summaryrefslogtreecommitdiff
path: root/source/know/concept/boussinesq-wave-theory/index.md
blob: e5fd43353b1508eae45f9c49d772398dcca4075e (plain)
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
---
title: "Boussinesq wave theory"
sort_title: "Boussinesq wave theory"
date: 2023-01-07
categories:
- Physics
- Mathematics
- Fluid mechanics
- Fluid dynamics
layout: "concept"
---

In fluid mechanics, **Boussinesq wave theory**
consists of several equations to describe waves on a liquid's surface.
It was the first attempt to explain the nonlinear phenomenon of solitons,
which were not predicted by the linear theories existing at the time.



## Boundary conditions

Consider the [Euler equations](/know/concept/euler-equations/)
for an incompressible fluid with negligible [viscosity](/know/concept/viscosity/):

$$\begin{aligned}
    \va{g} - \frac{\nabla p}{\rho}
    &= \pdv{\va{u}}{t} + (\va{u} \cdot \nabla) \va{u}
    \qquad \qquad
    \nabla \cdot \va{u}
    = 0
\end{aligned}$$

We rewrite the former using the velocity potential $$\Psi$$ for $$\va{u}$$
and the gravitational potential $$\Phi$$ for $$\va{g}$$,
such that $$\va{u} = \nabla \Psi$$ and $$\va{g} = - \nabla \Phi$$.
We also use a vector identity:

$$\begin{aligned}
    - \nabla \Phi - \frac{\nabla p}{\rho}
    &= \nabla \pdv{\Psi}{t} + \frac{1}{2} \nabla |\va{u}|^2 + (\nabla \cross \nabla \Psi) \cross \va{u}
\end{aligned}$$

Recall that the curl of a gradient is always zero, so the last term disappears.
Integrating in space yields integration constants $$C(t)$$ and $$p_0$$,
the latter representing atmospheric pressure:

$$\begin{aligned}
    \pdv{\Psi}{t} + \frac{1}{2} |\va{u}|^2
    &= -\Phi + \frac{p_0 - p}{\rho} + C
\end{aligned}$$

Consider a rectangular channel of depth $$h$$
extending infinitely far along the $$x$$-axis,
with a finite width along the $$y$$-axis.
We choose our coordinate system so that
$$z = 0$$ is the equilibrium water level,
and the bottom is at $$z = -h$$.
Let $$\eta(x, t)$$ be the deformation of the surface,
for which we want to find a wave equation.
All quantities are assumed to be constant in $$y$$,
so we only consider a 2D flow $$\va{u} = \big(u^{(x)}, u^{(z)}\big)$$.
At the surface $$z = \eta$$ we thus have:

$$\begin{aligned}
    \pdv{\Psi}{t} + \frac{1}{2} \bigg( \big(u^{(x)}\big)^2 + \big(u^{(z)}\big)^2 \bigg)
    &= - g \eta + \frac{p_0 - p}{\rho} + C
\end{aligned}$$

Where $$g \eta = \Phi$$ with $$g \approx 9.81 \:\mathrm{m}/\mathrm{s}^2$$ on Earth.
Later, we will differentiate this formula in $$x$$,
so we can already set $$C = 0$$ now, since it will vanish then anyway.
Furthermore, we assume that at the surface $$z = \eta$$
the pressures are in equilibrium $$p_0 = p$$, leaving:

$$\begin{aligned}
    \boxed{
        \pdv{\Psi}{t} + \frac{1}{2} \bigg( \big(u^{(x)}\big)^2 + \big(u^{(z)}\big)^2 \bigg) + g \eta
        = 0
    }
\end{aligned}$$

This is called the **free surface boundary condition**.
Obviously, if $$z = \eta$$, then $$\eta \!-\! z = 0$$.
Taking the [material derivative](/know/concept/material-derivative/)
of this fact gives the following relation:

$$\begin{aligned}
    0
    = \frac{\mathrm{D}}{\mathrm{D} t} (\eta - z)
    = \pdv{\eta}{t} + \pdv{z}{t} + \va{u} \cdot \nabla \eta - \va{u} \cdot \nabla z
\end{aligned}$$

Since $$\eta$$ only depends on $$x$$ and $$t$$,
this becomes the **kinematic boundary condition**:

$$\begin{aligned}
    \boxed{
        \pdv{\eta}{t} + u^{(x)} \pdv{\eta}{x} - u^{(z)}
        = 0
    }
\end{aligned}$$

The equations will be derived from these two fundamental boundary conditions.



