\) when \(h \rightarrow \infty\) ) | \(F=\frac{P}{4 \pi^2 \rho} k \mathrm{e}^{k z}\) |
\(k\) | root of \(G=0\) | \(k=\alpha^2\) |
\(\partial_\alpha G\) | from \(G(\alpha, \beta)=\alpha^2-k\) | \(\frac{\partial G}{\partial \alpha}=\frac{2 \alpha^2-1}{\alpha}\) |
\(\kappa\) | \(\kappa=-{\partial^2_{\bar\beta}G}/{\partial_{\bar\alpha} G}\) | ~ |
note that the last line comes from \[ \begin{aligned} \partial_{\bar{\beta}} & =(-\sin \theta, \cos \theta)\binom{\partial_\alpha}{\partial_\beta} \\ \partial_{\bar{\beta}}^2 & =(-\sin \theta, \cos \theta)\binom{\partial_\alpha}{\partial_\beta} \cdot(-\sin \theta, \cos \theta)\binom{\partial_\alpha}{\partial_\beta} \\ & =\sin ^2 \theta \partial_\alpha^2+\cos ^2 \theta \partial_\beta^2-2 \sin \theta \cos \theta \partial_{\alpha \beta} \end{aligned} \] substituting the deep-water second derivatives
\[ G_{\alpha \alpha}=\frac{2 \alpha^2+1}{\alpha^2}, \quad G_{\alpha \beta}=\frac{\beta}{\alpha^5}, \quad G_{\beta \beta}=\frac{\beta^2}{\alpha^6}-\frac{1}{\alpha^2}, \] The following equations yield these expressions \[ G=\alpha^2-k \quad \text { and as}\quad G=0 \rightarrow \quad \alpha^2=k\tag{N6} \]
e.g.
\[ \begin{aligned} \partial_\beta^2 G & =\partial_\beta^2\left(\partial^2-k\right) \\ & =\partial_\beta\left(-\frac{\beta}{k}\right) \\ & =-\frac{k-\left(\frac{\beta}{k}\right) \beta}{k^2} \\ & =-\frac{\alpha^2-\frac{\beta^2}{\partial^2}}{\alpha^4} \\ & =\frac{\beta^2}{\alpha^6}-\frac{1}{\alpha^2} \end{aligned} \] base on the preceding derivation, it follows that: \[ \beta=-\alpha\bigl(2\alpha^{2}-1\bigr)\tan\theta \tag{N7} \] By combining equation \((N6),(N7)\), we arrive at the following result:
From equation \((18)\) we can see that outside this wedge our approximation gives a disturbance of at most \(O\left(r^{-1}\right)\). However, from previous theories (Ursell, F. 1960), we know that the disturbance there is actually \(O\left(r^{-3}\right)\). The two values of \(\alpha_m\) in (17) correspond to the diverging and transverse wave systems identified in earlier theories, depending on whether the + or - sign is chosen.
For precise wave-crest plotting, we express \(\alpha_m=k \cos \phi\) and \(\beta_m=k \sin \phi\), transforming the condition \(G=0\) into:
\[ k=\sec ^2 \phi \] at constant phase, we have \[ \begin{equation} \alpha_m x+\beta_m y=k r \cos (\theta-\phi)=k p=N=\mathrm{constant} \end{equation} \] Note that \(N\) represent the phase number.
Notes
Q: why is \(\mathbf{k}\cdot \mathbf{r}=N\) the constant phase
A:Full (spatio-temporal) phase of a plane wave.
A 2D plane wave traveling in direction \(\phi\) can be written (e.g. in phasor form) as
\[ \Psi(x, y, t)=A \exp [i(k r-\omega t)] \quad \text { where } \quad r=\sqrt{x^2+y^2}, \quad \theta=\arctan (y / x) . \]
Equivalently, if the wavevector is \(\mathbf{k}=(k \cos \phi, k \sin \phi)\), one often writes
\[ \Psi(x, y, t)=A \exp [i(\mathbf{k} \cdot \mathbf{r}-\omega t)]=A \exp [i(k \cos \phi x+k \sin \phi y-\omega t)] . \]
In that form, the full phase is
\[ \Phi(x, y, t)=\underbrace{k \cos \phi x+k \sin \phi y}_{k r \cos (\theta-\phi)}-\omega t . \]
What is a "line of constant phase"?
A "line of constant phase" in the real-space \((x, y)\) plane is simply the locus of points \((x, y)\) where
\[ \Phi(x, y, t)=\text { some fixed value }=\Phi_0 . \]
But if you fix the time at \(t=t_0\), then
\[ \Phi\left(x, y, t_0\right)=k r \cos (\theta-\phi)-\omega t_0=\underbrace{k r \cos (\theta-\phi)}_{\text {spatial part }}-\underbrace{\omega t_0}_{\text {constant (for fixed } \left.t_0\right)} \] Since \(\omega t_0\) is the same constant everywhere at that one instant, asking for
\[ \Phi\left(x, y, t_0\right)=\Phi_0 \]
is equivalent to asking
\[ k r \cos (\theta-\phi)=\left(\Phi_0+\omega t_0\right)=\text { some new constant. } \]
In other words, at a fixed \(t\), the \(-\omega t\) term just shifts all phases by the same amount.
Q: why parallel to \(\mathbf{k}\)
A: any line of the form \[ \alpha_m x+\beta_m y=N \]
has a normal vector \(\left(\alpha_m, \beta_m\right)\).
🚩Explanation
Define the function
\[ F(x, y)=a x+b y . \]
The given line is the set of points \((x, y)\) for which
\[ F(x, y)=c(\text { constant }) . \]
In multivariable calculus (or simply two-dimensional geometry), the gradient
\[ \nabla F=\left(\frac{\partial F}{\partial x}, \frac{\partial F}{\partial y}\right) \]
always points in the direction of greatest increase of \(F\), and is by definition orthogonal to the level sets of \(F\). Since
\[ \frac{\partial F}{\partial x}=a, \quad \frac{\partial F}{\partial y}=b, \] we have
\[ \nabla F=(a, b) . \]
Thus \((a, b)\) is normal to the curve \(F(x, y)=c\). in our notation, \[ \alpha_m=k \cos \phi, \quad \beta_m=k \sin \phi \]
and the wave-vector \(\mathbf{k}\) itself is
\[ \mathbf{k}=\left(\alpha_m, \beta_m\right) . \]
Hence the line's normal is exactly \(\mathbf{k}\).
