Essential Graduate Physics by Konstantin K. Likharev - HTML preview

PLEASE NOTE: This is an HTML preview only and some elements such as links or page numbers may be incorrect.
Download the book in PDF, ePub, Kindle for a complete version.

Page 32 of 68

Image 1739

Image 1740

Image 1741

Image 1742

Image 1743

Image 1744

Image 1745

Image 1746

Image 1747

Image 1748

Image 1749

Image 1750

Image 1751

Image 1752

Image 1753

Image 1754

Image 1755

Image 1756

Image 1757

Image 1758

Essential Graduate Physics

EM: Classical Electrodynamics

H

,

(7.106)

t dr

I

C

where I is the total current flowing along the conductor (or rather its complex amplitude). But, as was mentioned above, in the TEM wave the ratio Et/ Ht of the field components participating in these two integrals is constant and equal to Z = (/)1/2, so that Eqs. (105)-(106) give the following simple relation between the “global” variables of the conductor:

 / 

I

.

(7.107)

Z

1/2

k

This important relation may be also obtained in a different way; let me describe it as well, because (as we will see below) it has an independent heuristic value. Let us consider a small segment dz

<<  = 2/ k of the line’s conductor, and apply the electric charge conservation law (4.1) to the instant values of the linear charge density and current. The cancellation of dz in both parts yields

 ( z, t)

I

 ( z, t)

 

.

(7.108)

t

z

If we accept the sinusoidal waveform, exp{ i( kz – t)}, for both these variables, we immediately recover Eq. (107) for their complex amplitudes, showing that this relation expresses just the charge continuity law.

The global equation (108) may be made more specific in the case when the frequency dependence of  and  is negligible, and the transmission line consists of just two isolated conductors –

see, e.g., Fig. 17. In this case, to have the wave localized in the space near the two conductors, we need a sufficiently fast decrease of its electric field at large distances. For that, their linear charge densities for each value of z should be equal and opposite, and we can simply relate them to the potential difference V

between the conductors:

( z, t)

C ,

(7.109)

V ( z, t)

0

where C 0 is the mutual capacitance of the conductors (per unit length) – which was repeatedly discussed in Chapter 2. Then Eq. (108) takes the following form:

V

 ( z, t)

I

 ( z, t)

C

 

.

(7.110)

0

t

z

Next, let us consider the contour shown with the red dashed line in Fig. 18 (which shows a different cross-section of the transmission line – by a plane containing the wave propagation axis z), and apply to it the Faraday induction law (6.3).

I

I ( z, t)

I ( z, t) 

dz

z

V

V ( z, t)

d  L I ( z, t) dz

V ( z, t) 

dz

0

z

Fig. 7.18. Electric current, magnetic flux, and

voltage in a two-conductor transmission line.

dz

Chapter 7

Page 33 of 68

Essential Graduate Physics

EM: Classical Electrodynamics

Since, in the coarse-grain approximation, the electric field inside the conductors (in Fig. 18, on the horizontal segments of the contour) vanishes, the total e.m.f. equals the difference of the voltages V

at the ends of the segment dz, while the only sources of the magnetic flux through the area limited by the contour are the (equal and opposite) currents  I in the conductors, we can use Eq. (5.70) to express the flux. As a result, canceling dz in both parts of the equation, we get

I

 ( z, t)

V

 ( z, t)

L

 

,

(7.111)

0

t

z

where L 0 is the mutual inductance of the conductors per unit length. The only difference between this L 0

and the dc mutual inductances discussed in Chapter 5 is that at the high frequencies we are analyzing now, L 0 should be calculated neglecting the magnetic field penetration into the conductors. (In the dc case, we had the same situation for superconductor electrodes within their coarse-grain, ideal-diamagnet description.)

The system of Eqs. (110) and (111) is frequently called the telegrapher’s equations. Combined, they give for any “global” variable f (either V, or I, or ) the usual 1D wave equation, 2

2

f

f

L C

 0 ,

(7.112)

2

0

0

2

z

t

which describes dispersion-free TEM wave’s propagation. Again, this equation is only valid within the frequency range where the frequency dependence of both  and  is negligible. If this is not so, the global approach may still be used for sinusoidal waves f = Re[ fexp{ i( kz – t)}]. Repeating the above arguments, instead of Eqs. (110)-(111) we get a more general system of two algebraic equations

C V kI ,

L I kV ,

(7.113)

0 

0 

in which L 0   and C 0   may now depend on frequency. These equations are consistent only if 2

k

1

L 0 C 0

L C

  .

(7.114) product

0

0

 2

2

v

invariance

Besides the fact we have already known (that the TEM wave’s speed is the same as that of the plane wave), Eq. (114) gives us the result that I confess was not emphasized enough in Chapter 5: the product L 0 C 0 does not depend on the shape or size of line’s cross-section, provided that the magnetic field’s penetration into the conductors is negligible). Hence, if we have calculated the mutual capacitance C 0 of a system of two cylindrical conductors, the result immediately gives us their mutual inductance: L 0 =

/ C 0. This relationship stems from the fact that both the electric and magnetic fields may be expressed via the solution of the same 2D Laplace equation for the system’s cross-section.

With Eq. (114) satisfied, any of Eqs. (113) gives the same result for the following ratio: 1/ 2

V

L

Transmission

0

Z

,

(7.115) line’s TEM

W





I

Impedance

C 0 

which is called the transmission line’s impedance. This parameter has the same dimensionality (in SI units – ohms, denoted ) as the wave impedance (7),

Chapter 7

Page 34 of 68

Image 1759

Image 1760

Image 1761

Image 1762

Image 1763

Image 1764

Image 1765

Image 1766

Image 1767

Image 1768

Image 1769

Image 1770

Image 1771

Image 1772

Image 1773

Image 1774

Image 1775

Image 1776

Image 1777

Image 1778

Image 1779

Image 1780

Image 1781

Image 1782

Image 1783

Image 1784

Image 1785

Image 1786

Image 1787

Image 1788

Image 1789

Image 1790

Image 1791

Image 1792

Image 1793

Image 1794

Image 1795

Image 1796

Image 1797

Image 1798

Essential Graduate Physics

EM: Classical Electrodynamics

1/ 2

E

  

Z

   ,

(7.116)

H

  

but these parameters should not be confused, because ZW depends on the cross-section’s geometry, while Z does not. In particular, ZW is the only important parameter of a transmission line for its matching with a lumped load circuit (Fig. 19) in the important case when both the cable cross-section’s size and the load’s linear dimensions are much smaller than the wavelength.48

Z

I

W

Z ()

L

V

Fig. 7.19. Passive, lumped

termination of a TEM

transmission line.

z

z

0

Indeed, in this case, we may consider the load in the quasistatic limit and write

V ( z )  Z () I ( z )

,

(7.117)

0

L

0

where ZL() is the (generally complex) impedance of the load. Taking V( z,t) and I( z,t) in the form similar to Eqs. (61) and (62), and writing the two Kirchhoff’s circuit laws for the point z = z 0, we get for the reflection coefficient a result similar to Eq. (68):

Z ()  Z

L

W

R

.

