Sei sulla pagina 1di 192

O.

Kusch

Computer-Aided
Optical Design
of Illuminating and Irradiating Devices

ASLAN Publishing House


Moscow 1993

2
O.K. Kusch
Computer-Aided Optical Design of Illuminating and Irradiating Devices

ISBN 5-87793-001-X

O.K. Kusch, 1993


translated by V.N. Stepanov
-
125167, , 60-67
. (095) 151-52-28
_________________________________________________
05.12.1993. 6088 1/16.
. . .. 11,00. ...12,00.
1500 . - .
. 062738

142110, ., ., 25

3
CONTENS
INTRODUCTION

CHAPTER ONE
GEOMETRICAL OPTICS OF REFLECTORS AND REFRACTORS
1.1. Elements of Linear and Matrix Algebra
1.2. Surface Theory Elements

9
15

1.3. Laws of Reflection and Refraction. (Snells Laws)


1.4. Ray and Surface Intersection

21
27

1.5. Ray Tracing for Discrete-Point Representation


of a Surface

32

1.6. Surface Synthesis by Using Fermats Principle


1.7. Matrix Application in Geometric and Optical
Calculations
References

35
42
50

CHAPTER TWO
METHODS FOR CALCULATING LIGHT DISTRIBUTION
2.1. Light-Ray Methods

51

2.2. Light Distribution Created by Ellipsoid or Hyperboloid


With a Point Source

58

2.3. Inverse-Ray Methods


2.4. Inverse-Ray Method in Analytical Representation

64
83

2.5. Calculation of Illuminance by Wieners Scheme


2.6. The Method of Elementary Maps

99
107

2.7. Probabilistic Simulation in Design of Lighting and


Optical Fixtures

112

4
2.8. Calculation of Light Distribution from Lambertian
Sources

119

2.9. Models of Real Sources and Materials


2.10. Graphical Representation of Luminous Field

129

Produced by Lighting Fixture


2.11. Calculating Light Source Indicatrix for Specified

135

Luminous Intensity
References

138
142

CHAPTER THREE
INVERSE PROBLEM IN OPTICAL SYSTEM DESIGN
3.1. Point-Source Methods and Algorithms

145

3.2. Methods and Algorithms of Solving Inverse Problem


for Lengthy Sources

154

3.3. Calculating Schemes for Reflector Optimization


3.4. Optimization of Focal Parameter of Reflector
3.5. Calculation of Milling-Cutter Trajectory
References

166
168
170
172

ANNEXES
Annex 1
Annex 2

173
177

Annex 3
Annex 4

182
188

5
To the memory of
V.D. Komissarov and N.G. Boldyrev
remarkable Russian mathematicians
and illuminating engineers.

INTRODUCTION

The computers, presently, have germinated new opportunities in design of


illuminating and irradiating devices. These clever machines led to
improvement of old and stimulated the birth of new methods in design of
optical systems for illumination and irradiation. Renewal of old undeservingly
forgotten methods is not a tribute to dull tradition, but it is governed by
intention to find the most efficient practice adequate to rapid development and
broad application of lighting devices with various light sources. It should be
noted that Russian school has deep roots in this field. V.N.Chikolev established
first bricks at the theory of design of lighting fixtures on the eve of the century.
Fundamental works were published in 20-40s by V.A.Fock, N.G.Boldyrev,
A.A.Gershun, V.D.Komissarov, and in 50-70s by N.N.Ermolinskii,
N.A.Karyakin, V.V.Trcmbach. It is of great importance to follow the ideas that
could have made a link between different methods that exist today. To authors
mind, the principles of geometrical optics and light field theory* can serve this
uniting base.
Of course, geometrical and optical principles of design and calculation
should be set forth in the modern language of vector-and-matrix algebra.
Fermats variational principle is one of the basic statements in geometrical
optics. It enables to synthesize optical systems most easily and elegantly.
The point source is also one of the main concepts in geometrical optics.
Its role is similar to that of the point mass in theoretical mechanics.
Distribution of point sources in space is interpreted in modern theory as a
distribution of Diracs -functions. The point source concept enables to
connect differential properties of an optical surface, Gaussian curvature, in
particular, with distribution of reflected intensity.
Another important concept in optical calculations is the ray since it terms
all photometric values that affect a detector being sensitive to quadratic
constituent of a light field: illuminance, spherical illuminance, luminous
*

The author is familiar with similar views of Dr. O.N.Stavroudis and Dr. S.Cornbleet on fundamental role of
geometrical optics in optical systems design.

6
intensity, etc. Fock and Wieners theorems arc fundamental in calculation of
intensity.
It seemed very important to classify the problems of optical design by
separating them in two groups: direct and inverse problems.
Determination of intensity* in luminous field produced by a luminaire at
the proximate zone (the illuminance) or at the remote zone (the luminous
intensity), when geometrical and luminance parameters of the optical system
arc known, constitute the direct problems (the analysis problems).
Following Baltes**, we can describe an optical system by a set of spaceand-luminance characteristics L {L1 , L2 ,..., Ln } , while a resulted intensity
distribution by a set E {E1 , E2 ,..., En } , we specify the direct problem in
terms of mapping F: LE. The inverse mapping F-1:EL, generally, is a
solution of inverse problem. In the theory of light-fixtures design an inverse
problem means the determination of mirror surface shape that provides a
prescribed intensity distribution and meets a chosen optimizing criteria. Inverse
problems usually lead to the necessity of solving integral and differential
equations.
There are two approaches to solving direct problems. The first deals with
computation of intensity within a ray tube related to a point source. The result
is considered as a superposition of intensities produced by these point sources.
The second approach is based upon calculation of ray luminance while tracing
the rays through the optical system and applying the laws of geometrical optics
and the most general equations of light-field theory. By tracing a multitude of
rays, the luminance at the points of exit pupil of a fixture, and consequently, in
accordance with Mangens law, a desirable intensity characteristic can be
determined.
Elements of linear and matrix algebra as well as foundations of surface
theory are given in the First Chapter. The laws of geometrical optics, the laws
of reflection and refraction, Fermats principle and its applications in synthesis
of optical surfaces arc to be found here.
The Second Chapter deals with the methods of calculating luminous
distribution of illuminating and irradiating dcviccs, i. c. with direct problems
and their solutions. The main equations of ray method based on the pointsource conception are cited. The ideas of inverse-ray method are described in
details. The latter being the most accurate method for calculating various
optical systems with sources of finite dimensions. Its analytical version enables
*

**

Under the term intensity we mean power characteristics of a field: illuminance, irradiance, or luminous
intensity.
Inverse Source Problems in Optics, lid. by 11.P. Unites with a Foreword by J.-F. Moser. 1984.

7
to construct effective algorithm and better understand concentrating properties
of specular reflectors.
The method of elementary maps, which is being used in Russian school of
illuminating engineering when dealing with Lambertian sources, has been also
incorporated. Examples on application of contour integration and light-field
methods developed by Boldyrev and Gershun are presented as well.
The most notable addition herein is the paragraph on direct statistical
simulation of lighting fixtures, i.e. on application of Monte Carlo methods.
The problem has been stated on restoration of local characteristics of
radiation in a light source through its integral output (the inverse problem).
This is actual, since it enables to define the input parameters more exactly, and
thus to increase the accuracy of direct problem solution.
The subject matter in the Third Chapter touches theoretical and practical
spectrum of inverse problem. Classical results for point and linear sources are
presented. The conditions of algorithm convergence within a domain being
free from critical points have been analyzed. The solutions for the sources with
finite dimensions arc given.
Annexes contain the texts of C-programs for solving various direct and
inverse problems (solution of nonlinear equation, spline interpolation,
integration of a system of differential equations).
The author hopes that this book will serve to invoke new ideas and unite
the optical design of lighting devices with the general methodology of modern
optics and illuminating engineering.
The author expresses deep gratitude to V.N. Stepanov who translated the
book and assisted in its publishing.
Author

9
CHAPTER ONE
GEOMETRICAL OPTICS OF REFLECTORS
AND REFRACTORS
1.1. ELEMENTS OF LINEAR AND MATRIX ALGEBRA
For unification wc shall use the matrix notation accepted in courses for
matrix algebra [1]. A vector is represented as a column [2]

x1
x
2
.
x
.
.

x n
The operation of transposition, which changes columns for rows and vice
versa, is denoted by upper index T. In that representation quantities xi are the
coordinates of vector x, and the latter will be written in a matrix form:

x x1 x 2 ...x n

Vector is usually expressed via basis: we choose a conventional set of


vectors e1, e2, e3,..., en and express all other vectors as
n

x x1e1 x 2 e2 ... x n en xi ei

(1.1)

i 1

Basis is a system of linearly independent vectors, hence,

a e

i i

0 that

i 1

leads to ai=0 for i 1, n . Basis vectors form a coordinate system, where


numbers are projections of vector on corresponding axes.

10
The scalar product of vectors x and y is
n

x y ( x, y ) xT y xi yi ,
i 1

where xi and yi, are the coordinates of vectors x and y. The length (absolute
value) of a vector is computed as
n

x x T x xi2 .
i 1

Usually, basis is formed by a set of mutually orthogonal unit vectors

ei , i 1, n , for which the following equalities are valid


T
ei e j ij ;

e j

(e ij ) 2 1 ,
i 1

where ij , is Kroneckers symbol, ij = 1 for i=j and ij = 0 for ij.


For example, in three-dimensional space the system of vectors

e1 [1 0 0]T , e2 [010]T , e3 [0 01]T forms the basis.


Really,

1
0
0 1
1 0 2 1 2 0 2 0 ,
0
0
1 3
hence 1=2=3=0
Matrix A describes linear mapping of vector x into vector y that can be
written as y = Ax. The linearity of transformation A means that the following
equalities hold true
A(x) Ax ,

A(x1 x2 ) Ax1 Ax2 .


Substituting decomposition of x from Eq. (I.I), we get
y Ax xi Aei .

Now, Aei , is a new vector, which can be decomposed in the form

Aei ji ei (decomposition by columns of matrix A).


i

11
By changing the order of adding in the sums, we get

y xi a ji ei a ji xi ei y ji ei ,
i
i
i i
i

hence,

yi a ji xi .

(1.2)

Thus, (he coordinates of vector y, unlike for the transformation of basis


vectors, are obtained by multiplying rows of matrix A by the column-vector [x1
x2xn]T.
Matrix A is defined as

a11
a
21
A

am1

a12
a 22

... a1n
... a 2n
.
.

a mn

am 2

Matrix A = {aij} has m rows and n columns.


Further, we introduce a unit matrix I, for which AI=IA=A, and an inverse
matrix, such that A-1 A=I.
The rules for operating with matrices follow from the properties of linear
operators. For example, the rule of multiplication of matrices can be obtained if
the rule for multiplication of operators is applied:
( AB) x A( Bx) .
The k-th element of vector Bx is

kj

x j , [see Eq. (1.2)], then the i-th

element of vector A(Bx) will be

a b
ik

kj

x j x j aik bkj .
j

Thus, it follows that the element of a new matrix C is a product of the i-th
row of matrix A by the j-th column of matrix B:
cij aik bkj
(1.3)

Let us now consider how vector coordinates and linear operator


components arc transformed if basis is changed. We have a basis { ei' }
expressed in terms of unprimed system by relation

12

ei' ji e j ,
j

where n of elements ji evolve the coefficients of transforming matrix


that carries out the transition from one coordinate system to another.
Let us consider an arbitrary vector having coordinates xi and xi' in two
systems, respectively. Then
x x 'e ' x ' e e x ' ;
i

ji

ji

hence,

x j ji xi' ,
i

and that is equivalent to matrix product

x x` .
-1

(1.4)

Multiplying Eq. (1.4) by inverse matrix , we obtain


x` 1 x .
(1.5)
Orthogonal matrix P can be defined from the condition
PT P PPT I ,
where I is the unit matrix. It is an obvious property of such matrices that P1
= PT, hence, Eq. (1.5) takes the form
x` T x .
(1.6)
Operators also change their form during transition to another basis. Let
y=A`x` and y=Ax in two different systems, then
1 y y` 1 Ax 1 Ax` .
Consequently, the rule for matrix transformation will take the form
A` 1 A .
(1.7)
Note, that during the transition to another basis, the determinant does not
change, i. e. det(A')=det(A).
We also emphasize the rule: if C = AB, then
C T ( AB)T BT AT .
(1.8)
Expression (1.8) directly happens out from Eq. (1.3). Let us now consider
few examples.

13
Example I. Matrix of Rotation
We choose on a plane a system
of
two
orthogonal
vectors
e1 [1 0]T , e2 [0 1]T and rotate
it by an angle anticlockwisely
(Fig. 1.1). In accordance with the
rule of constructing matrix of linear
transformation
we
find
the
transformed vectors of the basis:

Fig. 1.1. Rotation of unit square

cos
e1 Ae1

sin ,
sin
e2 A e2

cos

Expressions being found are the columns of matrix of rotation Mrot:


Mrot:
cos sin .
(1.9)
M rot

sin cos
Matrix of rotation Eq. (1.9) is orthogonal, because after rotation by angle (clockwisely), vector x returns to its initial position. Sign change from + to in Eq. (1.9) shifts orthogonal matrix into transposed one. Matrices of rotation
on the plane commutate, i. e.
M rot (1 ) M rot ( 2 ) M rot ( 2 ) M rot (1 ) .
Rotation is unitary, i. e.
det( M rot ) 1 .
For example, let us rotate a triangle
ABC by 90 about the origin of
coordinates (Fig. 1.2)
[3], i. e.
4 2
0 1 3
1 0 1 1 1

1 1 1 .

3 4 2

Fig.1.2. Two-dimensional rotation by 90

14
Example 2. Matrix of Reflection
We consider mirror reflection with
respect to bisectrix y = x. It is easy to
show for this case that e`1 [0 1]T ,
e`2 [1 0]T , and the law of reflection is
defined by the matrix

0 1
A
.
1 0
Let us find mirror image of triangle
ABC (Fig. 1.3):
0
1

1 8
0 1

7
3

6 1

2 8

3
7

2
6

Fig. 1.3. Two-dimensional reflection with


respect to y=x

Similar to matrix of rotation, matrix of reflection does not change metrical


properties of figures, i. e. det( A) 1 .
Generally speaking, when applying operator A to vector x, we get vector
Ax, which differs from x. Vector, for which Au = u is called an eigenvector u
with eigenvalue . In case of space rotation about axis OZ, any vector directed
along the axis of rotation will be an eigenvector with eigenvalue of +1.
Let us assume eigenvectors ei , as a basis for matrix A, that is,
Aei a ji e j
j

and
y xi Aei i xi ei ,
i

then

ji

e j e j .

Thus,
i, i j .
a ji
0, i j
Matrix A takes, so called diagonal form: numbers i, stand along its
diagonal, while the rest places are occupied by naughts.
From the previous example we define the matrix of reflection if we
consider a new basis, which is being rotated in relation to the initial basis by
the angle of 45:

15
2

2
2
2

2 .
2
2

Wc carry out multiplication of matrices in accordance with the law


of similarity [see Eq. (1. 7) ]:
2

A 2
2
2
1

2 1 0
2 0 1
2

2
2
2
2

2 1 0 .
2 0 1
2

Obviously, there arc two eigenvectors of reflection: the vector directed


along the straight line y = x, and the vector, which is orthogonal to this line.
1.2. ELEMENTS OF SURFACE THEORY
Parametrical representation is the most advisable way for description and
study of surfaces. Let there be a vector-function of two scalar variables within
Cartesian basis j , i, k :

r r (u, v ) x(u, v )i y (u, v ) j z (u, v )k .


A tangent vector to the curve r =
r(u,vo), where v0 is a constant, is a multiple
to vector ru r / u . Similarly, a tangent
vector to the curve r=r (u0, v) is a multiple
to vector rv r / v . A plane, which is
tangent to these curves (parametric
curves) contains both said vectors,
therefore, a normal to the surface at the
point uo,vo is a multiple to their vector
product (Fig. 1.4). A unit normal vector is
expressed as
n [ru * rv ] / ru * rv . (1.10)
Matrices of the first G and the second
D fundamental forms

Fig.1.4. Tangent vectors and coordinate


lines ru and rv

16
[4] play a substantial role in differential geometry of surfaces:

ru2 ru rv
G
,
2
rv ru rv
,
nruu nru v
G
,
nrv u n rvv

(1.11)

where

ruu 2 r / u 2 ; ruv 2 r / uv .
An area of surface clement is equal to the area of parallelogram
constructed on vcctors ruu and rvv , i. e.
S ru * rv uv .
We note, that
2

ru * rv ru rv (ru rv ) g11 g 22 g12 g 21 G ,


where gij are the elements of the first fundamental form and G det(G ) . An
area of surface specified by the domain R of variables u, v can be calculated as
follows

S G

1/ 2

du dv .

A curve on the surface r (u,v) can be specified when parameters u and v are
replaced by a pair of functions in terms of a new parameter, say t: u=u(t), v =
v(t) or by expression U = U(t), where U [u (t ) v (t )]T .
A tangent vector to this curve is
r ru u rv v AU ,
(1.12)
where

x
u
y
A
u
z
u

x
v
y
.
v
z
v

When the surface is intersected by a plane containing the main normal


n o and the tangent vector r AU Fig. 1.5), a curve being

17
produced is called the normal section. A
normal curvature kn of a surface is
defined as follows [4]

U T DU
kn T
. (1.13)
U GU
For any other curves having at a
given point the similar direction and the
normal no the theorem of Meusner is
valid (Fig. 1.5)

k cos k n const.
The directions, for which the normal
curvature is minimum or maximum, are
termed the principal directions. From Eq. Fig. 1.5. Tangent vector AU to curve C
(1.13), by differentiating with respect to at point M; normal section Co; curvature
radius R=Ro cos
u and v , we obtain

(d11 k n g11 )u (d12 k n g12 )v 0,


(d 21 k n g 21 )u (d 22 k n g 22 )v 0,
(1.14)
By eliminating u and v from the equations, we find

G k n2 ( g11 d 22 d11 g 22 2 g12 d12 )k n D 0 ,

(1.15)

whence, the minimum and maximum curvature values can be calculated. The
principal directions in curvature are available when kn is eliminated from Eq.
(1.14). We arrive to quadratic equation

(d11 g12 d12 g11 )u 2 (d11 g 22 d 22 g11 )uv


(d12 g 22 d 22 g12 )v 2 0 ,
(1.16)
whence, the ratio u : v for the principal directions is determined. From Eqs.
(1.12) and (1.13) it follows that k1k2 then the principal directions arc
orthogonal.
Calculation of Gaussian and mean curvature is of special interest:

1
K k1k 2 , H (k1 k 2 ) .
2
In agreement with the features of the roots of Eq. (1.15)

18

D
G

, 2 H g11d 22 d11 g 22 2 g12 d12 .

(1.17)

Example
Let a surface be formed by rotation of a curve C

x (u ), z (u )
lying in the plane XOZ, about the axis OZ (Fig. 1.6). Let v be an angle of
rotation. The expression for the surface of revolution is

r (u ) cos v i
(u ) sin v j (u )k
or

r (u )e(v) (u )k .

(1.18)

where e (v) cos vi sin vj is the unit vector. The lines of v=const are the
meridians in the surface, and the lines of u =const are the parallels. Now the
elements of matrices G and D can be found. Differentiating Eq. (1.18), we
have
ru `(u )e(v ) `(u )k;

ruu ``(u )e(v) ``(u )k;


rv (u )e(v / 2);
rv (u )e(v).
Whence,
g11 ru2 `2 `2 ;
g12 ru rv 0;
g 22 rv2 2 (u );
1/ 2

G ( g11 g 22 )1 / 2 ( `2 `2 )1 / 2 .
Thus, the meridians and parallels form orthogonal
grid on the surface of revolution (g12 = 0).
We find the normal vector and coefficients dij
Fig. 1.6. Curvalurc lines on
a surface of revolution

19

xu
xv

yu
yv

zu
zv

1/ 2

`k `e(v)
;
( `2 `2 )1 / 2

` `` ` ``
;
( `2 `2 )1 / 2
d12 ruv n 0;
d11 ruu n

2 1 / 2

d 22 `( ` ` )

(1.19)

Determining the principal directions from Eq. (1.14), we obtain


d11 d 22
uv 0
g11 g 22
hence, u o or v o , i. e. the principal directions coincide with meridians
and parallels on the surface of revolution. Assuming in Eq. (1.14) first v o
(the meridian direction), and then u o (the parallel direction), we have,
correspondingly
d
` `` ` ``
k1 11
,
g11 ( `2 `2 ) 3 / 2
(1.20)
d 22
`
k2

.
g 22 ( `2 `2 )1 / 2
We choose the curve x=x(z) (u=z) for meridian. From Eqs. (1.20) we have

k1

x``
,
(1 x`2 ) 3 / 2

k2

1
.
x(1 x`2 )1 / 2

From geometrical considerations it is obvious that the curvature of normal


section
is
equal
to
the
meridian
curvature.
Since
2 1/ 2
d 1 / k 2 x (1 x` ) x / cos , d is the length of normal segment
between the surface and the intersection point of normal with the axis (Fig.
1.6).
Let us calculate Gaussian and mean curvatures:
K

`( ` `` ` ``)
,
( `2 `2 ) 2

( ` `` ` ``) `( `2 `2 )
H 0.5
.
( `2 `2 ) 3 / 2

(1.21)

20
An important specific case should be noted, when a surface is formed by
rotation of a curve termed in polar coordinates R=R(u). Here (u)=R sin u,
(u)=R cos u. Further, we find

1/ 2

R sin u ( R 2 Ru2 )1/ 2 .

By definition of derivative

tan (u ) Ru / R ,
where is an angle between the normal and the axis.
Thus,

1/ 2

R 2 sin u / cos(u ) .

(1.22)

and the normal vector

n [sin cos v sin sin v cos ]T .

(1.23)
Using Eq. (1.20), the curvatures in meridian and sagittal sections can be
found

R 2 2 Ru2 RPuu
sin (u )
k1
, k2
.
2
2 3/ 2
( Ru R )
r sin u

(1.24)

The latter equation has an obvious geometrical meaning and can be derived
directly from Meusners theorem.
The divergence of a light beam which is reflected from a mirror segment,
as it will be shown below (see Ch. 2), depends on the principal curvatures at
the point of reflection. We find these values for paraboloid of revolution. A
parabola equation in polar coordinates can be termed as follows
r 2 f /(1 cos u ) f cos 2 (u / 2) ,
(1.25)
where f is the focal distance of parabola, and u is counted from the focal axis.
The principal meridian curvature can be conveniently calculated when
applying Eq. (1.24). Differentiating Eq. (1.25) and substituting the result into
Eq. (1.24), after transformations, we obtain
k1 (1 / 2 f ) cos3 (u / 2).
(1.26)
The curvature in the other principal section can be found easily by using
Meusners formula
k 2 (1 / 2 f ) cos(u / 2).
(1.27)

21
1.3. LAWS OF REFLECTION AND REFRACTION
(SNELLS LAWS)
The Law of Reflection
Let ao be the direction of incident ray, a be the direction of reflected ray,
and n be the normal to the surface at the point of incidence (Fig. 1.7 a). Since
ao , a , and n are the unit vectors, the law of reflection is represented as
follows:
(1) the vectors ao , a , n are coplanar, i. e. they lie in one plane;
(2) the angle of incidence is equal to the angle of reflection, i. e. the
following equation is true
ao n an.
(1.28)
In linear algebra terms the first statement means that ao , a , and n are
linear-dependent. Therefore, the scalar quantities , , exist such that they are
not equal to zero simultaneously; this leads to
a o a vn 0.
(1.29)
Since Eq. (1.28) is symmetrical, Eq. (1.29) can be represented in the form
ao a n ,
(1.30)
where is a new scalar constant.

Fig. 1.7. On deduction of laws of reflection (a) and refraction (b) in vector form

22
Multiplying Eq.(1.30) by n scalarly, and taking account of Eq. (1.28), we
get

2a o n .
Consequently, the law of reflection is defined as follows
a a o 2n (ao , n ) .
(1.31)
Equation (1.31) allows to solve the initial problem to find the direction
of the reflected ray when the incident ray and the normal at the mirrors point
are known. Here Eq. (1.31) is inversible. If we change a for ao equation
(1.31) remains unchanged. This allows to analyze a mirror fixture either in
direct or inverse ray-tracing mode (sec Ch. 2).
The vector operation in Eq. (1.31) can be represented in the matrix form
a ( I 2 E )a o ,
(1.32)
where

n x2
1 0 0

I 0 1 0; E n y n x
n z nx
0 0 1

nx n y
2
y

n
nz n y

nx nz

n y n z .
nz2

We put the law of reflection as follows

2n x2 1 2n x n y
2n x n z

[a x a y a z ]T 2n y n x 2n 2y 1 2n y n z * [aox aoy aoz ]T ,


2n z nx
2n z n y 2n z2 1

(1.33)
or, denoting the matrix as Mrefl,

a M refl a o .
The matrix Mrefl depends on basis choice. In the natural basis connected
with the point of ray incidence, where the axis OZ coincides with n , the
matrix Mrefl, as it can be easily checked, will take a simple diagonal form.
Really, it follows from Eq. (1.31) that ao = n is an eigenvector of the
linear transform Mrefl with the eigenvalue of -1, and any vector, which is
orthogonal to n , is an eigenvector with the eigenvalue of +1. Hence, the
matrix Mrefl takes the form

23

M refl

1 0 0
0 1 0 .
0 0 1

(1.35)

The main advantage of using matrices is the possibility of multiplying


them, so that computing automation for the ray path through several reflecting
surfaces can be realized [5]. For instance, for the double specular reflection the
vector coordinates are determined by matrix product
M refl ( I 2 E1 )( I 2 E 2 ) ,
(1.36)
where E1 and E2 are matrices related to the normal vectors of the first and the
second mirrors, correspondingly.
The Law of Refraction
According to the first law of rcfraction the vectors
coplanar (Fig. 1.7b), i.e. it is true that

ao , a , and n are

a a o n ,

(1.37)

where , are the constants to be determined.


The second law of refraction stales that q sin =sin , where q=v1/v2 is the
relative refractive index, and can be termed in vector form as follows
[a0 n ] 1 / q[a n ] .
(1.38)
Multiplying Eq. (1.37) by n vectorially, and taking account of Eq. (1.38),
we find that =q. We rewrite Eq. (1.37) as follows
a qa o n,
(1.39)
Since ao , a , and n are unit vectors, the square of the above equation
yields

2 2qp (q 2 1) 0,
where p= ao * n .
Its solution is

qp (qp ) 2 (q 2 1) .
The law of refraction is represented in the form

a qao {qao n 1 q 2 [1 (ao , n ) 2 ] }n ,

(1.40)

24

Fig. 1.8. Ray tracing through system of refracting surfaces:


a - incident and reflected rays; b - propagation of ray after refraction

The formulas in Eqs. (1.31) and (1.40) arc applicable for programming a
computer. Only the following has to be noted. The equation (1.40) docs not
hold true when q 1 / 1 (ao , n )

2 *

, this relates to the total internal

reflection. The upper sign before square root has been choscn in Eq. (1.40); it
can be proved when noting that for q= -1 Eq. (1.40) has to describe the law of
reflection [see Eq. (1.31) ].
Evidently, Eq. (1.40) is nonlinear, so it can not be presented in the matrix
form. But if we take bounded range of small angles , so that sin , the law
of refraction transforms into linear form, and use of matrices becomes possible.
A paraxial ray in the system with rotational symmetry is defined by two
parameters: the linear coordinate, that is the height where the ray intersects the
reference tangent plane; and the angle between the ray and the axis (Fig. 1.8
a). Let the ray coordinates hk-1 and k-1 corresponding to the (k-l)-th plane be
known. Within paraxial space Eq. (1.40) appears as follows
k k k 1 ,
(1.41)
where k vk 1 / vk is the relative rcfraction coefficient. Then the path of the
refracted ray is described as follows
k hk 1 k 1 (1 k ) k k 1 ,
(1.42)
where is the curvature of the refracting surface.
The refraction termed in matrix form

Since ( a o , n ) = cos , the critical angle is determined by sin cr =v2/v1. Consequently, the total internal
reflection takes place when a ray travels from the higher density medium to the lower density medium
(v1>v2).

25

1
0 hk 1
hk
1(1 ) .
k k
k
k k 1

(1.43)

After refraction the ray transfer from the (k-1)-th to the k-th surface can be
expressed by the following formula (see Fig. 1.8 b)
hk hk 1 k Tk ,
(1.44)
where Tk is the geometrical distance along the ray path. We can write Eq.
(1.40) in matrix form

hk 1 Tk hk 1
0 1 .
k
k 1

(1.45)

Now, when calculating the ray path for a set of surfaces, we consider the
process to be a sequence of matrix products [6].
Reflection and Refraction Problems
Problem I. Let the directions of incident and reflected rays be known. A
normal at the point of incidence has to be determined. The task arises when we
want to construct a profile of reflector knowing its angle function (ray-tracing
function).
Multiplying scalarly Eq. (1.31) by a , we find
1/ 2


a 0 , n 1 a0 , a
2

Solving Eq. (1.31) with respect to n ,

we obtain
1 / 2

n a0 a 2[1 a , a0 ]

. (1.46)
In case of refraction, solution of similar
problem can be obtained by using Eq.
(1.39) [see Fig. 1.9]. Unlike for reflection,
the solution here exists not for all a and
ao .
Problem 2. A ray is sequentially
reflected by two flat mirrors with the
normals n1 and n2, and a common bound m.
We have to find out how its direction
changes after two reflections.

Fig. 1.9 Determining normal vector to


refracting surface with the aid of
cyclotomic diagram kn na a 0 , where
k, n are constants.

26
After the first reflection a ray travels at the direction

a1 a k1n1
where k1 is a constant. Now a1 becomes an incident ray with respect to the
second mirror, and the second reflected ray is

b a k1n1 k 2 n2
where k2 is a constant.
Since bm=am, an angle between a ray and a bound remains unchangeable.
This holds true for arbitrary number of reflecting and refracting flat bounds.
Note, that similar result can be obtained by multiplying matrices from Eq.
(1.37).
Problem 3. A light ray traveling at the
direction I1 is being refracted when it strikes a
glass with refractive index n=l/q (Fig. 1.10).
Having passed the glass, it has to be partly
reflected by a mirror bound and to exit
through the refracting surface along the
direction I2 . The task is to find the orientation
of a mirror bound nr if the normal n1 of a front
surface is prescribed.
After refraction an incident ray turns to

J1 qI2 k1n1.
After reflection it obtains the direction

J2 qI2 k 2 n1.
For the normal nr, we obtain

Fig. 1.10. Determining normal vector to


mirror bound in optical wedge.

nr k 3 ( J1 J2 )
Parameters k1 and k2 are known, while constant k3 can be found from Eq.
1.46.

27
1.4. RAY AND SURFACE INTERSECTION
The task of determining the point of intersection between a ray and a
surface is included as a separate geometric block in the programs package for
computing light fixtures. The task has to be solved in two cases: (1) when
defining the point, where the ray is incident to a radiance-transforming surface
(refracting or reflecting), and finding the transformed ray; (2) when searching
for a point where the ray strikes a stretched light source (in the programs based
on inverse-ray tracing).
Without loss of generality, while considering these tasks, we may reduce
ourselves to observation of surfaces defined by equations of not higher than the
second order.
The second order surface is described by a quadratic form. By choosing the
coordinate system, the latter can be translated to the diagonal form [1], The
surface equation is as follows [7]
px x 2 p y y 2 pz z 2 1 ,
or in the matrix form

S T RS 1 ,

(1.47)

where s=[xyz]T, and R is a diagonal matrix

x
R
0

0
.