## Boussinesq approximation

Let us take a Taylor expansion of the velocity potential $$\Psi(x, z, t)$$
at the bottom $$z = -h$$:

$$\begin{aligned}
    \Psi(x, z)
    = \Psi(x, -h) + (z + h) \: \Psi_z(x, -h) + \frac{(z + h)^2}{2} \: \Psi_{zz}(x, -h) + ...
\end{aligned}$$

Because the fluid is incompressible, this can be rewritten.
Laplace's equation tells us:

$$\begin{aligned}
    \nabla \cdot \va{u}
    = \nabla^2 \Psi
    = 0
    \qquad \implies \qquad
    \:\:\:\Psi_{zz}
    = - \Psi_{xx}
\end{aligned}$$

Which we use for the expansion's second-order term.
Similarly, for the fourth-order term:

$$\begin{aligned}
    0
    = \nabla^2 \nabla^2 \Psi
    &= \Psi_{zzzz} + \pdvn{2}{}{z} \Psi_{xx} + \pdvn{2}{}{x} \Psi_{zz} + \Psi_{xxxx}
    \\
    &= \Psi_{zzzz} + \pdvn{2}{}{x} \Psi_{zz} - \pdvn{2}{}{x} \Psi_{xx} + \Psi_{xxxx}
    \\
    &= \Psi_{zzzz} - \Psi_{xxxx}
\end{aligned}$$

And so on for higher orders.
In the Taylor expansion, all even derivatives can be rewritten in this way,
and all odd derivatives can be split into a single $$\ipdv{}{z}$$ and an even $$x$$-derivative:

$$\begin{aligned}
    \Psi(x, z)
    &= \Psi + (z \!+\! h) \Psi_z - \frac{(z \!+\! h)^2}{2} \Psi_{xx}
    - \frac{(z \!+\! h)^3}{6} \pdv{}{z} \Psi_{xx} + \frac{(z \!+\! h)^4}{24} \Psi_{xxxx} + ...
    \\
    &= \bigg(\! \Psi - \frac{(z \!+\! h)^2}{2} \Psi_{xx} + \frac{(z \!+\! h)^4}{24} \Psi_{xxxx} - ... \bigg)
    \!+\! \bigg(\! (z \!+\! h) \Psi_z - \frac{(z \!+\! h)^3}{6} \pdvn{2}{}{x} \Psi_{z} + ... \bigg)
\end{aligned}$$

By definition $$\Psi_z = u^{(z)}$$,
but the bottom is solid, so at $$z = -h$$ we need $$\Psi_z = 0$$,
leaving:

$$\begin{aligned}
    \boxed{
        \Psi(x, z)
        = \Psi(x, -h) - \frac{(z \!+\! h)^2}{2} \Psi_{xx}(x, - h) + \frac{(z \!+\! h)^4}{24} \Psi_{xxxx}(x, - h) - ...
    }
\end{aligned}$$

This result is exact for an inviscid incompressible fluid,
but once the Taylor series is truncated at a finite number of terms,
this is known as the **Boussinesq approximation**.
In effect, this removes all $$z$$-derivatives from the problem,
and will enable us to describe the surface dynamics
based on $$\Psi$$'s behaviour near the channel's bottom.