(7.118)

Z ()  Z

L

W

This formula shows that for the perfect matching (i.e. the total wave absorption in the load), the load’s impedance ZL() should be real and equal to ZW – but not necessarily to Z.

As an example, let us consider one of the simplest (and most practically important) transmission lines: the coaxial cable (Fig. 20).49

 ,  b a

0

a

Fig. 7. 20. The cross-section of a coaxial cable

with (possibly, dispersive) dielectric filling.

For this geometry, we already know the expressions for both C 0 and L 0,50 though they have to be modified for the account of arbitrary dielectric and magnetic constants, and the magnetic field’s non-penetration into the conductors. As a result of this (elementary) modification, we get the formulas, 48 The ability of TEM lines to have such a small cross-section is another important practical advantage.

49 It was invented by the same O. Heaviside in 1880.

50 See, respectively, Eqs. (2.49) and (5.79).

Chapter 7

Page 35 of 68

Image 1799

Essential Graduate Physics

EM: Classical Electrodynamics

2

Coaxial

C

,

L

ln( b / a) ,

(7.119) cable’s

0

ln( b / a)

0

2

C 0 and L 0

illustrating that the universal relationship (114) is indeed valid. For the cable’s impedance (115), Eqs.

(119) yield a geometry-dependent value

  1/2

b

ln( / a)

b

ln( / a)

Z   

Z

Z .

(7.120)

W

  

2

2

For the standard TV antenna cables (such as RG-6/U, with b/ a ~ 3, /0  2.2), ZW = 75 , while for most computer component connections, coaxial cables with ZW = 50  (such as RG-58/U) are prescribed by electronic engineering standards. Such cables are broadly used for the transmission of electromagnetic waves with frequencies up to 1 GHz over distances of a few km, and up to ~20 GHz on the tabletop scale (a few meters), limited by wave attenuation – see Sec. 9 below.

Moreover, the following two facts enable a wide application, in electrical engineering and physical experiment, of coaxial-cable- like systems. First, as Eq. (5.78) shows, in a cable with a << b, most energy of the wave is localized near the internal conductor. Second, the theory to be discussed in the next section shows that excitation of other ( H- and E-) waves in the cable is impossible until the wavelength  becomes smaller than ~( a + b). As a result, the TEM mode propagation in a cable with a

<< b < / is not much affected even if the internal conductor is not straight, but bent – for example, into a helix – see, e.g., Fig. 21.

Fig. 7.21. A typical traveling-wave tube:

(1) electron gun, (2) ac input, (3) beam-

focusing magnets, (4) wave attenuator,

(5) helix coil, (6) ac output, (7) vacuum

tube, (8) electron collector. Adapted from

https://en.wikipedia.org/wiki/Traveling-

wave_tube under the Creative Commons

BY-SA 3.0 license.

In such a system, called the traveling-wave tube (TWT), a quasi-TEM wave propagates with velocity vc along the helix’s length, so that the velocity’s component along the cable’s axis may be made close to the velocity u << c of the electron beam moving ballistically along the tube’s axis, enabling their effective interaction, and as a result, a length-accumulating amplification of the wave.51

Another important example of a TEM transmission line is a set of two parallel wires. In the form of twisted pairs,52 they allow communications, in particular long-range telephone and DSL Internet 51 Despite the current prevalence of semiconductor devices in electronics, TWTs are still used in satellite TV and radio systems, because they may work at very high microwave power – e.g., up to 200W at 20

GHz and pulsed 50W at 200 GHz. Very unfortunately, in this course, I will not have time/space to discuss even the (rather elegant) basic theory of such devices. The reader interested in this field may be referred, for example, to the detailed monograph by J. Whitaker, Power Vacuum Tubes Handbook, 3rd ed., CRC Press, 2017.

52 Such twisting, around the line’s direction axis, reduces the crosstalk between adjacent lines, and the parasitic radiation at their bends.

Chapter 7

Page 36 of 68

Essential Graduate Physics

EM: Classical Electrodynamics

connections, at frequencies up to a few hundred kHz, as well as relatively short, multi-line Ethernet and TV cables at frequencies up to ~ 1 GHz, limited mostly by the mutual interference (“crosstalk”) between the individual lines of the same cable, and the unintentional radiation of the wave into the environment.

7.6. Waveguides: H and E waves

Let us now return to Eqs. (100) and explore the H- and E-waves – with, respectively, either Hz or Ez different from zero. At the first sight, they may seem more complex. However, Eqs. (101), which determine the distribution of these longitudinal components over the cross-section, are just the 2D

Helmholtz equations for scalar functions. For simple cross-section geometries, they may be readily solved using the methods discussed for the Laplace equation in Chapter 2, in particular the variable separation. After the solution of such an equation has been found, the transverse components of the fields may be calculated by differentiation, using the simple formulas,

i

i

k

E

k E

kZ n

H

H

k H

n

E

(7.121)

t

(

)

z

t

z

z

t

z ,

 ( 

) ,

2

t

2  z

t

z

z

t

z

k

k

Z

t

t

which follow from the first line of Eqs. (100).53

In comparison with the boundary problems of electro- and magnetostatics, the only conceptually new feature of Eqs. (101) is that they form the so-called eigenproblems, with typically many solutions ( eigenfunctions), each describing a specific wave mode, and corresponding to a specific eigenvalue of the parameter kt. The good news here is that these values of kt are determined by this 2D boundary problem and hence do not depend on kz. As a result, the dispersion law ( kz) of any mode, which follows from the last form of Eq. (102),

1/ 2

Universal

2

2

k k

dispersion

 

z

t

  2 2

2

v k  

,

(7.122)

z

c 1/ 2

relation

  

is functionally similar for all modes. It is also similar to that of plane waves in plasma (see Eq. (38), Fig.

6, and their discussion in Sec. 2), with the only differences that the speed in light c is generally replaced with v = 1/()1/2 – the speed of the plane or TEM waves in the medium filling the waveguide, and that

p is replaced with the so-called cutoff frequency

  vk ,

(7.123)

c

t

specific for each mode. (As Eq. (101) implies, and as we will see from several examples below, kt has the order of 1/ a, where a is the characteristic dimension of the waveguide’s cross-section, so that the critical value of the free-space wavelength   2 c/ is of the order of a.) Below the cutoff frequency of each particular mode, such wave cannot propagate in the waveguide.54 As a result, the modes with the 53 For the derivation of Eqs. (121), one of these two linear equations should be first vector-multiplied by n z. Note also that this approach could not be used to analyze the TEM waves, because for them kt = 0, Ez = 0, Hz = 0, and Eqs. (121) yield uncertainty.