In this definition, \(p\) denotes the length of the perpendicular dropped from the origin to the tangent line of constant phase. This perpendicular aligns with \(\mathbf{k}\), forming angle \(\phi\) with the \(x\)-axis (see Figure).
in terms of \(p\) and \(\phi\) the equation of a line is
\[ \begin{equation} x=p \cos \phi-\frac{\mathrm{d} p}{\mathrm{~d} \phi} \sin \phi ; \quad y=p \sin \phi+\frac{\mathrm{d} p}{\mathrm{~d} \phi} \cos \phi . \end{equation} \]
which gives
\[ p=N / k=N \cos ^2 \phi \]
Notes
Q: how to get equation (20) A: For one fixed value of the phase \[ N=\alpha_m x+\beta_m y=k r \cos (\theta-\phi) \]
the normal (wave-number) vector is
\[ \mathbf{k}=\left(\alpha_m, \beta_m\right)=k(\cos \phi, \sin \phi) \]
Hence every crest-line belonging to that phase is the straight line whose normal makes the angle \(\phi\) with the \(x\)-axis and whose signed distance from the origin is
\[ p(\phi)=\frac{N}{k}=N \cos ^2 \phi \]
Writing the line in its normal form gives the one-parameter family
\[ F(x, y, \phi)=x \cos \phi+y \sin \phi-p(\phi)=0 \tag{N8} \]
The physical crest is the envelope of the family (N8): it is the set of points where two neighbouring lines, with parameters \(\phi\) and \(\phi+d \phi\), intersect. In differential geometry the envelope of a family \(F(x, y, \phi)=0\) is obtained by solving \[ F(x, y, \phi)=0, \quad \frac{\partial F}{\partial \phi}(x, y, \phi)=0 \]
Compute the partial derivative:
\[ \frac{\partial F}{\partial \phi}=-x \sin \phi+y \cos \phi-\frac{d p}{d \phi} . \]
Now the system is
\[ \left\{\begin{array}{l} x \cos \phi+y \sin \phi=p \\ -x \sin \phi+y \cos \phi=\frac{d p}{d \phi} \end{array}\right. \]
Then, we determine the solution pair \((x,y)\) using this equation.
The line of constant phase is equation \((20)\). Solution (16) fails at \(\theta= \pm \theta_c\) because \(\kappa=0\), despite the presence of a single stationary point \(\left(\sqrt{\frac{3}{2}}, \mp \frac{\sqrt{3}}{2}\right)\). Here, the sign change in \(\kappa\) produces a \(\frac{\pi}{2}\) phase offset between the diverging and transverse wave systems."
Notes
Q: why \(\kappa=0\) at the cusp angle \(\theta= \pm \theta_c\)
- For every direction \(\theta\) except the cusp value there are two distinct stationary points \(\left(\alpha_m, \beta_m\right)\) : one has \(\kappa>0\) (this gives the transverse waves) and the other \(\kappa<0\) (the diverging waves).
- At \(\theta=\theta_c\) those two points coalesce. In catastrophe/caustic language this is a fold; in optics it would be called a cusp.
Mathematically the coalescence is detected by the extra condition
\[ \left.\frac{\partial^2}{\partial \beta^2}[\alpha x+\beta y]\right|_{G=0}=0 \quad \Longrightarrow \quad G_{\beta \beta}=0 . \]
Because \(\kappa\) is proportional to \(G_{\beta \beta}\), we obtain
\[ \kappa=0 \text { at } \theta= \pm \theta_c \text {. } \]
The specific numerical values given in the paper,
\[ \left(\alpha_m, \beta_m\right)=(\sqrt{3 / 2}, \mp \sqrt{3} / 2), \]
satisfy both \(G=0\) and \(G_{\beta \beta}=0\).
To find a solution when \(\abs{\theta}=\theta_c\) we must replace \((14)\) by \[ \begin{equation} \bar{\alpha}=\bar{\alpha}_m+\frac{1}{6}\left(\frac{\mathrm{~d}^3 \bar{\alpha}}{\mathrm{~d} \bar{\beta}^3}\right)_m\left(\bar{\beta}-\bar{\beta}_m\right)^3 \end{equation} \]
When \(|\theta|\) is near to \(\theta_c\) the solution (28) is not adequate because \(\kappa\) is small. Here both the \(\kappa\) and \(\mathrm{d}^3 \bar{\alpha} / \mathrm{d} \bar{\beta}^3\) terms of (22) and (36) must be retained when approximating for \(\bar{\alpha}\), and there are additional difficulties as the two points of stationary phase are close together.
Notes
Q: Why the ordinary stationary-phase formula breaks down near the Kelvin angle \(\theta_c\)
For a fixed observation angle \(\theta\) the outer integral in the far-field expansion is written (after the residue step) \[ I(r)=\int_{-\infty}^{\infty} A(\bar{\beta}) e^{i r \Phi(\bar{\beta})} d \bar{\beta}, \quad \Phi^{\prime}\left(\bar{\beta}_m\right)=0 \]
Standard situation \(\left(|\theta| \ll \theta_c\right)\). Expand the phase about a stationary point \(\bar{\beta}_m\) :
\[ \Phi(\bar{\beta})=\frac{1}{2} \kappa\left(\bar{\beta}-\bar{\beta}_m\right)^2+\frac{1}{6} \Phi^{\prime \prime \prime}\left(\bar{\beta}_m\right)\left(\bar{\beta}-\bar{\beta}_m\right)^3+\ldots \]
If the curvature \(\kappa=\Phi^{\prime \prime}\left(\bar{\beta}_m\right)\) is \(O(1)\), the quadratic term dominates and the usual stationary-phase estimate
\[ I(r) \sim A\left(\bar{\beta}_m\right) e^{i r \Phi\left(\bar{\beta}_m\right)} \sqrt{\frac{2 \pi}{|\kappa| r}} \quad(r \rightarrow \infty) \]
is adequate.