Let us define an equation for the ray originating from the point
s0 [ xo yo z o ]T at the direction g [ g x g y g z ]T .

s s0 gl ,

(1.48)

where l is the ray length counted from the point so.


Substituting Eq. (1.48) into Eq. (1.47), we get a quadratic equation in
unknown parameter l
Al 2 2 Bl C 0 .
(1.49)
The coefficients of equation are [8]
2
2
2
A g T Rg p x g x p y g y p z g z ;

B s0T Rg p x g x x0 p y g y y0 p z g z z0 ;
2

C s0T Rs 0 p x x0 p y y0 p z z 0 1.

(1.50)

28
We find a solution of quadratic equation, using the following formula

l ( B D ) / A ,
(1.51)
where D is the discriminant in Eq. (1.49); D=B2-AC.
Let the ray start at the point (x0, y0, z0). Then the coordinates of a point
where it intersects the surface arc specified by

x x0 lg x ,
y y0 lg y ,
z z 0 lg z .
Arithmetic condition for existence of intersection. In most cases this
checking is sufficient to continue program computations, e. g. when calculating
luminous intensity of a fixture.
The sign in Eq. (1.51) has to be chosen. We consider two cases: (1) an
outside intersection of a surface; (2) an inside intersection (Fig. 1.11). It is
clear that g T n 0 in the first case, and g T n 0 in the second case, where n is
the normal vector at the point of intersection. We get the normal vector by
differentiating Eq. (1.47)
n = Rs .
(1.53)
Substituting Eq. (1.48) into Eq. (1.53), wc obtain the scalar product
g T n g T R ( s0 gl ) g T Rs0 lg T Rg B lA.
(1.54)
Substituting now the expression for l from Eq. (1.51) into Eq. (1.54), we
have

gT n D .
We have to choose the positive sign near the square root in Eq. (1.51) for
inside intersection (see Fig. 1.11a), and the negative

Fig.1.11. Intersection of ray with surface: a - from outside, ( g , n ) 0 ;


b - from inside, ( g , n ) 0

29
sign in the opposite case (Fig. 1.11b).
We find now the coefficients A, B, C and the normal vector for a surface
of revolution having the axis OZ. Wc denote 0=x=y, 1=z. Then the
expressions arc reduced to
A 0 ( g x2 g 2y ) 1 g z2 ;
(1.55)
B 0 ( g x x0 g y y0 ) 1 g z z0 ;
C 0 ( x02 y02 ) 1 z02 1.
The value of pn in Eq. (1.55) has a meaning
of a square of curvature in the section
z 0, 0 1 / R02 , where Ro is the radius of
curvature.
We define conic sections in terms of
semiaxes: the minor b and the major a, while
0 1 / b 2 , 1 1 / a 2 (Fig. 1.12). The matrix R
has a following form
0
1

,
Rb 1

q 2
0
2

where q 2 (b / a ) 2 .
We take the parameters a and e (e is the
eccentricity)
as
basic:
2
2
2
2
2
q 1 e , b a (1 e ). The condition
e<1 represents an ellipsoid; e>1 and b2<0,
q2<0, a hyperboloid; e=1, a paraboloid. The
coefficients A, B, and C, now take the form

Fig. 1.12. Sections of conic surfaces:


(e<1 - ellips,
e=1 - parabola,
e>1 - hyperbola;
a, b are semiaxes;
p is focal parameter;
F1, F2 are foci)

A 1 (q 2 1) g z2 ;
B g x xo g y yo g z z o q 2 ;
C x02 y02 z 02 q 2 b 2 .
We find the normal at the point of intersection s=[xyz]T
0
1

n b 2 1

0
q 2

x
x
y b 2 y .


z
zq 2

(1.56)

30
The length of vector n

n b 2 ( x 2 y 2 z 2 q 4 )1/ 2 .
The unit normal vector

n [ x / M y / M qk z / M ]T ,

(1.57)

where M ( x 2 y 2 qk2 z 2 )1 / 2 ; bk b 2 ; q k q 2 .
We examine now some specific cases.
Cylinder
Let a=, b=Ro, q=0. The coefficients A, B, C take simple form

A 1 g z2 ;
B g x x0 g y yo ;

(1.58)

C x02 y02 R02 .


Obviously, if the point s0 [ xo yo z o ]T belongs to cylinder, then Eq. (1.49)
turns to linear which has a nontrivial solution
(1.59)
l 2 B / A.
The normal vector at the point of intersection
n [ x / R0 y / R0 0]T .
(1.60)
Sphere
Let b=Ro, q=1. From Eq.(1.56) we get

A 1;
B q x xo q y y 0 q z z 0 ;
C x02 y02 z 02 R02 .
The normal vector is directed along the radius towards the point s
n [ x / R0 y / R0 z / R0 ]T .

(1.61)

31
Cone
A cone may be considered as an
extreme of a hyperboloid under
a 0, b 0 , while k tan 1 / q .
Obviously, the focal length c=0
(Fig.1.13). Thus, we get

A 1 (qk 1)q z2 ;
B q x xo q y y0 q z z0 qk ; (1.62)
C x02 y02 z 02 qk .
The unit normal vector n [sec Eq.
(1.57)]

Fig. 1.13. Cone as limiting case of


hyperboloid (a0, b0)

n [ x / M 0 y / M 0 z / M 0 ]T .

(1.63)

Paraboloid
A paraboloid is a particular case of ellipsoid when the second focus tends
to infinity while focal parameter p=b2/a=const. Let e=1-, where is a small
given value, then 2c and b should be chosen as follows
2c 2ep / q 2 p / ;
(1.64)

b p / q p / 2 .

(1.65)

We relate deviation with the divergence of the rays originated from the
focus and reflected by the surface (quasi-paraboloid). The divergence of
reflected axial rays can be determined by differentiating the equation for conic
section termed in polar coordinates

p
,
1 e cos

that yeilds

da
2 sin

.
de 1 2e cos e 2
Having calculated the derivative at the point e=1, we get

da

tan ,
de
2

32
or

a tan

e.
2

(1.66)

Hence, if we take e=0,995, then for =90 we obtain the divergence of


reflected ray from the axis =0,3.
Finding the Point of Intersection with a Plane
In this case the surface is defined by linear form. The equation for a plane
being in parallel to the plane XOY and situated at the distance H from z = 0, is

s T k H .

(1.67)

Substituting Eq. (1.48) into Eq. (1.67), we have

s T k lg T k H ,
whence,

H soT k H z 0
.

gz
g T k

(1.68)

Substituting the expression for l into Eq. (1.52), we find the coordinates of
the intersection point.
1.5. RAY TRACING FOR DISCRETE-POINT
REPRESENTATION OF A SURFACE
Evidently, we may represent a surface of revolution in the following form

f ( x2 y 2 ) z 0 ,

(1.69)

or in vector terms

f (u ( s )) s T k 0 ,

(1.70)
2

where, as usual, s is the vector of a surface point and u x y .


Solving jointly Eqs. (1.49) and (1.70), we get an equation in l
T
f (u ( s0 gl )) ( s0 gl )k 0 .

(1.71)

33
Let a generatrix curve f (x) have a discrete-poinl assignment, i. e. it is
defined by arrays zi , xi , i 1, m . To solve Eq. (1.71), we must know
description of a function f (x) at an arbitrary interval [ xi , xi 1], i 1, m 1 ,.
Thus, at any interval [ xi , xi 1], i 1, m 1 the curve is substituted by its
representation f f where

f ( xi ) zi , f ( xi 1) zi 1, i 1, m 1 . (1.72)
Interpolation in terms of Lagranges polinomial Lm-1(x) led through m
points is a usual method for analytical representation of discrete function. But
when the power is growing the Lagranges polinomial may perform strong
oscillations between the nodes (Runges effect), that leads to unwanted results
in ray-tracc calculations. To avoid oscillations, piecewise smooth interpolation
by polinomials of low power (not higher than the third power) is applied.
Interpolation by segments of a parabola (Ermolinskiis method) and by
segments of a circle (Boldyrevs method) is widely used. The disadvantage of
these methods lies in unsmoothness of curves and their derivatives at the
connection points. Interpolation by cubic splines is efficient and widely used
method [9, 10] that provides a conjugation between curves and derivatives up
to the second order, i. e. to the coincidence of curvatures at the gluing points.
Thus, within an arbitrary interval [xi, xi+1] we have representation
f ( i ) q ( i ) ( x) ,
(1.73)
where
q (i ) ( x ) zi bi ( x xi ) ci ( x xi ) 2 d i ( x xi )3 ,
(1.74)
while splines q (i ) ( x) meet the condition in Eq. (1.72). Consequently, we have
a solution of the equation [see Eq. (1.71)]
T
q (i ) (u ( s0 ql )) ( s0 ql )k 0, i 1, m .

(1.75)
Obviously, the left side of it depends on the parameter l in a complex way,
that is why the equation has to be solved by the step-by-step approximation
method.
Annex 1 contains the computing program, which is a C-realization of
Forsythes ZEROIN program [11]. The subroutine-function, which presents an
array of parameters, can be included in the formal parameters list of a new
program version, in contrast to that of FORTRAN program. It is reached by
assigning the ZEROIN-type MATH_EXT.H structure and the description of
FUN_ZEROIN-type. VU-\ H

34
The function ZEROIN uses the formal parameter of ZEROIN-type, which
is a pointer onto a working function and a parameter vector being applied. In
this way, the explicit data transfer from the main function to the working one
bypassing the global level is provided.
Generally speaking, the cubic spline representation of a specular surface is
sufficient for calculating the light distribution. Further increase in profile-curve
smoothness, for example, when persistence of the third derivative is required,
docs not improve the smoothness in luminous intensity curves. Really, as it is
shown in Ch. 2, the normal illuminance in an infinite pcncil-bcam section at an
arbitrary distance from the mirror point is fully dependent on the normal vector
direction and the principal curvatures at the mentioned point, in other words on
the first and second derivatives.
The calculation of the normal at the intersection point is based on the
following considerations. We differentiate the surfacc equation (1.70) and
obtain

F F u

k 0.
s u s

(1.76)

Translating Eq. (1.76), we find the normal vector in the following form
T

x
x

n fu
fu
1 ,
u
u

where, as it was earlier, u x2 y2 , and we denote f u f / u . After


normalization of vector n, we have
T

fu
x
fu
y
1
n
. (1.77)
2 1/ 2
2 1/ 2
2 1/ 2
(1 f u ) u (1 f u ) u (1 f u )
The function f (x) is described by a set of splines, it is twice-differentiated
within the interval [x1,xm], therefore, for an arbitrary u there exists a
representation
f u q`i (u ) bi 2ci (u ui ) 3d i (u ui ) 2 .
(1.78)
Calculation of the first and second derivatives may be carried out by the
program SEVAL if we insert the following operators [11]

SP B( I ) DX * (2. * C ( I ) 3. * DX * D( I ));
SPP 2. * C ( I ) 3. * DX * D ( I ).
In the C-version of the program SEVAL the function value and

35
the first and second derivatives are being returned. The pointer to the structure
containing the arrays of interpolated function, the coefficients of interpolating
function, and the type of interpolation (linear or spline) are being transmitted
through the heading as well [see Annex 2]. The reflected ray is determined
according to the law of specular reflection [see Eq. (1.31) ]
s 2n (n, g ) g ,
where n is calculated in accordance with Eq.(1.78)
1.6. SURFACE SYNTHESIS BY USING FERMATS PRINCIPLE
The Fermats principle enables to enlighten focusing and concentrating
properties of a surface most completely [15]. It is a known fact that for a path
of any light ray passing between two fixed points A and B the following
relationship is true
B

ds 0 ,

(1.79)

where (s ) is the refractive index related to the path.


For a discrete series of refracting surfaces, when the refractive index is
approximated by a pieccwisc linear function, we obtain
k sk 0 ,
(1.80)

where all the surfaces are included in


the sum.
Let the rays being radiated from the
point F1 converge at the point F2 (Fig.
1.14). We calculate the difference
between the optical pathlengths of two
close rays, and assume that the rays
travel in the air before and after
reflection. Comparing the ray path F1P
and F1Q (Fig. 1.14), we obtain

L (r r ) ( )
(r ) 0

Fig. 1.14. Optical paths of two close rays


converging into point (triangles PQR and PQT
are congruent)

36
and this yields r .
This equation unambiguously defines a curve (cllips) that collects rays at
its real focus

r const 2a
From congruence of triangles PQR and PQT we obtain (Fig. 1.14)
sin i = sin r (the law of reflection),
and

PQ cos i PQ cos r .
The latter equality yields

(1.81)
If the rays arc gathered at a point of imaginary focus, this results in
equation
r 0 ,
whence, r const 2a , i. e. hyperbola equation.
Hence, r , whether we consider elliptic or hyperbolic surfaces.
Generally, dealing with a collection of reflecting surfaces, we have

0,

(1.82)

where all surfaces are included in the sum, and the sign choice is made in
accordance with the said-above remark.
Similarly, using Eq. (1.82) for a number of reflectors, we find

r
k

0,

(1.83)

where we choose + for hyperbolic, and - for elliptic surfaces.


Equation (1.83) being taken with a sign + can be applied to a ray tracing
through a series of refracting surfaces.
We now give the examples of designing (synthesis) optical surfaces by
applying Eqs. (1.82) and (1.83).

37
Parabolic mirror
Let a surface be required that
transforms the flux of a point light source
into a beam of parallel rays (Fig. 1.15).
According to Fig. 1.15, we have rd dz .
Knowing that z r sin , we obtain a
differential equation with separable
variables
d (ln r ) tan( / 2)d .
Integrating the equation under the
initial condition of r ( 0) f , we
obtain

2f
r
1 cos

(1.84)

Fig. 1.15. Optical properties of parabola in


differential form

Spherical lens
We consider a spherical refracting surface which separates the mediums
with refractive indexes 1 and 2 (Fig. 1.16). We find the condition under
which the paraxial rays passing from the first medium into the second one arc
being focused.
According to Fermats principle the optical pathlengths are being equal for
all the rays traveling from the point O to the point O' [16]. An optical
pathlength over-run for an arbitrary ray is 1r1 2r2 , and it has to be equal
to an optical length surplus of a central ray, that is ( 2 1 )x .
Let a height where a ray intersects a surface be h=1. Supposing that the
rays are close to the axis, we obtain
l l 2 d 2 (l d )(l d ) 2lr1 .
Thus,

Fig.1.16 Spherical lens

38

r1 1 / 2l , r 1 / 2l `, x 1 / 2 R ,
where R is the radius of a surface.
We obtain the focusing condition as follows

1 v v 1

l l`
R
or

1 v 1

l l` f

(1.85)

where f=R/(v-1) is the lens focal length, and v 2 / 1 is the relative


refractive index.
We have to note, that the signs rule accepted in optics [13] is used in Eq.
(1.85).
Taking v=1 in the second term in Eq. (1.85), we gel the well-known lens
formula.
Spherical mirror
Setting v=-1 in Eq. (1.85), we get an equation valid for spherical mirror

1 v 2

l l` f

(1.86)

A linear extension of an image is determined by relation (1/l), and an area


extension is a square of linear one, or
M ( R /(2l R)) 2
(1.87)
Hence, a concave mirror (R>0) amplifies luminous intensity of a light
source [17]. For the curvature radius of R = 21 the amplification is infinite, and
this is a property of a parabolic reflcctor [see the formula for meridian section
curvature in Eq. (1.26) when u=0].
It is clear, that occurrencc of infinity at caustic point can be explained by
approximate nature of the optics. Using the laws of geometrical optics, we
assume that the strength gradient of electromagnetic field in considered region
is small in comparison with the field itself [13|. Obviously, this condition is
violated near focal points or caustics.
Though being restricted to some extent, the approach in terms of
geometrical optics is often useful because of its simplicity and inversion-ability
of formulas (transfer from direct to inverse

39
problems). For example, an elementary analysis using Eq. (1.87) shows that
uniformly spreaded defects of a flat mirror in the form of small concavities and
convexities cause the sharpening of luminous intensity curve, and this fact is
confirmed by photometric data [171.
Hyperbolic lens
A lens is required which, having
the refractive index v, transforms a
flux of a point source in the air into a
beam of parallel rays (Fig. 1.17).
Examining Fig. 1.17, we obtain a
differential
equation
dr=v dx.
Accounting that x=r sin and
separating the variables, we get

d (ln r )

v sin
d ,
v cos 1

from where, integrating under the


condition v(0)=f we have

Fig. 1.17. Hyperbolic lens

f (v 1)
v cos 1

(1.88)

i. e. for v>1 we get a hyperbola equation.


When a lens which transforms the rays traveling from a medium with
higher density to a medium with lower density is required, we have to assume
v1/v in above equation. Obviously, we get a lens with elliptic profile [15].
To find an equation of reflecting surface, we have to assume v=-1 in Eq.
(1.88), and then we get a parabola equation (1.84).
Double-surfaced reflector (Cassegrain and Gregorys schemes)
After two reflections the light rays run in parallel to the axis (Fig. 1.18).
The second surface is assumed to be parabolic. Accounting Eq. (1.83), we
write down r1 d1 r2 `d 2 dz . We assume that the first surface and the
parabolic surface have a common centre, so that dz= d=p d2. Hence, we
define an equation for the first surface r1 d1 r2 d 2 , and this is the
equation of conic section [15]. Precisely, the upper sign (+) defines a hyperbola

40
in Cassegrains scheme; and the
lower (-), an ellips in Gregorys
scheme.
One important thing has to be
mentioned. The obtained-above
curves, for instance, described by
conic section equations, arc the only
curves
that
fit
the
stated
requirements for rays transformation.
It follows from the fact, that they
have been resulted as solutions of
ordinary differential equations under
given initial conditions.
The conditions of Cauchys
theorem are met here, so these
solutions are unique.

Fig. 1.18. Scheme with double reflection

Ray-tracing function
In contrast with the previous examples, we have a surface specified by a
profile curve r=r(), and the task is to reconstruct a relationship between
corresponding incident and reflected rays (ray-tracing function) =(). We
determine the function =() for conic sections assigned in polar terms
r p /(1 e cos ) . We take a derivative dr /(r d ) e sin /(1 e cos ) .
Setting it equal to the right side of Boldyrevs equation
dr / r d tan(( ) / 2) ,.
(1.89)
we obtain

tan( / 2) tan( / 2)

1 e cos 2e cos 2 ( / 2)
1 e cos 2e cos 2 ( / 2)

and finally,

tan( / 2)

1 e
tan( / 2)
1 e

(1.90)

From Eq. (1.90) it follows that angles a have to be assumed as positive, i.


e. in case of ellips (e<1) the rays arc being directed towards the optical axis;
and as negative in ease of hyperbola (e>l), when the rays are being directed
from the optical axis. Hence, the double sign in Eq. (1.90) is not necessary.

41
Conic sections formulas
For further convenience we give the formulas for conic sections in
different coordinate systems.
(1) Cartesian system. The coordinate origin is situated at a symmetry
centre of a curve (canonical representation)

x2 y2

1.
a 2 b2

(1.91)

(2) The origin is matched with a vertex of a curve

y 2 2 px (1 e 2 ) * x 2 .

(1.92)

(3) Polar system. The pole is matched with a focus of a curve

p
.
1 e cos

(1.93)

The parameters in given formulas have the following meanings: a and b


are half-axes, a>b;

c 2 a 2 b 2 , c b,
where c is the distance between the focus and the centre; e c / a is the
eccentricity; p b 2 / a a(1 e 2 ) is the parameter of a conic section.
Remarks
(1) The sign (+) in Eq. (1.91) corresponds to ellips; (-), to hyperbola;
(2) Parabola is specified by parameter p or by focal length f=p/2.
Cofocal conic sections are defined by the following formula

x2
y2

1,
a 2 b2

(1.94)

where is a family parameter. It follows from the definition c2=(a2+)-(b2


+ )=a2-b2 that curves have one and the same focus.

42
1.7. MATRIX APPLICATION IN GEOMETRIC AND
OPTICAL CALCULATIONS
We consider now how matrices arc used in lighting calculations.
Spherical Coordinates of a Point
The position of a point in space or on a surface is often described by
spherical coordinates: the radius-vector r, the polar angle , and the azimuth
angle (Fig. 1.19). A vector p coincident by direction with a vector OP of the
length r can be termed as a result of two rotations: (1) by the angle in the
plane XOY; (2) by the angle in the plane ZOY (Fig. 1.19). To determine
matrices of rotation we have to find the coordinates of the revolved basis
vectors, which form the columns in the matrix of rotation.*
For the projections of basis vectors e1 , e2 and e3 on the plane XOY
(Fig.1.19) we obtain
e1 [cos sin 0]T ;

e2 [sin

cos

0]T ;

e3 [ 0
0
1]T .
Consequently, the matrix of rotation by the angle is
cos sin 0
M sin cos 0 .
0
1
0

(1.95)

For the projections of basis vectors


e1 , e2 , e3 on the plane ZOY (Fig. 1.19) wc
obtain
e1 [1
0
0 ]T ;

Fig. 1.19. Coordinates of point P in space


and rotation of basis by angle

We mean the left multiplication of matrices.

e2 [0 cos

sin ]T ;

e3 [0 sin

cos ]T .

Consequently, the matrix of rotation


by the angle is

43

0
1

M 0 cos
0 sin
The resulting matrix is M M M .

0
sin .
cos

According to the rule of multiplication of matrices, we obtain


cos cos sin sin sin
M sin cos cos sin cos .
0
sin
cos

(1.96)

(1.97)

Note, that M M M M . Multiplying the matrix of rotation M by


the vector k [0 0 1]T , we find the vector p :

p [sin sin sin cos cos ]T ,


hence, as OP rp ,
OP [r sin sin r sin cos r cos ]T
or

x r sin sin ,

y r sin cos , z r cos .

(1.98)

Rotation of Lighting Fixture Axis


Let a lighting fixture (LF), considered as a point source, be placed at the
coordinates origin with its axis directed along the axis OZ. It is usually
photometrically tested at a plane, which is perpendicular to the LF axis. Wc
take the photometric plane being at a
distance z=1.
We assume that the LF is aimed
at a point P(x, y) on the plane z=H
(Fig.1.20). To find the LF luminous
intensity at the direction rs, we do the
following: (1) determine the angles
of rotation and from Eq. (1.98);
(2) apply the matrix of inverse
1
rotation M to the vector rs, and
consider the coordinates of obtained
(unrevolved) vector on the plane z=1.

Fig. 1.20. Determination of ray coordinates


on projection plane

44
Since matrix M is orthogonal and unitary [1], its inverse matrix is equal

to the transposed one, i.e. to M


1

. Hence,

cos
cos sin
sin sin

sin
cos cos
sin cos

0
sin .
cos

(1.99)

After rotation defined by matrix M , the vector rs, (Fig. 1.20) turns to

rB M 1rs
or
1
[ x` y` h]T M
[ X Y H ]T .

(1.100)

The coordinates of vector rB in the plane h=1 can be obtained from


Eqs.(1.99) and (1.100) in the following form [18]

m11 X * m12Y * m13


;
m21 X * m23Y * m33

m21 X * m22Y * m23


,
m13 X * m23Y * m33

(1.101)

where X*=X/H; Y*=Y/H; mij are the matrix elements.


The vector coordinates in the plane h=1 may be regarded as the projection
coordinates of a point. The formulas (1.101) describe the most general view of
coordinate translation in an adduced plane. In particular case, when =0 and
900-, Eq.(1.101) transposes into the known expression from
Sapozhnikovs method [19].
If the luminous intensity is prescribed in the plane h=1 as a matrix Jij i. e.
discretely with spacings of hx and hy, then the candlepower related to the LF
aiming point can be determined by reading from the array J(i,j) with i=[x/hx]+
1, j=[y/hy]+1, where expression being put in square brackets denotes an integer
part of a number.
Earlier we saw that the matrix of reflection takes the most simple form
being represented in the basis, where one of the vectors is directed along a
surface normal at a point of ray incidence. This is because the basis consists of
eigenvectors of reflection matrix Mrefl from Eq. (1.34). Defining the normal at
a point of incidence in terms of spherical coordinates, i. e. by the angles and
(Fig. 1.21), we obtain the expression similar to Eq. (1.98)

45

n [sin sin sin cos cos ]T .

(1.102)
The vector n described by Eq. (1.102) can be resulted from the translation
of the basis axis vector OZ ( k ) into the vector n . Consequently we pass to
another coordinate system, to a new vector basis e1 , e2 , e3 e.i, where e3 is
directed along the normal; e2 , along the meridian =const; and e1 , along the
parallel =const (Fig.1.21).* If a reflecting surface is produced by rotation of a
curve r=r() about the axis OZ, then ( ) , and the parallel is defined by
the line =const. Obviously, the coordinates of an arbitrary vector within the
1
basis e1 , e2 , e3 are determined by applying inverse matrix M similar to that
from Eq. (1.99)

sin
0
cos

(1.103)
M cos sin cos cos sin .
sin sin sin cos cos
Let an incident ray at point P be set as so [ sox soy soz ]T . In the local
1

system of coordinates it is denoted as vector s`o


1
s`o M
so ,

(1.104)

1
where matrix M
is defined by Eq. (1.103).

After reflection, the ray travels at the direction assigned by vector s [see
Eq. (1.36)]

Fig.1.21 Local coordinate system at mirror point

Vectors

e1 , e2 , e3

compose the right triple.

46

s` [s `ox s`oy s `oz ]T ,

(1.105)

o determined from Eq.


where s `ox , s `oy , s`oz are coordinates of vector s`
(1.104).
The matrix from Eq, (1.103) enables to solve all the tasks connected with
reflection from a specular surface. Let an incident ray be set as so [0 0 1]T .
Then according to the rule of matrix multiplication, this ray can be defined in
the local coordinate system
1
s `o M
[0 0 1]T [0 sin cos ]T .
Obviously, the reflected ray is

s` [0 sin cos ]T .
In the main coordinate system the reflected ray is defined as

s M s` [ sin 2 sin sin 2 cos cos 2 ]T .


We have got the result well-known in optics; when a normal is being
rotated by angle , reflected ray rotates by the double angle 2 . The formulas
(1.103)-(1.105) are suitable for automated calculations of ray-paths.
When transferring to the local coordinate system, some additional
considerations have to be accounted. For example, within inverse ray-tracing
method when searching for an intersection point between a ray and a light
source, it is more convenient to translate a ray into the coordinate system
associated with a light source, thus presenting a light source surface in the most
simple form.
Coordinate System Associated With a Light Source
Let a light source be rotationally
symmetric. It is convenient to direct the
basis vector, say e3 , along the axis of
symmetry (Fig, 1.22).
As it follows from said above [see Eq.
(1.99)], in order to define the matrix of
rotation it is sufficient to find the coordinates of novel basis vectors within
original
coordinate
system
XOZ
associated
with
reflector;
these
coordinates give the columns of desired
matrix.

Fig.1.22. Displacement of coordinate system


into point O' associated with light source

47
Let e3 [u n w]T , where u 2 v 2 w 2 1 . In the absence of other
considerations it is suitable to choose e1 [v u 0]T . It is easy to prove that

e1 is orthogonal to e3 . Vector e2 is orthogonal to e1 and e3 as well. Therefore,


e2 can be specified by vector product between e3 and e1
e2 [e1 e3 ] ,
or
i
e2 u

j
v

v u

k
.
w
0

Expanding the determinant, we get

e2 [uw vw w 2 1]T .
Hence, the matrix of transfer to the coordinate system associated with a
light source takes the form
uw
u
v

(1.106)
M s u
vw
v .
0 w 2 1 w
Let a point P be assigned by vector Rp=[xp yp zp]T in the coordinate system
associated with reflector (the main system). The origin of the local coordinate
system is r0 [ x0 y0 z 0 ]T . Then in the local coordinate system this point is
defined by equations
rsp R p r0 [ x p x0 y p y0 z p z 0 ]T ;

r `sp M s rsp ,
where Ms is calculated from Eq. (1.106).
Light Source Image Produced by Reflecting Surface
Reflection is a linear transformation, as it is shown in paragraph 1.3.
Consequently, if a and b are arbitrary vectors, while p and q are scalar
constants, then
(1.107)
M refl ( pa qb) pM refl a qM refl b .
Since every vector can be defined within the basis, then, according to Eq.
(1.107), its transformation by a specular reflector can be found.
Let a surface of revolution be assigned by angles and . The basis of
rectangular coordinate system is specified by vectors

48

Fig. 1.23. Image of light source produced bv element of specular surface:


a - orientation of source, mirror, and pictorial plane;
b - images of basis vectors

e1 , e2 , e3 (Fig. 1.23). We direct the normal being al the point , outwards the
surface; the angle between the normal and the axis OZ is (see Fig. 1.21).
Substituting the coordinates of vector n into Eq. (1.34), we obtain the matrix
of reflection in the following form
2 sin 2 sin 2 1
sin 2 sin 2
sin 2 sin

M refl sin 2 sin 2


2 sin 2 cos 2 1 sin 2 cos . (1.108)
sin 2 sin
sin 2 cos
2 cos 2 1

We find now the images of vectors et and e1 and e2

e`1 M refl e1 2 sin 2 sin 2 1 sin 2 sin 2

e`2 M refl e2 sin 2 sin 2

2 sin 2 cos2 1

sin 2 cos . (1.110)


T

sin 2 sin ; (1.109)


T

For the points Ph ( 0) and Ph ( 90 0 ) [see Fig. 1.231 the images of


vectors e1 and e2 on the plane being parallel to XOY (the plane of observation
or pictorial plane) take the form
0]T , 0;
[1
(1.111)
e`1
[ cos 2 0]T , 90 0 ;
[0
cos 2 ]T , 0;
(1.112)
e`2
[0
1]T , 900 ;
Thus, according to Eqs. (1.111) and (1.112), vectors may change
orientation, and their length may be subjected to projective shortening. Let us
find the angle of rotation of vector e1 after

49
reflection (Fig. 1.23 b). Taking the coordinates of e`1 from Eq. (1.109), we
obtain

sin 2 sin 2
.
tan

e`1x 2 sin 2 cos 2 1


e`1 y

(1.113)

Transformations in Eq. (1.113) lead to

tan

(1 cos 2 ) tan
.
1 tan 2 cos 2

(1.114)

The maximal angle of rotation can be obtained by differentiating Eq.


(1.114)

tan max

1
.
cos 2

(1.115)

Equations (1.109) and (1.110) describe rotation of elementary map of a


linear source being perpendicular to the optical axis of reflector with arbitrary
geometry. They are similar to that of paraboloid, if we take / 2 [20].
For reflector points specified discretely with steps h and h, the matrix of
reflection [Eq. (1.108)] can be determined in advance. The use of matrix
multiplication [see Eq. (1.35)] enables to create effective algorithm for
determining light source images on an arbitrary pictorial plane.
For a light source, in particular, having rectangular form, wc may delete
the restriction on its dimensions l and h [20] (they supposed to be small
enough). Any contour point of rectangle may be assigned by vector
s te1 e2 ,
where 0 t 1 ; 0 h ; t and are some scalar constants.
In accordance with Eqs. (1.107), (1.109), and (1.110), we obtain full
description of elementary maps for the points Ph and Pv that belong to a
paraboloid

Ph : s x t ,

s y cos ,

Pr : s x cos , s y .

(1.116)

Note, that in the course of linear mapping described by matrix Mrefl,


straight lines transform into straight lines, and a trace of elementary image on a
plane is rectangular as well. If a grid of test points is specified on the pictorial
plane, then it is easy to develop an algorithm that makes record when an
elementary image covers a given test point (a node).