This expression for $$\Psi$$ gives the flow components,
where we define $$f(x, t) \equiv \Psi_x(x, -h, t)$$
as the value of $$u^{(x)}$$ at the bottom:

$$\begin{aligned}
    u^{(x)}(x, z)
    = \pdv{\Psi}{x}
    &= f(x) - \frac{(z \!+\! h)^2}{2} f_{xx}(x) + \frac{(z \!+\! h)^4}{24} f_{xxxx}(x) - ...
    \\
    u^{(z)}(x, z)
    = \pdv{\Psi}{z}
    &= - (z \!+\! h) f_x(x) + \frac{(z \!+\! h)^3}{6} f_{xxx}(x) - ...
\end{aligned}$$

The Boussinesq approximation is the basis of many other shallow-water wave theories,
most notably the [Korteweg-de Vries equation](/know/concept/korteweg-de-vries-equation/).



## Two coupled equations

Inserting this result (without truncating)
into the kinematic boundary condition with $$z = \eta$$:

$$\begin{aligned}
    0
    &= \pdv{\eta}{t} + \pdv{\eta}{x} \bigg( f - \frac{(\eta \!+\! h)^2}{2} f_{xx} + ... \bigg)
    + \bigg( (\eta \!+\! h) f_{x} - \frac{(\eta \!+\! h)^3}{6} f_{xxx} + ... \bigg)
\end{aligned}$$

And into the free surface boundary condition after differentiating it with respect to $$x$$:

$$\begin{aligned}
    0
    &= \pdv{u^{(x)}}{t} + \frac{1}{2} \pdv{}{x} \bigg( \big(u^{(x)}\big)^2 + \big(u^{(z)}\big)^2 \bigg) + g \pdv{\eta}{x}
    \\
    &= \pdv{}{t} \bigg( f - \frac{(\eta \!+\! h)^2}{2} f_{xx} + ... \bigg)
    + \frac{1}{2} \pdv{}{x} \bigg( f - \frac{(\eta \!+\! h)^2}{2} f_{xx} + ... \bigg)^2 \\
    &\qquad + \frac{1}{2} \pdv{}{x} \bigg( \!-\! (\eta \!+\! h) f_x + \frac{(\eta \!+\! h)^3}{6} f_{xxx} - ... \bigg)^2 + g \pdv{\eta}{x}
    \\
    &= \pdv{}{t} \bigg( f - \frac{(\eta \!+\! h)^2}{2} f_{xx} + ... \bigg)
    + \frac{1}{2} \pdv{}{x} \bigg( f^2 + (\eta \!+\! h)^2 (f_x^2 - f f_{xx}) + ... \bigg) + g \pdv{\eta}{x}
\end{aligned}$$

Switching to a shorter notation for derivatives,
we now have the following set of equations:

$$\begin{aligned}
    0
    &= \eta_t + \eta_x \bigg( f - \frac{(\eta \!+\! h)^2}{2} f_{xx} + ... \bigg)
    + \bigg( (\eta \!+\! h) f_{x} - \frac{(\eta \!+\! h)^3}{6} f_{xxx} + ... \bigg)
    \\
    0
    &= \pdv{}{t} \bigg( f - \frac{(\eta \!+\! h)^2}{2} f_{xx} + ... \bigg)
    + \frac{1}{2} \pdv{}{x} \bigg( f^2 + (\eta \!+\! h)^2 (f_x^2 - f f_{xx}) + ... \bigg) + g \eta_x
\end{aligned}$$

Now we must decide which terms to keep, i.e. where to truncate the expansion.
Let us therefore introduce the characteristic length scales $$\eta \sim a$$ and $$x \sim \lambda$$,
and assume that the water is shallow compared to the waves' length ($$\lambda \gg h$$ by a lot),
and that the waves' amplitude is small compared to the channel's depth ($$h \gg a$$).
Specifically, we assume:

$$\begin{aligned}
    \frac{a}{h}
    \gg \frac{h}{\lambda}
    \gg \frac{a}{\lambda}
\end{aligned}$$