Near the critical (Kelvin) angle \(\left(|\theta| \rightarrow \theta_c\right)\)
- \(\kappa \rightarrow 0\) - the quadratic term is no longer larger than the cubic.
- The two stationary points \(R_1, R_2\) coalesce (distance \(O\left(r^{-1 / 3}\right)\) ).
Consequences: 1. The factor \(\sqrt{1 /|\kappa|}\) in (28) diverges. 2. The cubic term \(\frac{1}{6} \Phi^{\prime \prime \prime}\left(\bar{\beta}_m\right)\left(\bar{\beta}-\bar{\beta}_m\right)^3\) can no longer be neglected.
Remedy: third-order (Airy-type) uniform expansion.
Retain both terms when re-scaling \(\bar{\beta}-\beta_0=r^{-1 / 3} s\) : \[ \Phi(\bar{\beta})=\frac{1}{2} \kappa\left(\bar{\beta}-\beta_0\right)^2+\frac{1}{6} \Phi^{\prime \prime \prime}\left(\beta_0\right)\left(\bar{\beta}-\beta_0\right)^3 \Longrightarrow I(r) \approx r^{-1 / 3} \int A\left(\beta_0\right) e^{i\left(\frac{1}{2} \bar{\kappa} s^2+\frac{1}{3} \bar{\mu} s^3\right)} d s \]
an Airy integral that yields a finite \(O\left(r^{-1 / 3}\right)\) amplitude and smoothly connects the transverse and diverging systems as \(\theta \rightarrow \theta_c\).
Notes
Q: Why \(\kappa \rightarrow 0\) as \(\theta \rightarrow \theta_c\), (deep-water, pure-gravity waves)
1 Two governing equations for a stationary point 1. Dispersion curve
\[ G(\alpha, \beta)=\alpha^2-k=0, \quad k=\sqrt{\alpha^2+\beta^2} \Longrightarrow k=\alpha^2, \beta^2=\alpha^2\left(\alpha^2-1\right) . \]
- Stationary (normal-alignment) condition
\[ G_\beta=\tan \theta G_\alpha \Longrightarrow \beta=-\alpha\left(2 \alpha^2-1\right) \tan \theta \]
Eliminating \(\beta\) gives a single algebraic relation for \(\alpha\) :
\[ \left(2 \alpha^2-1\right)^2 \tan ^2 \theta=\alpha^2-1 \tag{N9} \]
For a fixed \(\theta\) inside the Kelvin wedge, (N9) has two real roots \(\alpha_{1,2}(\theta)\).
- Explicit curvature From the rotated second derivative one obtains
\[ \kappa(\theta, \alpha)=\frac{\cos \theta\left(1-6 \alpha^2 \sin ^2 \theta\right)}{\alpha\left(2 \alpha^2-1\right)}\tag{N10} \]
- Critical angle The two stationary points coalesce when (A) has a repeated root, which occurs at
\[ 1-8 \tan ^2 \theta=0 \Longrightarrow\left|\theta_c\right|=\arctan \frac{1}{2 \sqrt{2}} \approx 19.47^{\circ}, \quad \alpha_c^2=\frac{3}{2}, \quad \beta_c^2=\frac{3}{4} . \] 4. \(\kappa\) vanishes at \(\theta_c\) Insert \(\alpha_c, \theta_c\) into (N10):
\[ 1-6 \alpha_c^2 \sin ^2 \theta_c=1-6\left(\frac{3}{2}\right)\left(\frac{1}{2 \sqrt{2}}\right)^2=0 \Longrightarrow \kappa\left(\theta_c\right)=0 \]
More generally,
\[ \kappa(\theta)=\frac{\cos ^3 \theta}{\sqrt{2} \sin \theta} \sqrt{1-8 \tan ^2 \theta} \xrightarrow{\theta \rightarrow \theta_c} 0 \]
The method of stationary phase then gives \[ \begin{equation} \eta \sim\left\{\frac{-4 \pi \sqrt{3}\Gamma\left(\frac{4}{3}\right)F \cos \theta}{\partial G / \partial \alpha}\left(\frac{6}{r\left|\mathrm{~d}^3 \bar{\alpha} / \mathrm{d} \bar{\beta}^3\right|}\right)^{\frac{1}{3}} \sin (\alpha x+\beta y)\right\}_{\substack{\alpha=\alpha_m \\ \beta=\beta_m}} . \end{equation} \] When \(|\theta|\) is near to \(\theta_c\) the solution (16) is not adequate because \(\kappa\) is small. Here both the \(\kappa\) and \(d^3 \bar{\alpha} / d \bar{\beta}^3\) terms must be retained when approximating for \(\bar{\alpha}\), and there are additional difficulties as the two points of stationary phase are close together.
The waves along the line \(\theta=0\) have also given difficulty in previous methods. Here the transverse system gives the contribution
\[ -\frac{P}{\rho} \sqrt{\frac{2}{\pi x}} \mathrm{e}^z \sin \left(x+\frac{1}{4} \pi\right) \]
to \(\eta\), and the diverging system gives a term in which \(\alpha_m, \beta_m \rightarrow \infty\), since \(\mathrm{d} \alpha / \mathrm{d} \beta \rightarrow 0\) on \(G=0\) as \(\alpha, \beta \rightarrow \infty\). Thus as \(\theta \rightarrow 0\) the diverging wave amplitude \(\rightarrow \infty\) and its wave length \(\left(2 \pi / \sqrt{\alpha_m^2+\beta_m^2}\right) \rightarrow 0\);
Notes
1 Stationary-phase condition
\[ G_\beta=\tan\theta\,G_\alpha \;\xrightarrow{\;\theta\to0\;}\; G_\beta=0 . \]
Along the deep-water gravity dispersion curve \(G=0\), the equation \(G_\beta=0\) can be satisfied only for \(\alpha,\beta\to\infty\); therefore the stationary point associated with the diverging system moves off to infinity.