50
REFERENCES
(1) Streng G. Linear Algebra and Its Applications. Academic Press, New York San Francisco - London, 1976.
(2) Rice J.R. Matrix Computations and Mathematical Software. Mc Graw Hill
Book Comp., St.Louis - San Francisco, 1981.
(3) Rogers D.F., Adams J.A. Mathematical Elements of Computer Graphics. Mc
Graw Hill Book Comp., St.Louis - San Francisco, 1980.
(4) Faux I.D., Pratt M.J. Computational Geometry for Design and Manufacture.
Ellis Horwood Ltd., 1979.
(5) ONaill E.L. Introduction to Statistical Optics. Addison-Wesley Publ.Comp.,
Reading (Massachusetts) - London, 1963.
(6) Gerrard A., Burch J.M. Introduction to Matrix Methods in Optics. Wiley Interscience Publ., London - New York - Sydney - Toronto, 1978.
(7) Korn G.A., Korn T.M. Mathematical Handbook for Scientists and Engineers.
2nd edition. McGraw-Hill Book Comp., 1968.
(8) Rodionov S.A. Computer-Aided Design of Optical Systems. L.:
Mashinostroenie, 1982 (in Russian).
(9) Alberg J.H., Nilson E.N., and Walsh J.L. The Theory of Splines and Their
Applications. Academic Press, New York - London, 1967.
(10) Stockmar A. Spline-Funktionen-eine neue Methode zur Darsiellung von
Lichtstarkeverteilungen.// Lichttechnik, 1975, N8, s.320-324.
(11) Forsythe G.E., Malcolm M.A., Moler C.B. Computer Methods for
Mathematical Computations. Prentice-Hall, N.J. 1977.
(12) Kusch O.K., Sofronov N.N. Computer Calculation of Specular Reflectors by
Using Splines.// Svetotekhnika, 1985, N4, p. 12-13 (in Russian).
(13) Born M., Wolf E. Principles of Optics, 4th cd., Pergamon Press, Oxford,
1970.
(14) Stavroudis O.N. The Optics of Rays, Wavefronts, and Caustics. Academic
Press, New York - London, 1976.
(15) Cornbleet S. Microwave Optics. Academic Press. New York - San Francisco
- London, 1976.
(16) Feynman R.P., Leighton R.B., Sands M. The Feynman Lectures on Physics.
Vol. 1. Addison-Wesley Publ. Comp., Massachusetts, Palo Alto, London, 1963.
(17) Benford F. Studies in the Projection of Light. General Electric Review, 192326.
(18) Vakhromeeva L.A., Bugaevskii L.M., Kazakova Z.L. Mathematical Cartography. M. Nedra, 1986. (in Russian).
(19) Knorring G.M. Lighting Calculations in Installations of Artificial
Illumination. M.-L., Energiya, 1973. (in Russian).
(20) Karyakin N.A. Lighting Devices of Floodlighting and Projection Types, M.
Vysshaya Shkola, 1966. (in Russian).

51

CHAPTER TWO
METHODS FOR CALCULATING LIGHT DISTRIBUTION
2.1. LIGHT-RAY METHODS
Light-ray methods can be originated from two propositions [1]:
(1) The property of luminous field to form light-tubes that transmit one
and the same luminous flux through their arbitrary cross-section. This is
expressed by the following equation: divE=0;
(2) The existence of a family of orthogonal sections in a light tube
(Maluces theorem).
We consider a light-tube falling at a mirror surface and cutting an area
du dv there (Fig. 2.1). The mirror-surface element receives a luminous flux

d Io

I[ru rv ]
du dv ,
r2

(2.1)

where Io is the luminous intensity at the tube; r=r(u, v) is the ray equation
in curvilinear coordinates u, v related to the surface.
During reflection, a part of luminous flux is being lost. The rest of
luminous power spreads within the other light-tube that ends at

Fig.2.1. On energy balance within incident and reflected tubes

52
the target surface by an element I1[ Ru Rv ] du dv, where I1 is the unit vector
of reflected ray. Consequently, the normal illuminance at the target surface is
[2], [3]:

E n I o

I0 [ru rv ] *
r 2 I1[ Ru Rv ]

(2.2)

Expression for the normal illuminance at the target surface can be


presented as follows
En Eo D (u , v) ,
(2.3)

D (u , v)

I0 [ru rv ]
I1[ Ru Rv ]

where D (u , v ) is the coefficient equal to the ratio between illuminances in


incident and reflected beams; E0 is the illuminance at mirror surface.
Hence, the multiple D(u,v) is the Jacobian of the transformation dS1dS2
[4].
Now we determine a field in remote zone. A vector corresponding to a
point at the target plane is

R r I1 .
Partial derivatives of vector R are

Ru ru I1u , Rv rv I1v ,

(2.4)

Substituting Eq. (2.4) into Eq. (2.2), we obtain

E n I o

I0 [ru rv ]
.
r 2 I1 (ru I1u )(ru I1v )

(2.5)

Multiplying Eq. (2.5) by and assuming , we obtain the value for


intensity at the remote field

I E0

I0 [ru rv ]
.
I1[ I1u I1v ]

Since E-value is positive, mixed product should be taken with relevant sign (+) or (-).

(2.6)

53
Formula (2.6) defines the field at the remote zone in terms of proximate field.
We set r r0 0 I0 . Substituting r into Eq. (2.5) and supposing 0, we
obtain

I I 0

I0 [ Iu Iv ]
.
I1[ I1u I1v ]

(2.7)

If the candlepower pattern I 0 (u , v) of a point source is known, expressions


in Eqs. (2.5)-(2.7) allow to determine the intensity at the remote and proximate
fields. Light-tubes directions can be obtained by applying the law of reflection

I1 I0 2n ( I0 , n ) .

(2.8)

Surfaces S1 and S2 in Eqs.(2.2), (2.5), and (2.7) are prescribed in


parametric form

r x(u, v) y (u, v) z (u , v ),
R X (u , v ) Y (u , v ) Z (u, v),

(2.9)

whence, we obtain

ru [ xu yu zu ]T , rv [ xv yv zv ]T
Ru [ X u Yu Z u ]T , Rv [ X v Yv Z v ]T .

(2.10)
Equations (2.3), (2.5)-(2.10) make up the closed sequence for calculating
the reflected field intensity when the point-source candlepower distribution is
known.
We must note, that similar equations based on geometric representation of
luminous field arc being used for calculating scalar density of diffused energy
reflected from antennas [5], when designing simulators of thermal radiation
[6], and in the wave theory [7]. Spherical coordinates u=, v= are often being
chosen as reflecting surface parameters. Since surface equation is prescribed,
consequently, the pathways equations for reflected rays a=a (,) and =
(,) are known as well (where a and are spherical coordinates of reflected
ray). The unit vectors for incident and reflected rays are as follows

I0 sin sin
I sin sin
1

sin sin

cos ,

sin cos

cos

Differentiating Eq. (2.11) and substituting the results into Eq. (2.7),

(2.11)

54
after transformations we obtain an expression known from [8]
D( , )

sin
a

sin a
a

(2.12)

The use of obtained expressions for computing the intensity of reflected


field is connected with the coordinate representation of vectors and surfaces.
When the number of surfaces is large, the other formulas are preferred; the
formulas, which present intensity in invariant form, i. e. as a function of
geometric and differential properties of reflecting surface, and parameters of
incident and reflected rays.
Formulas for calculating coefficient D
Table 2.1
Source type
Spatial source:
proximate
zone

remote zone
Linear source:
proximate
zone
remote zone
Spatial source:
proximate zone
remote zone
Linear source
proximate zone
remote zone

Vector form
Coordinate form
Parametric representation
x

l1 x

l1 y

l1 z

xu

yu

zu : X u

Yu

Zu

xv

yv

zv

Yv

Zv

x
( x 2 y 2 z 2 ) 1/ 2 xu
xv

y
yu
yv

l1 x
l1 y u
l1 y v

l1z
l1z u
l1z v

I0 [ru rv ]
I1[ Ru Rv ]

(x y z )

I0 [ I0 u I0 v ]
I [ I I ]
1

1u

1v

2 1 / 2

( ru , I0 )
( Ru , I1 )

( x 2 y 2 ) 1/ 2

( I0u , I0 )
( I1u I1 )

( x 2 y 2 ) 1/ 2

Xv

z l1x
zu : l1x u
zv l1 x v

xxu yyu
X u l1x Yu l1 y
loxu x lo yu y
l1xu l1x l1 yu l1 y

Invariant representation
( K refl 2 2 H refl 1) 1
~
~
[1 2 0 cos (2 H k tan 2 )] 4 02 K ]1
1

1
1
2
1

( 0 )
0 R1 cos

2 0
1

1 cos

Remark. We have E=I0D at the proximate zone, and J=IoD at the remote
zone. We take an absolute value of D.

55
Equations for the value F=l/D, known as divergence, were obtained in a
general form by V.A. Fock [9], and are as follows*
F K refl 2 2 H refl 1 ,
where

1
~ 2 cos
K refl 4 K
(2 H k tan 2 ) 2 ;

0
1
H refl cos (2 H k tan 2 )
;
0

(2.13)

1
~ 1 1
1
cos 2 sin 2
, k
, H
.

R1R2
2 R1 R2
R1
R2
~ ~
The following denotions are used in Eq.(2.13): K , H are the Gaussian
~
K

and mean surface curvatures at the point of ray incidence; R1, R2 are the radii
of principal curvatures of a reflecting surface at the point of ray incidence; is
the angle between the plane, which contains the ray, and the plane of the
principal normal section; k is the curvature in the section containing the ray;
is the angle of ray incidence; 0 is the distance between the light source and the
intersection point; is the distance along the reflected ray up to the point at the
target surface.
For the remote zone () the divergence takes more simple form [10]

~ ~
~
F 1 2 0 cos (2 H k tan 2 ) 4 02 K .

(2.14)
The ratio D=F =I/I0, where F is calculated from Eq. (2.13), may be
defined as an indicatrix of reflected radiation (or as a direction diagram) in the
remote zone. It is an advantage that Eqs. (2.13, 2.14) do not depend on the
choice of coordinate system, because all the values here have a particular
physical meaning. Given formulas describe reflected field intensity in the case
when the light source possesses the spatial intensity distribution of Io(u, v).
Complete summary of D-factor expressions for various radiators, including
linear source I0=Io(u), and surfaces representations is given in Table 2.1 (where
indices u, v denote partial derivatives with respect to corresponding
coordinates).
-1

Similar equations for calculating the amplification factor of a specular surface are known in lighting
engineering [8].

56
Example
Let us consider the calculation of the illuminance pattern created by
paraboloid with a light source in its focus. Exact formula for calculating
illuminance from paraboloid can be obtained from Eq. (2.13). We have (see
Ch. 1) 0 f cos 2 ( / 2) , R1 2 f cos 3 ( / 2) , R2 2 f cos 1 ( / 2) .
Substituting these equalities into Eq. (2.13), we find the divergence to be
F 1 . Therefore, when I0=1 ang p0=1, it yeilds

En

1
1

2 cos 4 .
2
0
f
2

(2.15)

Using transformation =2arctan (x/2f), Eq. (2.15) yeilds a function with


respect to x-coordinate

En

1
.
f [1 ( x / 2 f ) 2 ]2
2

(2.16)

According to theory, the value of En does not change when the distance
from the optical system increases. It means that the light-tube originated from
paraboloid has a beam shape inside of which a constant amount of energy
flows. It is seen from Table 2.1 and Eq. (2.13) that in the reflected field the
points where F=0 (the caustic points) may appear. In the programs based on
ray methods the special means have to be foreseen to prevent computer
overflow when caustics occur, and corresponding messages signalling about
caustics have to be printed.
Light Distribution Created by Surface with Linear (Filament) Source
Let 0 0 ( ) be the equation for the profile of cylindric surface.
Obviously, the profile curve is identically defined (when the initial point
0 ( ) is prescribed) if the ray-tracing function a=a() is known.
According to the formulas from Table 2.1, we determine the normal
illuminance

En

I
,
0 (1 / 0 1 / 2 /( R cos a))

(2.17)

Supposing differentiability of a(), we find the curvature radius of the


profile

57

2 0
,
(1 a ) cos a

(2.18)

where a =da/d.
Substituting (2.18) into Eq. (2.17), we get

En

I
.
0 a

(2.19)

We consider some specific cases in defining profile curves:


(1) A straight line a 2 0

1 ; En I /( 0 ) .
(2) A parabola const
0 ; En I / 0 .
It follows that parabola as a profile curve can be defined in two equivalent
ways (for I=const):
(1) as a curve which originates a parallel beam of light;
(2) as a curve which provides the normal illuminance being inversely
proportional to the radius-vector of a section point for cylindric surface, and to
the second power of radius-vector for rotationally symmetric surface.
Location of caustics can be found from the following equation
(1 / 0 1 / 2 /( R cos )) 0 ,
(2.20)
or from equation being linear with respect to
0 0 .
(2.21)
The above-stated equations can be obtained directly from the Fermats
principle [see 1.6].
Assuming the angles to be small and cosa=1 (paraxial approximation), we
get the equation of focusing [Eq. (1.81)].
Correspondingly, we obtain

0 R cos a
,
2 0 R cos a

(2.22)

and

0
,
a

(2.22)

58
From Eq.(2.23), in particular, it is seen that all the singular points of
parabola appear in infinity. The parameters , R, cosa depend on surface
variable u only. To specify a surface through ( ) -characteristic, we may set
up u=.
The caustics surface is described as follows
R(u ) r (u ) (u )a (u ) ,
(2.24)
where r is the radius-vector of reflector point, a is the vector of reflected ray.
By tending in Eq. (2.24), we obtain the indicatrix of the reflected
light

I
,

(2.25)

which is in complete agreement with the flux balance.


For point-discrete assignment of a profjle curve the most correct way is to
specify the function ( ) , because only single differentiation of Eq.
(2.25) is needed. This situation occurs when, e. g. the functions r r ( ) and
( ) are obtained as inverse problem solutions, and the direct
calculation, has to be carried out.
Similar results can be obtained in gerteral Case, when a surface is specified
by a pair of parametric functions (u , v) , (u , v) which describes
the paths of reflected rays. Location of caustics is defined by the roots of
square equation
K (u , v ) 2 (u, v) 2 H (u, v) (u, v) 1 0 ,
(2.26)
while the equation of caustic surface is as follows
R(u, v) r (u , v) (u , v )a (u , v ) .
2.2. LIGHT DISTRIBUTION CREATED BY ELLIPSOID OR
HYPERBOLOID WITH A POINT SOURCE
Reflectors used for illumination can be divided into two groups:
hyperboloids and ellipsoids. Different kinds of ellipsoids are deep and narrow,
they converge the rays towards the optical axis; while hyperboloids are wide
and shallow, and diverge the rays outwards the axis.
After passing through the point of secondary focus the beam of ellipsoid
appears like that of hyperboloid; so to say, optical equivalence takes place [11].
For rather wide beam, the size of a

59
lighl source effects only on the resulting angular width of a lightspread curve,
thus the point-source theory can be used for analysis of ellipsoids and
hyperboloids [11, 12].
Returning to Fig. 1.14, we find
r sin sin a .
(2.27)
Taking into account Eqs. (1.81) and (2.27), we obtain a differential
equation that defines a relationship between the angles a and (the ray path)

da sin a

.
d sin

(2.28)

Light Amplification in Remote Zone


Now we determine the amplification factor Kam yielded by a surface.
Taking Eq. (2.12), we obtain for axially symmetric surface

sin
,
sin ada / d

or accounting Eq. (2.28)

(sin / sin a) 2 .

(2.29)

We calculate the quantity


2

a
1 tan 2 / 2
1 e
2
1 tan
1

(1 2e cos e 2 ) .
tan
2
2
1

e
2
(
1

e
)

Further we find

1 e

2
tan (1 e) 2
2 tan a / 2
1 e
2
sin a

,
1 tan 2 a / 2 (1 tan 2 / 2)(1 2e cos e 2 )
or

sin a sin

1 e2
.
1 2e cos e 2

Hence, we obtain from Eq. (2.29) (sec [14]):


2

K am

1 2e cos e 2
.

2
1

(2.30)

60
Equation (2.30) combined with Eq. (2.22) completely defines luminous
intensity distribution of hyperboloid or ellipsoid in the remote zone within the
whole beam section except for its boundaries, where edges are not so sharp as
it follows from Eq. (2.30).
Equation (2.30) can be represented in equivalent form [11]:
K am ( p / r )2 .
(2.31)
Consequently, for a convex hyperbola with Kam<1 the reflected light beam
is more diffused and less intensive than the initial radiation of a light source.
The observed picture is inverse to that of a concave hyperboloid: light beam
has minimal intensity in the centre and reaches maximum value equal to the
light source intensity at the edges, provided that reflector is expanded to
infinity [12].
For cylindric reflecting surface with a linear light source we obtain [14], in
much the same way as per Eq. (2.30), the following

K am

1 2e cos e 2
,
1 e2

(2.32)

where the upper sign refers to elliptical reflectors, and the lower sign, to
hyperbolic ones.
Light Amplification in Proximate Zone
To determine the illuminance we
apply Eq. (2.13). Futher, we have to find
angle v between the radius-vector and the
tangent to a conic section curve (Fig.
2.2).
Differentiating the radius-vectors, we
obtain
1

dr
tan v r
.
d
Since r p /(1 cos ) , we have
tan v (1 cos ) / e sin ,
Fig. 2.2. On deduction of amplification
factor in proximate zone

61
and futher,

tan v
1 e cos

;
2
1/ 2
(1 tan v)
(1 2e cos e 2 )1 / 2
1
e sin
cos v

.
2
1/ 2
(1 tan v)
(1 2e cos e 2 )1 / 2

sin v

We find the radius of curvature in the meridian section by applying known


formulas (see Ch. 1):

(r 2 r `2 )3 / 2
Rm 2
.
r 2r `2 rr ``

(2.33)

Now we calculate

r 2 r `2 p 2 (1 e cos ) 4 (1 2e cos e 2 );
r ``r p 2 (1 e cos ) 4 [e cos (1 e cos ) 2e 2 sin 2 ];
r 2 2r `2 p 2 (1 e cos ) 4 [(1 2e cos e 2 cos 2 ) 2e 2 sin 2 ];
r 2 2r `2 r ``r p 2 (1 e cos ) 3 .
Substituting the obtained expressions into Eq. (2.33), wc get

1 e cos
Rm p
1 2e cos e 2

1/ 2

p sin 3 v .

Applying Meusners theorem and setting the radius of a parallel equal to x,


we obtain the radius of curvature in the sagittal section
Rs x / cos ,
where is the angle between the normal at the mirror point and the
parallel (Fig. 2.2). We find

(900 ) (v 900 ) v ;
e sin
cos(v ) cos v cos sin v sin
cos
(1 2e cos e 2 )
1 e cos

sin sin (1 2e cos e 2 )1 / 2


(1 2e cos e 2 )1 / 2

62
Thus,
1

Rs p1 e cos sin (1 2e cos e 2 )1 / 2 sin 1 p sin 1 v .


The Gaussian curvature at a point is

~
K 1 / Rm Rs sin 4 v / p 2 .

The mean curvature is

~ 1 1
1 1
H

(sin v sin 3 v) .
2 Rm Rs 2 p
Applying reflection formulas, we determine

1
~ 2 cos ~
K refl 4 K
(2 H k tan 2 ) 2 ,
R0
R0
where v 900 is the angle of incidence; R0 r ; 0

cos 2 sin 2
1
k

, cos sin v , tan cot v .


Rm
Rs
Rm
We find the expression in brackets

1
2 sin 2 v
~
sin 3 v
.
cos (2 H k tan 2 ) sin v (sin v sin 3 v)
cot 2 v
p
p
p

Futher, we have
2


4r
r
1 e2
~
;
K refl r 1 4 Kr 2 sin 2 v 1 2 sin 2 v
2
p
p

1 2e cos e
~
2
1 e2
H refl r r sin v(2 H k cot 2 v ) 1 r sin 2 v 1
p
1 2e cos e 2
2

The divergence is

F K refl R 2 2 H refl R 1 ( K am 1)2 ,


where

K am

1 e2
r
; .
2
1 2e cos e
R

63
Thus, the illuminance at arbitrary distance from the focus is

E F 1En

En
r

1 e2

2
R 1 2e cos e

(2.34)

where En is the normal illuminance of a mirror clement produced by a light


source.
We consider some specific cases:
(1) the candlepower indicatrix or directional diagram in remote zone,
;
2

1 2e cos e 2
;
J ER
1 e2

(2) the illuminance provided by a paraboloid (e = 1)


E E0 J / r2 ;
(3) the caustic (we suppose that F= 0 or Kam =1)

K am

1 e2
r
.
2
1 2e cos e
R

From Eq. (2.23) wc have Kam = r/p, R=p, i.e. the second point is at the
focus of elliptic reflector. Since Kam is a nonnegative value, the equality holds
true only for e1;
(4) spherical reflector (e=0)
F ( 1)2 I [( R r ) / r ]2 ,
that is the intensity deterioration produced by spherical surface (the same result
can be obtained from geometrical considerations);
(5) there is a complete equivalence between ellipsoid and hyperboloid in
the infinity. At close distance (for the same -values) F is greater for e>1, i.e.
the beam of hyperboloid is dispersed at a greater state.

64
2.3. INVERSE-RAY METHODS
Occurcnce of infinilc illuminance al some field points (caustics), when the
ray formulas (2.13) are applied, is caused by breaking the rules in usage of
geometrical terms [10].
Actually, in a fixture with real light source the averaging of luminance
over the exit pupil lakes place. Formally, any separate point of the exit pupil
can make a definite contribution to illuminance, acting like a -source, i.e. a
source with infinite luminance.
In accordance with the known expression for the illuminance at the point
P, we have

L
cos1 cos 2 dS ,
R2
and for the luminous intensity at the direction a0
E ( P)

I (a0 ) L cos dS ,

(2.35)

(2.36)

where L is the luminance of a flashed area dS of reflector when viewed from


the point P along the direction a0 ; R is the distance from the area element dS
to the point P; p is the reflectance (further we set up p=1).
Integration in Eqs. (2.35) and (2.36) is carried out over the total reflector
surface or over the exit pupil.* The integral here may be treated in sufficiently
broad sense, including the -function representation of luminance distribution.
We suppose that there exists a system of nonlinear coordinates u, v; i.e. the
surface equation takes the form r=r(u,v) or r=[x(u,v) y(n,v) z (a, v)]T. We
rewrite Eqs. (2.35) and (2.36) in the following form

1
L(u , v )(an , n )(a0 , n p ) H (u, v)du dv;
R2
I (a0 ) L(u , v )(a0 , n ) H (u, v)du dv;

E ( P)

(2.37)
(2.37)

where H(u,v) is the Jacobian of transformation to nonlinear coordinates u, v;


n p is the unit normal vector to the area P; R Pm . The vector a0 in Eq.
(2.38) is external parameter for the integral of luminous intensity, while the
viewing vector a0 in Eq. (2.37), generally, is a function of current reflector
point P.
On the surface we introduce a grid with cell dimensions hu and
*

In accordance with Maxwells principle [15].

65
hv along u and v coordinates, respectively

ui uo (i 1)hu , i 1, N u ;
vi vo (i 1)hv , i 1, N v . ;
For each node we assign the value ksw=1 if a point is perceived as bright,
and ksw= 0 in the opposite case. The discrete analogs of Eqs. (2.37) and (2.38)
applicable for computer calculations take the form
Nu Nv

E ( P) k sw
i 1 j 1

L(ui , vi )
(aoij , nij )(a0ij , n p ) H (ui , v j )hu hv ;
Rij2

(2.39)

Nu Nv

I (a0 ) k sw L(ui , vi )(aoij , nij ) H (ui , v j )hu hv ;

(2.40)

i 1 j 1

Expressions (2.39) and (2.40) give assessments of E and I values with


accuracy related to the intervals of discretization hu and hv. These relations arc
applicable cither to symmetric or asymmetric surfaces if we possess an
independent procedure for calculating the radiance factor ksw at an arbitrary
point i, j.
Calculation by using the inverse-ray method [16-20]
Realization of the inverse-ray method includes the following stages:
(1) description of luminance and geometric properties of a light source
body;
(2) description of an optical surface including computation of normals and
the matrix of reflection;
(3) parametrization and discretization of the optical surface;
(4) discretization of the target surface (the pictorial surface);
(5) description of the inverse ray;
(6) computation of a ray path after reflection (determination of the inner
ray);
(7) testing whether a ray intersects a light source (ray incidence);
(8) reading of the ray luminance.
At the first stage we specify the following: the equation of luminous
surface of a source in the local coordinate system X's Y's Z's that is F'(r's)=0; the
transfer vector -r0 (Fig. 2.3); and the matrix Ms of transition from global to the
local coordinate system associated with a light source. The equation of
luminous

66

Fig.2.3. Principle of calculation by inverse-ray tracing


1-exit-pupil plane; 2-reflector; 3-light source

surface in the global coordinate system XYZ is known, that is F(S) = 0, hence

rs M s1r `s ;
S rs ro .
Discretization of optical surface (the third stage) is a necessary step in
creating a numerical model based on the inversc-ray method. Since the grid
cells arc chosen, the error of computation depends on the numerical
interpolation and integration methods being applied. Note, that rather than
dividing a surface into elements u v , we may apply discretization of an
exit pupil area Scos [see Eqs. (2.35), (2.36)]. This relates to the imposition
of a rectangular grid on the exit-pupil plane projection observed from the
direction a0 , followed by scanning the grid rows and testing if nodes are
luminous or not (Fig. 2.3). An algorithm of searching for luminous points,
actually, is similar to one that has been applied in television; the latter is based
on the image expansion into an ordered matrix of luminous spots arranged into
the rows and columns (Fig. 2.4). Similar algorithms, known as ray-tracing
algorithms, are widely used in computer graphics for picturing various
photometric objects [21-23].
Stages 5-8 are being carried out within every cycle of ray tracing and are
destined for calculation of radiance factor ksw and the

67

Fig. 2.4. Algorithm for searching


flashed points over reflector
(0end, 03600);
direct tracing;
inverse tracing

luminance at given direction. Here follows the algorithm for calculating these
values:
Step 1. We find the reflected ray (signs i, j are omitted) at the nodal point
of reflector
a 2n (a0 , n ) a0 ,
(2.41)
or in scalar terms

ax 2nx sp a0 x ;
a y 2n y sp a0 y ;
az 2nz sp a0 z ,
where sp (a0 , n ) .
Step 2. We test whether the reflected ray a intersects a light source. Here,
the equations for the reflected ray and a surface have to be solved
simultaneously

s r al ,
F ( s) 0 ,
where a is defined from Eq. (2.41); r is the nodal point vector; l is the ray
parameter (see Ch. 1).
To establish incidence of a ray is often sufficient when dealing with
sources of uniform luminancc. If ksw=1 and radiation of a source is perplexed,
then one more step is needed.
Step 3. To find the point of intersection and calculate the luminance by
using the radiance indicatrix Ls Ls ( s , s ) . In this case, we have to turn into
a light-source coordinate system by

68
applying matrix transformations and find spherical coordinates of a ray a (Fig.
2.3), and then get the luminance values from appropriate tables or formulas.
Algorithms for Surfaces of Revolution
We consider an optical system with
a specularly reflecting surface having a
shape of a surface of revolution (Fig.
2.5). Let OZ be the geometric axis of
reflector; dS be the mirror element of
the reflector or the point of reflection
(the calculated position of dS is in the
right half-plane ZOY); xo=0, yo, zo be the
coordinates of the element dS centre;
and a be the polar angles for incident
and reflected rays; be the polar angle
of the normal to the mirror element, due
to symmetry of reflector the normal n
Fig. 2.5. Parameters of specular reflection
lies in the plane ZOY.
surface of revolution
The angles and a are related
according to the law of specular
reflection
2 ,
here the angles are positive when counted at clockwise direction. The
viewing vector in the axially symmetric system becomes
a0 [sin a sin sin a cos cos a ]T ;
the normal vector is
n [sin sin sin cos cos ]T .
From these equations it is obvious that any relationship between a0 and n
keeps unchanged provided that [24]:
(1) the point of observation is fixed: 0 , a=const, but the ray of
observation moves along the parallel =const, =var;
a0 [0 sin a cos a]T , n [sin sin sin cos cos ]T .

69
(2) the point on reflector is fixed: =const, =0, but the point of
observation moves in such a way that =const, =var, i.e.,
a0 [sin a sin sin a cos cos a ]T , n [0 sin cos ]T .
Existence of mapping a=a(), = that characterizes ray paths in symmetrical
specular system, enables to rearrange the latter relationships in terms of
variables and . This situation is illustrated in Fig. 2.6. The observer leaves
the initial point and moves along the parallel a=const viewing the element ,
under various angles . (The scheme is efficient when a lighting fixture with
regularly diffusing reflector is under design).
We note, that if a source shows complex luminance performance, then the
only way to find the luminance distribution over elementary mapping (EM) is
to scan by using inverse-ray tracing. The most important role here plays a
remark made by V.D. Komissarov: an outer-field ray, i. e. a reflected ray, is
one-to-one related with an inner-field ray by the law of reflection, hence, if
reflected ray is prescribed, the corresponding incident ray is defined as well.
First of all, it is advisable to deal with the rays of inner-field. We present
Komissarovs scheme for determining ray luminancc within the EM for highpressure mercury fluorescent lamp (Fig. 2.7):

Fig. 2.6. Cone of observation [ray A(,)A']

Fig. 2.7. Scheme for determining ray luminance


in course of inverse tracing

70
(1) to fix the direction a;
(2) while assigning various , to define the outer inverse ray

(a 0), sin a

A [cos sin cos a]T ;


(3) to determine the inner inverse ray

I A ` A 2(A , n )n ,
(4) to calculate the polar L and asimuth L angles of the inner ray;

tan L I y / I x , cot L I 2 I 2 x I 2 y ;
(5) to determine a point where ray intersects lamp surface, and to find the
luminance by using the indicatrix and obtained values of L and L (Fig. 2.8).
This scheme is included in the program based on the inverse-ray principle;
so, there is no need to prepare tables of luminance distribution over the EM
trace [25]. The EM boundaries arc determined in -cycle automatically. A
point where the luminance becomes equal to zero indicates the boundary value
bnd.
There are some algorithm modifications based on inverse-ray principle, so
the choice has to be done by taking into account different, often conflicting
requirements, such as the universality and the most rapid operation.
To organize the programmers work effectively, it is advisable to have a
set of standardly arranged interchangeable modules that perform similar
functions, c. g. geometric modules that calculate the luminous intensity of a
tube source. Below, we describe the algorithms for computing the luminous
intensity.

Fig. 2.8.
Reading ray luminance
from indicatrix of mercury vapor
fluorescent lamp
(MVFL)

71
1. Discretely discrete variant
The search of flashed area (the source image) is being carried out by
making fixed steps wife respect to and (Fig- 2.9). After passing to the
spherical coordinates in Eqs. (2.38) and (2.40), we find scalar product (a0 , n ) :

(a0 , n ) cos a cos sin a sin cos( ) .


For the surface of revolution the polar angle and the Jacobian H ar the
functions of argument , hence, -cycle should be designed as the most rapid
and internal; while -cycle should be external. By analogy with Eq. (2.40), we
obtain estimate I (a0 )
I (a0 )
cos a H i cos i k sw Lij sin a H i sin i ksw Lij cos( ) .
h h
i
j
i
j
Generally, for arbitrarily positioned source the search along the parallel
=const has to be run from o=0 to k=360. In doing so, double contribution
of equivalent points =0 and = 360 must be avoided.