We also introduce characteristic horizontal velocity scales $$u_0$$ and $$f_0$$
respectively at the surface and at the bottom.
Finally, let there be a characteristic time scale $$\lambda / u_0$$ for the surface dynamics.
Inserting all these scales into the two boundary conditions yields:

$$\begin{aligned}
    0
    &\sim \frac{a u_0}{\lambda}
    + \frac{a}{\lambda} \bigg( f_0 - \frac{(a \!+\! h)^2}{2 \lambda^2} f_0 + ... \bigg)
    + \bigg( \frac{(a \!+\! h)}{\lambda} f_0 - \frac{(a \!+\! h)^3}{6 \lambda^3} f_0 + ... \bigg)
    \\
    0
    &\sim \frac{u_0}{\lambda} \bigg( f_0 - \frac{(a \!+\! h)^2}{2 \lambda^2} f_0 + ... \bigg)
    + \frac{1}{2 \lambda} \bigg( f_0^2 + \frac{(a \!+\! h)^2}{\lambda^2} f_0^2 + ... \bigg)
    + \frac{g a}{\lambda}
\end{aligned}$$

Intuitively, we expect that $$u_0 \gg f_0$$, and this is indeed true:
from a linearization of this problem (given in the next section),
it turns out that $$f_0 / u_0 \sim a / h$$ and $$u_0 \approx \sqrt{g h}$$.
Multiplying the former equation by $$1 / u_0$$
and the latter by $$\lambda / u_0^2$$ this leads to:

$$\begin{aligned}
    0
    &\sim \frac{a}{\lambda}
    + \frac{a}{\lambda} \bigg( \frac{a}{h} - \frac{(a \!+\! h)^2}{2 \lambda^2} \frac{a}{h} + ... \bigg)
    + \bigg( \frac{(a \!+\! h)}{\lambda} \frac{a}{h} - \frac{(a \!+\! h)^3}{6 \lambda^3} \frac{a}{h} + ... \bigg)
    \\
    0
    &\sim \bigg( \frac{a}{h} - \frac{(a \!+\! h)^2}{2 \lambda^2} \frac{a}{h} + ... \bigg)
    + \frac{1}{2} \bigg( \frac{a^2}{h^2} + \frac{(a \!+\! h)^2}{\lambda^2} \frac{a^2}{h^2} + ... \bigg)
    + \frac{a}{h}
\end{aligned}$$

The smallest term we will include is $$a h^2 / \lambda^3$$;
anything smaller (specifically containing $$a^2 / \lambda^2$$) will be discarded.
Of course, this decision is arbitrary:
higher-order approximations exist for deeper water and/or taller waves,
but we stick with Boussinesq's original choice, leaving:

$$\begin{aligned}
    0
    &= \eta_t + \eta_x f + (\eta \!+\! h) f_{x} - \frac{h^3}{6} f_{xxx}
    \\
    0
    &= \pdv{}{t} \bigg( f - \frac{h^2}{2} f_{xx} \bigg)
    + \frac{1}{2} \pdv{}{x} \big( f^2 \big) + g \eta_x
\end{aligned}$$

Rearranging this gives the **Boussinesq equations**
for nonlinear waves on shallow water:

$$\begin{aligned}
    \boxed{
        \begin{aligned}
            \pdv{\eta}{t} + \pdv{}{x} \Big( (\eta \!+\! h) f \Big)
            &= \frac{h^3}{6} \pdvn{3}{f}{x}
            \\
            \pdv{f}{t} + f \pdv{f}{x} + g \pdv{\eta}{x}
            &= \frac{h^2}{2} \pdv{}{t} \pdvn{2}{f}{x}
        \end{aligned}
    }
\end{aligned}$$

If we instead took $$\lambda$$ to be even larger compared to $$h$$,
the right-hand sides of these equations would have vanished,
yielding a form of the so-called *shallow water equations*.



## Single equation

We would like to combine these two equations into a single one for $$\eta$$,
but their nonlinear nature makes it hard to eliminate $$f$$ directly.
To get around this, Boussinesq opted to make a lower-order version of these equations,
to use as a guide for some additional approximations
to help handle the higher-order version.