2 Second derivative
\[ \kappa=-\frac{G_{\bar\beta\bar\beta}}{G_{\bar\alpha}} \sim\frac{1}{\alpha_m}\qquad(\alpha_m\to\infty). \]
Because the stationary-phase prefactor contains \(\sqrt{2\pi/|\kappa|}\propto\alpha_m^{1/2}\), the diverging-wave amplitude diverges.
3 Wavelength
\[ k=\sqrt{\alpha_m^{2}+\beta_m^{2}}\sim\alpha_m\to\infty \;\;\Longrightarrow\;\; \lambda=\frac{2\pi}{k}\to0 . \]
4 Physical interpretation
The simultaneous limits amplitude \(\to\infty\) and wavelength \(\to0\) are a mathematical artefact of idealising the surface load as a point force. Distributing the pressure smoothly over a finite area filters out the very high wave numbers and removes the divergence of the diverging system as \(\theta\to0\).
🚩 Why do \(\alpha_m,\beta_m\to\infty\) as \(\theta\to0\)?
4.1 Governing equations
Equation | Deep-water form |
---|---|
Dispersion curve \(G=0\) | \(\beta^{2}=\alpha^{2}\bigl(\alpha^{2}-1\bigr)\) |
Stationary condition \(G_\beta=\tan\theta\,G_\alpha\) | \(\beta=-\alpha\bigl(2\alpha^{2}-1\bigr)\tan\theta\) |
4.2 Eliminate \(\beta\)
Set \(x=\alpha^{2}\) \((x>1)\). Substituting the stationary condition into the dispersion relation gives
\[ x-1=(2x-1)^{2}\tan^{2}\theta . \]
4.3 Limit \(\theta\to0\)
Put \(T=\tan\theta\) \((T\to0)\). If \(x\) remained \(O(1)\), the right-hand side would be \(O(T^{2})\ll1\) and could not balance \(x-1=O(1)\). Hence \(x\) must increase as \(T\) decreases. For \(x\gg1\), \[ x\simeq4x^{2}T^{2} \;\;\Longrightarrow\;\; x\simeq\frac{1}{4T^{2}} =\frac{1}{4\tan^{2}\theta} \xrightarrow{\theta\to0}\infty . \] Thus \[ \alpha_m^{2}\sim\frac{1}{4\tan^{2}\theta}\to\infty, \qquad \alpha_m\sim\frac{1}{2\tan\theta}\to\infty . \]
4.4 Divergence of \(\beta_m\)
From the stationary condition
\[ \beta=-\alpha\bigl(2\alpha^{2}-1\bigr)\tan\theta , \] and with \(\alpha_m\sim(2T)^{-1}\), \[ \beta_m\simeq-2\alpha_m^{3}\tan\theta \sim-\frac{1}{4\tan^{2}\theta} \xrightarrow{\theta\to0}-\infty . \]
Hence
\[ \theta\to0 \;\Longrightarrow\; \alpha_m,\;|\beta_m|\;\to\infty . \]
4.5 Two positive roots of the algebraic equation
The algebraic equation
\[ x-1=(2x-1)^{2}\tan^{2}\theta , \qquad x=\alpha^{2},\;\theta\ll1 \] has two positive roots:
Small root (transverse wave). Write \(x=1+\delta\) with \(\delta=O(\varepsilon)\), \(\varepsilon=\tan^{2}\theta\): \[ 0=\delta-(1+2\delta)^{2}\varepsilon \approx\delta-\varepsilon-4\delta\varepsilon \;\Rightarrow\; \delta\approx\varepsilon(1+4\varepsilon), \] giving \(x_{\mathrm T}\approx1+\varepsilon\) and \(\alpha_{\mathrm T}\approx1\) (finite).
Large root (diverging wave). For \(x\gg1\), \[ x=4x^{2}\varepsilon \;\Longrightarrow\; x_{\mathrm D}\approx\frac{1}{4\varepsilon}, \qquad \alpha_{\mathrm D}\sim\frac{1}{2\tan\theta}\to\infty . \]
Only the diverging-wave root produces the divergent amplitude.
4.6 Why only the diverging wave diverges
- Transverse wave: \(\kappa=O(1)\), amplitude \(\propto r^{-1/2}\) remains finite.
- Diverging wave: \(\kappa\sim1/\alpha_{\mathrm D}\propto\tan\theta\to0\), hence \(\sqrt{2\pi/|\kappa|}\sim\alpha_{\mathrm D}^{1/2}\to\infty\); simultaneously \(k\sim\alpha_{\mathrm D}\to\infty\), so \(\lambda\to0\).
Therefore the diverging-wave amplitude blows up at \(\theta=0\) because its stationary root is forced to extremely large wave-number by the point-load model.
the effect of surface tension
for the case infinite depth, the poles of the inner integral are now given by \[ \begin{equation} \alpha^2-k-\sigma k^3=0 \end{equation} \] If we again write \(\alpha=k \cos \phi, \beta=k \sin \phi\), (23) can be solved for \(k\) :
\[ \begin{equation} k=\frac{\cos ^2 \phi \pm \sqrt{\cos ^4 \phi-4 \sigma}}{2 \sigma} \end{equation} \]
These two solutions will be referred to as capillary waves (upper sign) and gravity waves. The two curves given by (24) meet at points \(M\), say, where \(k_M=\sigma^{-\frac{1}{2}}\), \(\cos \phi_M=(4 \sigma)^{\frac{1}{4}}\).
Notes
❓Q: \(k_M=\sigma^{-\frac{1}{2}}\), \(\cos \phi_M=(4 \sigma)^{\frac{1}{4}}\).
A: Consider the deep-water dispersion relation that includes surface tension, \[ G(\alpha, \beta)=\alpha^2-k\left(1+\sigma k^2\right)=0, \quad k=\sqrt{\alpha^2+\beta^2}, \quad \sigma=\frac{T}{\rho U^2}, \]
and introduce the polar decomposition \(\alpha=k \cos \phi\).Rewriting \(G=0\) as a quadratic equation in \(k\) gives
\[ \sigma k^2-k \cos ^2 \phi+1=0 . \]
Its two roots, quoted in the paper as Eq. (24), are
\[ k_{ \pm}(\phi)=\frac{\cos ^2 \phi \pm \sqrt{\cos ^4 \phi-4 \sigma}}{2 \sigma} \]
where the "+" sign corresponds to the capillary branch and the "-" sign to the gravity branch.