Fig. 2.9. On determination of flashed area (image)


for symmetrical surface with coaxial cylindric source (a=6, =0)

72
Since the surface of revolution is given, it is advisable to prepare in
advance the tables or arrays which contain the values of Hicosi, and further
just read and interpolate the needed quantities. Scanning along the parallel
=const (Fig. 2.9) with the steady step h enables to design a calculation
program applicable to light sources with arbitrary luminance distribution.
Here we use the concept of filling factor known from the theory of
projector design [26].
Assuming a to be small, we obtain from Eq. (2.38)
I (a ) cos a H i cos i k sw h h .
i

By definition, the filling factor of a zone is

cz sw / .
For discrete series of ksw values, we obtain
kswh ns 1 ,
cz

N 1
where ns is the number of flashed points along the line =const; N is the
total number of steps in -cycle.
Hence, we obtain
I (a) cos a H i cos i czi h .
2. Discretely continuous model
The program will be more accurate if an iteration procedure is applied in
-cycle for searching the image boundaries.
For =const, we find the derivative of I with respect to [Eq. (2.38)]
up

1 I
J (n, ao ) Ld cos Ld
H
l
up

sin a sin L cos( )d .


l

where integration limits are determined in each -cycle.


When all -cycles are performed, the total luminous intensity is

(2.42)

73
determined by summing up all the zonal candlepowers
N

I (a0 )

J H h ,
i

(2.43)

i 1

where N [( k 0 )] 1 ; k and 0 are the boundaries of reflector. When


tracing the rays backwards along the line const with the step h, the
search is being continued until the ray meets the source image. Discovery of
the image edge is indicated by transitions

S (k sw 0) S (k sw 1);
S (k sw 1) S (k sw 0),
where S(...) is a search state.
If we take current value of ksw and its foregoing value kp, then each state of
the search can be coded by a number

k F 2k sw k p ,
while

S (0,1) k F 1;
S (0,1) k F 2;
S (0,0) k F 0;
S (1,1) k F 3.
When the first two cases occur, the procedure of computing the roots of
equation has to be included into the program. Subroutine ZEROIN from
Forsythes collection is the most suitable. The mentioned subroutine is based
on the bisection method, reliability of which is urgent for the case, when
nothing is known about convergence of the process.
In the third and fourth cases the search is being continued until its state has
changed or the upper boundary =360 has been reached. A step, with which
the search is conducted, is selected empirically; it should not be too small,
otherwise the seach will be too time-consuming; and it should not be too large,
or else the target can be missed.
The scheme of computation, when applying Eqs. (2.42) and (2.43), is
suitable for any arbitrarily directed light source with prescribed luminance
distribution. If a light source has axial symmetry, and

74

Fig. 2.10. Flashed area in case of transversal cylinder (a=0, =0)

Fig.2.11. Flashed area in case of transversal cylinder (a=60, =0)

75

Fig. 2.12. Flashed area in case of transversal cylinder (a=60, =900)

symmetry axis coincides with one of coordinate axes, then the image symmetry
should be accounted. Taking a tube which is perpendicular to the symmetry
axis, e. g. for a=0, we obtain the image having two axes of symmetry; these are
1=0 and 2=90 (Fig. 2.10). When 0 and the plane of observation is defined
by up=0, the search has to be performed from =0 up up=180 (Fig. 2.11); for
up=90, from =90 up to up=270 (Fig. 2.12).
For a source with uniform luminance L=L0=1, Eq. (2.42) can be
transformed to the following form
m

J i cos a[cos i ( up( k ) l( k ) )]


k 1

m
(k)
sin a[sin i sin( up
) sin( l( k ) )],
k 1

where the chance that several images m exist along the parallel =const is
accounted.
We note, that the total luminous intensity can be calculated by applying
efficient procedure of numerical integration [27], which

76
allows (when J does not change sharply) to use rather large step of
quantization (see remarks of V.D. Komissarov [24]).
3. Analytical algorithm
For some special cases which have theoretical and practical significance, e.
g. for a surface of revolution with a thin coaxial cylinder of uniform
luminance, the inverse-ray tracing procedure enables to find analytical
expressions for image boundaries at the equatorial plane =0 [28]. The first
model of observation is applied in order to determine the boundary () when
the ray of observation moves along the parallel =const until reflected ray
becomes tangent to the surface of cylinder. The condition of tangency of the
ray (nc , a 0) to the infinite cylinder of radius rc yeilds a square equation in
s=cos:

P2 ( s ) sin 2 a (sin 2 a p 2 sin 2 2 ) s 2 0,5 p 2 sin 2 2a sin 4s


p 2 (1 cos2 a cos 2 2 ) s 2 sin 2 sin 2 a 0 ,
(2.44)
where p = rc/r.
Flashed points in section =const are defined by inequality P2(s)>0 (these
points lie above the axis in Fig. 2.13a). For a source of infinite length there are
two roots of equation (2.44). For a tube with finite length we have two
equations that define the upper up and lower l boundaries in the meridian
plane =0 (Fig. 2.13b)
f up sin( 2 a ) Q sin( 2 a ) 0 ;
(2.45a)

f1 sin( 2 a ) Q sin( 2 a ) 0 ;

Fig. 2.13. On determination of flashed points for


=const in equatorial (a) and meridian (b) planes (- - - <0)

(2.45b)

77
where

arctan(rc / lc ) ;

Q rc2 lc2 / r .

Equations (2.45) are transcedental and can be solved by numerical methods


[27, 29]. An arbitrary point pertaining to a multitute of flashed points (the

image) has to meet the following conditions concurrently


f up ( ) 0 ; f l ( ) 0 ,

(2.46)

where and fup, are calculated from Eq. (2.40).


To complete the description of boundaries we should have carried out a
detailed analysis for the second image (Fig. 2.13a and 2.14). But the symmetry
eases the algorithm. The solution in Eq. (2.44) is mirror-symmetric with
respect to angle , that is
( 2) (a, const ) (1) (a, const ).
(2.47)
For the plane =, the analysis of image equations being similar to Eq.
(2.45) shows (Fig. 2.14> that
( 2 ) (a, ) (1) (a, 0) ,
where (1) is the solution of Eq. (2.45).

Fig. 2.14. Boundary angles of


flashed area for given viewing
direction

78
Thus, the false roots for the second boundary can be excluded by applying
inequalities (2.46) if the sign before is changed to opposite. The following
operator is suitable for this operation
IF(FN(FI).LT.0.AND.(FN(FI) * FW(FI).GT.0)) NROOT = 0,

(2.48)

where NROOT denotes the number of roots for the second boundary. Now we
can calculate the derivative of the luminous intensity at a fixed angle . We
(1)
take Eq. (2.38) with m=2, =0, up
= 0, l( 2 ) =:

J i cos a{cos [ (1) ] NROOT ( ( 2) )]}

sin a[sin (sin (1) NROOT sin ( 2) )] ,

(2.49)
where NROOT=1 if the conditions of Eq. (2.46) are satisfied; and
NROOT=0 in the opposite case.
We cite the main steps of the algorithm for one a-cycle.
Step 1. Using Eq. (2.45), find the boundaries up(0),l/(0). Then in the
cycle for i=up+(i-1)h do steps 2-4.
Step 2. Taking the first root from Eq. (2.44), find the boundary ().
Step 3. After calculating NROOT, go to operator (2.48). If NROOT=1, set
(2) =-(1)(-), then go to step 4; if NROOT = 0, proceed to the next step.
Step 4. Applying Eq. (2.49), find Ji
Step 5. Calculate the integral of luminous intensity [see Eq. (2.43)]
I cos a Hi cos i sin a H i sin i si ,

where ci and si are the multipliers that stand before cos and sin in Eq.
(2.49); the value of Hi can be calculated.
Organization of Programs
The inverse-ray method enables to reach high flexibility of application
package, where each separate module (subprogram) can be used in different
programs. For example, the geometrical module PIPE, that varifies intersection
between a ray and cylindric source, can be used in described-above searching
algorithm. Since this module is addressed each time the inverse ray is traced, it
should be specialized for a certain source type (uniform tube, uniform sphere,
metal-halide lamp, mercury vapor lamp, etc.)

79
We consider three programs for calculating luminous intensity, which are
based on inverse-ray method: TASDD, TASDN, and TASAN (The Tube:
Axially Symmetrical System Discretely Discrete Variant; Discretely
Continuous Variant; Analytical Variant). Specialized program for symmetric
optical system shows a better performance than the universal program. This
fact is clearly seen from Table 3, where programs TASAN and TASDD are
compared.
Thorough logic of TASDN allows to use larger step of quatizalion h
without loss of accuracy but with gain of time. Furthermore, this leads to small
increase (by 10 KByte) of required memory. The program TASDN in
comparison with TASDD uses more simple formula (2.38) for uniform
sources, but, as it was noted, the likely scheme can be used for nonuniform
sources as well.
Table 2.2
Characteristics of programs for calculating luminous intensity produced
by optical system with a tubular light source
Algorithm
TASDD
TASDN
TASAN

lime of calculation, rcl.


val.
3.44
1.73
1.00

Memory volume, KByte


56
68
64

Cylindrical system
The scheme of calculation used for surfaces of revolution can be easily
extended onto other surfaces, for instance, onto cylindrical surface. Arithmetic
and luminance program modules can be used without changes. Every point of
cylindrical reflector is assigned by coordinates , x (Fig. 2.15), where the linear
variable x plays the role of .
If the profile curve is specified, say r=r(), then the expressions for n and
H take the following form
T
n 0 sin cos ;
H r sin cos 1( ),

80

Fig.2.15. Flashed area of cylinrical reflector

where the angle of a normal is defined by equation

tan( )

1 dr
.
r d

Applying Eq. (2.32), we obtain the expression for luminous intensity


up

xup

I (a0 ) (n, a0 )d L( , x)dx.


l

(2.50)

xl

Substituting the expression for the scalar product ( n , a ) into Eq. (2.50), we
get
up

xup

I (a0 ) cos a H cos d L( , x )dx


l

xl

up

xup

sin a cos H sin L( , x)dx,


l

(2.51)

xl

where 1 and up are the lower and upper boundaries of an image in the
transversal plane.
We find now the equations for the flashed area at the transversal plane =0
when uniform tubular source is used. Assuming that a source image lies within
the outline dimensions of reflector, and considering Fig.2.15, we obtain the
equations for calculating distances

81
H1, H2 and the difference H:

01 1;
02 2 ;
H1 r1 sin(1 a );
H 2 r2 sin( 2 a );
H H 2 H1 ,

(2.52)

where 01, 02 are the orientations of axial rays at the points 1 and 2 .
The first two equations in (2.52) prescribe the points 1 and 2 implicitly
and can be solved by numerical approximate methods (note, that =arcsinR/r).
Having found 1 and 2, we proceed to determining r1 and r2, and then H1, H2
and H.
The inverse-ray method is the most suitable for calculating the spatial
characteristics of luminous field created by specular reflectors [1, 30, 31]

E p L( A ) ( A )d ,

where L( A ) is the luminance of reflector along the ray A (Fig. 2.16);

( A ) is the value function at the direction A ; is the solid angle associated


with reflector flashed area seen from the point P (Fig. 2.16a). Obviously,
similar expression can be obtained for Ev, if N p is assumed to be the unit
vector being in parallel to OY (Fig. 2.16b). Spatial irradiance can be defined as
the irradiance of a plane being perpendicular to the axis of elementary beam, i.
e., when N p = A (Fig. 2.16c).
For horizontal irradiance

( A ) N p , A ,
where N p is the unit normal vector to the horizontal plane (Fig. 2.16b).
For automated calculation of Eh produced by cylindrical reflectors we use
the following expression (see Fig. 2.16a)
xmax max

Eh

x0

A sin Ay cos
r d dx
L( A ) x
( A, N p )
. (2.53)
2
Rp
cos( )

82

Fig. 2.16. On determination of spatial characteristics of illuminance produced by specular


reflector

Fig. 2.17. Diagrams of horizontal illuminance produced by parabolic-cylindrical reflector


(a) and luminaire SGS01 (b)

83
The luminance L( A ) is determined by inverse-ray scanning. Figure 2.17
shows the pattern of horizontal illuminance produced by a paraboliccylindrical reflector and specially designed irradiator, both operating with the
metal-halide lamp. The former provides a field of high intensity over small
area of 0.8 m2, while the latter spreads the flux over area of 7.8 m2.
2.4. INVERSE-RAY METHOD IN ANALYTICAL REPRESENTATION
Determining flashed area boundaries of a symmetric narrow specular zone
It would be useful to describe the inverse-ray method by considering
narrow specular zone and its flashed area. The inverse-ray calculating
procedure consists of the following stages:
(1) generating inverse ray in the inner space of the optical system;
(2) finding the point of intersection between a ray and a light source.
We present the law of reflection in the following form
a 2n (a0 , n ) a0 .
(2.54)
We observe a motionless mirror element while moving in the outer space
within a cone having the aperture of . The mobile vector of the ray of
observation is

a0 sin a sin

sin a cos

cos a .
T

The normal vector to the mirror element is

n [0 sin

cos ]T .

The matrix of reflection corresponding to Eq. (2.54) takes the following


form

0
1

0 cos 2
0
sin 2

M refl
sin 2 .
(2.55)
cos 2
We define the coordinates of the inverse ray a [ ]T by using matrix
product a M refl a 0 :

84

sin a sin ;
cos 2 sin a cos sin 2 cos a ;
sin 2 sin a cos cos 2 cos a.

(2.56)

The action of matrix (2.55) can be visually represented as imaginary


rotation by 180 of the cone of observation about the normal to mirror element.
As a result of this rotation each reflected (in direct tracing) ray matches with
the corresponding incident ray (Fig. 2.18). This rotation can be represented as a
superposition of two rotations:
(1) rotation about the axis OZ by 180;
(2) rotation about the axis OX by 2 (Fig. 2.18).
These rotations are described by matrices:

1 0 0
M 1 0 1 0;
0 0 1
0
0
1
M 2 0 cos 2
sin 2 .
0 sin 2 cos 2
The general rule for composing matrices: each column of a matrix contains
the coordinates of the corresponding new basis vector in the old basis. It is
easy to prove that
M refl M 1M 2 .
Let us set up the task of determining the flashed area of a narrow mirror
zone at =const. Obviously, the boundary points of this zone correspond to
viewing rays that are tangent to the surface

Fig. 2.18. Specular reflection as superposition of two rotations: a - in plane XOY; b - in


plane ZOY.

85
of a light source. We consider some specific cases that have practical
significance.
Cylindrical source with uniform luminance
Equation for the tangent plane that contains the vector a and passes
through the point A (Fig. 2.19) is
yo sin r rc ,
or
y02 2 rc2 ( 2 2 ).
(2.57)
We transform the expressions (2.56) by introducing the following
substitutions

sin

2t
1 t 2

;
sin

; t tan .
2
2
1 t
1 t
2

After transformations we get

2t
;
1 t2
k1 k 2t 2
sin a
;
1 t2
k k t2
sin a 3 42 ,
1 t

sin a

where

k1 sin( 2 a ) / sin ;
k2 sin( 2 a) / sin ;
k3 cos(2 a) / sin ;

(2.58)

k4 cos(2 a ) / sin ;
Fig. 2.19.
Tangency condition for ray with
respect to cytindrie source

86
Substituting these expressions into (2.57), we obtain

2t k 1 k 2t 2 ,

(2.59)

where

k 1 k 1 tan ; k 2 k 2 tan ;

tan r

rc
y02 rc2

Equation (2.59) has the roots *

tan

i
ki

2
1 1 k1k 2 .

(2.60)

i 1, 2

Where the angles 1 and 2 are counted off from the opposite sides of the
axis OY. If k 1 k 2 1 , then 1=2=180; therefore, the narrow mirror zone is
completely bright. We find the angle cr that fits this condition

k1 k 2 1 ,
sin( 2 a) sin( 2 a) sin 2 a / tan 2 r .
Transforming this equation, we obtain (sec Fig. 2.20)

sin cr (rc / r )(sin 2 / sin ) .

(2.61)
This nontrivial result can be obtained by applying geometrical conceptions
of the method of elementary maps (EM).
For cr, if an arc-tube is long enough, all the elements of narrow
specular zone are bright once at a time though they may have nonuniform
luminance. We may call it the zone of complete glow. When >cr, the image
in the mirror zone gels broken no matter how long is the luminous cylinder.

Fig. 2.20. Calculating cr wilh the aid of spherical triangle.

It is easy to prove that Eq.(2.60) is equivalent to Eq.(2.44)

87
If we set =/2 (the zone of paraboloid having its focus at the centre O),
the previous formula yields
sincr=rc/r.
(2.62)
Note that this formula is cxact (compare with [26]).
Using geometrical terms, we can simplify all considerations accounting the
fact that imaginary rotation of the EM by 180 about the normal of mirror
element leads to coincidence of corresponding reflected and incident rays (Fig.
2.21a). After rotation, the planes which are tangent to the EM match as well
(Fig. 2.21b). If we pay attention to the rotated axes of the grid, we find out that
the tangent planes cut the intercept lan2 out of the axis OX, and the intercept
sin2 tany out of the axis OY.
Straight-line equation in segments on the plane X'O'Y' being the map of the
grid plane is as follows (see Fig. 2.21b)

Fig. 2.21. Principle of correspondence in specular reflection: a - congruent beams AB and


A'B' match after rotation by 180 about normal; b - rotation of cone of observation into internal
space; plane XOY transforms into X'O'Y'
(OS is tangent plane to cylinder; OS=sin2tan; O'B= tan2).

88

X`
Y`

1,
tan 2 sin 2 tan
or

Y ` (tan 2 X `) cos 2 tan r .

(2.63)
The cone of observation with aperture produces a circular trace of radius
tdna on the plane X'Y'. Multiplying Eq. (2.63) by cot, we obtain for the
intercepts on the axis Y'

Y `i cot (sin 2 cot cos 2 ) tan r

sin( 2 )
tan r .
(2.64)
sin

i 1, 2

We denote 1=2-, 2=2+. Sines ratios are known* to be the


proportionality factors between incident and reflected luminous intensities of a
point light source (Fig. 2.22):

Fig. 2.22. Angular dimensions 1 and 2 of cylindrical source image on surface of narrow
specular zone: 1 and 2 are point sources being characterized by factors of amplification k1 and k2,
respectively.

In accordance with the law of enegy conservation, considering pmir= 1.

89

sin 1
sin 2
, k2
.
sin a
sin a
Values k 1 and k 2 from Eq. (2.58) get
k1

the following interpretation: they are the


tangent coordinates of the EM trace on the
plane of projection (Fig. 2.22).
Luminous filament may be considered
as a collection of elements z. In case of
cone-shaped zone, factors ki are common
for all elements. Obviously, the zonal
luminous intensity subordinates to the law
of 1/sin*.
Here the performance of a specular Fig. 2.23. Source images within internal surface
of narrow specular zone and flat band.
zone, being a curvilinear surface
combined with a thin light source, appears
most evidently.
We consider the images of a thin cylindric source in the specular zone of
reflector and in the flat specular band having the same width and inclination as
a reflector zone (Fig. 2.23). The width of the source image in the flat band is
equal to the width 2rc of the source itself, the width of the image produced by
curvilinear surface is equal to 2y0 sin. For a thin source we obtain
approximately (Fig. 2.24)

Fig. 2.24. Possible locations of viewing cone and source image: a - image of ring with gap;
b - image of continuous ring.

This explains the tendency for sharpening and unsteadiness of luminous intensity curve
produced by reflector with linear sources.

90
that is, the source image lengthens in k1 times in comparison with that of (he
flat mirror; reflected luminous intensity increases in k1 times as well. While an
arc-tube is thin, the luminance pattern extends almost evenly. Now it is easy to
obtain from Eq.(2.64) an exact equation with respect to . Just note only that
the points, where a straight line intersects a circle (-parallel) of radius tan, fit
the condition
x` tan cos , y ` tan sin .
(2.65)
Excluding x' and y' from Eqs. (2.64) and (2.65), we obtain the quadratic
equation (2.59).
It is clear from geometry that for a=acr a secant from Eq. (2.63) turns into
a line being tangent to a circle, and k 1 k 2 =1 (Fig. 2.24).
Image dimensions in meridional plane =0 or =0 can be determined if
we consider Fig. 2.22. Placing the origin O at the middle of a source with
length of 2lc, we find

lc
sin 2 0,
r
l
sin( 1 ) c sin 1 0.
r
sin( 2 )

(2.66)

Implicit equations (2.66) define


image boundaries 1 and up, on the
positive side of the axis y. Similar pair of
equations can be obtained for negative a.
These equations, together with
additional
equation
sinim=k sinr,
completely define the image of a thin
source in angular coordinates ,.
Sometimes, instead of angular
coordinates, it is suitable to operate with
Cartesian coordinates on the plane being
orthogonal to the line of observation, so
to say pictorial plane. We rotate the
coordinate system about the axis OX so,
that the axis OZ matches with the line of
Fig. 2.25. Projection of source image onto observation (Fig. 2.25). The matrix of
pictorial plane
rotation is as follows

91

0
0
1
0 cos a sin a .

0 sin a cos a
The coordinates x', y' of a point in
pictorial p lane Q (Fig.2.25) are represented
in terms of x, y coordinates in the old system:
x`=x ;
y`=y cos-z sin
The rectangular coordinates of an Fig. 2.26. Exit aperture of reflector with
cylindrical source projected
image for a thin source model are

yim i

xim 2rc k ;
ri sin( i a ),

i 1, 2

(2.67)

onto pictorial plane


1 - image of lower edge of reflector;
2 - image of upper edge;
3 - image of source;
4 - image of flashed area.

where 1 and 2 are the lower and upper boundaries that are determined by
applying Eq. (2.66). By using Eq.(2.67), the contours of source image and also
the lower and upper edges of reflector can be projected on a plane (Fig. 2.26).
The luminous intensity of reflected beam for given viewing angle is defined
by the flashed area that is not shaded by a light source.
We estimate an error associated with described thin source model. Wc
rewrite Eq. (2.63) in the following form

sin k sin r cos 2 (1 1 sin 2 )

sin r
1 sin 2 r

or, denoting sin/sin = u,

u 1 sin 2 k cos 2 (1 1 u 2 sin 2 r ) ,


Applying expansion of radicand into series, we obtain

u1 1 sin 2 r ... k cos 2 0 u 2 sin 2 ....


2
2

We express the variable u in terms of series with small parameter:

u k m sin 2 r

92

and find the correction m:

cos 2 2
1

(k m sin 2 r )1 sin 2 r ... k


(k ...) sin 2 r ,
2
2

2
k k
whence m
cos 2 . Hence,
2
cos 2 2
uk
(k k ) sin r ;
2
k k2
sin k sin r cos 2
sin 3 r .
2
Application of formula sin = k sin leads to relative error

1 k
cos 2 sin 2 r .
2

(2.68)

Graphs k (, ) show limited resources for enhancement of luminous


intensity from circular specular zone. The maximum of amplification factor is
reached at =45+/2 and is equal to 1/sin. So, we may consider that with
increase of a amplification factor descends according to the law of 1/sin
(Fig.2.27). For =/2 the value of k becomes zero; in this case the reflected
beam (in backward direction) runs in parallel to the optical axis without
intersecting a light source.
In the first approximation the graph k [()] depicts a contour of a flashed
area for the chosen angle of
observation a (when >cr).
Fig. 2.27. Family of amplification
factor functions:
1 =200;
2 =300;
3 =400;
4 =500;
5 =600;

93
Ellipsoidal light source
Ellipsoid is specified by equation
(Fig. 2.28)

x2 y2 z 2
2 1 . (2.69)
b2
a
The ray of observation is assigned
by the following equations

x 0 ;
y y0 ;

Fig. 2.28. On deduction of equation for


enveloping cone at point of mirror
element A(0, yo, zo); a is viewing vector.

z z0 .

The condition of intersection between a ray and a surface from Eq. (2.69)
yeilds

y02 2y0 2 ( 2 2 ) z02 2z0 2 2 )

1,
b2
a2
or
2
2
2 2 2
z0 y0 y0 z0

2 2

1 0.

2
b 2
b2 b2 a2
a
a

We find the discriminant of quadratic equation


2

2
z 2 2 2 2
y z y
D 02 0 2 } 02 02 1
2 .
2
a b
a
a
b
b

The expression D=0 is the equation of wrapping cone, where , , are


current coordinates of a cone with the origin at the point (0 yo zo) (Fig. 2.28).
Intersection of this cone with the cone of observation rotated by the angle 2
gives the boundary angles of the flashed zone. We put the wrapping cone
equation in the following form
2

2
2
2
y0 z0
2

b2
a2
a 2
b

(2.70)

94
where

K 2 y02 / b 2 z 02 / a 2 1.
We calculate the right-side value from Eq. (2.70), substituting Eq. (2.58)

2 2 2
sin 2 a 4t 2 (k1 k2t 2 ) 2 (k3 k 4t 2 ) 2

.
2

b2
a
(1 t 2 ) 2
b2
a2

Setting up y0=r sin, z0 = r cos, we calculate the left side


2
y 0 z 0 r sin a sin ( k1 k 2 t 2 ) cos k 3 k 4 t 2
2

.
b2
a
1 t 2
b2
a2

Substituting found expressions into Eq. (2.70), we obtain the quadratic


equation in u= t2:
Au 2 2 Bu C 0,
(2.71)
where

A k 22 f1 k 42 f 2 k 2 f 3 k 4 ;
B k1k 2 f1 k 3 k 4 f 2 0,5k1k 4 k 2 k 3 f 3 0,5 f 4 ;
C k12 f1 k 32 f 2 k1k 3 f 3 ,
and here, in its turn

f 1 cos 2 0 a 2 / r02 ; f 2 sin 2 0 b 2 / r02 ;


2
a 2
a
2
2
f 3 sin 2 0 ; f 4 4 sin 0 cos 0
b
r0

Since u2=tan2(/2), the positive solution of Eq. (2.71) is chosen. The points
that define the boundaries of flashed area of a narrow zone can be found by
solving

tan( i / 2) uo

i 1, 2

Let us consider specific cases:


(a) Extremely oblate ellipsoid (disk), a=0. We obtain for this case: *

Note that direct application of the limit 0 in Eq. (2.71) is impossible.

(2.72)

95

Fig. 2.29. Marginal flashed points of specular zone in meridian section for spherical (a) and
disk (b) sources.

f1 cos 2 , f 2 sin 2 (b / r ) 2 , f 3 sin 2 , f 4 4 cos 2 ;


(b) Sphere (a = b):

f1 cos 2 (a / r ) 2 , f 2 sin 2 (a / r ) 2 , f 3 sin 2 , f 4 4[1 ( a / r ) 2 ].


We test the obtained formulas by examining the flashed area from axial
direction a=0. In this case, the luminosity of a zone is determined by the sign
of discriminant

D [(a / r )2 sin 2 2 (b / r ) 2 cos 2 2 sin 2 (2 )].


For a spherical source the luminosity of a zone is determined by the sign of
discriminant (Fig. 2.29a)
D (a / r )2 sin 2 (2 ).
(2.73)
For a disk source (Fig. 50b)

D b 2 cos 2 2 r 2 sin 2 (2 ).
Taking paraboloid for reflector and matching its focus with the centre of a
light source, we obtain, obviously, a completely bright zone, i.e. D>0.

96
Light beam of parabolic projector with spherical sourcc
(1) We find the luminous intensity produced by a paraboloid when a light
source is placed in its focus. The equation for a paraboloid with the focus
situated in the origin of coordinate system is
x 2 y 2 4 f ( f z ).
(2.74)
Equation for the normal becomes
n [2 x 2 y 4 f ]T / M ,
2

where M 2 x y 4 f .
Inverse ray is assigned by the vector

a0 [sin a 0 cos a ]T .
The inner ray has a direction defined by the vector

a [2 xQa sin a 2 yQa 4 fQa M 2Qa ] / M 2 ,

(2.75)

where Qa x sin a 2 f cos a.


Obviously, after reflection the axial ray a0 [0 0 1]T runs at the direction
ax/az=x/z, i. e. passes through the origin of coordinate system (the optical
property of parabola). We suppose that the light source has spherical form with
a small radius R, where R/f< 1.
If viewing angle is fixed, then applying Eq. (2.73), we obtain the equation
for the tangent inner ray
sin=R/r.
(2.76)
Keeping in mind the parabola properly r=2f-z, and applying Eq. (2.76), we
find z-coordinate of the boundary bright point of reflector
z 2 f R / sin a .
(2.77)
Substituting Eq. (2.77) into paraboloid equation (2.74), we obtain the
equation for cylinder
x 2 y 2 4 f 2 ( R / af 1) .
Hence, the luminous intensity at the direction is

I (a) 4Lo f 2 ( R / af 1) 4Lo f 2 (amax / a 1)


where L0 is the luminance of a light source; max=R/f.

97
Since the luminous intensity is nonnegative and max is the total angle of
radiation, then I(max)=0.
Since D>0 in Eq.(2.73), the lighting aperture is seen totally bright in the
range of 0min=R/rmax. Hence, the expression for calculating the luminous
intensity of paraboloid with a small spherical light source having the
luminance L0 is as follows [26]

tan 2 ( max/ 2),


I (a )

amax / a 1,
2
4Lo f

0,

0 a amin ;
a min a amax ;
a amax ,

(2.78)

where rmax f / cos 2 ( max / 2) ; 2 max is the acceptance angle of reflector.


It can be easily proved that at the point a=amin the connection of functions
is provided.
(2) We find the luminous intensity of a paraboloid at rather remote, but
still finite distance from reflector [32],
Let xo, zo be the coordinates of a fixed point of observation, and
x0 / x 1, z 0 / z 1; x, z the coordinates of paraboloid.
The inverse ray is defined as follows

a0 [ x0 x 0 z0 z ]T / M 1 ,
where M 1

( x x0 ) 2 ( y y0 )2 .

The inner reflected ray emerges an angle with the axial ray
sin a tan a ( x0 x ) /( z0 z ) .
(2.79)
Substituting Eq. (2.79) into Eq. (2.77), we find the coordinates of the
boundary bright point of reflector
z 2 f R( z0 z ) /( x0 x ) .
(2.80)
Expanding the denominator from Eq. (2.80) into series in small values of
x/xo, we get

z 2f

R
x
(1 ) ,
a
x0

(2.81)

where a x0 / z0 .
Substituting Eq. (2.81) into equation of paraboloid, we obtain
2

R
R
x 2 f
y 2 4 f 2 1 2 2 .
ax0
a x 0

af

(2.82)

98
This is the equation of cylinder shifted from the axis by 2 fR /( ax0 ) . The
luminous intensity depends on the area of cylinder from Eq. (2.82)

a
R2

I 4f 2 Lo max 1 2 2 ,
a x 0
a
where amax= R/f.
Equation (2.83) presents an expansion of I(a) into asymptotic series in
powers of l/x0. When x0 it turns into Eq. (2.78), i.e. the expression in
parentheses is the null element of series.
The procedure of calculating the luminous intensity at remote distances
was comprehensively developed by F. Benford and N. A. Karyakin.
Candlepower distribution of paraboloid with a light source placed in the focus
is a function of distance r from reflector. The distance of projector light-beam
formation l0 is such, that the law of square distances can be applied.
We may assume approximately
lo xmax / amin ,
(2.84)
where xmax is the radius of mirror.
Equation (2.83) can be presented as follows
I I 4 f 2 Lo R 2 / a2 x02 ,
(2.85)
where additional term defines an error associated with application of the law of
square distances. The relative error in terms of the axial luminous intensity is

R2
.
tan 2 ( max / 2)a 4 z 02

(2.86)

The error is maximal for =min. Combining min=R sinmax/xmax with Eq.
(2.84), we obtain
2
amin

R sin max
l0

Equation (2.86) now transforms into

l02
4 sin 4 ( max / 2) z02

(2.87)

If we set the permissible error for photometrical measurements =per, then


a desirable distance of photometrical testing can be

99
obtained

z per

l0
2

2 sin ( max / 2) per

(2.88)

For max= 90, we get a simple formula

z per

l0
.
per

(2.88)

Equations (2.87), (2.88), and (2.89) can be used to test the programs of
computing illuminance distribution produced by specular reflector.
2.5. CALCULATION OF ILLUMINANCE BY WIENERS SCHEME
According to Wieners scheme, the
elementary illuminance at the point of
observation A is calculated as a light-vector
projection on a small area that contains the
said point (Fig. 2.30) [33, 34]
d 2 E ( L d )n ,
(2.90)
where L is the luminance of a ray hitting a
small area; n is the normal to a small area;
d is the vector of a solid angle having a
vertex at the point A.
Fig. 2.30. Determination of illuminance
in accordance with Wiener's scheme
Dividing the space on elementary
solid angles and summing up their (illuminance at point A is proportional to
d cos).
contributions [Eq. (2.5.1)], we find the
resulting illuminance at the viewing point

E Ld n

(2.91)

We assume the system to possess axial symmetry and direct OX along the
symmetry axis. We bring the point A to the origin of the polar coordinate
system. Without loss of generality, the following can be set up

100

n s [cos n 0 sin n ]T ,

d [cos a sin a sin sin a cos ]T sin a da d ,


where n is the polar angle of a normal at the point A; , are the spherical
coordinates of a ray.
(a) We set n=0, then the x-th illuminance projection (horizontal
illuminance) is

Ex

1
L sin 2a da d ,
2

(2.92)

where (, ) specifies the limits of integration related to the dimensions


of a domain percepted as bright from the viewing point A.
(b) We set n=/2, then the z-th illuminance projection (vertical
illuminance) is

Ez L sin 2 a cos da d .