54 An interesting twist in the ideas of electromagnetic metamaterials (mentioned in Sec. 5 above) is the so-called

-near-zero materials, designed to have the effective product  much lower than 00 within certain frequency ranges. Since at these frequencies the speed v (4) becomes much lower than c, the cutoff frequency (123) virtually Chapter 7

Page 37 of 68

Image 1800

Image 1801

Image 1802

Image 1803

Image 1804

Image 1805

Image 1806

Image 1807

Essential Graduate Physics

EM: Classical Electrodynamics

lowest values of c present special practical interest, because the choice of the signal frequency 

between the two lowest values of the cutoff frequency (123) guarantees that the waves propagate in the form of only one mode, with the lowest kt. Such a choice enables engineers to simplify the excitation of the desired mode by wave generators and to avoid the unintentional transfer of electromagnetic wave energy to undesirable modes by (virtually unavoidable) small inhomogeneities of the system.

The boundary conditions for the Helmholtz equations (101) depend on the propagating wave type. For the E-modes, with Hz = 0 but Ez  0, the condition E = 0 immediately gives E

 ,

0

(7.124)

z C

where C is the inner contour limiting the conducting wall’s cross-section. For the H-modes, with Ez = 0

but Hz  0, the boundary condition is slightly less obvious and may be obtained using, for example, the second equation of the system (100), vector-multiplied by n z. Indeed, for the component normal to the conductor surface, the result of such multiplication is

k

H

ik (H )  i

n E

.

(7.125)

z

t n

z

t

z

Z

n

n

But the first term on the left-hand side of this relation must be zero on the wall surface, because of the second of Eqs. (104), while according to the first of Eqs. (104), the vector E t in the second term cannot have a component tangential to the wall. As a result, the vector product in that term cannot have a normal component, so that the term should equal zero as well, and Eq. (125) is reduced to

H

z

 0 .

(7.126)

C

n

Let us see how all this machinery works for a simple but practically important case of a metallic-wall waveguide with a rectangular cross-section – see Fig. 22

y

b

H t

Fig. 7.22. A rectangular waveguide, and the

E

transverse

field

distribution

in

its

0

a

fundamental mode H 10 (schematically).

x

In the natural Cartesian coordinates shown in this figure, both Eqs. (101) take the simple form

 2

2

E , for E -

modes,

 2

k

f  ,

0

where f

z



(7.127)

2

2

t 

  x

y

H , for H - modes.

z

From Chapter 2, we know that the most effective way of solution of such equations in a rectangular region is the variable separation, in which the general solution is represented as a sum of partial solutions of the type

vanishes. As a result, the waves may “tunnel” through very narrow sections of metallic waveguides filled with such materials – see, e.g., M. Silveirinha and N. Engheta, Phys. Rev. Lett. 97, 157403 (2006).

Chapter 7

Page 38 of 68

Essential Graduate Physics

EM: Classical Electrodynamics

f X ( x) Y ( y) .

(7.128)

Plugging this expression into Eq. (127), and dividing each term by XY, we get the equation, 1

2

d X

1 2

d Y

2

k  0,

(7.129)

2

2

t

X dx

Y dy

which should be satisfied for all values of x and y within the waveguide’s interior. This is only possible if each term of the sum equals a constant. Taking the X-term and Y-term constants in the form (– k 2

x ) and

(– k 2

y ), respectfully, and solving the corresponding ordinary differential equations,55 for the eigenfunction (128) we get

f   c cos k x s sin k x c

k y s

k y

k k k ,

(7.130)

x

x

x

x 

cos

sin

y

y

y

y

2

2

2

,

with x

y

t

where the constants c and s should be found from the boundary conditions. Here the difference between the H-modes and E-modes kicks in.

For the H-modes, Eq. (130) is valid for Hz, and we should use the boundary condition (126) on all metallic walls of the waveguide, i.e. at x = 0 and a; and y = 0 and b – see Fig. 22. As a result, we get very simple expressions for eigenfunctions and eigenvalues:

nx

my

H

cos

cos

,

(7.131)

z

H

nm

l

a

b

1/ 2

2

2

n

m

n

m

 

 

k

, k

,

(

that

so

k )

(7.132)

x

y

t nm

 2 2

k k

x

y 1/ 2 

      ,

a

b

 a

b  

where Hl is the longitudinal field’s amplitude, and n and m are two integer numbers – each of them arbitrary besides that they cannot be equal to zero simultaneously.56 Assuming, just for certainty, that a

b (as shown in Fig. 22), we see that the lowest eigenvalue of kt and hence the lowest cutoff frequency (123) are achieved for the so-called H 10 mode with n = 1 and m = 0, and hence with Fundamental

mode’s

( k ) 

,

(7.133)

t 10

cutoff

a

thus confirming our prior estimate of kt.

Depending on the a/ b ratio, the second-lowest kt (and hence c) belongs to either the H 11 mode with n = 1 and m = 1:

1/ 2

1/ 2

2

 1

1

a

( k )   

 1

     ( k ) ,

(7.134)

t 11

2

2

t 10

a

b

b

 

or to the H 20 mode with n = 2 and m = 0:

55 Let me hope that the solution of equations of the type d 2 X/ dx 2 + k 2

x X = 0 does not present any problem for the

reader, at least due to their prior experience with problems such as standing waves on a guitar string, wavefunctions in a flat 1D quantum well, or (with the replacement xt) a classical harmonic oscillator.

56 Otherwise, the function Hz( x,y) would be constant, so that, according to Eq. (121), the transverse components of the electric and magnetic field would equal zero. As a result, as the last two lines of Eqs. (100) show, the whole field would be zero for any kz  0.

Chapter 7

Page 39 of 68

Essential Graduate Physics

EM: Classical Electrodynamics

2

( k ) 

 (

2 k ) .

(7.135)

t 20

t 10

a

These values become equal at a/ b = 3  1.7; in practical waveguides, the a/ b ratio is not too far from this value. For example, in the standard X-band (~10-GHz) waveguide WR90, a  2.3 cm ( f c  c/2 

6.5 GHz), and b  1.0 cm.

Now let us have a look at the alternative E-modes. For them, we still should use the general solution (130) with f = Ez, but now with the boundary condition (124). This gives us the eigenfunctions

nx

my

E

sin

sin

,

(7.136)

z

E

nm

l

a

b

and the same eigenvalue spectrum (132) as for the H modes. However, now neither n nor m can be equal to zero; otherwise, Eq. (136) would give the trivial solution Ez( x,y) = 0. Hence the lowest cutoff frequency of TM waves is achieved at the so-called E 11 mode with n =1, m = 1, and with the eigenvalue given by Eq. (134), always higher than ( kt)10.

Thus the fundamental H 10 mode is certainly the most important wave in rectangular waveguides; let us have a better look at this field distribution. Plugging the corresponding solution (131) with n = 1

and m = 0 into the general relation (121), we easily get

k a

x

( H )   z

i

H sin

,

( H )  ,

0

(7.137)

x 10

l

y 10

a

ka

x

( E )  ,

0

( E )  i

ZH sin

.