Coincidence point of the two branches
The two branches meet where the discriminant vanishes, \[ \cos ^4 \phi-4 \sigma=0 \quad \Longrightarrow \quad \cos \phi_M=(4 \sigma)^{1 / 4} \]
Substituting \(\cos ^2 \phi_M=2 \sigma^{1 / 2}\) into either branch yields
\[ k_M=\frac{\cos ^2 \phi_M}{2 \sigma}=\frac{2 \sigma^{1 / 2}}{2 \sigma}=k_M=\sigma^{-1 / 2} \]
Hence the two dispersion curves intersect at the point \(M\) given by
\[ k_M=\sigma^{-1 / 2}, \quad \cos \phi_M=(4 \sigma)^{1 / 4}, \]
which identifies the wavenumber and propagation direction at which the capillary and gravity wave branches coincide.
When \(\cos ^4 \phi<4 \sigma\) there are no real solutions of the equation \(G=0\), and the poles will be at complex values of \(\bar{\alpha}\); any waves in this range will therefore have an exponential decay factor when the first integration has been carried out, and they will be neglected.
When the outer Fourier integral for the free-surface elevation is closed in the lower half of the complex \(\bar{\alpha}-\) plane, the far-field contribution is obtained from the poles of the integrand,
\[ \eta(r)=\int_{-\infty}^{\infty} \frac{F(\bar{\alpha}, \bar{\beta})}{G(\bar{\alpha}, \bar{\beta})} e^{i \bar{\alpha} r} d \bar{\alpha} \quad \longrightarrow \quad \sum_p \operatorname{Res}_{\bar{\alpha}=\bar{\alpha}_p}[F / G] e^{i \bar{\alpha}_p r}, \quad r=\sqrt{x^2+y^2} \]
where each pole satisfies \(G\left(\bar{\alpha}_p, \bar{\beta}\right)=0\) and lies below the real axis:
\[ \bar{\alpha}_p=a-i b, \quad b>0 \]
The exponential factor for such a pole is
\[ e^{i \bar{\alpha}_p r}=e^{i a r} e^{-b r} \]
so the spatial attenuation is governed solely by the imaginary part \(b\) : the amplitude decays as \(e^{-b r}\).
For certain values of the transverse wavenumber \(\bar{\beta}\) the root of \(G=0\) approaches the real axis, i.e. \(b \ll 1\). In this limit the decay length \(1 / b\) becomes very large and the associated mode behaves almost like a freely propagating wave, producing only weak attenuation in the physical field. Conversely, for other \(\bar{\beta}\) the pole sits deeper in the lower half-plane, \(b=O(1)\); the factor \(e^{-b r}\) then suppresses the contribution rapidly and the corresponding mode is essentially evanescent.
Hence, when the poles generated by \(G=0\) lie close to the \(\bar{\alpha}\) axis, the imaginary part \(b\) is small and the exponential decay of the residue term is correspondingly weak, giving a slowly attenuating wave component in the far field.
The tangent at \(M\) passes through the origin, making an angle \(\phi_M\), say, with the \(x(\alpha)\) axis; the wave crests therefore make an angle \(\frac{1}{2} \pi-\phi_M\) with the \(x\) axis since they are perpendicular to the tangent (k). Thus the velocity normal to the wave crest at points \(M\) is \[ U \sin \left(\frac{1}{2} \pi-\phi_M\right)=U \cos \phi_M=(4 T g / \rho)^{\frac{1}{4}}=U_{\min .} \]
Note
Why does the tangent to the dispersion curve \(\mathrm{G}=0\) at the capillary-gravity intersection \(M\) pass through the origin?
Parametric form of the curve, put
\[ \alpha=k \cos \phi, \quad \beta=k \sin \phi, \]
so that the deep-water capillary-gravity relation
\[ G(\alpha, \beta)=\alpha^2-k\left(1+\sigma k^2\right)=0 \]
becomes the quadratic
\[ \sigma k^2-k \cos ^2 \phi+1=0 . \tag{N12} \]
Its two roots (upper = capillary, lower = gravity) are
\[ k_{ \pm}(\phi)=\frac{\cos ^2 \phi \pm \sqrt{\cos ^4 \phi-4 \sigma}}{2 \sigma} \] Coincidence point \(M\), the branches meet where the discriminant vanishes:
\[ \cos ^4 \phi_M=4 \sigma \Longrightarrow k_M=\sigma^{-1 / 2}, \quad \cos \phi_M=(4 \sigma)^{1 / 4} \]
Tangent direction at \(M\), treat \(k(\phi)\) as a parametric curve; its tangent is
\[ \frac{d \alpha}{d \phi}=k^{\prime} \cos \phi-k \sin \phi, \quad \frac{d \beta}{d \phi}=k^{\prime} \sin \phi+k \cos \phi \]
From (N12),
\[ k^{\prime}=-\frac{2 k \cos \phi \sin \phi}{2 \sigma k-\cos ^2 \phi}\tag{N13} \] At \(M\), the denominator in (N13) vanishes, so \(\left|k^{\prime}\right| \rightarrow \infty\). Hence
\[ \left.\frac{d \beta}{d \alpha}\right|_M=\lim _{\left|k^{\prime}\right| \rightarrow \infty} \frac{k^{\prime} \sin \phi+k \cos \phi}{k^{\prime} \cos \phi-k \sin \phi}=\tan \phi_M\tag{N14} \]
Line through the origin, the radial line \(OM\) has slope
\[ \frac{\beta_M}{\alpha_M}=\frac{k_M \sin \phi_M}{k_M \cos \phi_M}=\tan \phi_M \]
which equals (N14). Therefore the tangent at \(M\) is collinear with the radius vector \(OM\); i.e. it passes through the origin and makes the angle \(\phi_M\).