(2.93)

Hence, the problem is reduced to finding the limits of integration. It can be


solved by using the inverse-ray method.
Calculation of illuminance produced by elliptical reflector
Elliptical specular reflector with a light source placed in its focus is widely
used in radiant-heating installations, simulators of thermal flux, and radiative
furnaces. For example, the calculation of the thermal field produced by a
furnace with a spherical xenon lamp is considered in [31].
If a point source is placed at the focus of ellipsoid, then, according to the
ray theory, the caustics point with infinite illuminance occurs at the second
focus. In case of a source with finite dimensions, an enlarged source image
corresponding to some illuminance distribution appears in the second focus.
Observing the reflector from chosen point, we see flashed only such mirror
points that reflect inner rays towards the source. Flashed area corresponds to
the solid angle which is cut on a surface of unit sphere surrounding the point
on a target plane (Fig. 2.30).
Algorithm of computing the illuminance consists of the following stages:
(1) formation of inverse-ray vector
a0 [cos a sin a sin sin a cos ]T ;

101
(2) calculation of coordinates of a point where the ray intersects an
ellipsoidal surface, and of an inner ray as well;
(3) testing whether the inner ray intersects a light source, and calculating
the ray luminance;
(4) calculation of integrals of illuminance from Eqs. (2.92) and (2.93).
We consider algorithm of computing horizontal illuminance Ex at the focal
spot of ellipsoid when a spherical source with uniform luminance is used.
Stages 2 and 3 can be realized by applying formulas described in Ch. 1. All
stages are organized in program cycles: the inner cycle with respect to angle
with a step h, the outer, with respect to angle a with a step ha. The choice of
step ha, that provides necessary accuracy, is important. Sincc boundary rays arc
of the most luminous value, inaccuracy in determining the flashed area
boundaries may cause significant errors. Searching algorithm must be
organized in two stages, similar to that, used for solution of equations; first, to
separate the roots, and then to refine them by applying numerical methods. In
this connection, there exists the inner procedure which identifies whether a
point is inside (ksw= 1) or outside (ksw=0) the domain.
The boundaries of flashed area can be determined by applying the
bisection method until the needed accuracy is reached. We note, that while
determining a-boundaries in a-cycles, -boundaries are not refined, and the
search is being continued until the first bright point is found (ksw=1).
Figure 2.31a (curve 1) shows the illuminance distribution at the focal plane
of elliptical reflector. The reflector has the following parameters: the interfocal
distance 2c=992.4, the focal parameter p=175, the diameter of blind aperture
Db=350, the acceptance angle 2max = 180. The diameter of spherical source
is 2R=2 (all dimensions are given in milimetcrs). To provide more uniform

Fig. 2.31. Modelling of elliplic reflector with spherical source: a - irradiance distribution in
focal spot (1) and in exit aperture of lightguide (2); b - indicatrices for ideal ellipsoid (3) and
system with spherical source (4).

102
illuminance distribution from elliptical reflector, a lightguide is placed in its
focus.
Figure 2.31 (curve 2) shows illuminance distribution in transversal section
of the lightguide with the radius rs =10 mm at the distance of ls=80 mm.
The indicatrix (candlepower curve) is invariant characteristic of the
lightguide. Computation of indicatrix is carried out in the same cycles as in the
program for illuminance. The luminance distribution L(, z0) within the rays
cone of a=const is being found in -cycle for the point z=z0.
The averaging along gives a value of filling factor (as per the second
model of observation), or

ka ( z 0 )

La ( , z0 )d .

The elementary luminous intensity produced by the circular ring dz is


equal to dI(a)=2 kazo dz0, while the total luminous intensity is determined by
integration along lightguide radius:
rS

I (a) 2 La ( , z0 )d dz0 .

(2.94)

0 0

Figure 2.31b shows the indicatrices, both for an ideal ellipsoid and a
system consisting of elliptical reflector and lightguide.
Calculation of illuminance produced by reflector with plane elements and
Lambertian source
In this case, the luminance L being a constant value can be removed
beyond the integral sign in Eqs. (2.91)-(2.93), thus the illuminance calculation
is reduced to integration along the contour formed by the rays on the unit
sphere

E L dn

(2.95)

This expression can be put in equivalent form

E L / 2 da n

(2.96)

where da is the vector of the planar angle formed by the side-surface


gcneratrixcs of a solid angle cone. For the ease when the contour is specified
by a discrete series of rays wc expand the expression

103
for the solid-angle vcclor in lernis of the planar-angle vector.
Let two close rays have the polar coordinates (1 , 1), (2 , 2). We form
the vectors:

a1 [sin a1 sin 1 sin a1 cos 1 cos a1 ]T ;


a2 [sin a2 sin 2 sin a2 cos 2 cos a2 ]T ;
The vector vp=a1 a2 has the direction da and the following components:

vp1 sin a1 cos a2 cos 1 sin a2 cos a1 cos 2 ;


vp2 sin a2 cos a1 sin 2 sin a1 cos a2 sin 1;

(2.97)

vp3 sin a1 sin a2 sin( 1 2 ).


The absolute value of vp is equal to sin(da), i. e.

da arcsin( vp )
2

(2.98)

where vp vp1 vp2 vp3 .


Hence, the light vector can be found by summation

E 0.5 L (vpi / vp )dai ,

(2.99)

where the summation is carried out for each i-th pair of close rays.
Since the mirror is flat, the position of source image is exactly determined.
This enables to find angular dimensions of the image when the point of
observation is prescribed.
We consider an example of calculating the illuminance from a trough
reflector with Lambertian ellipsoidal light source (say, MVFL). The reflector
consists of planar plates laid out along the generatrix (Fig. 2.32).
In coordinate system of the source image the latter is defined by equation

X 2 Z2 Y2
2 1
b2
a
In the same coordinate system the position of observer is xo, yo, zo. In
accordance with the procedure described in paragraph 2.3, wc find the equation
for wrapping cone
2

x0 X z 0 Z y0Y 2
Y2
2
2
2 2

K
X

Z
b

b2
a 2
a 2

(2.100)

104

Fig. 2.32. On illuminance calculation from plane element by Wiener's scheme.

105
where K 2 x02 y02 z02 1 ; x 0 x0 / b; y 0 y0 / a; z 0 z0 / b.
Assuming X X / b; Y Y / a; Z Z / b, we get the equation in new
terms (superposing the cone vertex with the system origin)

K 2 X Y Z x0 X y0Y z 0 Z 0

(2.101)
We introduce the plane of triangle: an observer a horizontal element of
reflector. The equation of the plane is Z m X . To find the boundary rays in
the plane, we consider the intersection between the plane and the wrapping
cone. We set up

X r 1 ; Z r m.
(2.102)
The third coordinate can be determined from the following equation
2

K 2 (1 m 2 Y r ) ( x 0 m z 0 y 0 Y r )2 0 ,

(2.103)

the solution of which yeilds

Y r y0

x0 mz0
2

K 2 (1 m 2 ( z 0 m x 0 ) 2
2

x0 z0 1

(2.104)

x0 z 0 1

Equations (2.102)-(2.104) present the coordinates of boundary rays. For


special case of distribution along the axis y 0 0 we obtain the following
equations with the preceding notations (without bar)

X r bK ;
Yr a 1 m 2 ( z 0 m x 0 ) 2 ;
Z r mX r ,
2

(2.105)

K x0 z 0 1
The condition of discriminant positiveness in expression for constant K
means that the viewing point is outside the ellipsoid. Really, from the condition
|x|>b or |z|>a immediately follows that ( x02 / b) ( z02 / a) 2 1 . Obviously, the
range of possible m-values is determined by the condition that the radicand in
expression for Y is equal to zero.
Solving the quadratic equation, we obtain
2

m1, 2 ( z0 x0 K ) /( x 0 1) ,
where K is the same as in Eq. (2.105).

(2.106)

106
In order to find the angular coordinates of the image rays we have to turn
into the coordinate system of observer. This can be done by multiplying the
vector [Xr Yr Zr]T by the matrix of reflection (see the description of the inverseray method), where the latter is defined by the position of the normal to the
planar element.
We omit the details, supposing that the reader can restore them easily. We
only note that in this specific case n [cos 0 sin ] , where is the
inclination angle of the planar mirror to the axis
Now, since xr, yr, zr are the ray coordinates in the observers system, the
angular coordinates are

tan r yr / xr ;
tan ar xr2 yr2 / z r .

(2.107)

To find the illuminance we apply Eq. (2.99) or Eqs. (2.92) and (2.93).
When composing a program, one has to take into account that the planar
specular clement plays the role of diaphragm for the observer. Since element
bounds form a convex envelope on the unit sphere, it is not difficult to design
an algorithm that records if a ray hits within the prescribed domain '. After
choosing the direction of going around the integration contour, we register
whether a ray occurs within the domain ' (Fig. 2.33). The first and the last
hittings extract an active contour C2. It is closed by the segment C1 lying inside
the source image (Fig. 2.33).
Similar scheme is applicable for calculating
the illuminance produced by reflector with other
Lambertian sources (e.g. metal-halide lamp). In
this case the condition a must be set in Eqs.
(2.100)-(2.104), thus eliminating the components with Y and y0. The finite dimensions of the
source along the Y-axis have to be accounted in
final expressions that define the boundary rays.
Say, yr 0.5 l , where l is the source length.

Fig. 2.33. Going around flashed


area contour of element

107
2.6. THE METHOD OF ELEMENTARY MAPS
The melhod of elementary maps (EM) is based upon the following
assumptions [11, 25, 26 ]:
(1) a source radiates in accordance with Lamberts law;
(2) the luminance is evenly distributed or slightly varies over source
surface (or its substituted, in the latter case the surface can be divided in a
series of small areas with uniform luminance;
(3) the EM trace on a tangential plane is described by a simple geometrical
figure (rectangular, ellips, etc.) [35].
The first two assumptions show that a direct ray travel is used within the
EM method.
According to the EM method, a point M of reflector is considered to be
bright when viewed from direction I if EM associated with this point contains
a ray traveling along said direction.
To carry out flashed points extraction it is more handy to operate with a
plane. This leads to rather simple equations of EM traces on the plane
(assumption 3). The choice of curvilinear coordinate plane, where further
operations with EM are being held, is a complicated problem [36].
In Ch. 1 we found out that regular reflection converts any linear
combination of vectors into linear combination of their images. Hence, if EM
is represented by a discrete set of vectors {S1, S2,,Sn}, then a trace of EM on
the tangential plane is a polygon <A1, A2,,An>, where

Ai M refl Si , i 1, n .
It is easy now to describe a procedure of determining whether an arbitrary
point of target plane is inside or outside the EM trace.
The stages of calculating the illuminance by using the EM melhod are as
follows [37]:
(1) description of geometrical properties of a light source;
(2) description of EM;
(3) quantization of EM;
(4) quantization of optical surface:

i 0 (i 1)h
i ( j 1)h

i 1, N
i 1, N

0 i max ;
0 i 2 ;

108
(5) quantization of target plane:

xk (k 1)hx , k 1, N x

X min x X max ;

yl (l 1)hy , l 1, N x Ymin y Ymax .


(6) definition of EM trace on target plane;
(7) testing whether grid nodes xk, yi are inside EM trace; if NOT, then
jump to the next point or node i, j; if YES, then record into an array the
i

weight factor W k l , which is a projection of solid angle w` on the target plane;


(8) calculation of illuminance integral
ekl
Wki l .

Description and Quantization of Elementary Map


We consider, e. g. the EM of ellipsoid. First, we have to find an equation
of enveloping cone for the light source. The cone vertex is at the point A(x0,yo,
zo). A unit vector of the ray AS or AS` may be defined as
p [sin a cos sin a sin cos a ]T ,
where , are the polar and azimuth angles for the mirror point; AO is the
axis.
We determine the condition of intersection between a ray and ellipsoid
being defined by the following equation

( x 2 y 2 ) / R 2 z 2 / q 2 1.
For ease we turn to affine space having presented the transformation in the
form given in Table 2.3.
Angular dimension of EM is defined by the formula
sin ( 02 02 ) 1 / 2 .
Axial ray vector, obviously, is

p `0 2
0
0 02

.
02 02

109
Table 2.3
Cartesian space
Point x, y, z
Angles ,
Ellipsoid

x 2 y2 z2
2 1
R2
q

Affine space

z
x
y
, ,
q
R
R
q
Angles , tan a` tan a
R
Point , , ;

Sphere

2 2 2 1
Vector

Ray vector

p [sin a`cos sin a`sin cos a`]T

The condition of intersection can be written as follows

sin sin ,

arccos( p 1` , p 0` ) / 2 ,

we have

sin [ p ` p 0 ] [ 0 sin a`sin 0 cos a`


0 sin a`cos 0 sin a`sin ]( 02 02 )1 / 2 ,
whence

(02 02 ) sin 2 02 sin 2 (0 cos a`cos 0 sin a`) 2 1 ,


or

(0 cos 0 tan a`) 2 (1 tan 2 a`)(1 02 sin 2 ).


The equation of enveloping cone is quadratic with respect to tana'

[0 (1 02 sin 2 )] tan 2 a`2 0 0 cos tan a` 02 0,


and, returning to real space, we obtain
2

q [ 0 (1 02 sin 2 )] tan 2 a 2 0 0 cos q tan a 02 R 2 0 . (2.108)


Giving n meanings to , and solving quadratic equation (2.108), we obtain
a table of values , i that define a discrete sequence of EM vectors:
si [sin a cos sin a sin i cos a]T .

110
Test of Incidence Between a Point and Elementary Map Trace
Significant reduction of testing time can be obtained by defining
rectangular envelope for the EM trace in terms of four numbers [Xmin, Xmax, Ymin,
Ymax]. To determine an envelope it is sufficient to look through the list of
polygon vertices F=A1, A2, ... , An on the plane:
X min min( X 1 , X 2 ,..., X n ), X max max( X 1 , X 2 ,..., X n ), etc.

Fig. 2.34. Testing incidence between point W and figure F:


a - point W is outside polygon F; b - point W is inside polygon F.

Let R be the ray that originates from the point W and passes though the
polygon F. We may choose the centre of arbitrary segment Ai Ai+1 and draw the
ray WAm. Now, we determine the number of intersections between the ray and
all the ribs i=1, n-1. If the number of intersections between R and F is odd,
then the point is inside the polygon F; if the number is even, then it is outside
the polygon F (Fig. 2.34a,b). This algorithm is applied in computer graphics
for testing the incidence of a point lo an object, which is approximated by
polygon F in pictorial plane. Another efficient algorithm is based upon
calculation of angles sum when bypassing a polygon [38].
Modified Scheme of the Method of Elementary Maps
Vogl et al. [37] suggested an algorithm named the field patch mode. In
this method the aperture is divided into a grid of elemental patches that can be
considered as pinholes, and ray beams (EM) correspond to each of these
pinholes (Fig. 2.35). The light distribution on the target (pictorial) plane due to
energy coming through the pinholes is evaluated, while inclination of
corresponding area of mirror surface determines the position of illuminated
patch (EM trace) on the target surface. Test grid is selected on the target

111
plane and then conlrubution of elemental patches at nodal points is calculatcd
by recording covering of points by EM traces. This method enables to
consider illuminating beams as being independent from each other. The model
is based on the following statement from theoretical photometry [38]: if a
system consisting of two arbitrary openings (diaphragms) in nontransparent
screens is placed in front of a stretched background of uniform luminance
which radiates in accordance with Lambertian law, then the luminous flux
that passes through both diaphragms will be equal to the product between the
luminance L and the measure of rays set N. It follows from the statement that
illuminance at a point of target surface is proportional to the area of
illuminating diaphragm [39].
Though computer interpretations of EM method appeared comparatively
not too long ago [40], this method was known in lighting practice as far back
as 30s.

Pig. 2.35. Scheme of illuminating plane by elementary beams (EM):


a - aperture discretization in accordance with VogI et al.;
b - discretization of reflector.

112
2.7. PROBABILISTIC SIMULATION IN DESIGN OF LIGHTING AND
OPTICAL FIXTURES
In order to design a reflector with regular and diffused (mixed) reflection,
one can use the methods discussed above, e. g. the inverse-ray method [41].
But when diffusion shows a complex nature [42] or when dealing with
interreflections, e. g. in slit lightguides [43], application of these methods is
hampered.
A method given below is based on direct mathematical modeling of an
object and its operation. Propagation of the radiation from a light source
towards a target surface in illuminating installation; or into the outer-space
from lighting fixture can be imitated. Obviously, we restrict ourselves to a
direct problem, i.e. to calculation of luminous field characteristics
(illuminance, candlcpower, etc.), when parameters of installation or lighting
fixture are prescribed. This is in contrast to inverse problem, where parameters
of optical system have to be calculated to ensure required lighting distribution
in a prescribed zone.
Mathematical simulation of a light propagation is based on probabilistic
representation of photometric variables. For instance, a luminous flux falling
onto a small area A is directly proportional to the probability PA of hitting this
area, i. e. A=PAs, where s is the luminous flux emitted by a light source.
Hence, the problem reduces to finding PA. To solve the latter task, a
mathematical experiment has to be carried out: the radiance emitted by a light
source is represented by a set N of discrete portions of luminous flux or light
rays [44]. Then a random choice is realized from the set, and each ray obtains
the initial weight w=1. Next, we trace a traveling path of each ray in a
considered space. After reflection from bounding surfaces, the weight of ray w
decreases in accordance with reflectance factor . When the i-th ray hits an
area A, wi is recorded. After constructing the trajectories for all N rays, we
N

obtain N the sum

w , the ratio of which to N gives a statistical estimate P


i

i1

of desired probability PA, i.e. PA P A (1 / N ) wi [45].


i 1

By applying described procedure, one can obtain the estimates of any


photometric variable, c. g. the illuminance estimate averaged over the area A
is

EA

s N
wi ,
AN i 1

(2.109)

113
and the estimate of luminous intensity averaged within arbitrary solid
angle (, ) (where , are the angles that define the axis direction of a
solid angle) is

I ( , )

s N
wi ,
AN i 1

(2.110)

where the sum is determined for the rays leaving a light fixture and
appearing within .
Simulation as a process can be divided into three stages [46, 47]:
(1) generation of random rays emitted by a light source;
(2) calculation of ray paths;
(3) registration of rays on target plane.
We consider these stages separately, and take for illustration a rectangular
parallelepiped with reflecting walls, inside of which a source with known light
distribution is placed (Fig. 2.36). Such object may be regarded as a lighting
fixture, where walls serve as a reflector, and floor, as an aperture. Shape
of reflecting surfaces has no principal significance.

Fig. 2.36. Scheme of ray tracing within parallelepiped.

114
(I) Generation of rays. For generality of description we use the term of
phase-space [48], where every point q is defined in arbitrarily chosen basis

i, j, k by the radius-vector r of ray origin, i. e. the point of light emittance


from a source or the point of reflection; by the unit vector of direction s ; and
by the rays weight w. In other words, each point qi, in a space defines a ray, i.
e. its location (ri), orientation (si), and power (wi). The choice of qi. i. e. ri, and
si (wi is prescribed or calculated), is determined by the probability distribution
P(q) that specifies the location of q within the interval [qo, qk] from the phasespace being considered. In order to calculate q we use the Monte Carlo
equation
q

*
dp(q~ ) ,

(2.111)

q0

which links the location probability of q within the interval [qo, qk] with the
random number uniformly distributed within the interval [0 1]. In this case
dp(q~) is equal to the ratio of ray flux d(q~ ) to the flux carried by all rays
from the interval [q0, qk].
Hence,
q

d( q~ ) / .

(2.112)

q0

Solution of Eq. (2.117) with respect


to the upper limit q determines the coordinates (r, s ) of a random ray q. Random
numbers are produced by appropriate
generator. Usually, it is a standard
computer program [49].
We consider an example of
generating a random ray from cylindrical
light source with Lambertian side surface
(Fig. 2.37). An elementary flux emitted
by a small area dA R d dz within a
solid angle d 2 sin d d is equal
to
Fig. 2.37. On modeling of random
ray emilted by light source.
*

Here and below index marks a variable of integration.

115

d 4 LR cos sin 2 d dz d d ,

(2.113)

hence, the total flux of a light source is


2

l /2

0 z l / 2

/2
4

2 2 LRl .

(2.114)

0 / 2

As it is clear from Eq. (2.113), the coordinates of random ray, which


carries a flux d 4s , are determined by four parameters: the first pair ( and z)
defines the point r, while the other (, ), the direction of ray take-off from a
light source. Since the parameters are independent, after substituting Eqs.
(2.113) and (2.114) into Eq. (2.112), the latter splits in four equations. Their
solutions with respect to the upper limit of integration express desired
parameters in terms of appropriate random numbers :

2 ;
z l ( z 0,5);
2 sin 2 4 ;

(2.115)

arcsin .
Further, we express unknown quantities r and s in terms of found
parameters

r R cos i R sin j z k;
s sin cos( )i sin sin( ) j cos k .

(2.116)

(2.117)

For axially symmetrical lighting fixture, the candlepower of which is given


in the form I(, )=I0 cosma, and location of which is specified by vector ro
(Fig. 61), the direction of random ray is

s sin a cos i sin j cos a k .

(2.118)

a arccosm1 a ; 2 .

(2.119)

where
If, while designing lighting installation, we need to account the joint effect
of all light fixtures, then random choice of a fixture has to be done before
generating a random ray. In this case, it is suitable to represent a population of
lighting fixtures in the matrix containing m n elements. Then the choice of a
fixture is reduced to drawing a random element, the indices of which i=1, m
and j=1, n determine the location coordinates of appropriate fixture, or

116
the vector roij,; (Fig. 2.36). Then, by using a pair of random numbers m and n,
we calculate these indices according to the following formulas: i=[(m+1)m], j=
[(n+1)n], where square brackets mean the integer part of a number.
The described procedure can be applied to calculation of luminous
intensity distribution of a light fixture as well.
(2) Ray trajectory construction. The sequence of ray states c(q1, q2,,
gi,) is determined on this stage, i. e. intersection points r and ray directions s
after reflections from surrounding surfaces. The appropriate procedure for
specular surface is described in Ch. 1.
In case of diffusion the total flux is substituted by a single random ray. In
this way the nonterminating branching of rays s can be avoided. Calculation
of s here is similar to the generation of a ray emitted by a light source,
because reflecting surface may be considered as a secondary source. Generally,
the indicatrix of reflection depends on a point location on reflecting surface, as
well as on a direction of incident ray s ; this peculiarity must be taken into
account. In case of mixed reflection with luminancc factor f ( s , s`) being
symmetrical with respect to the direction of specular reflection Ss, it is
advisable [42, 46] to turn to the local basis i`, j`, k`, where the unit vector k`
coincides with the normal n to the surface at the point of ray incidence, and the
unit vector j` is directed along the projection of vector Ss on a reflecting
surface (Fig. 2.38). In this basis, the random reflected ray s is defined by the
angles of diffusion: the polar angle p and the asimuth angle p (Fig. 2.39).
Generally, the parameters p and p are dependent,

Fig. 2.38. On calculation of reflected ray.

Fig. 2.39. On determination of


divergence angles p and p.

117
that is why their expressions in terms of random numbers and are defined
by applying conditional probability [48].
For obtained p and p the coordinates of vector s in the local basis take
the form

s x cos 0 sin a p cos p sin 0 cos a p ;


s y sin p sin p ;

(2.120)

s z cos 0 cos a p sin 0 sin a p cos p ,


where 0 is the angle of ray s` incidence (Fig. 2.39).
For the indicatrix f ( s , s`) being symmetrical with respect to n, i.e. when
p and p are independent, s is defined from Eq. (2.118).
The weight of reflected ray is equal to w=p(r, s )w', where p(r, s ) is the
integral reflectance of a surface at the point of ray incidence r at the direction
s , w' is the initial weight.
Important procedure at this stage is the search of a surface, with which a
ray intersects (r, s ). For rectangular parallelepiped the condition whether a ray
hits a ceiling (C), walls (W1-W4) and target plane (TP) are given in Table 2.4,
where rx, ry, rz are the coordinates of a point r where a ray intersects with
appropriate plane.
Table 2.4
Conditions of rays hitting the plane
C
Wl
W2
W3
W4
TP

sz 0
0 rx
0 ry b

sy 0

0 rx
0 rz c

sx 0
0 ry b

0 rz c

sy 0

0 rx
0 rz c

sx 0
0 ry b

0 rz c

sz 0
0 rx
0 ry b

Obviously, to hit an axially symmetrical reflector bounded by the exit


aperture of radius Rcx and the throat of radius Rth a ray must fit the following
conditions
rx2 ry2 Rcx2 , s z 0
(2.121)

rx2 ry2 Rth2 , s z 0

(2.122)

The trajectory terminates in the following cases: (1) a ray hits the
prescribed surface (the space); (2) the condition w wmin is fulfilled, where
wmin is a minimal weight of a ray set up in advance. When there is a need to
take into account a reflection from the target surface, then even after the ray
strikes the target surface, path construction is continued.

118
(3) Registration of rays. As a rule, the illuminance distribution E(x, y) on
the target plane or the candlepower distribution serves as the main output
characteristic. In order to calculate statistical estimates of these characteristics
wc divide the target plane or the light fixture in cells A (). After tracing all
N

of N rays, a sum

w , is accumulated in each cell. Further, by applying


i

i 1

Eq.(2.109), or Eq. (2.110), we calculate the E j or I j distribution.


Characteristics like efficiency of a light fixture, luminous flux utilization in
lighting installation, luminance of reflecting surface, projections of light
vector, etc. can be calculated in similar way. To evaluate the accuracy of
results obtained by Monte Carlo method is very important. The matter is, that
the results, c. g. E or I are not calculated analytically, but they are statistical
estimates of E or I , correspondingly. Dispersion DE of desired value,
absolute error , and N are connected by the following relationship

N t 2 DE / 2 ,

(2.122)

where t is a coefficient that depends on the chosen fiducial probability .


Dispersion is determined by the following formula
2
2
s 1 N 2 1 N ,
DE
wi wi

N i1
A N 1 i1

(2.123)

Fig.2.40. Calculated candlepower curves under


specular (l), mixed (2) and diffuse (3) reflection
(luminous flux is equal to 1000 lm).

119
where sums

2
i

w and w
i

are being accumulated during the process of

calculation.
From Eq. (2.123) we see that in order to obtain a sufficient accuracy, a
great number of rays has to be used, and this is time-consuming. Therefore, to
decrease DE is very important. Various methods of lowering DE are
described in [46].
A. A. Korobko developed a program based on described algorithm. The
program enables to compute a candlepower distribution (Fig. 2.40) for
prescribed profile of axially symmetrical reflector, which may have three types
of reflection (specular, diffused, and mixed), and for the different lamp types
(metal-halidc, high-pressure sodium, mercury-vapor).
2.8. CALCULATION
OF
LAMBERTIAN SOURCES

LIGHT

DISTRIBUTION

FROM

Designing illuminating and irradiating installations, we must be able to


determine such characteristics of light sources as luminous flux, luminous
intensity, and illuminance [50]. Description of light field in terms of light-tubcs
enables to solve photometric problems easily [51].
Ellipsoidal source
We consider the calculation of characteristics of
Lambertian ellipsoid having half-axes and b (a>b)
(Fig. 2.41).
Luminous flux. We find the expression for
ellipsoid area and flux. Equation of ellipsoid in
parametrical form is

x b cost , z a sin t.
An area element is

d dx 2 dz 2 b 2 sin 2 t a 2 cos2 t dt.


An area element is

dS 2 x d 2 ab 1 e 2 sin 2 t dt ,
where e 2 (a 2 b 2 ) / a 2 .
Fig. 2.41. Projection of
ellipsoid along direction .

120
Further

dS
2b 1 (e 2 / a 2 ) z 2 .
dz
A surface area from z = z1 to z=z2 is
1/ 2

S1, 2

2
z
2ab 2 ez


e z1 a

ez
d
a

ab w2 1 w22 w1 1 w12 [arcsin( ew2 ) arcsin( ew1 )] ,(2.124)


e

where w1 z1 / a; w2 z 2 / a.
The total area is

1
S 2ab( 1 e 2 arcsin e).
e
Hence, the total flux of Lambertian ellipsoid having luminance Lo is

1
2 2 abLo ( 1 e 2 arcsin e).
e

(2.125)

Luminous intensity of ellipsoid


Inclined projection of ellipsoid on a plane yeilds the large axis (Fig. 2.41)

a` a 2 b 2 cot 2 .
A projection area is

S a`b b a 2 sin 2 b 2 cos2 .


The luminous intensity at the angle is

I L S L b A B cos 2 ,
where A (a 2 b 2 ) / 2; B (a 2 b 2 ) / 2.