(7.138)

x 10

y 10

l

a

This field distribution is (schematically) shown in Fig. 22. Neither of the fields depends on the coordinate y – the feature very convenient, in particular, for microwave experiments with small samples.

The electric field has only one (in Fig. 22, vertical) component that vanishes at the side walls and reaches its maximum at the waveguide’s center; its field lines are straight, starting and ending on wall surface charges (whose distribution propagates along the waveguide together with the wave). In contrast, the magnetic field has two non-zero components ( Hx and Hz), and its field lines are shaped as horizontal loops wrapped around the electric field maxima.

An important question is whether the H 10 wave may be usefully characterized by a unique impedance introduced similarly to ZW of the TEM modes – see Eq. (115). The answer is not, because the main value of ZW is a convenient description of the impedance matching of a transmission line with a lumped load – see Fig. 19 and Eq. (118). As was discussed above, such a simple description is possible (i.e., does not depend on the exact geometry of the connection) only if both dimensions of the line’s cross-section are much less than . But for the H 10 wave (and more generally, any non-TEM mode) this is impossible – see, e.g., Eq. (129): its lowest frequency corresponds to the TEM wavelength max =

2/( kt)min = 2/( kt)10 = 2 a. (The reader is challenged to find a simple interpretation of this equality.) Now let us consider metallic-wall waveguides with a round cross-section (Fig. 23a). In this single-connected geometry, the TEM waves are impossible again, while for the analysis of H-modes and E-modes, the polar coordinates {, } are most natural. In these coordinates, the 2D Helmholtz equation (101) takes the following form:

Chapter 7

Page 40 of 68

Image 1808

Image 1809

Image 1810

Image 1811

Image 1812

Image 1813

Image 1814

Image 1815

Image 1816

Image 1817

Image 1818

Image 1819

Image 1820

Image 1821

Image 1822

Image 1823

Image 1824

Image 1825

Image 1826

Image 1827

Image 1828

Image 1829

Image 1830

Image 1831

Image 1832

Image 1833

Image 1834

Image 1835

Image 1836

Image 1837

Image 1838

Image 1839

Image 1840

Image 1841

Image 1842

Image 1843

Image 1844

Image 1845

Image 1846

Image 1847

Image 1848

Image 1849

Image 1850

Essential Graduate Physics

EM: Classical Electrodynamics

 1    

1 2

E

E

2

, for - modes,

z





k

f

f

(7.139)

2

2

t

 ,

0

where

 

       

H , for H - modes.

z

Separating the variables as f = R() F(), we get

1 d dR

1

2

d

F

2

k  0





.

(7.140)

2

2

R d  d 

t

F d

But this is exactly the Eq. (2.127) that was studied in Sec. 2.7 in the context of electrostatics, just with a replacement of notation:   kt. So we already know that to have 2-periodic functions F() and finite values R(0) (which are evidently necessary for our current case – see Fig. 23a), the general solution must have the form given by Eq. (2.136), i.e. the eigenfunctions are expressed via integer-order Bessel functions of the first kind:

f

J

,

(7.141)

nm

n k

nm

 c cos 

n s sin 

n

n

n

  const Jnk

nm

cos n 0 

with the eigenvalues knm of the transverse wave number kt to be determined from appropriate boundary conditions, and an arbitrary constant 0.

(a)

(b)

 , 

R

 , 

R

0

0

Fig. 7.23. (a) Metallic and (b) dielectric

, 

waveguides with circular cross-sections.

As for the rectangular waveguide, let us start from the H-modes ( f = Hz). Then the boundary condition on the wall surface ( = R) is given by Eq. (126), which, for the solution (141), takes the form d

J

( )  ,

0

where

kR .

(7.142)

d

n

This means that the eigenvalues of Eq. (139) are

'

k k

nm

,

(7.143)

t

nm

R

where  ’nm is the m th zero of the function dJn()/ d. Approximate values of these zeros for several lowest n and m may be read out from Fig. 2.18; their more accurate values are given in Table 1 below.

Table 7.1. Zeros 

nm of the function dJn()/ d for a few lowest

values of the Bessel function’s index n and the root’s number m.

m = 1

2

3

n = 0

3.83171

7.015587

10.1735

1

1.84118

5.33144

8.53632

2

3.05424

6.70613

9.96947

3

4.20119

8.01524

11.34592

Chapter 7

Page 41 of 68

Image 1851

Image 1852

Image 1853

Image 1854

Image 1855

Image 1856

Image 1857

Image 1858

Image 1859

Essential Graduate Physics

EM: Classical Electrodynamics

The table shows, in particular, that the lowest of the zeros is  11  1.84.57 Thus, perhaps a bit counter-intuitively, the fundamental mode, providing the lowest cutoff frequency c = vknm, is H 11, corresponding to n = 1 rather than n = 0:

'  

H H J

.

(7.144)

z

l



cos   

1

11

0 

R

It has the transverse wave number is kt = k 11 =  11/ R  1.84/ R, and hence the cutoff frequency corresponding to the TEM wavelength max = 2/ k 11  3.41 R. Thus the ratio of max to the waveguide’s diameter 2 R is about 1.7, i.e. is close to the ratio max/ a = 2 for the rectangular waveguide. The origin of this proximity is clear from Fig. 24, which shows the transverse field distribution in the H 11 mode. (It may be readily calculated from Eqs. (121) with Ez = 0, and Hz given by Eq. (144).) E

H

Fig. 7.24. Transverse field components in the

fundamental H 11 mode of a metallic, circular

waveguide (schematically).

One can see that the field structure is actually very similar to that of the fundamental mode in the rectangular waveguide, shown in Fig. 22, despite the different nomenclature (which is due to the different coordinate system used for the solution). However, note the arbitrary constant angle 0, indicating that in circular waveguides, the transverse field’s polarization is arbitrary. For some practical applications, such degeneracy of these “quasi-linearly-polarized” waves creates problems; some of them may be avoided by using waves with circular polarization.

As Table 1 shows, the next lowest H-mode is H 21, for which kt = k 21 =  21/ R  3.05/ R, almost twice larger than that of the fundamental mode, and only then comes the first mode with no angular dependence of any field, H 01, with kt = k 01 =  01/ R  3.83/ R,58 followed by several angle-dependent modes: H 31, H 12, etc.

For the E modes, we may still use Eq. (141) (with f = Ez), but with the boundary condition (124) at  = R. This gives the following equation for the problem eigenvalues:

J

i.e. k

nm

,

(7.145)

n k

R

nm

  ,0

nm

R

where  nm is the m th zero of function Jn() – see Table 2.1. That table shows that the lowest kt is equal to

01/ R  2.405/ R. Hence the corresponding mode ( E 01), with no angular dependence of its fields, e.g.

57 Mathematically, the lowest root of Eq. (142) with n = 0 equals 0. However, it would yield k = 0 and hence a constant field Hz, which, according to the first of Eqs. (121), would give a vanishing electric field.