Physical implication: at \(M\) the phase velocity is parallel to the wave-number vector, and the group velocity-normal to the curve -attains its minimum value \(U_{\min }\) : the point marks the slowest capillary-gravity wave that can be generated.
Notes
wave crest and wave vector \(\mathbf k\)
- A wave crest is a line (in 2D) or surface (in 3D) where the wave displacement is at its maximum (peak).
- Wave crests are special examples of wave fronts, since a crest is also a constant-phase surface (often the phase is \(\pi/2\) or \(3\pi/2\), depending on the definition).
- Direction: The normal to the crest is likewise parallel to \(\mathbf{k}\), because the crest is a particular phase location (where the wave is at a maximum).
Parallelism of \(k_x\) to \(x\)-axis?
Typically, we don't explicitly state that \(k_x\) is parallel to the physical \(x\)-axis; the equivalence is implicit in the Fourier representation. Immediately after the transform is written as
\[ \eta(x, y, z)=\iint e^{i(\alpha x+\beta y)} \frac{F(\alpha, \beta)}{G(\alpha, \beta)} d \alpha d \beta \]
It remarks "It is evident from inspection that \(\alpha=k_x, \beta=k_{y .}\)." Once \(\alpha \equiv k_x\) and \(\beta \equiv k_y\) are identified, the axes of the \((\alpha, \beta)_{-}\)(i.e. \(k_{-}\)) plane are, by definition, parallel to the spatial \(\mathscr{x}\) - and \(\mathscr{y}\)-directions. Hence, when the discussion later refers to the tangent at the intersection point \(M\) making an angle \(\phi_M\) with the " \(x(\alpha)\)-axis," it is simply invoking this original identification; no additional statement about the parallelism of \(k_x\) and the physical \(x\)-axis is required.
Wave-Crest Orthogonality to the Wave Vector (no matter the Presence of Uniform Currents or not)
The wave number vector \(\mathbf{k}\) is, by definition, normal to isophase surfaces; hence wave crests remain perpendicular to \(\mathbf{k}\) even in a uniform background flow \(\mathbf{U}\). The current modifies phase and group velocities through a Doppler shift but leaves the geometric relationship between \(\mathbf{k}\) and the crest intact.
- Geometric Necessity
For a monochromatic plane wave
\[ \phi(\mathbf{x}, t)=\mathbf{k} \cdot \mathbf{x}-\omega t \]
wave crests satisfy \(\phi=\) const. The spatial gradient is
\[ \nabla \phi=\mathbf{k} \]
which is orthogonal to every isophase (crest) surface. This definition is independent of reference frame, isotropy, or the presence of a uniform current.
- Uniform Background Flow
With a steady current \(\mathbf{U}\), the intrinsic frequency \(\sigma\) in the dispersion relation is Doppler-shifted:
\[ \omega=\sigma+\mathbf{k} \cdot \mathbf{U} \]
The orientation of \(\mathbf{k}\) and thus the crest geometry are unchanged. Only the following vectors acquire an advective component:
a. Phase velocity: \(\mathbf{c}_p=\mathbf{U}+\frac{\sigma}{|\mathbf{k}|} \hat{\mathbf{k}}, \quad\)
b. Group velocity: \(\mathbf{c}_g=\mathbf{U}+\nabla_{\mathbf{k}} \sigma\).
- Directional Distinctions
Quantity | Symbol | Static water | With uniform U | Relation to crests |
---|---|---|---|---|
Wave vector | k | \(\perp\) crest | unchanged | normal |
Phase velocity | \(\mathrm{c}_p\) | \(\vert\vert k\) | tilted by U | advects phase |
Group velocity | \(\mathrm{c}_g\) | \(\vert\vert k\) | tilted by U | advects energy |
Thus, while energy packets may advance obliquely, the instantaneous crest orientation remains normal to \(\mathbf{k}\).
- Extension to Shear Flows
For spatially varying currents, \(\mathbf{U}=\mathbf{U}(\mathbf{x})\), the ray equations give
\[ \frac{d \mathbf{k}}{d t}=-(\nabla \mathbf{U})^{\top} \mathbf{k} . \]
The wave vector (and hence the local crest normal) evolves in time, but its orthogonality to the local crest is preserved at every instant.
- Conclusions
Wave crests are perpetually perpendicular to the wave number vector by geometric definition. Uniform currents alter phase and group velocities through advection yet do not tilt the crests relative to \(\mathbf{k}\). Misinterpretations generally arise from conflating wave-vector direction with energy or phase propagation directions.
why \(U_{\min}\)
the dispersion relation is
\[ \omega^2=g k+\frac{T}{\rho} k^3 \quad \text { (deep-water gravity-capillary) } \]
so that the group speed (velocity normal to the wave-crest)
\[ U_g(k)=\frac{d \omega}{d k}=\frac{g+3 T k^2 / \rho}{2 \sqrt{g k+T k^3 / \rho}} \]
Locate the extremum, Set \(d U_g / d k=0\) :
\[ \frac{d U_g}{d k}=0 \Longrightarrow k^2=\frac{\rho g}{T} \frac{1}{2} . \]
Hence the critical wavenumber is \[ k_M=\sqrt{\frac{\rho g}{2 T}}=\sigma^{-1 / 2}, \quad \sigma \equiv \frac{T}{\rho U^2} . \]
Evaluate group velocity \(U_g\) at \(k=k_M\)
\[ U_g\left(k_M\right)=\frac{g+3 T k_M^2 / \rho}{2 \sqrt{g k_M+T k_M^3 / \rho}}=\frac{g+3 g / 2}{2 \sqrt{g k_M+g k_M / 2}}=(4 T g / \rho)^{1 / 4} \equiv U_{\min } . \]
For \(k \ll k_M\) (pure-gravity range) \(U_g \sim \sqrt{g / 4 k} \rightarrow \infty\); for \(k \gg k_M\) (pure-capillary range) \(U_g \sim \sqrt{3 T / 4 \rho} k^{1 / 2} \rightarrow \infty\). Thus the stationary value at \(k_M\) is the global minimum.