(2.126)

121
Light vector
For ellipsoid having a centre at the
point (x, 0, z) (Fig. 2.42) we have the
folowing wrapping cone equation
2

x z
2 2
a
b
2
2

2
, (2.127)
cot 2

2
a 2
b
where cot 2 x 2 / a 2 z 2 / b 2 1;
, , are direction cosines of a ray
starting from point (x, 0, z) and being
tangent to ellipsoid.
Fig. 2.42. On construction of light-lines
Setting 0 , / tan in Eq.
produced by ellipsoid of uniform
luminance, and on calculation of absolute
(2.127), we obtain polar angles for
value of light-vector at point A;
boundary tangential rays (Fig. 2.42):
Hyperbola G divides fluxes into up and
l.

tan i

x2 b2
xz a 2 x 2 b 2 z 2 a 2b 2

(2.128)

i 1, 2

or
tan i

x2 b2
.
xz ab cot

(2.129)

i 1, 2

Equation (2.129) can be written in equivalent form


tan i

xz b 2 ( z 2 a 2 ) a 2 x 2
.
z 2 a2

(2.130)

i 1, 2

According to the definition, the light-vector is directed along the bisectrix


of an angle at the vertex (x, 0, z). We obtain
2 xz
(2.131)
tan(1 2 ) 2
.
2
( z a ) ( x2 b2 )

122
Moreover, we have an equation for cofocal hyperbolas

z2
x2

1,
a2 u b2 u
where b 2 u 2 a 2 .
A tangent to hyperbola or polar angle tangent is

dx
z b2 u
tan k

,
dz
x a2 u
whence
2 xz
tan( 1 2 ),
2
2 a u
2 b u
x 2
z 2
b u
a u
since
hyperbola
equation
yeilds
( a 2 u ) /(b 2 u ) a 2 u z 2 ,
(b 2 u ) /( a 2 u ) b 2 u x 2 . So, the light-vcctor is tangent to a hyperbola
which defines light-lines of ellipsoid. Focus of hyperbola is
z f a2 b2 .
tan 2 k

Here follows an algorithm for finding light-lines in arbitrary point at


meridian plane (x, 0, z) and flux fractions radiated at upper and lower
hemispheres up, l (Fig. 2.42, see also [381):
(1) we find hyperbola parameter it by solving the equation
u 2 ( x 2 z 2 a 2 b 2 )u a 2 b 2 cot 2 ;
(2) we calculate a point of intersection between a light-line and ellipsoid

z 02
a2 u

,
a 2 a 2 b2
(2.132)
x02
u b

;
b2 a 2 b2
(3) setting up w1 1, w2 z 0 / a in Eq. (2.124) we find l, and up=
1- l,.
To obtain the absolute value of light-vector we have to determine half-axes
A and B of elliptic disk in the plane being orthogonal to

123
the angle bisectrix at the vertex (x, 0, z), e. g. by using Focks formulas [52].
Generally, light sources, e. g. mercury-vapor fluorescent lamps (MVFL)
possess nonuniform luminance. In this case, calculation of illuminance can be
organized by using the inverse-ray algorithm. We put the origin of polar
coordinate system into the test point and organize a-and -cycles (see example
in paragraph 2.5).
If inverse-ray scanning is applied, the boundaries a1 and a2 are calculated
from the expressions

x b, a1 0, a 2 arctan

2bz
;
z a2

(2.133)

x b,
a1 , a2
- (see Eq. (2.116))
The value a f ( ) is being determined on each step inside the a-cycle
by solving the quadratic (with respect to cos) equation (2.108). If x b , then
the upper limit of integration along is equal to . If L(, ) is the ray
luminance, we obtain the light-vector components [see Eqs. (2.92) and (2.93)]:
a

Ez

1 2
sin 2a L(a, )d ;
2 a1
0
a2

(2.134)

E x sin 2 a L(a, ) cos d .


a1

Filament Light Source


We consider a filament with uniform luminance Lo, diameter , and length
2l. We find vertical Ez and horizontal Ex components of a light-vector at the
point Mo(xo, yo, zo) (Fig. 2.43). We determine the unit vector directed from the
source element dz towards the point Mo:

x0
m o
0
2
x0 ( z0 z ) 2

.
x02 ( z0 z ) 2
z0 z

(2.135)

The light-vector at the point Mo (Fig. 2.43) is

dE Lo

dz sin
m o ,
x ( z0 z ) 2
2
0

(2.136)

124

Fig. 2.43.
On calculation of light-vector
produced by filament
source with length of 2l.

o is calculated in accordance with Eq. (2.101).


where m
Further we find
l

E z L0 x0

( z0 z )dz
2

l [ xo

2 2
( zo z ) ]

L0 x0

l
xo dz
E x Lo xo

2
l [ xo ( zo z ) 2 ] 2

2 z 0l
; (2.137)
( x z l 2 ) 2 4 z 02l 2
2
0

2
0

(2.138)
l

xo2 zo2 l 2
1
2 xol
Lo x

arctan
.
2 z2 l2
o xo ( x 2 z 2 l 2 ) 2 4 z 2l 2 2 x 2
x

o
o
o
o
o
o

Light-line produced by Lambertian filament is hyperbola, i. e. a vector


having components Ez and Ex is tangential to hyperbola. We find the tangent to
hyperbola

125

z2 x2

1;
a2 b2
zdx xdx
2 ;
a2
b
2
b
x dx
0
k,
2
a
z0 dz
whence

x02
dz
a z 2 z02 x0 z0 ,
k
dx
2

2
0

(2.139)

dz
b 2 a 2k x0 z0 x02 .
dx
Light-line equation (omitting index 0) fits the condition
dz E z
2 xz

. (2.140)
2
2
2 2
( x z l ) 4 z 2l 2
2 xl
dx Ex
2
2
2
x z l
arctan 2
2 xl
x z2 l2
Equation for hyperbola passing through the point (xo, 0, zo) is
Z 2 a 2 ( z / x ) z `X 2 ,
(2.141)
where z' is calculated from Eq. (2.140) ; a,
b, from Eq. (2.139).
Obviously, the ratio between fluxes
into the lower and upper hemispheres is
up/1=(l-a)/(l+a) (Fig. 2.44). The
luminous intensity of Lambertian filament
at the direction with respect to the axis
OZ (Fig. 2.44) is
I 2 Lol sin I max sin . (2.142)
The total flux is obtained by
integration
Fig. 2.44. Light-line produced by
filament source with uniform luminance
(hyperbola) (a is line origin).

126

d 2 I max sin 2 d I max (1 cos 2 )d ,

1 I max (1 cos 2 )d 2 I max .

(2.143)

To find the illuminance produced by filament sources with nonuniform


luminance distribution, c. g. from mctal-halidc lamp, the inverse-ray algorithm
has to be constructed. In this case, the boundaries of search domain are
determined by polar angles

a1 arctan

z 0 lc
;
Rs x0

(2.144)

z l
arctan 0 c , x0 Rs ;
a2
x0 Rs

/ 2,
x0 Rs ;
where Rs is the source radius, lc is the source length; xo, z0 are the coordinates
of the viewing point.
The search boundaries along are
1 0, 2 arcsin Rc / x0 .
(2.145)
Lambertian Thin Tube
A source of this kind can be treated as a model for metal-halide or quartz
halogen lamps. Transforming slightly Eq. (2.138), we obtain the horizontal
illuminance

x2 z 2 l 2
1
z 1
z 1 (2.146)
E x ( z ) I max 2

arctan
arctan

2
2 2
2 2

2lx
x
x
(x z l ) 4z l
Often, e. g. when carrying out Fourier analysis, the coordinate
representation of Eq. (2.146) is more suitable than familiar expression for E in
angular terms.
For a point below tube centre, i. e. z = 0, we have
1
l
1
E x (0) I max 2 2 arctan
lx
x
x l

We assume that the source length is small, (l/x)0. We expand

(2.147)

127
the first term in Eq. (2.147) in scries with respect lo small values of
1/ x .
1
1
(1 2 4 ....)
x 2 l 2 lx
while the second term we present in the form
1
l
1
2 4
arctan
(1

...)
xl
x x
3
5
Hence, rejecting small values, we obtain E x (0) I max / x 2 i. e. the law of
reciprocal square distances is accomplished. That means that the linear source
can be substituted by the point source, provided that found error is accounted
[33].
Let a source be of infinite length, i.e. (x/l)0. The main term in expansion
of the first summand is equal approximately to l/l 2, i. e. it tends to zero.
Assuming I=Imax/l to be a specific luminous intensity, we obtain
Ex(0)=Io/x, i.e. the illuminance corresponds lo the law of reciprocal distances.
2

Ring-Like Filament of Uniform Luminance


An observer being at the parallel =const sees the ring of diameter dr in the
form of ellips with the following half-axes

a 0.5d r ; b a cos a.
Centres of ellips and ring may be considered as coincident. Equation of
ellips in parametric form is as follows

x a sin ; y b cos ,
where is an angle of so called eccentric anomaly. An arc segment of
ellips is

ds a 2 cos2 b 2 sin 2 d ,
or

ds 1 e 2 sin 2 d ,
where a

a2 b2
is the eccentricity of ellips.
a

128
In our case e=sina, so the length of a ring is
/2

S 4a

1 sin 2 a sin 2 d 4a E (a ),

where E(a) is the total elliptic integral of the second genus.


Luminous intensity of a ring with specific luminance L0 at the direction a
is

Ia 2d r Lo E (a) .

(2.148)

Since E (0) / 2 , Ia can be expressed as


/2

I a I a 0 0

i ( ) d
,
/2

(2.149)

where I a 0 d r Lo , i. e. the luminous intensity of a ring is equal to the


average value of indicatrix i ( ) 1 sin 2 a sin 2 being calculated along
the circular ring (02) and multiplied by the luminous intensity at the
normal direction.
The total luminous flux is
2

2 I a sin ada 4d r Lo E (a) sin a da .


0

(2.150)

To calculate Ia and we may use the expansion of E (e sin a ) into series


[53]
2

E (e ) 1 e 2 1 3 e 4 1 3 5 e 6

...
/ 2 2 1 2 4 3 2 4 6 5

(2.151)

A program for calculating E(e) with required accuracy can be composed


easily. The series converges like a geometrical sequence, so determination of
relevant number of elements in series can be carried out automatically.
Application to Problems of Reflector Design
Geometrical representation of a field produced by a light source leads to
rather simple solutions of photomctrical problems.
Direct problem. Let a source be in the form of uniformly radiating thin
tube or filament located along the vertical axis. We have to

129
estimate the flux fraction q which falls on reflector throat if initial point of
reflector has coordinates xo, zo, and filament length is 2l.
Here follows an algorithm of calculation;
(1) knowing the initial point and applying Eq. (2.140), we find the tangent
to hyperbolic light-line;
(2) using Eq. (2.139), we find the half-axis of hyperbola;
(3) we calculate q=(1-(/l)) x 100%.
For l=12; xo=100; zo=80, we obtain q=28.1%.
Inverse problem. For the same light source and prescribed throat radius xo
and corresponding flux loss, the location of initial point zo is required.
When x=xo is prescribed, the derivative from Eq. (2.106) becomes a
function of z. Then the expression z2- x z z' - a2 is a function of z as well. Thus,
a desired value of z is the rool of this function. Geometrically, the solution of
this problem means finding the intersection point between a hyperbola
originated from the point (0, ) and a vertical x=xo.
So, the following algorithm reaches the goal:
(1) to find the half-axis of hyperbola a=(1-0.01 q) l;
(2) to solve the equation z2 - x z z'x-a2=0 by using some numerical method
(e. g. by using the collection ZEROIN from Annex 1). Allowed range of z has
to be set in the program. In order to decrease losses obtained in previous
example, we choose q=8% and obtain zo=132.4.
2.9. MODELS OF REAL SOURCES AND MATERIALS
Models of metal-halide lamps
According to [54], the luminancc distribution along arc-tube of vertical
lamp is described by the following formula
L( ) L0 (mh / hg 1) exp[ n(r / rb ) 2 ](sin ) 0.6
(2.152)
where is the angle between the lamp axis and the viewing direction; r is the
distance between viewing direction a and the axis of cylindrical lamp body; h
is a current height, and h=0 indicates the centre of an arc-tube; Lo, rb, hg are the
constants depending on lamp power; Lo is the luminance at the centre of an arctube axis; rb is the inner radius of an arc-tube; hg is the apparent arc-tube
length; m, n are the coefficients in the relationship between the luminance and
the point position at an arc-tube (Tabic 2.5).

130
Tabic 2.5
Data on metal-halide lamps which are necessary for calculation of lighting
fixture
Characteristic
diameter a
length b
height of light centre c
Constant m
Constant n
Luminous flux, lm 103
Luminance at arc-tube centre,
cd/m2 x 106

Power, W
250
400
700
Lamp dimensions, mm
91
122
152
227
290
370
142
185
240
0.45
0.21
0.63
4
3.8
2.7
135
2
10.2
19
35
60
6.3

6.6

1000

2000

176
390
245
0.78
3.2
14.2
90

100
430
255
1.15
3.3

7.4

8.8

7.3

190

For horizontally positioned lamps the luminance along the axis is


uniformly distributed, but the discharge cord shifts upwards from the axis by
r0. The luminance model for this case is as follows

L( ) L0 exp[ a(r r0 ) / rb ](sin )0.5


2

(2.153)

Constants are given in Table 2.6.


Table 2.6
Lamp type
DRI 250-5
DRI 400-5
DRI 700-5
DRI 1000-5
DRI 2000-5
DRI 3500-5

Lo, Mcd/m2
7,1
6,8
6,8
7,0
7,6
10,5

Parameter value
ro, mm
0,5
1
1
0,5
1
1,5

4,26
4,13
2,07
1,72
1,25
1,20

The presence of sin is explained by partial transparency of discharge for


nonresonancc lines of scandium and for visible lines of mercury.

131
Module for luminance determination
We describe now the algorithm that
uses
Eqs. (2.152) and (2.153) and
determines the ray luminance for melalhalide lamp. Applying these formulas, we
suppose that the ray luminance docs not
depend on the angle of emission, i. e.
radiation follows Lambert's law.
Thus the luminance of inverse ray a
becomes a function of the point
coordinates zL and rL where a ray intersects
the cylindrical arc-tube (Fig. 2.45).
Equation for inverse ray is
s s0 al ,
where

a [a1 a2 a3 ]T ,

and

s0 [ xo yo zo ] is the radius-vector of
reflector point.
The condition of tangency between a
ray and a cylinder of radius rL yeilds

Fig. 2.45. On determination of ray


luminance for cylindric source

(a1 y0 a2 x0 ) 2 rL2 (a12 a22 ),


whence,

rL

a1 y0 a2 x0
a12 a22

(2.154)

The ray luminance is determined by the following procedure:


Step 1. We calculate xo, yo, zo and a1, a2, a3 in the local coordinate system
associated with a light source.
Step 2. We calculate the coordinates of a point where the ray intersects the
cylinder. They relate to the roots l1 and l2 of the square equation: l2 : l1 X, Y
Step 3. We find the scalar product
scal a1 X a2Y ,
if scal<0 then the intersection point corresponds to the root l1 (see Fig.
2.45) and

132

z L z0 a3l1.
In the opposite case,

z L z0 a3l2 .
Step 4. By using Eq. (2.154), we calculate the minimal radial distance
between the ray and the axis of an arc-tube.
Step 5. We calculate the ray luminance, by substituting r = rL and z = zL
into Eqs. (2.152) and (2.153). We take the values of r, Lo, m, and n from Tables
2.5 and 2.6.
Model of mercury-vapor fluorescent lamp (MVFL)
A luminous body in MVFL can be regarded as an ellipsoid of revolution
with major half-axis a and minor half-axis b (Fig. 2.46a). Surface luminance
indicatrix of MVFL has one plane of symmetry which coincidcs with the
corresponding meridian plane of a bulb [24]. The real indicatrix is not a surface
of revolution but still it is substituted by such a surface. The luminance at the
direction a is as follows

Fig. 2.46. Model of mercury-vapor fluorescent lamp; a - determination of eccentric


anomaly angle; b - reading luminance from bulb meridian.

133

L Ld Lm cos m

(2.155)
where Ld is the diffuse component of luminance indicatrix; Lm is the mixed
component of indicatrix; is the angle between a and the normal to the lamp
surface.
Generally, the power of cosine is not a constant value, it varies from 2 to 4,
according to [24]. The values of Ld and Lm are the functions of point location
on a bulb meridian, i. e. Ld= Ld/(), Lm=Lm(), where is an angle of eccentric
anomaly related to ellips coordinates by equations x=b cos , z=acos . The values are usually calculated and put on the bulb profile similar to marking of a
circular limb (Fig. 2.46b).
Calculating light source flux for piecewise linear specification of
candlepower curve
Let the function I=I() be changeable between the nodes o and 1
linearly

I Io

I1 I o
( 0 )
1 0

(2.156)

where I o I ( o ) , I1 I (1 ) .
We consider the following cases:
(1) Cylindrical system
We substitute Eq. (2.156) into the flux formula
1

F I d
0

and get

F {I o 0.5( I1 I o )}(1 o ).

(2.157)
Formula (2.157) is suitable for programming. It gives a good
approximation, and the result needs only small correction.
(2) Rotation ally symmetric system
For this case we have

F I sin d .
0

134
After substituting o, we obtain

I I
F ( I o cos o I1 cos 1 ) 1 0 (sin 1 sin o )
1 o
or

sin 1 sin o
(2.158)
F I o (cos o cos 1 ) ( I1 I 0 )cos 1
1 o

Calculating mean radius of dispersion indicatrix


There is often a need to obtain suitable approximation of dispersion
function expressed in terms of direct component of reflected or transmitted
light. The following function is applicable in this case
i io exp( a sin 2 ( / 2))
(2.159)
We find the luminous flux under the indicatrix

F 2io sin exp( a sin 2 ( / 2))d .


0

Translating the sine according to double angle formula and changing the
variables, we get
F (4 io / a)(1 exp( a sin 2 ( / 2)))
(2.160)
Example. Let an angle of dispersion be prescribed, say, 2. The task is to
ensure that the flux within the cone of = is equal to 70% of the flux within
the cone for =/2. This states a condition
1 exp( a sin 2 ( / 2)) 0.7(1 exp( a / 2)).
(2.161)
hence, we obtain the equation with respect to x=exp(-a/2)

1 x 2 sin

2 (

/ 2)

0.7(1 x ) .

or

0.7 x x q 0.3 0,
where q 2 sin 2 ( / 2) 1 cos .
Assuming the x-value to be very small, we find qx=0.3, whence a=323.06.

135
2.10. GRAPHICAL REPRESENTATION OF LUMINOUS FIELD
PRODUCED BY LIGHTING FIXTURE
Computer graphical representation of output characteristics is of great
significance [55, 56], since it enables to preestimate the quality of designed
lighting fixtures.
It is a known fact that the light field of a luminaire can be presented by a
family of light-lines or lubes. The main property of a light-tube is such, that a
constant luminous flux flows through its arbitrary cross-section. The picture of
luminous field allows, e. g. to estimate illuminance on any arbitrarily oriented
surface.
For axially symmetrical specular surface with coaxial source the luminous
field is separated into tubes, the walls of which arc the surfaces of revolution
with common axis. To make a picture of luminous field [57], one must
calculate a family of illuminance curves E(y) at different cross-sections z=const
(Fig. 2.47). Assuming that one and the same luminous flux flows within a
space between two lines, we can determine the coordinates of a tube from the
following equation
E ( y ) / S y (2.162)
where Sy is a cross-section area of
a tube.
Let the total number of tubes be
N, then the k-th tube coordinates in zi
section can be found from equality
( y ) k ,
or
yk

2 E ( zi , y ) y dy
0

k
,
N

(2.163)
where ( y ) denotes the flux
determined from the curve of
luminous flux growth for section zi;
k is the summarized luminous flux
from k tubes; o is the total luminous
flux determined by integrating from 0 to
90 (Fig. 2.47).

Fig. 2.47. Determining light-tube


coordinates in section zi.

136
Since finite dimensions of flashed area should be considered at close
distances from luminairc, the calculation of E(z,y)-curves becomes a laborious
task. Inverse-ray principle enables to consider nonuniform luminance
distribution of a light source, that is especially important when dealing with
high-pressure gas-discharge lamps.
It is a known fact that a specular reflector with metal-halide or highpressure sodium lamps shows harsh illuminance irregularities in a proximate
zone at points close to axis. Generally, integration of such harshly changing
functions, while determining tube coordinates [Eq. (2.163)], is a nontrivial
problem. Adaptive quadrature program, e. g. QUANC8 [29] solves the
problem best of all. The program automatically selects the step of integration
related to prescribed accuracy: fine at the vicinity of peaks, and rude at a
distance.
Solution of Eq. (2.163) can be found by applying subprograms of inverse
spline-interpolation SPLINE and SEVAL (Annex 2).
Below, we give a logic scheme of determining light-tube coordinates in
prescribed cross-section; when the illuminance distrubution in the cross-section
is found beforehand, e. g. by applying program CAMIL. The scheme can be
translated into computer program.
(1) To input: M, N, HY, YMAX, where M is the number of tubes, N is the
number of illuminance levels E(y); HY is the step with respect to y; YMAX is
the null of function E(y).
(2) To input: E(I), I=1, N, where E(I) is the array of illuminance values.
(3) To calculate: E(I)=E(I)*YI, YI=(I-1)*HY, I=1, N, i. e. the value of
integrand function that is recorded in storage location E(I).
(4) To calculate F(I), I=1, N. In order lo calculate the luminous fluxgrowth function one should apply subprogram QUANC8 [29]. While calling
QUANC8, one should set up the lower limit of integration A=0, and the upper
limit B=(I-1)*HY, 1=1,N. To compute intermediate values of integrand
function E(I), address to subprograms SEVAL and SPLINE.
(5) To calculate the step DF=F(N)/M.
(6) To calculate the tube coordinates S(K), K=1,M by using subprograms
SEVAL and SPLINE. To write ARG(I)=F(I), I=1,N into the arguments array;
to write FUN(I)=(I-1)*HY, I=1,N into the array of function values.
(7) To print: S(K), K=1,M; F(I)=1,M.
Figure 60 shows drafts of luminous field near aperture of specular reflector
with lamps DRI 400-5 and reflector lamps with arc-tubes. The drafts were
produced by computer.
From above-stated we may conclude the following:

137
(1) In the proximate zone the luminous field of lighting fixture with metalhalide lamp is characterized by increased solid density of energy. Luminous
flux transported through the tube which has a diameter approximately half of
that of luminairc aperture makes up to 60% of the total flux;
(2) The z-coordinate of flux concentration zone enables to identify the
type of candlepower curve;
(3) The luminous field picture enables to forecast the distance of
photometrical testing of lighting fixture. At the test distance light-lines differ
but negligibly from asymptotes of a point source;
(4) The least thermal load results for flat safely glass, since the density of
light-lines (Fig. 2.48) increases greatly with the growth of glass curvature, and
with approaching to the axis.

Fig. 2.48. Luminous field of specular luminaire with metal-halide lamp DRI-400: a candlepower curve of G type; b - candlepower curve of K type.

138
2.11 CALCULATING LIGHT SOURCE INDICATRIX FOR SPECIFIED
LUMINOUS INTENSITY
When geometry and luminance characteristics of a light source are set,
calculation of luminous distribution relates to direct problems. Sometimes an
inverse problem arises, that is to determine local characteristics, e. g. luminous
intensity indicatrix or luminance indicatrix when integral output parameters,
say, candlepower curves, are known. To solve these problems, one has to deal
with integral equations [58 ].
Ring-like source
Complex spiral luminous body is often substituted by torus or anchor. We
consider a ring segment having the direction
[sin cos 0] .
A ray of observation is
[sin 0 cos a] .
An angle between the element normal and the ray of observation is
sin sin a sin .
(2.164)
We introduce the condition that specific luminance (per unit of length)
depends on the angle
L = L.
The luminous intensity at the direction a is

dI a I ds cos ,
or, introducing ring diameter dr,

dI a 0.5 d r L 1 sin 2 a sin 2 d .


The total luminous intensity of a light source is
/2

I a 2d r

1 sin 2 a sin 2 d .

(2.165)

Setting L=Lo, we obtain Eq. (2.144). Wc introduce now the indicatrix of


luminous intensity

139

L
cos ,
Lo

hence
/2

/2

I a 4 0.5d r L0i d d r Lo

i d

/ 2 ,

where Lo is the mean luminance at the direction a=0. Then the luminous
intensity of a thin ring is
/2

/2

I a 4 0.5d r L0i d d r Lo
0

i d

/ 2 ,

(2.166)

or
/2

Ia Ia 0

i d

/ 2' ,

(2.167)

i. e. the luminous intensity of a ring is equal to the indicatrix value, averaged


along the circlc (0), and multiplied by the candlcpowcr at null-direction.
Expression (2.167) is an integral equation in i. This is Volterras equation
of the first genus, mainly the Abels equation [59] if reduced to canonical
form.
We substitute the variable in the integrand in Eq. (2.167)
d sin sin ad sin sin a cos d ,
(2.168)
whence

d sin
d sin
d sin
d sin

2
2
2
sin a cos sin a 1 sin 2
sin a sin sin a
sin 2 a sin 2
Now we recount the limits of integration

0
/2

sin 0
sin sin a

and obtain
/2

i d sin
2

sin a sin

Ja ,
2

(2.169)

140
where J a I a / I a 0 .
We translate Eq. (2.169) lo the following form
/2

sin

d (sin 2 )

J a .
sin 2 a sin 2

(2.170)

Equation (2.170) is similar to that of Abel. Within the function class Ja


with continuous derivatives in the interval [0, /2] equation (2.170) has unique
solution [60]:
cos 2
J `sin 2 a d (sin 2 a )
i
1 J (0)


,
sin sin 2
sin 2 sin 2 a
0

(2.171)

whence,
sin

i J (0) sin

J `sin a d (sin 2 a )

(2.172)
sin 2 sin 2 a
Taking into account that dJ / d (sin a ) (dJ / d ) / cos a , and having made
the change of variable sin a sin sin w , after manipulations we obtain

/2

i J (0) sin

J `a dw
.
cos a

(2.173)

If the curve Ja was obtained by measuring, errors are inevitable, and the
equation (2.173) becomes unstable due to differentiation. Therefore, we have
to use more stable quadratic forms [61].
Cylindrical source
We consider a cylinder with radius rc being small in comparison with its
length. Let L be the luminancc of unit area of cylindric surface at the direction
. The normal to cylinder element is
[cos sin 0] .
The ray of observation is
[sin a 0 cos a] .
The luminous intensity at the direction a is

141
/2

I a I a / 2

L cos d ,

(2.174)

where cos=sin cos; I=/2 is the luminous intensity of cylinder at the


direction =/2 for uniform unit luminance. Reducing the prescribed luminous
intensity curve to this value, we translate Eq. (2.174) in the following form
/2

L cos d J

(2.175)

The expression (2.175) is Abels integral equation. By changing variables


it can be reduccd to canonical form just in the same way as it was done above.
The solution is as follows

2 J (0)
L

cos

/2

J `a dw
,
cos a

(2.176)

where sin a cos sin w.


Setting J a sin a , from Eq. (2.176) we find L 1 . J a 1 , we get L
cos = 1. These correspond to variants of totally opaque and absolutely
transparent discharges in Gershuns model [62].

142
REFERENCES
(1) Boldyrev N.G. Theoretical Photometry. L.: LIOT, 1938.
(2) Komissarov V.D. Foundations of Calculating Specular and Prismatic Fittings
// Trans, of VEI, 1941, Issue 43, p. 6-61.
(3) Shealy D.L., Burkhard D.G. Analytical Illuminance Calculation in a
Multiinterface Optical System // Optica Acta, 1975, vol.22, 6, p. 485-501.
(4) The Computer in Optical Research Methods and Applications. Ed. by
B.R.Frieden. Springer-Verlag, Berlin - Heidelberg New-York, 1980.
(5) Cornbleet S. Microwave Optics. Academic Press. New York San Francisco
London, 1976.
(6) Mc.Dermil H., Horton T. Reflective Optics for Obtaining Prescribed
Irradiative Distribution from Collimated Sources // Applied Optics, 1975, vol. 13, 6,
p. 144-11450.
(7) Galchinskii B.Ya. Reflection and Refraction of Elastic Wave of Arbitrary
Shape from Curvilinear Bound between Mediums // USSR AN Reports, 1958, vol.118,
3, p. 458.
(8) Ermolinskii N.N., Komissarov V.D. Calculation of Specular Reflectors// VEI
Bull, 1934, 2, p. 21-26.
(9) Fock V.A. Fresnel's Laws of Reflection and Laws of Diffraction// UFN, 1948,
vol.36, issue 3, p. 308-327.
(10) Kravtsov Yu.A., Orlov Yu.I. Geometrical Optics of Non-Uniform Mediums.
M. Nauka. 1980.
(11) Benford F. Studies in the Projection of Light. General Electric Review, 192326.
(12) Heat Transfer and Spacecraft Thermal Control. Ed. by John W.L.Lucas. Jet
Labs. Pasadena, California, 1970.
(13) Boldyrev N.G. On Calculation of Asymmetrical Specular Reflectors//
Illumination of Industrial Enterprises. L.: LIOT, 1935, p. 179-186.
(14) Lebedev L.M., Morozov M.G., Raibman E.M. Irradiation in Cylindrical
Elliptic Concentrator // Optiko-mekh. promyshlennost', 1974, 3, p. 10-13.
(15) Volosov D.C., Tsivkin M.V. Theory and Calculation of Lighting Systems.
M.: Iskusstvo, 1960.
(16) Slyusarev G.G. Calculation of Optical Systems. L.: Mashinostroenie, 1975.
(17) Kusch O.K., Mitin A.I. Computer Calculation of Luminous Intensity
Distribution from Symmetrical Specular Surfaces with Extended Light Sources //
Svetotekhnika, 1982, 6, p. 5-8.
(18) Gavrilenkov V.A., Smolyanskii M.F., Trembach V.V, Computer Calculation
of Luminous Intensity Curves of Parabolic Reflector with Cylindric Light Source //
Svetotekhnika, 1982, 3, p. 15-16.
(19) Mazur J., Zagan W. Obliczanie odblysnikow zwierciadlany ch
samochodowych proektorow oswietleniowych, metoda odbie elementarnych //
Arch.Eleklrotechniki. 1983, vol. 32, p. 684-696.