58 The electric field lines in the H 01 mode (as well as all higher H 0 m modes) are directed straight from the symmetry axis to the walls, reminding those of the TEM waves in the coaxial cable. Due to this property, these modes provide, at  >> c, much lower energy losses (see Sec. 9 below) than the fundamental H 11 mode, and are sometimes used in practice, despite the inconvenience of working in the multimode frequency range.

Chapter 7

Page 42 of 68

Essential Graduate Physics

EM: Classical Electrodynamics

 

E E J

,

(7.146)

z

l

0  01

R

has the second-lowest cutoff frequency, ~30% higher than that of the fundamental mode H 11.

Finally, let us discuss one more topic of general importance – the number N of electromagnetic modes that may propagate in a waveguide within a certain range of relatively large frequencies  >> c.

It is easy to calculate for a rectangular waveguide, with its simple expressions (132) for the eigenvalues of { kx, ky}. Indeed, these expressions describe a rectangular mesh on the [ kx, ky] plane, so that each point corresponds to the plane area  Ak = (/ a)(/ b), and the number of modes in a large k-plane area Ak >>

Ak is N = Ak/ Ak = abAk/2 = AAk/2, where A is the waveguide’s cross-section area.59 However, it is frequently more convenient to discuss transverse wave vectors k t of arbitrary direction, i.e. with an arbitrary sign of their components kx and ky. Taking into account that the opposite values of each component actually give the same wave, the actual number of different modes of each type ( E- or H-) is a factor of 22  4 lower than was calculated above. This means that the number of modes of both types is A A

N

2

k

.

(7.147)

2

(2 )

Let me leave it for the reader to find hand-waving (but convincing :-) arguments that this mode counting rule is valid for waveguides with cross-sections of any shape, and any boundary conditions on the walls, provided that N >> 1.

7.7. Dielectric waveguides, optical fibers, and paraxial beams

Now let us discuss electromagnetic wave propagation in dielectric waveguides. The simplest, step-index waveguide (see Figs. 23b and 25) consists of an inner core and an outer shell (in the optical fiber technology lingo, called cladding) with a higher wave propagation speed, i.e. a lower index of refraction:

v v ,

i.e. n n , k k ,      .

(7.148)

at the same frequency. (In most cases the difference is achieved due to that in the electric permittivity, +

< -, while magnetically both materials are virtually passive: -  +  0, so that their refraction indices n, defined by Eq. (84), are very close to (/0)1/2; I will limit my discussion to this approximation.) The basic idea of the waveguide’s operation may be readily understood in the limit when the wavelength  is much smaller than the characteristic size R of the core’s cross-section. In this

“geometric-optics” limit, at distances of the order of  from the core-cladding interface, which provides the wave reflection, we can neglect the interface’s curvature and approximate its geometry with a plane.

As we know from Sec. 4, if the angle  of the wave’s incidence on such a plane interface is larger than the critical value c specified by Eq. (85), the wave is totally reflected. As a result, the waves launched into the fiber core at such “grazing” angles, propagate inside the core, being repeatedly reflected from the cladding – see Fig. 25.

59 This formula ignores the fact that, according to the above analysis, some modes (with n = 0 and m = 0 for the H

modes, and n = 0 or m = 0 for the E modes) are forbidden. However, for N >> 1, the associated corrections of Eq.

(147) are negligible.

Chapter 7

Page 43 of 68

Image 1860

Image 1861

Image 1862

Image 1863

Image 1864

Image 1865

Image 1866

Image 1867

Image 1868

Image 1869

Image 1870

Image 1871

Image 1872

Image 1873

Image 1874

Image 1875

Image 1876

Image 1877

Image 1878

Image 1879

Essential Graduate Physics

EM: Classical Electrodynamics

“cladding”  , 

“core”  , 

Fig. 7.25. Wave propagation in

a thick optical fiber at  > c.

The most important type of dielectric waveguides is optical fibers.60 Due to a heroic technological effort over three decades starting from the mid-1960s, the attenuation of such fibers has been decreased from values of the order of 20 db/km (typical for a window glass) to the fantastically low values of ~0.2 db/km (meaning virtually perfect transparency of 10-km-long fiber segments!), combined with the extremely low plane-wave (“chromatic”) dispersion below 10 ps/kmnm.61 In conjunction with the development of inexpensive erbium-based quantum amplifiers, this breakthrough has enabled inter-city and inter-continental (undersea), broadband62 optical cables, which are the backbone of all modern telecommunication infrastructure.

The only bad news is that these breakthroughs were achieved for just one kind of materials (silica-based glasses)63 within a very narrow range of their chemical composition. As a result, the dielectric constants   /0 of the cladding and core of practical optical fibers are both close to 2.2 ( n

 1.5) and hence very close to each other, so that the relative difference of the refraction indices, 1/ 2

1/ 2

n n

 

  

 

,

(7.149)

1/ 2

n

2

is typically below 0.5%. This factor limits the fiber bandwidth. Indeed, let us use the geometric-optics picture to calculate the number of quasi-plane-wave modes that may propagate in the fiber. For the complementary angle (Fig. 25)

 

 ,

sin

that

so

  cos ,

(7.150)

2

Eq. (85) gives the following propagation condition:

n

cos    1  .

(7.151)

n

60 For a comprehensive discussion of this vital technology see, e.g., A. Yariv and P. Yeh, Photonics, 6th ed., Oxford U. Press, 2007.

61 Both these parameters have their best values not in the visible light range (with wavelengths from 380 to 740

nm), but in the near-infrared, with the attenuation lowest between approximately 1,500 and 1,630 nm. As a result, most modern communication systems use two spectral windows – the so-called C-band (1,530-1,565 nm) and L-band (1,570-1,610 nm) within that range.

62 Each of the spectral bands mentioned above, at a typical signal-to-noise ratio S/ N > 105, corresponds to the Shannon bandwidth  f log2( S/ N) exceeding 1014 bits per second, some five orders of magnitude (!) higher than that of a modern Ethernet cable. The practically usable bandwidth of each fiber is somewhat lower, but a typical optical cable, with many fibers in parallel, has a proportionately higher aggregate bandwidth. A relatively recent (circa 2017) example is the C-band transatlantic (6,600-km-long) cable Marea, with eight fiber pairs and an aggregate useable bandwidth of 160 terabits per second.

63 The silica-based fibers were developed in 1966 by an industrial research group led by Charles Kao (who shared the 2009 Nobel Prize in physics), but the very idea of using optical fibers for long-range communications may be traced back at least to the 1963 work by Jun-ichi Nishizawa – who also invented semiconductor lasers.

Chapter 7

Page 44 of 68

Essential Graduate Physics

EM: Classical Electrodynamics

In the limit  << 1, when the incidence angles  > c of all propagating waves are very close to /2, and hence the complementary angles are  small, we may keep only two first terms in the Taylor expansion of the left-hand side of Eq. (151) and get

2

 2 .