In the above discussion of the terms' relative sizes,
let us instead choose $$a / \lambda$$ as the highest order to include,
thereby reducing the equations to:

$$\begin{aligned}
    \eta_t + h f_x
    = 0
    \qquad \qquad
    f_t + g \eta_x
    = 0
\end{aligned}$$

Respectively differentiating them with respect to $$t$$ and $$x$$,
and then substituting $$f_{xt}$$ in the former using the latter,
we get Lagrange's linear wave equation:

$$\begin{aligned}
    \pdvn{2}{\eta}{t} - g h \pdvn{2}{\eta}{x}
    = 0
\end{aligned}$$

It is well-known that such a problem has a general solution
consisting of an arbitrary forward-moving part $$\eta_{+}$$
and a backward-moving part $$\eta_{-}$$,
both going at a constant velocity $$\sqrt{g h}$$,
and neither of which change shape over time:

$$\begin{aligned}
    \eta(x, t)
    = \eta_{+}\big(x - \sqrt{g h} t\big) + \eta_{-}\big(x + \sqrt{g h} t\big)
\end{aligned}$$

Let us consider only forward-moving waves $$\eta(x \!-\! \sqrt{g h} t)$$,
such that we can rewrite $$t$$-derivatives as $$x$$-derivatives with a factor $$-\sqrt{g h}$$.
Our linearized free-surface equation thus becomes:

$$\begin{aligned}
    0
    = \pdv{\eta}{t} + h \pdv{f}{x}
    = - \sqrt{g h} \pdv{\eta}{x} + h \pdv{f}{x}
    \qquad \implies \qquad
    f
    = \sqrt{\frac{g}{h}} \eta + C
\end{aligned}$$

The integration constant $$C$$ can be removed by absorbing it into $$\eta$$.
Effectively, we have seen that, at least as a first-order approximation,
$$\eta$$ is proportional to $$f$$.
Note that this analysis justifies our earlier assumption
that $$f_0 / u_0 \sim a / h$$ and $$u_0 \approx \sqrt{g h}$$.

Armed with this knowledge, we return
to the higher-order equations after some rearranging:

$$\begin{aligned}
    \eta_{t} + h f_x + \pdv{}{x} \bigg( \eta f - \frac{h^3}{6} f_{xx} \bigg)
    &= 0
    \\
    f_{t} + g \eta_{x} + \frac{1}{2} \pdv{}{x} \bigg( f^2 - h^2 f_{xt} \bigg)
    &= 0
\end{aligned}$$

Inserting $$f = \sqrt{g / h} \eta$$ into both equations
and multiplying the latter by $$h$$ yields:

$$\begin{aligned}
    \eta_{t} + \sqrt{g h} \eta_x + \sqrt{\frac{g}{h}} \pdv{}{x} \bigg( \eta^2 - \frac{h^3}{6} \eta_{xx} \bigg)
    &= 0
    \\
    \sqrt{g h} \eta_{t} + g h \eta_{x} + \frac{1}{2} \pdv{}{x} \bigg( g \eta^2 - h^2 \sqrt{g h} \eta_{xt} \bigg)
    &= 0
\end{aligned}$$

Respectively differentiating by $$t$$ and $$x$$
and assuming a travelling wave $$\eta(x \!-\! \sqrt{g h} t)$$
such that we can rewrite $$\ipdv{}{t} = - \sqrt{g h} \ipdv{}{x}$$:

$$\begin{aligned}
    \eta_{tt} + \sqrt{g h} \eta_{xt} - g \pdvn{2}{}{x} \bigg( \eta^2 - \frac{h^3}{6} \eta_{xx} \bigg)
    &= 0
    \\
    \sqrt{g h} \eta_{tx} + g h \eta_{xx} + \frac{g}{2} \pdvn{2}{}{x} \bigg( \eta^2 + h^3 \eta_{xx} \bigg)
    &= 0
\end{aligned}$$

Subtracting the latter from the former yields the following equation containing only $$\eta$$:

$$\begin{aligned}
    \eta_{tt} + \sqrt{g h} \eta_{xt} - \sqrt{g h} \eta_{tx} - g h \eta_{xx}
    - \pdvn{2}{}{x} \bigg( g \Big( \eta^2 - \frac{h^3}{6} \eta_{xx} \Big) + \frac{g}{2} \Big( \eta^2 + h^3 \eta_{xx} \Big) \bigg)
    &= 0
\end{aligned}$$