143
(20) Ishchenko E.F., Zhilenkova N.V., Burov A.V. Computer-Aided Design of
Optical Lighting Systems // Svetotekhnika, 1992, 7/8, p. 5-8.
(21) Muigg P. Computer-unterstutzte Reflektor-berechnung // Licht, 1985, 7, p.
475-478.
(22) Whilted T. An Improved Illumination Model for Shaded Display // CACM,
23(6), June 1980, p. 343-349.
(23) Foley J.D., van Dam A. Fundamentals of Interactive Computer Graphics,
Addison-Weslev Publishing Company, 1982.
(24) Komissarov V.D. On Calculation of Specular Luminaires with DRL, Lamps
// Svetotekhnika, 1966, 5, p. 11-17.
(25) Trembach V.V. Lighting Devices. M.: Vysshaya Shkola, 1972.
(26) Karyakin N.A. Lighting devices of Floodlighting and Projection Types, M.
Vysshaya Shkola, 1966.
(27) McCracken D.D., Dorn W.S. Numerical Methods and Fortran Programming.
John Wiley & Sons, New York - London - Sydney, 1965.
(28) Kusch O.K., Rokhlina N.V. Analytical Calculation of Symmetrical Specular
Luminaires with the Aid of Inverse-Ray Method // Svetotekhnika, 1984, 3, p. 7-10.
(29) Forsythe G.E., Malcolm M.A., Moler C.B. Computer Methods for
Mathematical Computations. Prentice-Hall, N.J. 1977.
(30) Medvedev V.E., Parilskaya G.G. Calculation of Illuminance within Image //
Optika i Spektroskopiva, 1966, vol. 21, issue 5, p. 638-642; 1967, vol. 22, p. 819-823.
(31) Umarov G.Ya., Zakhidov R.A., Sokolova Yu.B. Calculation of Irradiance
Field of Elliptical Furnace with Xenon Lamp // Geliotekhnika, 1974, 3, p. 41-46.
(32) Boldyrev N.G. On Light Beam Formation of Parabolic Projector with
Spherical Light Source // Svetotekhnika, 1936, 5, p. 69-70.
(33) Sapozhnikov R.A. Theoretical Photometry, Energiya, 1972.
(34) Moon P. The Scientific Basis of Illumination Engineering. McGraw-Hill
Book Comp., New York - London, 1936.
(35) Nier J. Strahlungsstarkeverteilung eines Parabolscheinwefer mit
Halbkugelstrhler // Bosch Techn.Berichte, 5 (1975) 1.
(36) Giloi W.K. Interactive Computer Graphics. Prentice-Hall. Inc., Englewood
Cliffs, New Jersy, 1978.
(37) Vogl T., Lintner L.C., Pegis R.J. et al. Semiautomatic Design of Illuminating
Systems // Applied Optics, 1972, vol. II, 5, p. 1007-1009.
(38) Gurevich M.M. Introduction to Photometry. L.: Energiya, 1968.
(39) Weis B., Petry K., Willig A. Beleuchtungsstarkeberechnung fur ausgedehnte
Lichtquellen und ideale Spiegelreflektoren // Lichtforschung, 1981, 3, p. 89-96.
(40) Vasin E.G., Stepanov V.N. Computer Calculation of Headlamp Light
Distribution // Svetotekhnika, 1987, 5, p. 13-14.
(41) Smolyanskii M.F. Modelling of Luminance Characteristics of Filaments in
Optical Systems with Mixed Reflection and Transmission // Svetotekhnika, 1991, 5,
p. 8-10.
(42) Korobko A.A. Mathematical Model of Diffusion of Lightguide Optical Slit //

144
Svetotekhnika, 1983, 11, p. 8-10.
(43) Korobko A.A., Kusch O.K. Computer Calculation of Photometrical
Characteristics of Slit Lightguide by Monte Carlo Method // Svetotekhnika, 1979 3,
p. 9-11.
(44) Tregenza P.R. The Monte Carlo Method on Lighting Calculations // Light.
Res. & Techn., 1983, vol. 15, 4, p. 163-170.
(45) Kinameri K., Akazawa K., Awata M. The Monte Carlo Method in the
Predetermination of a Luminous Intensity Distribution // J.Light.& Vis.Env., 1986, vol.
10, 2, p. 57-66.
(46) Korobko A.A. Working-out, Investigation and Application of Mathematical
Model of Illuminating Devices with Slit Lightguides. Thesis. M. 1984.
(47) Basov Yu.G. On Application of Monte Carlo Method in Lighting
Calculations // Svetotekhnika, 1991, 4, p. 6-8.
(48) Marchuk G.l. et al. The Monte Carlo Methods in Atmospheric Optics.
Springer-Verlag, New York, 1980.
(49) Knulh D.E. The Art of Computer Programming. Vol. 2. Addison-Wesley,
Reading, Mass., 1969.
(50) Siegel R., Howell J.R. Thermal Radiation Heat Transfer. McGraw-Hill Book
Company, New York, 1972.
(51) Moon P., Spencer D.E. The Photic Field. Yhe MIT Press, Cambridge, Mass.,
1981.
(52) Fock V.A. Illumination from Surfaces of Arbitrary Shape // Trudy GOI,
1924, 28.
(53) Handbook of Mathematical Functions. lid. by M.Abramowitz and J.A.Stegun.
National Bureau of Standards. Applied Mathematics, Series 55. 1964.
(54) Sofronov N.N. About Luminance of Melal-Halide Lamps // Svetotekhnika,
1988, 12, p. 4-7.
(55) Lewin I. Computer Design of Luminaires // LD&A, 1977, 7-8, p. 26-31.
(56) Happe K. Reflektor-entwicklung am Bildschirm // Licht, 1990, 3(4), p. 303307.
(57) Boldyrev N.G., Tsyplyakov G.R. Luminous Field of Projector // Trudv GOI,
1939, vol. 13, issue 111, p. 20-24.
(58) Ivanov V.I. Determination of Local Radiation Ability of Axially Symmetrical
Source. In the book: Noil-Correct Inverse Problems in Nuclear Physics, Izdat.
Novosibirskogo otd. Akademii Nauk, 1977, p. 118-123.
(59) Krasnov M.L. Integral liquations. M.: Nauka, 1975.
(60) Podlivaev A.F., Shumakova O.M. Approximation of Filament Radiation by
Radiation from Cylindrical Surface. // OMP, 1986, 3, p. 7-10.
(61) Verlan A.F., Sizikov V.S. Integral Equations. Kiev. Naukova Dumka, 1986.
(62) Gershun A.A. Selected Works on Photometry and Illuminating Engineering.
M.: GIFML, 1958.

145

CHAPTER THREE
INVERSE PROBLEM IN OPTICAL SYSTEM DESIGN
3.1. POINT-SOURCE METHODS AND ALGORITHMS
General Equations
The inverse problem means determination of active surface shape of
reflector when intensity distribution or efficiency criterion is prescribed. The
inverse problem is considered to be well posed if its solution: (1) exists; (2) is
unique; (3) is stable when input data are slightly changed.
Existence and uniqueness of solution for specular surface with a point
source were proved by N. G. Boldyrev and V. D. Komissarov [1, 2], Equation
for nonsymmetrical specular surface is presented in the following form

sin( )d sinh( a t )dt


,
(3.1)
cos( ) cosh(a t )
radius-vector of a surface; t ln(cot( / 2)) ;

d (ln r )
where

is

the
a ln(cot(a / 2)) .
V. D. Komissarov showed that by changing the variables the equation can
be reduced to elliptical equation of Monge-Ampere. Recently, similar result
was once again discovered by Schruben [3, 4],
Still, Eq. (3.1) has to be supplemented with a flux equation stating that
reflected luminous flux is equal to incident flux multiplied by reflectance
factor
I `( , ) sin a da d I ( , ) sin d d ,
(3.2)
or
I `( , ) sin a D I ( , ) sin ,
(3.3)

146
where
a

D
.
a

The sign in Eq. (3.2) can be chosen arbitrarily. Keeping in mind that
luminous flux is a positive value, we must take the absolute value of
determinant in Eq. (3.3).

Remote Zone
Axially Symmetrical Surface
In this case we have

0,
0,
1.

Equation (3.1) yeilds

ln r
ln r
a
0,
tan
,

(3.4)

where and are positive when counted in clockwise direction (Fig. 3.1).
Hence, Eq. (3.3) becomes
I `(a ) sin ada pI ( ) sin d .
(3.5)
This equation assumes existence of two types of surfaces:
(1) a surface of elliptical type, where reflected ray intersects the axis of
symmetry (Fig. 3.1a);
(2) a surface of hyperbolical type, where reflected ray docs not intersect
the axis of symmetry (Fig. 3.1b).

Fig. 3.1. Elliptic (a) and hyperbolic (b) surfaces.

147
Surface type can be chosen in advance, if bounding conditions (e. g.
constraints on reflector diameter) are taken into consideration. Equations (3.4)
and (3.5) can be represented in the form

a
I ( ) sin

I `( a) sin a
r
a
r tan
.

(3.6a)
(3.6b)

The obtained system of equations is a consequence of supposition stating


that the laws of geometrical optics arc valid. Similar equations are studied in
optics, acoustics, theory of antennas [5, 6, 7]. These are differential equations
of the first order. Under initial conditions =0, a= a0, r=ro the system (3.6)
constitutes Cauchys problem. Since Eq.(3.6a) has a singularity at a=0, a curve
must start from angle a 0.
Expression (3.6a) is an equation with separable variables; it can be easily
integrated; thus the function of axial ray paths ao= ao() can be obtained. After
substituting into Eq. (3.6b), we gel the quadrature


a( )
r ro exp tan
d .
2
0

(3.7)

While composing an algorithm for solving the problem stated by system


(3.6) or for determination of r() by applying Eq. (3.7), the following stages
must be undertaken:
(1) normalization of fluxes, i. e. instead of given I(), the following has to
be considered

I `(a ) K am I (a ),
where
k

I ( ) sin d

K am

0
ak

I `(a) sin a da

a0

here is an acceptance angle of reflector, is corresponding polar angle


of reflected ray [the end of I()-curve];
(2) the own light of a source has to be taken into account, i. e.
I ``( a) I `( a) I 0 (a) is the required luminous intensity curve;

148
(3) function I ``(a ) is now substituted into Eq. (3.6a).
Cylindrical Surface
For this case we have equations [1, 8] similar to Boldyrevs equation (3.1)
a
I ( )
(3.6a)

I `(a)
r
a
.
(3.6b)
r tan

2
where I() is the luminous intensity of a source, I'() is a prescribed
luminous intensity in transversal plane.
Equations (3.8), as well as Eqs. (3.6), can be solved by methods of
numerical integration or be reduced to approximate quadratures. The source
image must rest within mirror length dimension [9], this is a condition of
applying Eqs. (3.8). Candlepower curve of a designed lighting fixture is
usually prescribed in the form of indicatrix with maximal radius equal to unit:
I ao f (a) . Therefore, the luminous intensity in absolute units of
measurement (cd) is I `(a ) K am I a0 I ( ) , where K am is a scale factor. If
candlepower of a lamp in transversal plane is I ( ) I o , then the specified
curve will be
I `(a ) I o ( K am I a0 1) ,
(3.9)
where K am is a factor of maximal amplification.
Integrating (3.8a) with account of Eq. (3.9) within maximal acceptance
angle max and maximal divergence of reflected rays amax , we obtain

K am ( max amax ) / amax ,

(3.10)

whence

max

1
( K a a ) ,
am max max
amax

where the function amax

0
a

da

can be found by numerical

integration.
Thus, assigning max, amax and reflectance po, we find the constant Kam
being the term of equation (3.8). Now Eqs. (3.8) are well integrable if a
condition of physical feasibility is met: K am I a 1.

149
Proximate zone
Axially symmetric system
Similar to flux equation expressed in terms of luminous intensity, we can
write an equation that defines the local balance of fluxes [10, 11]
2 E ( x ) x dx 2 I ( ) sin d .
(3.11)
Adding Boldyrevs equation to Eq. (3.11), we obtain a system of
differential equations

d
xE ( x )

;
dx
l ( ) sin
dr d
a

r tan
,
d dx
2

(3.12a)
(3.12b)

where relationship between a, x, o and z is

tan a (r sin x) /( R cos H ) .

(3.13)
Under initial conditions =0, x=xo, z=zo Eqs. (3.12) and (3.13) yield
desired profile curvc of axially symmetrical reflector presented in the table
form, x() and z(). The discussion on peculiarities in solving equations like
(3.12) can be found in Schrubens work [4].
E0 Ed 0 .
(3.14)
Cylindrical System
For this case the equations defining the profile curve are perfectly similar
to those applied to axially symmetric system and have the following form

d
E ( x)

;
dx
l ( )
dr
d
a
r
tan
.
dx
dx
2

(3.15a)
(3.15b)

For a light source we take the uniformly radiating tube (or filament) for
which

I I o , Ed Eo (1 x 2 )1.

150
where x x / H , E0 I 0 / H .
In accordance with Fig. 3.2, where d is an angle of divergence, we obtain
the following equation for the balance of integral fluxes
e

x1

[K

( 0)

am

2 1

E ( x ) E0 (1 x ) ]dx I ( ) d ,
0

x0

whence, taking E ( x) E0 1 , we obtain for c o m

K am

m d
,
x1 x 0

(3.16)

where x 0 x0 / H , x1 x1 / H .
Note, that Eq. (3.16) is similar to Eq. (3.10) by structure. Thus, we have
the following system of equations
2

d
[ K am (1 x ) 1 ]

;
dx

dr
d
a
r
tan
,
2
dx
dx

(3.17)

where Kam is calculated front Eq. (3.16).


The sign + fits the condition of ellipticity (see Fig. 3.2). The condition of
physical feasibility is equivalent to condition of positiveness of denominator
in the right side of Eq. (3.17): Kam>1, or

x1 x0
m d

(3.18)

(compare with corresponding condition for remote zone).

Fig.3.2. On energy balance in cylindrical


system.

151
Singular and Critical Points in Equations of Specular Surfaces
Here we consider the conditions of correctness for Cauchys problem
solution in more detail, when dealing with specular surfacc equations, say Eq.
(3.12). We write the first relationship from Eq. (3.12) as d / dx F ( x, ) .
For existence and uniqueness of solution the derivative dF (x, ) / must be
bounded within the whole range of -values from segment [', " ]. Therefore,
the luminous intensity curve of a light source must be a smooth function with
respect to an argument and without naughts within given interval. Almost all
real functions meet this requirement. But still, it is obvious that the point =0
is a singular point. If it is included in the range of reflector angles, the
uniqueness of solution can not be guaranteed.
To eliminate the singularity at null, one can solve the first equation
individually, taking into account that this is an equation with separable
variables.
Integrating, we obtain the flux balance equation (accounting the direct
flux)

sin i ( ) d {E ( x ) Ed ( x)}dx,
0

(3.19)

x0

where Ed(x) is the direct illuminance at setting range [x0, x1] of illuminance
curve.
The flux balance equation defines the relationship implicitly.* By the
implicit-function theorem this equation has a unique solution within the whole
x-range except for the point where the right-side integrand becomes equal to
zero, i. e. E ( x) Ed ( x).
Supposing E ( x) K am e( x) , we obtain an equation for singular points

K ame( x) Ed ( x) ,

(3.20)

where K am is the scale multiplier, e(x) is the prescribed illuminance


distribution.
Points that fit the condition (3.20) are the critical points of solution. The
scale multiplier can be found from expression

It can tic solved numerically, e.g. by applying ZEROIN complex (see Annex).

152

`1

sin i( ) d I ( ) sin d

K am

`0

x1

(3.21)

e( x) x dx

x0

where 0, 1 are the boundary angles of reflector; 0` , 1` are the boundary


angles of direct illumination, i. e.

0` arctan( x0 / H ),

1` arctan( x1 / H ).

Critical points of cylindrical system [Eq. (3.8)] are defined as follows


K am I 0 (a ) 1,
(3.22)
where Kam is calculated from Eq. (3.21) [for isotropic light source see Eq.
(3.10) ].
Duality of the Problem
There is an alternative approach to solution of the problem of profile
determination. One may first set an acceptance angle of reflector max (or 1)
and then find Kam from Eq. (3.21), where Kam is a parameter in the flux balance
equation. Solving Eq. (3.19), we find the ray-tracing function x=x(). Hence,
the right side of Boldyrevs equation is defined as a function of variables r and
[accounting Eq. (3.12) ]. On the other hand, we may prescribe Kam and then
find the angle max from the flux balance equation or solve the system of
equations (3.12).
It can be demonstrated that under the same conditions both approaches
lead to the same result.
Presence of critical points change the matter. Exclusion of these points
leads to narrowing the range [xo, x1] which specifies the illuminance curve,
and, consequently, changcs the parameter Kam. So, according to Eq. (3.21) the
angle of acceptance changes as well. Applying iteration method, one can find
the Kam value that fits admissible limits.
When designing the shape of reflector in luminaire for proximate
illumination, actually, Eqs. (3.15) or (3.18) are used as initial expressions.
Further, the relationship x() is determined explicitly by integrating
numerically the equation of specular surface. We note that each method of
integrating differential equations uses polynomial representations of function
within integration interval. For instance, the fourth order of Taylors truncated
formula is used in

153
Runge-Kuttas method. Precisely from this point of view the efficiency of
integrating methods has to be estimated.
Numerical Models for Software Package
Relevant condition for software package organization is a search of
common features and statements in modeling of different objects. Unificated
modules must be used where it is possible. All noted-above problems can be
solved by the unique method of numerical integration of differential equations,
namely, Adams-Fultons or Runge-Kuttas method.
Subroutine RUNGE from Annex 3 which integrates the system
`
yk f k ( x, y1 , y2 , ... yn ) of ordinary differential equations by applying RungeKuttas method with automatic step selection is strongly efficient. The
following parameters have to be assigned for program operation: the initial
value of independent variable x; the initial values of desired functions y[k]; the
order of system n; the procedure that calculates the right side of equations and
array of derivatives z[k].
C-version of the translated Algol program [12] is given in Annex 3. The
conception being laid into RUNGE complex is similar to that in ZEROIN.
Parameters in the right sides of differential equations are independent data type
and can be transmitted as pointers onto structure.
For reflecting systems being described (n=2), the corresponding
parameters are presented in Table 3.1. The sign before multiplier in Table 3.1
may be arbitrary (sec above).
Table 3.1
System variant

Variable

Function Function
y1
y2

z(1)

z(2)

Remote zone:
axially symmetric
system

axially symmetric
system

cylindrical
system

cylindrical
system
Proximate zone

I `( ) sin
l ( ) sin

r tan

I `( )
l ( )

r tan

xE ( x)
l ( ) sin

r z (1) tan

E ( x)
l ( )

r z (1) tan

154
3.2. METHODS AND ALGORITHMS
PROBLEM FOR LENGTHY SOURCES

OF

SOLVING

INVERSE

The Principles of Solution


Since the light field of a point source can be represented in the form of
independent light tubes, the solution of inverse problem exists. This principle
makes a foundation in vector method proposed by N. G. Boldyrev and A. A.
Gershun [13] for designing the shape of specular surface. Inversion of
operator, which calculates the intensity when the mirror shape is prescribed,
leads to the equation with respect to Gaussian curvature (sec Ch. 2), i. e. to
differential operator of the second order, that further can be reduced to a
system of two first-order equations.
As for a source with finite dimensions, the light-tube structure becomes
sufficiently complex, and reflected tube geometry can not be determined
without knowing the properties of reflecting surface.
We analyse image structure for given direction ao by sending a set of
parallel rays with a flat wave-front towards a reflector (Fig. 3.3a). The extreme
points of a flashed zone (an image) in a meridian plane are determined by
internal rays AoBo and A1B1 that are tangent to a source contour. Angle
coordinates ' and ' of these points can be found, e. g. when appropriate
formulas given in Ch. 2 are used.
We set an angle of observation a1 close to a0; a1 a0 a (Fig. 3.3b).
Transition to a1 > a0 results in image displacement

Fig. 3.3. Isogonal trajectories for a=ao (a) and a=a1 (b); and design of zero and consequent
zones of reflector.

155
by >0 under accepted ray-path scheme. Situation is repeated for a2>a1. We
call the straight lines H1( k ) , H 2( k ) containing all image points for given
direction a aend as isogonal directions (isogones). Isogones form original
tubes that do transfer light source images, but not luminous flux.
Hence, the luminous intensity curve of a reflector can be represented as a
sum of zonal (virtual) candlepowers;
N

I (a) I a( k ) ,
k 1

while

I (ak ) I ( k ) (ak ), I ( k 1) (ak ) 0 , k 1, 2 ...


Algorithm of Composing Specified Luminous Intensity Curve
By using found relations ak H1( k ) , H 2( k ) , we can construct an algorithm
of composing specified luminous intensity curve (LIC) J(a) through the points
a0 , a1 , ..., an (a0 0, a1 5o , a2 100 ... in Fig. 3.4).
Step 1. To determine an extreme point of the zero-zone B1 in such a way
that I (1) J a 0 .
Step 2. To compute the zero-zone LIC at the points a0 , a1 , ..., an and the
lack of luminous intensity at the point a1 :

dI (a1 ) J (a1 ) I (1) (a1 ).


Step 3. To find the point B2 so that the virtual candlepower yeilds

I (1) (a1 ) dI (a1 ).


Step 4. To calculate the first-zone LIC at the points a0 , a1 , ..., an and the
lack of luminous intensity
2

dI (a2 ) J (a2 ) I ( k ) (a2 ).


k 1

Cyclicity of calculations is obvious.


The said method resembles the well-known method of interpolating
continuous function in terms of a system of orthogonal functions. We note, that
known method of filling the LIC with zonal curves

156

Fig. 3.4. On algorithm of composing specified candlepower curve from set of zonal curves.

157
does not provide candlepower preservation during transition from one zone to
another, bccause zones with constant angular width have been regarded there.
In our ease we do not fix the zone angular dimension but prescribe the
directions ak (isogone family). Flashed zones overlap with each other, so, that
the point B1 of a new surface segment becomes the inner point of the first zone
(Fig. 3.3b).
Construction of Mirror Shape
We describe the method of constructing (synthesis) a surface from a family
of parabolas. Let the initial point Bo(xo, zo) be set (Fig. 3.3b). The straight line
Ao Bo can be regarded as a line, along which the focus F0 of parabola runs,
while the axis of parabola is in parallel with isogones H1, H2. Assigning the
lack of intensity dI(ao) and using successive approximations, we find the
point Fo of parabola B0B1 to satisfy the following inequality

I B1B2 (a0 ) I (a0 ) 1 ,


where is the prescribed error.
For a1 a0 the image extinguishes at point B0 and appears again at point

B1` (the isogone H1), while the internal ray moves from point A0 to point A1.
Now we may assume that segment A'A1 is a light source, and A'B1 is the foci
line of consequent parabola are B1B2 of the second zone.
Setting up dI(a1), we define focus point F1 in such a way that the following
inequality is valid

I B1B2 (a1 ) I (a1 ) .


The process is being continued until the complete image is constructed.
However, the extension of parabolic segments may appear impossible if
the prescribed curve possesses sharp overfalls.
Obviously, two solutions are possible within described method; within
elliptic class of reflectors when isogones turn into positive direction of angles a
(see Fig. 3.3); and within hyperbolic class of reflectors when isogones turn into
negative direction. The lack of luminous intensity is calculated in each cycle
by inverse-ray method (see Ch. 2).
Since segment A1B1 remains fixed during the extention of a parabola
segment, this procedure provides the smooth first-order connection of curves.

158
Cylindrical System
V. D. Komissarov proved that the solution for cylindrical system could be
obtained in finite form [9]. We suppose to reproduce the LIC at the transversal
plane. Let point B1 be the initial point of reflector profile or the succesive point
where calculations were stopped. Starting from the condition that virtual
luminous intensity has a prescribed value, we find parabola are B1B2.
Obviously, it is sufficient to set a projection H=H2-H1 (Fig. 3.5). We take an
advantage of the fact that optical path-ways for the rays towards the front of
arbitrary reflected wave, say C1C2, must be equal to each other. Consequently,
the following path-ways have to be equal: A1B1+B1C1=L and A2B2+B2C2=L.
From the triangle A1B1C1 we have

( A1 B1 A1 A1` ) cos(1 a ) OC1 OA1`,


or

A1 B1 R cot(1 a ) [ H 1 R sin 1 ( a)] cos 1 (1 a), (3.23)


whence

A1 B1 H 1 cos 1 (1 a) R tan 1 (1 a ) ,

Fig. 3.5. Construction of zero-zone profile (point B2) for cylindrical source.

(3.23)

159
or

A1 B1

tan(1 a )

cos 1 (1 a )

From the same triangle we obtain

B1C1 A1`C1 tan(1 a ) H 1 tan(1 a ) R cos 1 (1 a),


or

B1C1

H1

cos 1 (1 a )

tan(1 a)

From the triangle A2B2C2 we find

A2 B2 A2` A2 ( A2` O OC 2 ) cos 1 ( 2 a );


A2 B2 R cot( 2 a) [ R sin 1 ( 2 a ) H 2 ] cos 1 ( 2 a ),
whence

l2 A2 B2 H 2 cos 1 ( 2 a ) R tan( 2 a),


or

A2 B2

H2

tan( 2 a )

cos 1 ( 2 a)

Further

B2C2 A2` C2 tan( 2 a) H 2 tan( 2 a ) R cos 1 ( 2 a ),


or

B2C 2

H2

cos 1 ( 2 a )

tan( 2 a )

Hence, we obtain

L1 A1B1 B1C1 ( H1 R)[cos1 (1 a ) tan(1 a )]. (3.24)


Now we have an equation for determining parameters of point B2

( H 2 R)[cos1 ( 2 a) tan(1 a)] L1 ,


where L1 is calculated from Eq. (3.24).

(3.25)

160
With the aid of Eq. (3.25) we find angle 2 and then other elements:

sin( 2 a )

L12 ( H 2 R) 2
;
L12 ( H 2 R) 2

tan 2 R / l2 ;

2 900 2 2 ;
r2 2 L22 R 2 .
In order to pass over to the next parabola segment wc have to find the
initial point X of the first isogone. Wc introduce the following denotions (Fig.
3.6):
, r are the polar coordinates of point X in the coordinate system having
the centre at the point O;
, t are the polar coordinates in the local system having the centre at the
point F;
2 1 is an angular dimension of the source viewed from the point F;
is an angular dimension of segment FX viewed from the point O;
' is an angle defining the direction of the ray A1X after reflection.

Fig. 3.6. Extension of consequent zone to zero-zone.

161
From the triangle OFX we have

r2 OF 2 t 2 2t OF cos( 1 1 );

(3.26)

OF sin( 1 1 ) r sin .

(3.27)

`
1

From the triangle A OX we obtain

r sin( ` ) R.

(3.28)

We add parabola equation to Eqs. (3.26)-(3.28)

FB1 (1 cos 1 )
.
1 cos 1

(3.29)

Equations (3.26)-(3.29) define the coordinates , t , r , of the point X.


Polar coordinate of unknown point can be found from equation
r sin t sin( 1 1 ) ,
or
r cos( 2 1 ) t sin( 1 1 ).
(3.30)
From Figs. 3.5 and 3.6 we find the constituent parameters of Eqs. (3.26)(3.30):

FB1 A1 B1 A1 F A1 B1 R cot(1 2 ) / 2;

1 (1 2 ) / 2, 1 1

(3.31)

The virtual luminous intensity of subsequent segment must ensure the


following

hvirt h(a`) req h(a`) XB2 .


The described method of designing specular surface by composing
parabola segments provides smooth connection between curves up to their
second derivatives. From this point of view the method is similar to that of
Boldyrev.

162
Differential Equations for the Profile of Specular Surface
Like any other evolutionary process, generally, the procedure of designing
the profile of specular surface, that ensures given candlepower or illuminance
distribution, is described by a system of differential equations similar to those
proposed by Boldyrev. We derive these equations for axially symmetrical
system combined with a source stretched along the axis OZ (Fig. 3.7); it can be
a tubular source, in particular [14].
The first stage of calculation
The first segment, or the zero-zone of reflector, is determined by basing on
the condition of ensuring the prescribed luminous intensity Jao at the direction
a0. For this segment the mirror surface equation holds true (Fig. 3.7a)

dz
a0
tan
.
dx
2
We rewrite it in another form:

dz
dx
a0

tan
.
dx
dz s
2

(3.32)

where linear coordinate zs is counted along the light source from system
origin O, which is the centre of the source with the length of 2l.

Fig. 3.7. On deduction of differential equations for spccular surface: a - the first stage; b the second stage.

163
From the same figure we obtain

H x cos a0 z sin a0 ;

(3.33)

dH dx cos a0 dz sin a0 ;

(3.34)
Furthermore, we can suppose a relation between image projection dH and
corresponding displacement dzs of a ray along the source:
dH k aoT ( z s )dz s ,
(3.35)
where kao is the amplification factor.
The function T(zs) defines movement rate of internal ray along the source.
It characterizes the multivalencc of solution. For simplicity, we set up T(zs)=1.
From Eqs. (3.33) and (3.34) we can find dx/dzs and substitute it into Eq. (3.32).
After transformations we obtain

dz
sin( a0 ) / 2
k ao
;
dzs
cos( a0 ) / 2
dx
cos( a0 ) / 2
k ao
.
dzs
cos( a0 ) / 2

(3.36)

The right sides of Eq. (3.36) are the functions of x, z, zs, if we account the
following relationship

arctan

x
.
zs z

(3.37)

There exists the only solution of parametrical equations (3.36) under initial
conditions x=x0, z=zo, and it can be found by applying numerical methods. The
right sides of Eq. (3.36) can be reduced to parameters zs and z. Integrating Eq.
(3.35), we obtain

H H 0 k ao ( z s l ),
whence, accounting Eqs. (3.33) and (3.37),

arctan

H o k ao ( z s l ) z s sin ao
.
cos ao ( z s z)

(3.38)

Equation (3.36) together with Eq. (3.38) defines the profile of the zerozone for zs=-l, z=z0.
The value of kao can be found by successive approximations. In order to
obtain the first approximation we integrate Eq. (3.35)

164

dH k dz
ao

HO

whence

k ao

H ho H

.
2l
2lc

The kao-value can be estimated approximately if the image along direction is known. If >rc/ro, where rc is the tube radius, and ro is the radiusvector of initial point, then according to theory of direct calculation (see Ch. 2)
we obtain the expression for the flashed area

S fl H 2r

1
.
sin a

Assuming source luminance to be Lo, we get

k ao

J a sin a
.
4 Lo rc lc

(3.39)

For =0, we, obviously, have an assessment from below

k ao

J0
,
2 lc ( Rmax r0 sin 0 )

(3.40)

where Rmax is a radius of exit aperture.


Generally, internal rays corresponding to isogonal trajectories form a
caustic, the specific case of which is the focus (when parabolic zones are
selected).
At the first stage of calculation by inverse-ray method we find the LIC for
the zero-zone (compare with the case of cylindrical source), and this serves as
initial approximation for the second stage.
The second stage
On the second stage we add to already-designed surfacc segment another
one that reproduces the deficient of luminous intensity at given direction.
Constructed in this way, the zone becomes initial for the next zone, and the
calculation is being continued until the prescribed LIC is composed.
If a light source is substituted by elementary segment, the scheme

165
of ray tracing for added zone is similar to that of the zero-zone (Fig. 3.7b). This
corresponds to discussed-above composition from paraboloid zones.
We have the following relationships [see Fig. 3.7b]
dz s sin r da;
(3.41)

r [( x d / 2) 2 ( z l ) 2 ]1/ 2 ;
xd /2
tan
.
z lc

(3.42)
(3.43)

Taking into account Eq. (3.35), we obtain (omitting index 0)

dz
r sin( a ) / 2
k ao
;
da
sin cos( a ) / 2
dx
r cos( a ) / 2
k ao
.
da
sin cos( a) / 2

(3.44)

Equations (3.44) combined with Eqs. (3.42) and (3.43) give a solution
under initial conditions x=x1, a=a1, z=z1 (where parameters x, a, z characterize
the terminal point of the zero-zone).
Transition from to +d gives the following deficiency of luminous
intensity
dI a [ J `( a) I `(a)]da,
(3.45)
where I'() is a derivative of luminous intensity of reflector.
Since the deficiency of intensity can not be negative, we get
J `(a ) I `( a).
(3.46)
Condition (3.46) is similar to that for the kernel of Fredholms first-genus
integral equation. If, e. g. the kernel has a continuous derivative with respect to
argument, then the right side of equation must contain the continuous
derivative with respect to the same argument as well.
The peculiarities in solution of the system of equations are such, that we do
not know the relationship k=k(a) beforehand. We can only prescribe it
approximately in accordance with Eqs. (3.39) and (3.40)

k a ~ J (0)
for =0
k a ~ J (0) sin a
for
>0

166
Since the source possesses finite angular dimension , the construction of
profile has to be terminated at the point max, where the positiveness condition
of dI is violated [see Eq. (3.45)]. Smoothness of calculated profile and
curvature sign constancy are the important properties of obtained solution; this
result is ensured by the ray-tracing scheme being chosen.
3.3. CALCULATING SCHEMES FOR REFLECTOR OPTIMIZATION
(A) In fixtures where extended radiating sources are applied the profile of
reflector has to be chosen in a way that flux losses associated with return of
rcflcctcd rays back onto lamp bulb are reduced to minimum. Let a source be of
elliptic form (sec Ch. 2). We demand that marginal (tangent) ray after
reflection must follow the same direction. Consequently, a point of reflector
circumscribes an evolvent of ellips. Evolvent equation for a curve under
arbitrary parametrization r(t) is as follows

r `(t )
~
r r (t )
r `2 (t )

r `2 (t ) dt ,

or

~
x x (t )
~
z z (t )

x`(t )
2

x` z `
z `(t )

x`2 z `2

x`2 z `2 dt ,
(3.47)

x` z ` dt.

Substituting ellips equation (see Ch. 2) into Eq. (3.47), we obtain

~
x b{cos t

a sin t

[ E (e, t ) E (e, t0 )]};


a 2 cos 2 t b 2 sin 2 t
a cos T
~
z a{sin t
[ E (e, t ) E (e, t0 )]},
2
a cos 2 b 2 sin 2 t

(3.48)

where E(e, t) is the elliptic integral of the second genus; t0 is a parameter,


the choice of which depends on initial point of evolvent.