(7.152)

max

(Even for the higher-end value  = 0.005, this critical angle is only ~0.1 radian, i.e. is close to 5.) Due to this smallness, we may approximate the maximum transverse component of the wave vector as ( k )

k (sin)

k

 2 

k ,

(7.153)

t max

max

max

and use Eq. (147) to calculate the number N of propagating modes:

 2

R

  2 2

k

 max 

N  2

 kR2 .

(7.154)

2 

2

For typical values k = 0.73107 m-1 (corresponding to the free-space wavelength 0 = n = 2 n/ k  1.3

m), R = 25 m, and  = 0.005, this formula gives N  150.

Now we can calculate the geometric dispersion of such a fiber, i.e. the difference in the mode propagation speed, which is commonly characterized in terms of the difference between the wave delay times (traditionally measured in picoseconds per kilometer) of the fastest and slowest modes. Within the geometric optics approximation, the difference in time delays of the fastest mode (with kz = k) and the slowest mode (with kz = k sinc) at distance l is

l

k lz

l

l

l

n

l

t  

 

 

k  1 sin

1

.

(7.155)

z

 

c  

   









v

z

v

v

n

v

 

For the example considered above, the TEM wave’s speed in the glass, v = c/ n  2108 m/s, and the geometric dispersion  t/ l is close to 25 ps/m, i.e. 25,000 ps/km. (This means, for example, that a 1-ns pulse, being distributed between the modes, would spread to a ~25-ns pulse after passing a just 1-km fiber segment.) This result should be compared with the chromatic dispersion mentioned above, below 10 ps/kmnm, which gives dt/ l is of the order of only 1,000 ps/km in the whole communication band d

~ 100 nm. Due to this high geometric dispersion, such relatively thick (2 R ~ 50 nm) multi-mode fibers are used for the transfer of signals over only short distances below ~ 100 m. (As compensation, they may carry relatively large power, beyond 10 mW, without being damaged by the field.)

Long-range telecommunications are based on single-mode fibers, with thin cores (typically with diameters 2 R ~ 5 m, i. e. of the order of /1/2). For such structures, Eq. (154) yields N ~ 1, but in this case, the geometric optics approximation is not quantitatively valid, and for the fiber analysis, we should get back to the Maxwell equations. In particular, this analysis should take into explicit account the evanescent wave in the cladding, because its penetration depth may be comparable with R.64

64 The following quantitative analysis of the single-mode fibers is very valuable – both for practice and as a very good example of Maxwell equations’ solution. However, I have to confess that its results will not be used in the following parts of the course. So, if the reader is not interested in this topic, they may safely jump to the text following Eq. (181). (I believe that the discussion of the angular momentum of electromagnetic radiation, starting at that point, is compulsory for every professional physicist.)

Chapter 7

Page 45 of 68

Essential Graduate Physics

EM: Classical Electrodynamics

Since the cross-section of an optical fiber lacks metallic walls, the Maxwell equations describing them cannot be exactly satisfied with either TEM-wave, or H-mode, or E-mode solutions. Instead, the fibers can carry the so-called HE and EH modes, with both vectors H and E having longitudinal components simultaneously. In such modes, both Ez and Hz inside the core (  R) have a form similar to Eq. (141):

f f J k  cos n   

2

,

where k  2

k  2

k

2

,

0

and k   2  ,

(7.156)

l

n t

0 

t

z

where the constant angles 0 may be different for each field. On the other hand, for the evanescent wave in the cladding, we may rewrite Eqs. (101) as

2   2

.

(7.157)

t f

 ,

0

 2

where

 2

k  2

k

2

,

0

and k   2 

t

z

Figure 26 illustrates these relations between kt,  t, kz, and k; note that the following sum, Universal

relation

2

k  2

  2

 (   ) 

2

2 k  ,

(7.158) between

t

t

0

kt and  t

is fixed (at a given frequency) and, for typical fibers, is very small (<< k 2). In particular, Fig. 26 shows that neither kt nor  t can be larger than  [(- – +)0]1/2 = (2)1/2 k. This means that the depth  = 1/ t of the wave penetration into the cladding is at least 1/ k(2)1/2 = /2(2)1/2 >> /2 . This is why the cladding layers in practical optical fibers are made as thick as ~50 m, so that only a negligibly small tail of this evanescent wave field reaches their outer surfaces.

2

2

2

k k   (   )

0

2

2

k

t

t

Fig. 7.26. The relation between the transverse

exponents k

t and  t for waves in optical fibers.

2

2

k

k

2

2

k

k

z

In the polar coordinates, Eq. (157) becomes

 1    

1

2

2





,

(7.159)

t f

 0

2

2

       

- the equation to be compared with Eq. (139) for the circular metallic-wall waveguide. From Sec. 2.7, we know that the eigenfunctions of Eq. (159) are the products of the sine and cosine functions of n by a linear combination of the modified Bessel functions In and Kn, shown in Fig. 2.22, now of the argument  t. The fields have to vanish at   , so that only the latter functions (of the second kind) can participate in the solution:

f K ( ) cos n

(7.160)

n

t

 0 

Now we have to reconcile Eqs. (156) and (160), using the boundary conditions at  = R for both longitudinal and transverse components of both fields, with the latter components first calculated using Eqs. (121). Such a conceptually simple, but a bit bulky calculation (which I am leaving for the reader’s exercise), yields a system of two linear, homogeneous equations for the complex amplitudes El and Hl, which are compatible if

Chapter 7

Page 46 of 68

Essential Graduate Physics

EM: Classical Electrodynamics

 2

2

k J '

k K '

J '

K '

n

k

k

n

n

1 n

1

2  2

2

n

1

1 

 









,

(7.161)

2 

2

2



 2

2 

k J

K

k J

K

R

k

k

t

n

t

n

t n

t

n

t

t

t

t

where the prime signs denote (as a rare exception in this series) the derivatives of each function over its full argument: kt for Jn, and  t for Kn.

For any given frequency , the system of equations (158) and (161) determines the values of kt and  t, and hence kz. Actually, for any n > 0, this system provides two different solutions: one corresponding to the so-called HE wave, with a larger ratio Ez/ Hz, and the EH wave, with a smaller value of that ratio. For angular-symmetric modes with n = 0 (for whom we might naively expect the lowest cutoff frequency), the equations may be satisfied by the fields having just one non-zero longitudinal component (either Ez or Hz), so that the HE wave are the usual E-modes, while the EH

modes are the H-waves. For the H-modes, the characteristic equation is reduced to the requirement that the expression in the second parentheses on the left-hand side of Eq. (161) is equal to zero. Using the Bessel function identities J 0 = – J 1 and K 0 = – K 1, this equation may be rewritten in a simpler form: 1 J ( k R)

K R

1

t

1

(

)

1

t

 

.

(7.162)

k J ( k R)

K ( R)

t

0

t

t

0

t

Using the universal relation between kt and  t given by Eq. (158), we may plot both sides of Eq.

(162) as functions of the same argument, say,   ktR – see Fig. 27.