After cleaning up, this becomes the **Boussinesq equation** for the shape of travelling waves:

$$\begin{aligned}
    \boxed{
        \pdvn{2}{\eta}{t} - g h \pdvn{2}{\eta}{x} - g h \pdvn{2}{}{x} \bigg( \frac{3}{2} \frac{\eta^2}{h} + \frac{h^2}{3} \pdvn{2}{\eta}{x} \bigg)
        = 0
    }
\end{aligned}$$

Clearly, the assumption of non-deforming waves $$\eta(x \!-\! \sqrt{g h} t)$$
was essential to get this equation.
But what if solving it yields a wave without that property?
Can it be trusted?
Fortunately yes: the first two terms ($$\eta_{tt}$$ and $$g h \eta_{xx}$$),
were not affected by that assumption (this is easy to see),
and the others are small according to the characteristic scales:

$$\begin{aligned}
    0
    \sim \frac{a u_0^2}{\lambda^2} - g h \frac{a}{\lambda^2}
    - g h \frac{1}{\lambda^2} \bigg( \frac{3}{2} \frac{a^2}{h} + \frac{h^2}{3} \frac{a}{\lambda^2} \bigg)
\end{aligned}$$

After dividing out $$a / \lambda$$,
we see that the last two terms are roughly $$a / \lambda$$ and $$h^3 / \lambda^3$$,
meaning they are much smaller than the first two,
which are both on the order of $$h / \lambda$$.



## Dimensionless form

Let us non-dimensionalize the equation by introducing
dimensionless quantities $$\tilde{\eta}$$, $$\tilde{t}$$ and $$\tilde{x}$$:

$$\begin{aligned}
    \tilde{\eta}(\tilde{x}, \tilde{t})
    = \frac{\eta(x, t)}{\eta_c}
    \qquad \qquad
    \tilde{t}
    = \frac{t}{t_c}
    \qquad \qquad
    \tilde{x}
    = \frac{x}{x_c}
\end{aligned}$$

Where $$\eta_c$$, $$t_c$$ and $$x_c$$ are unspecified scale parameters.
We rewrite the Boussinesq equation with these quantities
by using the chain rule of differentiation,
and divide by $$\eta_c / t_c^2$$:

$$\begin{aligned}
    0
    &= \tilde{\eta}_{\tilde{t} \tilde{t}}
    - \frac{g h t_c^2}{x_c^2} \tilde{\eta}_{\tilde{x} \tilde{x}}
    - \pdvn{2}{}{x} \bigg( \frac{3 g \eta_c t_c^2}{2 x_c^2} \tilde{\eta}^2 + \frac{g h^3 t_c^2}{3 x_c^4} \tilde{\eta}_{\tilde{x} \tilde{x}} \bigg)
\end{aligned}$$

Now we must choose values for $$\eta_c$$, $$t_c$$ and $$x_c$$
such that the prefactors become simple constants.
Conventionally it is demanded that:

$$\begin{aligned}
    \frac{g h t_c^2}{x_c^2}
    = 1
    \qquad \qquad
    \frac{3 g \eta_c t_c^2}{2 x_c^2}
    = 3
    \qquad \qquad
    \frac{g h^3 t_c^2}{3 x_c^4}
    = 1
\end{aligned}$$

Solving this system of equations yields the following values for the scale parameters:

$$\begin{aligned}
    \eta_c
    = 2 h
    \qquad \qquad
    t_c
    = \sqrt{\frac{h}{3 g}}
    \qquad \qquad
    x_c
    = \frac{h}{\sqrt{3}}
\end{aligned}$$

And the Boussinesq equation is reduced to its standard dimensionless form:

$$\begin{aligned}
    \boxed{
        \pdvn{2}{\tilde{\eta}}{\tilde{t}} - \pdvn{2}{\tilde{\eta}}{\tilde{x}}
        - \pdvn{2}{}{x} \bigg( 3 \tilde{\eta}^2 + \pdvn{2}{\tilde{\eta}}{\tilde{x}} \bigg)
        = 0
    }
\end{aligned}$$

Many authors flip the sign of $$\tilde{\eta}_{\tilde{x}\tilde{x}\tilde{x}\tilde{x}}$$
to get the so-called "good" Boussinesq equation
(as opposed to the "bad" one above).
For fluid surface waves, this is unphysical,
but it makes the problem more well-behaved mathematically;
the details are beyond the scope of this article.