Notes
A steady ship wake can be characterised in two mathematically equivalent ways.
Viewpoint | Formal condition | Interpretation |
---|---|---|
Temporal (Doppler) | \(\omega_{\text {obs }}=\omega_0+\mathbf{U} \cdot \mathbf{k}=0\) | In the ship-fixed frame the wave pattern is timeindependent; the observed frequency vanishes. |
Spatial (phase speed) | \(c_{\mathrm{ph}} =U \cos \theta, \quad c_{\mathrm{ph}}=\frac{\omega_0}{k}\) | The phase speed equals the component of the ship speed along the wave-vector direction; phase fronts drift with the vessel. |
Starting from the deep-water capillary-gravity dispersion
\[ \omega_0^2=g k+\frac{T}{\rho} k^3, \]
the condition \(\omega_{\text {obs }}=0\) implies
\[ \omega_0=U k \cos \theta \quad \Longrightarrow \quad c_{\mathrm{ph}}=\frac{\omega_0}{k}=U \cos \theta . \]
Squaring this relation and writing \(\mathbf{k}=(\alpha, \beta)\) yields the curve
\[ G(\alpha, \beta)=\alpha^2-k\left(1+\sigma k^2\right)=0, \quad \sigma=\frac{T}{\rho U^2} \]
which the paper uses as the fundamental steady-wake condition. Thus "zero observed frequency" and "phase speed equals \(U \cos \theta^{\text {" }}\) are two expressions of the same requirement for a stationary ship wave pattern.
Geometric link with \(\phi_M\)
At the intersection point \(M\) of the two dispersion branches
\[ \cos ^4 \phi_M=4 \sigma, \text { so } \cos \phi_M=(4 T g / \rho)^{1 / 4} / U \]
Taking the component of the ship speed \(U\) normal to the wave-crest
\[ U \sin \left(\frac{\pi}{2}-\phi_M\right)=U \cos \phi_M=(4 T g / \rho)^{1 / 4}=U_{\min } \]
which reproduces the analytic minimum derived above.
So, here, because a Doppler shift adds only a linear term \(-\mathbf{U} \cdot \mathbf{k}\) to the dispersion relation, it translates the \(\omega(k)\) surface without altering its slope. The group speed, defined by \(\mathbf{c}_g=\nabla_k \omega\), therefore remains the derivative of the intrinsic frequency
\[ \omega_0^2=g k+T k^3 / \rho \]
even in the ship-fixed frame. Hence \(c_g(k)=\partial \omega_0 / \partial k\) can be compared directly with the projected ship speed \(U \cos \theta\) to determine which wavenumbers satisfy the steady-wake condition.
The radiation condition is
\[ \frac{\bar{\alpha} \cos \theta-\bar{\beta} \sin \theta}{\partial G / \partial \bar{\alpha}}>0 . \]
For the points of stationary phase, we have
\[ \begin{aligned} \partial G / \partial \alpha & =(\alpha / k)\left\{2 k-\left(1+3 \sigma k^2\right)\right\} \\ & =(3 \sigma \alpha / k)\left(k_{+}-k\right)\left(k-k_{-}\right) \end{aligned} \]
where \(k_{+}, k_{-} =(1 / 3 \sigma)(1 \pm \sqrt{1-3 \sigma})\)
Note that \(k_{-}<1\left(\sigma<\frac{1}{4}\right)\), while at \(k=k_{+}, \mathrm{d} \alpha / \mathrm{d} \beta=\partial G / \partial \beta / \partial G / \partial \alpha=\infty\). Hence \(k=k_{+}\) gives the points at which the curve \(G=0\) is parallel to the \(\alpha\) axis. Since \(k \geqslant 1\) for any \(\sigma, k=k_{-}\)is not a point on the curve. Now the radiation condition becomes
\[ \frac{k \cos \theta}{3 \sigma\left(k_{+}-k\right)\left(k-k_{-}\right)}>0 \]
which is equivalent to
\[ \left.\begin{array}{rlrl} \cos \theta & >0 & & k_{-}<1 \leqslant k<k_{+} \\ & <0 & & k>k_{+} . \end{array}\right\} \]
Notes
❓\(\sigma<\frac{1}{4}\)
since \(1-3 \sigma>0 \quad \sigma<\frac{1}{3}\), also we have \(\cos ^4 \phi-4 \sigma>0\), then
\[ \sigma<\frac{1}{4} \cos ^4 \phi \]
it says
\[ \sigma<\frac{1}{4} \]
❓\(k_{-} <1\)
we have
\[ \begin{aligned} k_{-} & =\frac{1}{3 \sigma}(1-\sqrt{1-3 \sigma}) \\ & <\frac{1}{3 \sigma} \\ & <1 \end{aligned} \]
❓\(k \geq 1\)
For the deep-water capillary-gravity relation
\[ G(\alpha, \beta)=\alpha^2-k\left(1+\sigma k^2\right)=0, \quad k=\sqrt{\alpha^2+\beta^2}, \quad \sigma>0 \]
one has
\[ \alpha^2=k\left(1+\sigma k^2\right), \quad \beta^2=k^2-\alpha^2(\geq 0) \]
Substituting the first into the second gives
\[ k^2-k\left(1+\sigma k^2\right)=k-1-\sigma k^2 \geq 0 \]
The right-hand side can be non-negative only if \(k-1 \geq 0\); hence
\[ k \geq 1 \]
Consequently, any real point on the curve \(G=0\) satisfies this bound, independent of the value of \(\sigma\). Because the lower root \(k_{-}(\phi)\) lies entirely in the interval \(k<1\), it does not represent a physical point on the dispersion curve, and only the upper root \(k_{+}(\phi) \downarrow\) atained in the subsequent radiation-condition analysis.
This decomposition plays a critical role in analyzing the radiation condition:
1. Sign analysis:
The expression \(\partial G / \partial \alpha = (3 \sigma \alpha / k)(k_{+}-k)(k-k_{-})\) explicitly reveals sign changes in the derivative, corresponding to the tangent directions along the \(G=0\) curve (Figures 4–5).