3

LHS

01

11

02

12

03

Fig. 7.27. Two sides of the characteristic

0

equation (162), plotted as functions of ktR,

RHS

for two values of its dimensionless

parameter: V = 8 (blue line) and V = 3 (red

line). Note that according to Eq. (158), the

argument of the functions K 0 and K 1 is

 3

2

2

0

5

10

tR = [V – ( ktR)2]1/2  (V – 2)1/2.

  k R

t

The right-hand side of Eq. (162) depends not only on , but also on the dimensionless parameter V defined as the normalized right-hand side of Eq. (158):

2

2

2

2

2

V

  (   ) R  2 k R .

(7.163)

0

(According to Eq. (154), if V >> 1, it gives twice the number N of the fiber modes – the conclusion confirmed by Fig. 27, taking into account that it describes only the H-modes.) Since the ratio K 1/ K 0 is positive for all values of the functions’ argument (see, e.g., the right panel of Fig. 2.22), the right-hand side of Eq. (162) is always negative, so that the equation may have solutions only in the intervals where the ratio J1/ J 0 is negative, i.e. at

  k R   ,   k R   ,... ,

(7.164)

01

t

11

02

t

12

Chapter 7

Page 47 of 68

Essential Graduate Physics

EM: Classical Electrodynamics

where  nm is the m-th zero of the function Jn() – see Table 2.1. The right-hand side of the characteristic equation (162) diverges at  tR  0, i.e. at ktR  V, so that no solutions are possible if V is below the critical value Vc = 01  2.405. At this cutoff point, Eq. (163) yields k  01/ R(2)1/2. Hence, the cutoff frequency of the lowest H mode corresponds to the TEM wavelength

2 R

2

 7

.

3

R

.

(7.165)

max

 1/2

1/ 2

01

For typical parameters  = 0.005 and R = 2.5 m, this result yields max ~ 0.65 m, corresponding to the free-space wavelength 0 ~ 1 m. A similar analysis of the first parentheses on the left-hand side of Eq.

(161) shows that at   0, the cutoff frequency for the E modes is similar.

This situation may look exactly like that in metallic-wall waveguides, with no waves possible at frequencies below c, but this is not so. The basic reason for the difference is that in the metallic waveguides, the approach to c results in the divergence of the longitudinal wavelength  z  2/ kz. On the other hand, in dielectric waveguides, the approach leaves  z finite ( kzk+). Due to this difference, a certain linear superposition of HE and EH waves with n = 1 can propagate at frequencies well below the cutoff frequency for n = 0, which we have just calculated.65 This mode, in the limit +  - (i.e.  << 1) allows a very interesting and simple description using the Cartesian (rather than polar) components of the fields, but still expressed as functions of the polar coordinates  and . The reason is that this mode is very close to a linearly polarized TEM wave. (Due to this reason, this mode is referred to as LP 01.) Let us select the x-axis parallel to the transverse component of the magnetic field vector at  = 0, so that Ex=0 = 0, but Ey=0  0, and Hx=0 0, but Hy=0 = 0. The only suitable solutions of the 2D

Helmholtz equation (that should be obeyed not only by the z-components of the fields but also their x-

and y-components) are proportional to J 0( kt), with zero coefficients for Ex and Hy: E  ,

0

E E J ( k ), H H J ( k ), H  , 0

for   R .

(7.166)

x

y

0

0

t

x

0

0

t

y

LP 01 mode

Now we can use the last two equations of Eqs. (100) to calculate the longitudinal components of the fields:

1

E

y

k

H

k

t

1

E

i

E J ( k )sin,

x

t

H

i

H J ( k ) cos ,

(7.167)

z

0 1

t

z

0 1

t

ik

y

k

ik

x

k

z

z

z

z

where I have used the following mathematical identities: J 0 = – J 1 , / x = x/ = cos, and / y = y/

= sin. As a sanity check, we see that the longitudinal component or each field is a (legitimate!) eigenfunction of the type (141), with n = 1. Note also that if kt << kz (this relation is always true if  << 1 – see either Eq. (158) or Fig. 26), the longitudinal components of the fields are much smaller than their transverse counterparts, so that the wave is indeed very close to the TEM one. Because of that, the ratio of the electric and magnetic field amplitudes is also close to that in the TEM wave: E 0/ H 0  Z Z+.

Now to satisfy the boundary conditions at the core-to-cladding interface ( = R), we need to have a similar angular dependence of these components at   R. The longitudinal components of the fields 65 This fact becomes less surprising if we recall that in the circular metallic waveguide, discussed in Sec. 6, the fundamental mode ( H 11, see Fig. 23) also corresponded to n = 1 rather than n = 0.

Chapter 7

Page 48 of 68

Essential Graduate Physics

EM: Classical Electrodynamics

are tangential to the interface and thus should be continuous. Using the solutions similar to Eq. (160) with n = 1, we get

k J ( k R)

k J ( k R)

t

1

E   i

t

E K ( ) sin ,

t

1

H i

t

H K ( ) cos , for   .

R (7.168)

z

k K ( R) 0 1 t

z

k K ( R)

0

1

t

z

1

t

z

1

t

For the transverse components, we should require the continuity of the normal magnetic field  Hn, for our simple field structure equal to just  Hx cos, of the tangential electric field E = Ey sin, and of the normal component of D

66

n =  En =  Ey cos. Assuming that - = + = 0, and +  -– we can satisfy these conditions with the following solutions:

J ( k R)

J ( k R)

E  ,

0

0

E

t

E K ( ),

0

H

t

H K ( ), H  ,

0

for   .

R (7.169)

x

y

K ( R) 0 0 t

x

K ( k R) 0 0 t

y

0

t

0

t

From here, we can calculate components from Ez and Hz, using the same approach as for   R: 1

E

y

J ( k R)

t

0

E

i

t

E K ( )sin,

z

ik

y

k K ( R) 0 1 t

z

z

0

t

(7.170)

1

H

J ( k R)

x

t

0

H

  i

t

H K ( ) cos, for   .

R

z

ik

x

k K ( R) 0 1 t

z

z

0

t

These relations provide the same functional dependence of the fields as Eqs. (167), i.e. the internal and external fields are compatible, but their amplitudes at the interface coincide only if

LP 01 mode:

J ( k R)

K ( R)

characteristic

1

t

1

k

t

 

.

(7.171)

t

equation

J ( k R)

t K ( R)

0

t

0

t

This characteristic equation (which may be also derived from Eq. (161) with n = 1 in the limit 

 0) looks close to Eq. (162), but functionally is much different from it – see Fig. 28. Indeed, its right-hand side is always positive, and the left-hand side tends to zero at ktR  0. As a result, Eq. (171) may have a solution for arbitrary small values of the parameter V defined by Eq. (163), i.e. for arbitrary low frequencies (large wavelengths). This is why this mode is used in practical single-mode fibers: there are no other modes with wavelengths larger than the max given by Eq. (165), so that they cannot be unintentionally excited on small inhomogeneities of the fiber.