## Soliton solution

Let us make an ansatz for $$\tilde{\eta}$$ that describes a wave
with a fixed shape propagating in the positive $$\tilde{x}$$-direction
at dimensionless velocity $$v$$:

$$\begin{aligned}
    \tilde{\eta}(\tilde{x}, \tilde{t})
    = \phi(\xi)
    \qquad
    \xi
    \equiv \tilde{x} - v \tilde{t}
    \qquad \implies \qquad
    \pdv{}{\tilde{t}}
    = -v \pdv{}{\xi}
    \qquad
    \pdv{}{\tilde{x}}
    = \pdv{}{\xi}
\end{aligned}$$

With this, the Boussinesq equation becomes a nonlinear ordinary differential equation:

$$\begin{aligned}
    0
    &= (v^2 - 1) \phi_{\xi \xi} - \pdvn{2}{}{\xi} (3 \phi^2 + \phi_{\xi \xi})
\end{aligned}$$

We abbreviate $$w \equiv v^2 \!-\! 1$$ and integrate twice,
introducing integration constants $$A$$ and $$B$$:

$$\begin{aligned}
    w \phi - 3 \phi^2 - \phi_{\xi \xi}
    = A \xi + B
\end{aligned}$$

We restrict ourselves to localized solutions
by demanding that $$\phi \to 0$$ for $$\xi \to \pm \infty$$.
This implies that also $$\phi_\xi \to 0$$ and $$\phi_{\xi \xi} \to 0$$,
meaning that we must set $$A = B = 0$$ to satisfy the equation at infinity.
The remaining terms are multiplied by $$\phi_\xi$$ to give:

$$\begin{aligned}
    0
    &= w \phi \phi_\xi - 3 \phi^2 \phi_\xi - \phi_{\xi \xi} \phi_\xi
    = \frac{1}{2} \pdv{}{\xi} \Big( w \phi^2 - 2 \phi^3 - (\phi_\xi)^2 \Big)
\end{aligned}$$

Integrating (and dropping the integration constant due to localization) yields:

$$\begin{aligned}
    (\phi_\xi)^2
    = \phi^2 (w - 2 \phi)
\end{aligned}$$

Because $$\phi_\xi$$ is real, we need the right-hand side to be positive,
so $$w > 2 \phi$$; for $$\phi \to 0$$, this means that $$w > 0$$.
This equation is similar to the one encountered when solving
the [Korteweg-de Vries equation](/know/concept/korteweg-de-vries-equation/)
and is integrated in the same way; look there for details.
The result is:

$$\begin{aligned}
    \boxed{
        \tilde{\eta}(\tilde{x}, \tilde{t})
        = \frac{w}{2} \sech^2\!\bigg( \frac{\sqrt{w}}{2} \big( \tilde{x} - \sqrt{1 \!+\! w} \: \tilde{t} - \tilde{x}_0 \big) \bigg)
    }
\end{aligned}$$

This is known as a **soliton**.
Reintroducing units by replacing $$\tilde{\eta} = \eta / \eta_c$$ etc.
leads to:

$$\begin{aligned}
    \boxed{
        \eta(x, t)
        = w h \sech^2\!\bigg( \frac{\sqrt{3 w}}{2 h} \Big( x - \sqrt{(1 \!+\! w) g h} \: t - x_0 \Big) \bigg)
    }
\end{aligned}$$

Note that Boussinesq's original calculation had $$(1 \!+\! w/2)$$ instead of $$\sqrt{1 \!+\! w}$$;
the former is simply a first-order approximation of the latter.
Recall that $$\sqrt{g h}$$ is the phase velocity of Lagrange's linear theory:
this shows that nonlinear waves are faster,
and speed up with amplitude.



## References
1.  J. Boussinesq,
    [Théorie des ondes et des remous qui se propagent le long