Radiation condition:
Through Equations (48)–(49):
\[ \frac{\cos \theta}{\partial G / \partial \alpha} > 0 \implies \operatorname{sign}\left( \partial G / \partial \alpha \right) = \operatorname{sign}(\cos \theta), \]
the factorization enables direct sign determination via comparison of \(k\) with \(k_{+}\) and \(k_{-}\) (e.g., \(\partial G / \partial \alpha > 0\) when \(k<k_{+}\)).Wave system separation:
- \(k<k_{+}\) corresponds to gravity waves (\(\cos \theta > 0\), propagating behind the pressure point).
- \(k>k_{+}\) corresponds to capillary waves (\(\cos \theta < 0\), which may appear ahead of the pressure point).
- \(k<k_{+}\) corresponds to gravity waves (\(\cos \theta > 0\), propagating behind the pressure point).
Why \(\nabla G\) is Parallel to \(r\)
#parallel
Let the dispersion curve be defined by \(G(\alpha, \beta)=0\). For an observation point \(\mathbf{r}=(x, y)=r(\cos \theta, \sin \theta)\) the Fourier kernel reads
\[ e^{i(\alpha x+\beta y)} \]
Rotate \((\alpha, \beta)\) through the angle \(\theta\) :
\[ \bar{\alpha}=\alpha \cos \theta+\beta \sin \theta, \quad \bar{\beta}=-\alpha \sin \theta+\beta \cos \theta, \]
so that
\[ \alpha x+\beta y=\bar{\alpha} r . \]
- Stationary point with respect to \(\bar{\beta}\)
Applying the method of steepest descents to the outer \(\bar{\beta}\)-integration gives
\[ \frac{\partial \bar{\alpha}}{\partial \bar{\beta}}=0 .\tag{J1} \]
- Relating \(\partial \bar{\alpha} / \partial \bar{\beta}\) to \(\nabla_k G\)
Along the curve \(G=0\) we have \(G_{\bar{\alpha}} d \bar{\alpha}+G_{\bar{\beta}} d \bar{\beta}=0\); hence
\[ \frac{d \bar{\alpha}}{d \bar{\beta}}=-\frac{G_{\bar{\beta}}}{G_{\bar{\alpha}}}\tag{J2} \]
Combining (J1) and (J2) yields the stationary-phase condition
\[ G_{\bar{\beta}}=0 .\tag{J3} \]
- Gradient parallelism
The Jacobian of the rotation is unity, so
\[ \binom{G_{\bar{\alpha}}}{G_{\bar{\beta}}}=\left(\begin{array}{cc} \cos \theta & \sin \theta \\ -\sin \theta & \cos \theta \end{array}\right)\binom{G_\alpha}{G_\beta} . \]
Condition (J3) gives
\[ -\sin \theta G_\alpha+\cos \theta G_\beta=0 \Longrightarrow \nabla_{(\alpha, \beta)} G \|(\cos \theta, \sin \theta)=\hat{\mathbf{r}} .\tag{J4} \]
- Result
Equation (J4) shows that a wave-number point contributes to the far field in direction \(\theta\) if and only if the normal to the dispersion curve \(G=0\) is parallel to the observation vector. In other words,
\[ \nabla G \| \mathbf{r} \quad \Longleftrightarrow \quad \text { normal direction angle }=\theta . \]
near point \(C\)
Around point \(C\), "Which side is 'ahead of the pressure point' ?"
The pressure source (or a ship) is placed at the origin and assumed to move with speed \(U\) in the \(+x\) direction. In the body-fixed frame the external flow appears to come from \(-x\); therefore
Physical sector | Observation angle \(\boldsymbol{\theta}\) | Meaning |
---|---|---|
Upstream (ahead) | \(\boldsymbol{\theta}= \pm \boldsymbol{\pi}\) | Along \(-x\), facing the incoming flow |
Downstream (astern) | \(\theta=0\) | Along \(+x\), behind the source |
(the convention is standard in Lighthill-type ship-wake studies, even if not reiterated).
How the radiation condition selects \(\theta= \pm \pi\) The radiation test reads
\[ \frac{\bar{\alpha} \cos \theta-\bar{\beta} \sin \theta}{\partial G / \partial \bar{\alpha}}>0 \Longrightarrow \alpha \cos \theta\left(\frac{\partial G}{\partial \alpha}\right)^{-1}>0 \tag{J5} \]
For the capillary branch in the limit \(\phi \rightarrow 0^{+}\)
\[ \left(\beta \rightarrow 0^{+}, k \gg 1\right): \]
- \(\alpha>0\) and very large;
- a direct evaluation gives \(\partial G / \partial \alpha<0\). Insert the signs into (J5):
\[ (+) \cos \theta(-)>0 \Longrightarrow \cos \theta<0 \Rightarrow \theta=\pi(\text { or }-\pi) . \]
Hence the only observation angles that survive the radiation test are \(\theta= \pm \pi\), the two directions straight upstream.
Physical picture
- As \(\phi \rightarrow 0^{+}\)the wave vector is nearly parallel to the \(\alpha\)-axis, and the group velocity \(\mathbf{c}_g\) points toward \(-x\) because \(\partial G / \partial \alpha<0, \partial G / \partial \beta<0\).
- Requiring outward radiation forces one to keep the same-direction phase-speed branch; the radiation condition effects precisely this choice.
- Therefore the first observable capillary waves appear at \(\theta= \pm \pi\).
As \(\phi\) increases the normal rotates, the crest lines sweep around the hull, and finally become asymptotic to the Kelvin angles
\[ \theta= \pm\left(\frac{\pi}{2}-\phi_M\right) \text { downstream. } \]
Summary
"The radiation condition shows that the capillary wave starts ahead of the pressure point" means: in the large- \(k\) section of the dispersion loop only the normals pointing to \(-x(\theta= \pm \pi)\) satisfy (J5), hence they form the first segment of the radiating capillary wave.
capillary waves likes something a cross between a parabola and a hyperbola; the gravity wave-crests are more like true hyperbolae