It is easy to use the Bessel function approximations by the first terms of the Taylor expansions (2.132) and (2.157) to show that in the limit V  0,  tR tends to zero much faster than ktR  V:  tR

2exp{-1/ V } << V. This means that the scale  c  1/ t of the radial distribution of the LP 01 wave’s fields in the cladding becomes very large. In this limit, this mode may be interpreted as a virtually TEM wave propagating in the cladding, just slightly deformed (and guided) by the fiber’s core. The drawback of this feature is that it requires very thick cladding, to avoid energy losses in its outer (“buffer” and

“jacket”) layers that defend the silica layers from the elements and mechanical damages, but lack their 66 This is the core assumption of this approximate theory, which accounts only for the most important effect of the small difference of the dielectric constants 

2

2

2

2

2

+ and –: the difference between ( k+ – k z ) = kt > 0 and ( k- – k z ) = –

 2 t < 0. For more discussion of the accuracy of this approximation and some exact results, the interested reader may be referred either to the monograph by A. Snyder and D. Love, Optical Waveguide Theory, Chapman and Hill, 1983, or to Chapter 3 and Appendix B in the monograph by Yariv and Yeh, that was cited above.

Chapter 7

Page 49 of 68

Essential Graduate Physics

EM: Classical Electrodynamics

low optical absorption. Due to this reason, the core radius is usually selected so that the parameter V is just slightly less than the critical value Vc = 01  2.4 for higher modes, thus ensuring the single-mode operation.

10

RHS

LHS

Fig. 7.28. Two sides of the

0

characteristic equation (171) for the

LP

01 mode, plotted as a function of

ktR, for two values of the

dimensionless parameter: V = 8

(blue line) and V = 1 (red line).

 100

5

10

  k R

t

In order to reduce the field spread into the cladding, the step-index fibers discussed above may be replaced with graded-index fibers whose dielectric constant  is gradually and slowly decreased from the center to the periphery.67 Keeping only the main two terms in the Taylor expansion of the function

() at  = 0, we may approximate such reduction as68

 ()   ( )

0 

2

1  ,

(7.172)

where   –[( d 2/ d2)/2]=0 is a positive constant characterizing the fiber composition gradient.

Moreover, if this constant is sufficiently small ( << k 2), the field distribution across the fiber’s cross-section may be described by the same 2D Helmholtz equation (101), but with a space-dependent transverse wave vector:69

 2

2

  k

t

t

 f  ,0

where

(7.173)

2

2

2

2

2

k ()  k ()  k k (0)  k

.

t

z

t

0 2

2



,

and k 0

2

   00

Surprisingly for such an axially-symmetric problem, because of its special dependence on the radius, this equation may be most readily solved in the Cartesian coordinates. Indeed, rewriting it as 2

2

 

2

k (0)

2

k

,

(7.174)

t

0  2

2

x y f  0

2

2

 x

y

and separating the variables as f = X( x) Y( y), we get 67 Due to the difficulty of fabrication of graded-index fibers with wave attenuation below a few dm/km, they are not used as broadly as the step-index ones.

68 For an axially-symmetric smooth function (), the first derivative d/ d always vanishes at  = 0, so that Eq.

(172) does not have a term linear in .

69 This approach is invalid at arbitrary (large)  because in the macroscopic Maxwell equations, (r) is under the differentiation sign, and the exact Helmholtz-type equations for fields have additional terms containing .

Chapter 7

Page 50 of 68

Essential Graduate Physics

EM: Classical Electrodynamics

1

2

d X

1 2

d Y

2

k ( )

0

2

k

,

(7.175)

t

0 x y

2

2

 2 2 0

X dx

Y dy

so that the functions X and Y obey similar differential equations, for example 2

d X

k k

x X

(7.176)

2

 2 2

x

0 2

,

0

dx

with the separation constants satisfying the following condition:

2

2

2

2

k k k (0)  k

k .

(7.177)

x

y

t

0 2 z

The ordinary differential equation (176) is well known from quantum mechanics, because the stationary Schrödinger equation for one of the most important basic quantum systems, a 1D harmonic oscillator, may be rewritten in this form. Its eigenvalues are very simple:

( 2

k )  k

,

(7.178)

x

n

0 1/22 n  1,

( 2

k )  k

y m

0 1/22 m  1,

with n, m  ,

0 ,

1 ,...

2

but the corresponding eigenfunctions Xn( x) and Ym( y) are expressed via not quite elementary functions –

the Hermite polynomials.70 For most practical purposes, however, the lowest eigenfunctions X 0( x) and Y 0( y) are sufficient, because they correspond to the lowest kx,y, and hence the lowest

 2

k

,

(7.179)

t 0

2

2

( k )

( k )

2 k 0 

min

x 0

y 0

  1/2

and the lowest cutoff frequency. As may be readily verified by the substitution to Eq. (176), the eigenfunctions corresponding to this fundamental mode are also simple:

k0 1/2 2

x

X ( x)  const  exp

,

(7.180)

0



2

and similarly for Y 0( y), so that the field distribution follows the Gaussian function

k0 1/2 2

2

  

  

1/ 2

f ()  f (0) exp

  f (0)exp

,

 with a  1/ k

0 

,

(7.181)

0

0

0

  1/4

2

2

 2 a

where a >> 1/ k(0) has the sense of the effective width of the field’s extension in the radial direction, normal to the wave’s propagation axis z. This is the so-called Gaussian beam, very convenient for some applications.

The Gaussian beam (181) is just one example of the so-called paraxial beams, which may be represented as a result of modulation of a plane wave with a wave number k, by an axially-symmetric envelope function f(), where   { x, y}, with a relatively large effective radius a >> 1/ k.71 Such beams give me a convenient opportunity to deliver on the promise made in Sec. 1: calculate the angular momentum L of a circularly polarized wave propagating in free space, and prove its fundamental relation to the wave’s energy U. Let us start from the calculation of U for a paraxial beam (with an 70 See, e.g., QM Sec. 2.9.

71 Note that propagating in a uniform medium, i.e. outside of grade-index fibers or other focusing systems, such beams gradually increase their width a due to diffraction – the effect to be analyzed in the next chapter.

Chapter 7

Page 51 of 68

Essential Graduate Physics

EM: Classical Electrodynamics

arbitrary, but spatially-localized envelope f) of a circularly polarized wave, with the transverse electric field components given by Eq. (19):

E E f

E   E f

(7.182a)

x

0

 cos ,

y

0

 sin ,

where E 0 is the real amplitude of the wave’s electric field at the propagation axis,   kz –  t +  is its total phase, and the two signs correspond to two possible directions of the circular polarization.72

According to Eq. (6), the corresponding transverse components of the magnetic field are

E

E

0

H  

f

.

(7.182b)

x

sin ,

0

H

f

y

 

cos

Z

Z

0

0

These expressions are sufficient to calculate the energy density (6.113) of the wave,73

E E

H H

0 

2

2

x

y

0 

2

2

x

2

2

2

2

E f

E f