167
If a source has a circular form (e=0), then evolvent equation lakes the
form*
~
x a{cos t (t t 0 ) sin t}; ~
z a{sin t (t t 0 ) cos t}. (3.49)
(B) It is a known fact that paraboloid makes the most concentralion of
luminous flux emitted by sources of small angle dimensions. Light-tubes
associated with such sources represent a system of concentric cones.
Demanding reflected flux to be aimed at prescribed direction or towards fixed
point, we impose certain conditions on normals at points of reflector; this leads
to differential equations with known solutions - they are paraboloid and
ellipsoid. For extended sources light-tubes are more perplexed. For a source of
uniform luminance they comprise a family of hyperboloids with tangenls being
the bisectrices of corresponding angles between radius-vectors connecting
marginal points of a light source and a point
of reflector (Fig. 3.8).
We demand a light-vector which is
tangential to hyperbola to run in parallel to
the axis after reflection.
It is easy to see that the similar scheme
of ray-tracing corresponds to more dense
packing of elementary maps on a plane near
the point =0. The bisecrix of an angle at the
vertex forms an angle with the axis equal to
(1+2)/2 (Fig. 3.8).
Obviously, the polar angle of normal is
equal to (1+2)/4. Hence, the profile Fig. 3.8. On construction of bisectroid.
equation of the optimal reflector with
elliptical source takes the following form
[see Ch. 2]

dz
2 zx
.
tan arctan 2
2
2
2
dx
(z a ) (x b )
4

(3.50)

If the source has a filament form of 2l length, then Eq. (3.50) yeilds [15]

dz
1
2 zx

tan arctan 2 2
.
2
dx
z l x
4
*

(3.51)

Reflectors with profiles described by Eq. (3.49) are used, in particular, in luminaires with
fluorescent lamps.

168
Under prescribed initial conditions x=xo, z=zo, Eqs. (3.50) and (3.51)
define the only curves lhat meet the requirement of maximum concentration
along axial direction [16].
3.4. OPTIMIZATION OF FOCAL PARAMETER OF REFLECTOR
In floodlight design a throat diameter 2a of rcflcctor and an exit-aperture
diameter 2b arc usually determined by constructive, thermal and lighting
requirements. Under given constraints the choice of focal parameter p of
paraboloid, that provides the maximal utilization factor of luminous flux, is
difficult.
Let a source be placed at a parabola focus loeated at the centre of polar
coordinate system (Fig. 3.9). Polar angles that define the throat and the margin
of reflector are as follows

0 arccos[( p 2 a 2 ) /( p 2 a 2 )];
e arccos[( p 2 b 2 ) /( p 2 b 2 )].

(3.52)

The flux falling onto reflector within angles o and e is a function of focal
parameter
e ( p )

( ) 2

I ( ) sin d ,

(3.53)

o ( p)

where I() is candlepowcr curve of a source.


The necessary condition for reaching maximum of utilization factor yeilds

d
0.
dp

(3.54)

Fig. 3.9. Parameters of paraboloid


reflector.

169
Differentiating Eq. (3.53), we get

d
d
d
2 [ J ( c ) sin c c J ( o ) sin o o ] 0 .
dp
dp
dp

(3.55)

It follows from Eq. (3.52) that

d o
sin o d c
sin e

'

,
dp

dp

(3.56)

hence, the condition (3.54) can be transformed

I (0 ) sin 2 0 I (c ) sin 2 e .

(3.57)
Equation (3.57) can be regarded as an equation in p, if we express 0=0(p)
and e=e(p) explicitly from Eq. (3.56)
We add to Eq. (3.57) one more equation expressing the fact that the points
corresponding to 0 and e lie at the same parabola

1 cos o
1 cos e
b
.
sin o
sin e

(3.58)

If Eqs. (3.57) and (3.58) have a single solution p0 that meets the condition

d 2
dp 2

0,
po

then p0 defines the desired parabola.


If the function I() is symmetrical with respect to =/2, it becomes easier
to solve Eq. (3.57). For analytical calculations the following representation of
I() is suitable
N

I ( ) g m sin m .
m 0

Then Eq. (3.57) takes the following form


N

(sin m 2 0 sin m2 e ) 0 .

(3.59)

m 0

If all gm are nonnegative, then all solutions of Eq. (3.58) meet the condition
cos 0 cos e 0 .

170

Fig. 3.10. Relation between flux falling


onto reflector and focal parameter p: p1 = a,
po ab , p2 b;

1 I()-cos;
2 I()-cos-1;
3 I()-sin.

This equation yeilds the following solution


(3.60)
p o ab .
If all gm are nonnegative, then Eq. (3.58) can have different solutions, but
for majority of reasonable curves I() the solution being found is the global
maximum (Fig. 3.10).
Obtained results are applicable when angular dimensions of a source do
not exceed the angle a , within which flux concentration is required. In case
when this condition is violated, there is a chance to choose an optimal in this
sense parameter p by considering relationships F=F(p) and ( p )
simultaneously.
When overall dimensions of reflector are specified, the optimization of
paraboloid focal parameter enables to increase utilization factor of luminous
flux. Conversely, when utilization factor is prescribed, it allows to decrease the
overall dimensions of reflector.
3.5. CALCULATION OF MILLING-CUTTER TRAJECTORY
If a surface is treated by cutting instrument with round head, the centre of
instrument moves along parallel surface which is apart from the former at a
distance equal to milling-cutter radius R. Surface parallax also has to be taken
into account when specifying tolerance on thickness of metal sheet.
If surface is set in parametric form, i. e. r=r(u, v), then the corresponding
surface which is displaced at a distance d along normal n(u, v) can be described
as follows
r` r (u , v ) d n (u, v).
(3.61)
If G and G' are the matrices of the first fundamental forms of said surfaces,
then

171

G n` G (1 dk1 )(1 dk1 )n

(3.62)

where k1 and k2 are the principal curvatures of initial surface.


The radii of the principal curvatures in biased surface [17] are

k1` k1 (1 dk1 ); k 2` k 2 (1 dk 2 ).

(3.63)
If surface is set in terms of its sections then a software based on spline
representation of curves (see Annex) is suitable for calculating milling-cutter
trajectory. Number of points Ne or point spacing h can be adopted as the input
data in this case.
Algorithm for calculating milling-cutter trajectory is as follows:
(1) to introduce arrays of section coordinates X(I), Y(I), I 1, N ;
(2) to prepare spline coefficients for input arrays;
(3) to find the number of layout points (for given step) N=(X\N] X[l])/(h
+ 1);
(4) for each interpolated point (x, y) in a cycle with respect to ke 1, N to
do the following:
(a) to find the normal n = [nx, ny ]T;
(b) to calculate the coordinates of parallax surface

x` x Rnx ; y ` y Rny .
The calculated points define linear segments of milling-cutter trajectory.
Nevertheless, the milling-cutter radius, obviously, must not excced the minimal
curvature radius of a surface.
Appropriate reference list on computer-aided manufacture of surfaces with
regard of tolerances and other restrictions can be found in [18].

172
REFERENCES
1. Boldyrev N.G. About Calculation or Asymmetrical Specular Reflectors
//Svetotekhnika, 1932, 7, p. 7-8.
2. Komissarov V.D. The Foundations of Calculating Specular Prismatic
Fittings // Trudy VEI, Issue 43, 1941, p. 6-61.
3. Schruben J.S. Formulation of a Reflector-Design Problem for a Lighting
Fixture // J.Opt.Soc.Am. 1972, Vol. 62, 12, p. 1498-1501.
4. Schruben J.S. Analysis of Rotationally Symmetric Reflectors for
Illuminating Systems // J.Opt.Soc.Am. 1974, Vol. 64, N9 I, p.55-58.
5. Keller J.B. The Inverse Scattering Problem in Geometrical Optics and
the Design of Reflectors // JRE Trans, of Antennas and Propagation, Vol. AP7, Apr. 1959, 2, p. 146-149.
6. Kinber B.B. The Solution of Inverse Problem in Geometrical Acoustic //
Acoustic Journal, 1955, Vol. I, 3, p. 221-225.
7. Cornbleel S. Microwave Optics. Academic Press. New York San
Francisco London, 1976.
8. Wolber W. Berechnung vir Reflektoren fur beliebigc Lichtverteilungen
// Lichttechnik, 1970, Vol. 22, 12.
9. Komissarov V.D. On Calculation of Cylindric Specular Reflectors for
Incandescent Lamps with Tungsten-Iodide Cycle // Materialy IX nauchnotekhnich. konferentsii po osvetit. priboram, Ternopol', May 1969, M.:
Informelektro, 1969, p. 69-86.
10. Stanioch W. Calculation of Reflecting and Refracting Optical
Elements // Light. Res & Tcchn. 1991, Vol. 3, 3, p. 145-149.
11. Rymov A.I., Skoblova V.I. Melhod of Calculating Reflector Profile in
Imitators of Solar Radiation // Svetotekhnika, 1978, 3, p. 3-5.
12. The Library of Algorithms lb-50b. Reference Book. M.: Sovetskoe
Radio, 1975.
13. Boldyrev N.G., Gershun A.A. Vector Method of Calculating
Symmetrical Specular Reflectors // Svetotekhnika, 1936, 1, p. 7-9.
14. Korobko A.A., Kusch O.K. Construction of Mirrow Surface of a
Luminaire with Lengthy Light Source // Svetotekhnika, 1982, 3, p. 3-6.
15. Inventor's Certificate 1227908, USSR, MKI F21 V7/04. Irradiating
Installation // Discoveries and Inventions, 1986, 16.
16. Gehring A.P., Holten P.A.J. A System Approach to Incandescent
Reflector Lamp Development // J. of IES, Summer 1990.
17. Willmore T.J. Differential Geometry. Oxford University Press, 1958.
18. Faux I.D., Pratt J. Computational Geometry for Design and
Manufacture. Ellis Horwood Ltd., 1979.

173
ANNEX 1
D.Yu. Chepelevskii has composed program texts given below.
PROGRAM COMPLEX ZEROIN
*****************************************************************************
*
MATM_EXT.H Header file for declarations of mathematical routines
*
*
and macros.
*
*
*************************************************************************
#ifndef DEFMATH_EXT
/* Prevent second reading of
*/
#define DEF_MATH_EXT 1
/* these definitions.
*/
#include ath.h
/*______________________Definitions_________________________
*/
/* 2*Pi.
*/
#define M_2PI 6.28318530717958647692
/* Pi.
*/
#define M_PI 3.14159265358979323846
/* Pi/2.
*/
#define M_PI_2 1.57079632679489661923
#define BEGIN {
/* Left bracket.
*/
#define END }
/* Right bracket.
*/
/*_______Constants for use with AllcCurv (the option value)____
*/
# define ALLOC_ARG 1
#define ALLOC_FUN 2
#define ALLOC_B
4
#define ALLOC_C
8
#define ALLOC_D
16
#define ALLOC_BCD (ALLOC_B I ALLOC_C I ALLOC_D)
/*________Constants for type of curve interpolation______________
*/
#define INTER_LINEAR 0 /* Linear interpolation.
*/
#define INTER_SPLINE 1 /* Spline interpolation.
*/
/*__________________Macros______________________________
*/
/* Conversion from radians to degrees.
*/
#define DEG(x) ( (x)*180./M_PI )
/* Conversion from degrees to radians.
*/
#define RAD(x) ( (x)*M_PI/180. )
/* Whether the value is odd.
*/
#define ODD(x) ( (x) & 0x0001 )
/* Whether the value is even.
*/
#define EVEN(x)
( !ODD(x) )
/* Choice of maximum value.
*/
#define MAX(x,y)
( ( (x) > (y) ) ? (x) : (y) )
/* Choice of minimum value.
*/
#define MIN(x,y)
( ( (x) < (y) ) ? (x) : (y) )
/* Square of value.
*/
#define SQR(x) ( (x)*(x) )
/* Cube of value.
*/
#define CUBE(x)
( (x)*(x)*(x) )

174
/* Sign of value. Returns -1 if x 0,
/* 0 if x = 0 or I if x > 0.
#define
SIGN(x) ( ((x) >0) ? (1) : ( ((x) < 0) ? (-1) : (0) ) )
/* Inversion of two values.
#define /*
INVERS (a,b, tmp) tmp = a; a = b; b = tmp
/*______________________________Definitions of data types
typedef struct
/* Data structure of coefficients arrays
{
/*for spline interpolation.
int init_cfs;
/*Flag of coefficients initiation.
unsigned last_knot;
/*The knot number from last call.
/* If coefficients are initiated, flag
/* is equal to I.
Double
*b;
/* Pointer to the array.
Double
*c;
/* Pointer to the array.
Double
*d;
/* Pointer to the array.
} SPLN_COEFS;
typedef struct
/* Data structure for curve description.
{
Char * name;
/*Name of the curve.
Int
option;
/*Bitwise OR-ing of the following:
/*ALLOC_ARG allocate arguments array
/*ALLOC_FUN allocate function value array
/*ALLOC_BCD allocate spline coefficients
/*arrays.
Int_typeinter;
/*Type of interpolation.
Unsigned
nmb;
/*Number of points on the curve.
Double
*arg;
/*Array of curve arguments.
Double
*fun;
/*ray of curve function values.
SPLN_COEFS cfs;
/* Structure of spline interpolation coefficients.
} CURVE;
/*
/* Type of working function for
/* use in the function Zeroin.
typedef double FUN_ZEROIN( double arg, void *prm );
typedef struct
/* Data structure for calculating the
{
/* root in the function Zeroin.
FUN ZEROIN
*fun;
/* Pointer to the working function used
/* in the function Zeroin for root
/* calculation
Void
*prm;
/* Pointer to the parameters structure
/* used in the working function.
} ZEROIN;
/*
/* ___________________________Function declarations_________
void InitCurv
( CURVE* );
/* Initializes pointers to the arrays.
int AllcCurv
( CURVE* ) ;
/* Allocates arrays for curve.
void FreeCurv
( CURVE* ) ;
/* Frees arrays allocated with AllcCurv.
unsigned BnrSrch
(double,
/*Finds the left knot.
CURVE* );
/*
void Line
( CURVE* ) ;
/*Initializes coefficients for linear
/*interpolation.

*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/

175
void

Spline

( CURVE * );

double

Seval

void

Quadr

int

Zeroin

( double,
CURVE *,
double *,
double * );
(CURVE *,
double [] );
(double,
double,
ZEROIN *,
double,
double * );

#endif

/* Initializes coefficients for spline


/* interpolation.
/* Chooses function value from the curve
/* using spline or linear interpolation.
/* Finds values of first and second
/* derivative.
/* Integrates the function.
/*
/* Calculates argument value when
/*function is equal to zero.
/*
/*
/*
/* Ends #ifndef DEF_MATII_EXT

*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/

#include malh_ext.h
/ ********************************************************************************
* Name Zeroin
- finds the root of equation with one variable.
*
*
*
* Synopsis error
- Zeroin( ax,bx, f, tol, root );
*
*
int error
- Code of error;
*
*
double ax,bx
- Left and right borders of initial interval.
*
*
ZEROIN *f
- Pointer to the structure with pointer to the
*
*
working function (that allows to calculate
*
*
f(x) for any point on the interval [ax,bx])
*
*
and pointer to the parameters structure used
*
*
in the working function.
*
*
double tol
The desirable length of indefinite interval.
*
*
double *root}
Pointer to the root value.
*
*
*
*
Description
This function is intended for calculation of real zero of *
*
function f(x) = 0. Zcroin combines reliability of bisection
*
*
method and speed of chord method. It decreases the length
*
*
of indefinite interval until |b-a| < tol + 4*eps*fabs(b).
*
*
After this b is equalized to value of root.
*
*
*
*
Zeroin represents C-realization of Forsythes fortran program.
*
*
*
* Returns error Error code is equal to -1 or 1 if function
*
*
values on the borders of the interval have the same sign
*
*
and the root cannot be found. Otherwise 0 is returned.
*
******************************************************************************** /
int
Zeroin ( double ax, double bx, ZEROIN *f, double tol, double *root )
BEGIN
double a,b,c,d,e;
double eps=1.,
/* Relative machine accuracy.
*/
toll-2.;
/* High border of error.
*/

176
double fa,fb,fc;
double xm,p,q,r,s;

/* Va!ues of the function.

*/

/* Calculate relative machine accuracy.

*/

/* Find function values on the borders.


a = ax; b = bx; fa = (f->fun)(a,f->prm); fb = (f->fun)(b,f->prm);
/* If function values on the borders
/* less then permitted error, root is
/* equated to the value of interval
/* border (a,b).
if( fabs(fa) < tol ) { *root = a; return( 0 ); }
if( fabs(fb) < tol ) { *root - b; return( 0 ); }
/* If function has the same sign on the
/* borders return its sign.
if( SIGN(fa)*SIGN(fb) > 0 ) return( SIGN(fa) );
/*__________________________________Main loop___________________________
do {
c = a; fc = fa; d = b - a; e = d;
Ll: if( fabs(fc) < fabs(fb) ) {
a = b; b = c; c - a;
fa = fb; fb - fc; fc = fa;
}
/* Check up the permitted error
toll - 2.0*eps*fabs(b) + 0.5*tol; xm - 0.5*(c-b);
if( (fabs(xm) <= toll) I I (fb 0.0) ) {
*root = b;
return( 0 );
}
if( (fabs(e) >= toll) && (fabs(fa) > fabs(fb)) ) {
/* Reverse the square interpolation.
q = fa/fc; r = fb/fc; s = fb/fa;
p = s*( 2.0*xm*q*(q-r)-(b-a)*(r-1.0) );
q = (q - 1.0)*(r - 1.0)*(s - 1.0);
if (a- c ) {
/* Linear interpolation.
s * fb/fa; p = 2.0*xm*s; q = 1.0 - s;
}
if( p > 0.0 ) q *= -1.;
p =* fabs(p);
/* Whether the interpolation is
/* acceptable.
if( (2.0*p) < (3.0*xm*q-fabs(toll*q)) && (p<fabs(0.5*q*e)) ) {
e = d; d = p/q; goto L2;
}
}
d = xm; e = d;
/* Bisection.
L2:
a = b; fa - fb;
if( fabs(d) > toll ) b += d;
else b += (xm > 0) ? fabs(toll) : (-fabs(toll));
fb = (f->fun)(b,f->prm);
} while( (fb*(fc/fabs(fc))) > 0.0 );
goto Ll;
END

*/

while( toll > 1.) { eps *= 0.5; toll - 1. + eps; }

*/
*/
*/
*/

*/
*/
*/

*/

*/

*/

*/
*/

*/

177
ANNEX 2
PROGRAM COMPLEX SPLINE
#include math_ext.h
********************************************************************************
* Name
Line
initializes coefficients for linear interpolation.
*
* Synopsis
void Line( crv );
*
CURVK *crv
Pointer to the structure of curve arrays.
*
* Description
This function calculates coefficients for linear interpolation:
*
B and C.
*
*
Arguments of the curve must be in strictly increasing order.
*
* Returns
Nothing.
********************************************************************************
/*________________________Internal definitions ______________________
#define N crv->nmb
/* The number of points on the curve.
#define X crv->arg
/* Array of curve arguments.
#define Y crv->fun
/* Array of curve function values.
#define B crv->cfs.b
/* Arrays of interpolation coefficients.
#define C crv->cfs.c
/*
#define D crv->cfs.d
/*
void Line
BEGIN
unsigned i;

*
*
*
*
*
*
*
*
*
*
*
/
*/
*/
*/
*/
*/
*/
*/

( CURVE *crv )

/* Whether coefficients were initialized.

*/

/* Set flag that coefficients are


/* initialized.

*/
*/

if( crv->cfs.init_cfs | | N < 2 ) return;


for( i=0; i < N-1; I++) {
B[I]=(Y[i+1]-Y[i])/(X[i+l]-X[i]);
C[i] - D[i]= 0;
}
B[N-1] = B [N-2];
C[N-I] D[N-1] = 0;
crv->cfs.init_cfs = 1;
END

178
#include math_ext.h
/ **********************************************************************************
* Name
Spline
initializes coefficients for spline interpolation.
*
*
*
* Synopsis
void Spline( crv );
*
*
CURVE *crv
Pointer to the structure of curve arrays.
*
*
*
* Description
This function calculates coefficients for cubic spline
*
*
interpolation: B,C and D.
*
*
*
*
Arguments of the curve must be in strictly increasing order.
*
*
*
*
Spline represents C-realization of Forsythe's fortran program.
*
*
*
* Returns Nothing.
*
/ **********************************************************************************
/*______________Internal definitions_____________________
#define
N
crv->nmb
/* The number of points on the curve.
#define
X
crv->arg
/* Array of curve arguments.
#define
Y
crv->fun
/* Array of curve function values.
#define
B
crv->cfs.b
/* Arrays of spline coefficients.
#define
C
crv->cfs.c
/*
#define
D
crv->cfs.d
/*

*/
*/
*/
*/
*/
*/
*/

void Spline( CURVE *crv )


BEGIN
unsigned ib,i;
double
t;

/* Counters.
/* Variable for temporary use.

*/
*/

/* Whether coefficients were initialized

*/

/* Construction of three diagonal system.

*/

/* Border conditions.

*/

if( crv->cfs.init_cfs I I N < 2 ) return;


if( N !=2 ) {
D[0] - X[1] - X[0];
C[1] - (Y[1]-Y[0]) / D[0];
for( i=l; i < N-1; i++) {
D [i]=X[i+l] - X[i];
B [i]=2 * (D[i-1]+D[i]);
C[i+1] - (Y[i+l]-Y[i]) / D[i];
C[i] = C[i+1] - C[i];
}
B[0] - -D [0];
B[N-1] * -D[N-2];
C[0] - C[N-1] = 0.;

if(N !=3 ) {
C[0] C[2] / (X[3]-X [1]) - C[1] / (X[2]-X[0]);
C[N-1] = C[N-2] / (X [N-1]-X [N-3]) C[N-3] / (X[N-2]-X[N-4]);
C[0] = C[0] * D[0] * D[0] / (X[3] -X[0]);

179
C[N-1] *= -D[N-2] * D[N-2] / (X[N-l]-X[N-4]);
}
/* Direct passage.

*/

for(i=1; i<N; i++ ) {


t=D[i-1]/B[i-I];
B[i]-= t * D[i-1];
C[i]-= t * C[i-1];
}
C[N-1] /= B[N-1];
/* Indirect substitute.
for( ib=0; ib < N-1; ib++) {
i = N - 2 - ib;
C[i] = (C[i] - D[i] * C[i+1]) / B[i];
}
B[N-1] - (Y[N-I]-Y[N-2]) / D[N-2]+D[N-2] * (C[N-2]+2. * C[N-1]);
/* Calculation of polynomial
/* coefficients.
for (i=0; i<N-l; i++ ) {
B[i] = (Y[i+1] - Y[i]) / D[i] D[i] * (C[i+1]+2.*C[i]);
D[i] = (C[i+l]-C[i]) / D[i];
C[i] = 3*C[i];
}
C[N-1] = 3. * C[N-1];
D[N-1] = D[N-2];
}
else
{
B[0] = (Y[1]-Y[0]) / (X[1]-X[0]);
C[0] = D[0] = C[1] = D[1]=0.;
B[1] = B[0];
}
/* Set flag that coefficients are
crv->cfs.init_cfs - 1;
/* initialized.
END

*/

*/
*/

/*
/*

180
#include math_ext.h
********************************************************************************
* Name
BnrSrch
- finds the left knot of the interval.
*
* Synopsis
last_knot
- BnrSrch( u, crv );
*
unsigned 1ast_knot - The knot number from last call.
*
double u
- Argument value where function is
*
calculated.
*
CURVE *crv
- Pointer to the structure of curve arrays.
*
* Description
This function finds the number of left knot of interval where
*
spline is calculated.
*
* Returns
The number of left knot.
********************************************************************************
/* ___________________________Internal definition__________________________
#define
N
crv->nmb
/* Thc number of points on the curve.
#define
X
crv->arg
/* Array of curve arguments.
unsigned BnrSrch( double u, CURVE *crv )
BEGIN
unsigned i=crv->cfs.last_knot;
/* Number of the left knot of the
/* interval where spline is calculated.
unsigned j,k;
/* Counters.
if( i >= N-I ) i=0;
/* If u doesnt belong to the interval
/* calculated in the previous call, use
/* the binary search
if((u<X[i]) | | ( u > X[i+1] ) ) {
i = 0; j = N;
do {
/* The binary search.
k = (i+j) >> 1;
if(u < X[k] ) j - k;
else i = k;
} while( j > i+1);
}
return( crv->cfs.last_knot=i);
END

*
*
*
*
*
*
*
*
*
*
*
*
/
*/
*/
*/

*/
*/
*/
*/
*/
*/

*/

181
#include math_ext.h
/********************************************************************************
* Name
Seval
calculates value of function, first and second
*
*
derivative at the point u.
*
* Synopsis fun_value = Seval ( u, crv, drvl, drv2 );
*
*
double fun_value
Interpolation value at the point u.
*
*
double u
Argument value where function is
*
*
calculated.
*
*
CURVE *crv
Pointer to the structure of curve arrays.
*
*
double *drvl
Value of first derivative.
*
*
If value of first derivative isnt
*
*
necessary drvl can be equal to NULL.
*
*
double *drv2
Value of second derivative.
*
*
If value of second derivative isnt
*
*
necessary drv2 can be equal to NULL.
*
* Description This function calculates cubic spline or linear value at the
*
*
point u using interpolation coefficients B, C, D. Before using
*
*
interpolation coefficients must be initialized in the function
*
*
Spline or Line (depends on type of interpolation). Seval also
*
*
calculates values of first and second derivative at the point u.
*
*
Seval represents C-realization of Forsythe's fortran program.
*
* Returns Interpolation value at the point u.
*
********************************************************************************
/*_______________________Internal definitions_____________________
*/
#define
N
crv->nmb
/* The number of points on the curve.
*/
#define
X
crv->arg
/* Array of curve arguments.
*/
#define
Y
crv->fun
/* Array of curve function values.
*/
#define
B
crv->cfs.b
/* Arrays of spline coefficients.
*/
#define
C
crv->cfs.c
/*
*/
#define
D
crv->cfs.d
/*
*/
double SevaK double u, CURVE *crv, double *drvl, double *drv2 )
BEGIN
unsigned i;
/* Counter.
*/
double dx;
/* Initialize interpolation coefficients.
*/
if( !crv->cfs.init_cfs ) {
if( crv->type_inter = INTER_LINEAR ) Line( crv );
else if( crv->type_inter = INTER_SPLINE ) Spline( crv );
}
i = BnrSrch( u, crv );
/* Find the nearest knot.
*/
dx = u-X[i];
/* First derivative.
*/
if( drvl ) *drvl = B[i] + dx*( 2.*C[i] + 3.*dx*D[i]);
/* Second derivative.
*/
if( drv2 ) *drv2 = 2.*C[i] + 6.*dx*D[i];
/* Function value.
*/
return( Y[i] + dx*( B[i] + dx*( C[i]+dx*D[i] ) ) );
END

182
ANNEX 3
PROGRAM COMPLEX RUNGE.
/*********************************************************************************
*
*
*
RUNGE.H Header file for Runge-Kutta program complex.
*
*
*
**********************************************************************************
#ifndef
DEF_RUNGE
/* Prevent second reading of
*/
#define
DEF_RUNGE 1
/* these definitions.
*/
/*__________________________Definitions____________________________
*/
#define
BEGIN {
/* Left bracket
*/
#define
END
}
/* Right bracket.
*/
#define
NDE
5
/* Maximum number of differential
*/
/* equations.
*/
#define
MIN_ORD_DBL -308
/* Order of minimal positive doublc
*/
/* value.
*/
/*_____________________Definitions of data types________________________
*/
/* Type of working function calculating
*/
/* values of right parts of differential
*/
/* equations system for use in the
*/
/* function Runge.
*/
typedef void FUN_RUNGE( double, double [ ], double [] void * );
typedef struct
{
int
nmb_eq;
FUN_RUNGE *fun;
void
} RUNGE;

*prm;

/* Data structure for use in the


/* Runge-Kutta program complex.
/* Order of system of differential
/* equations.
/* Pointer to the working function
/* calculating values of right parts.
/* Pointer to the parameters structure
/* used in the working function.
/*

/*__________________________Function declarations __________________________


double
Compa
(double,
/* Calculates absolute value of mantissas
double,
/* difference.
double );
/*
int
Expon
(double );
/* Calculates decimal order of floating
/* point value.
void
RklStep (doubIe,
/* Integrates system of equations on the
double [ ],
/* single step.
RUNGE *,
/*
double,
/*
double *,
/*
double [ ] );
/*
void
Runge ( double,
/* Integrates system of differential

*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/
*/

183
double | |,
RUNGE *,
int,
double,
double | |,
double,
double );
#endif

/* equations.
/*
/*
/*
/*
/*
/*
/* Ends #ifndef DEF_RUNGE

*/
*/
*/
*/
*/
*/
*/
*/

#include <math.h>
#include runge.h
/ ********************************************************************************
* Name
Expon
calculates decimal order of a variable
*
*
with floating point.
*
* Synopsis
order = Expon( x );
*
*
int order
Decimal order of value x.
*
*
double x
Floating point variable.
*
* Description
This function is included in the Runge-Kutta program complex.
*
*
When x <> 0, Expon calculates decimal order of value x.
*
*
If x = 0, function returns order of minimal positive double
*
*
value.
*
* Returns
Decimal order of floating point variable for use in the
*
*
function Compa.
*
******************************************************************************** *
int Expon( double x )
BEGIN
if( x ) return( (int) (0.4342944819 * log(fabs(x))) + 1 );
return ( MIN_ORD_DBL );
END

184
#include math_ext.h
#include runge.h
/ ********************************************************************************
* Name Compa calculates the absolute value of mantissa
*
*
difference
*
* Synopsis double Compa( a, b, c );
*
*
double a,b,c Arbitrary values with floating point.
*
*
*
* Description This function is included in the Runge-Kutta program complex.
*
*
It calculates absolute value of mantissa difference between
*
*
the values a and b when their orders are equated to the
*
*
largest order of v9h1es a,b and c.
*
*
*
* Returns Absolute value of mantissa difference between variables a and b.
*
******************************************************************************** /
double Compa( double a, double b, double c )
BEGIN
int ae,be,ce;
/* Decimal orders of values a,b and c.
*/
ae = Expon( a );
/* Decimal order of value a.
*/
be = Expon( b );
/* Decimal order of value b.
*/
ce = Expon( c );
/* Decimal order of value c.
*/
ae= MAX( MAX(ae,be), ce );
/* Choose maximal order.
*/
return (fabs(a-b)/pow(10.,ae) );
/* Calculate mantissa difference.
*/
END

185
#include runge.h
/********************************************************************************
* Name
RklSlcp
- integrates system of equations on the
*
* Synopsis
void RklStep (x,y,f,h,xh,yh)
*
*
double x
- Start point of integration.
*
*
double y []
- The vector of initial values of sought functions.
*
*
RUNGE *f
- Pointer to the data structure with pointer
*
*
to the working function calculating values of right parts of
*
*
differential equations system and pointer to the parameters
*
*
structure used in the working function.
*
*
double h
- Step of integration.
*
*
double *xh
- Pointer to the stop integration point.
*
*
double yh[]
- Vector of solutions at the point *xh.
*
*
*
* Description
This function integrates system of differential equations on the single step.
*
*
RklStep is used in the Runge-Kutta program complex.
*
* Returns
Nothing
*
********************************************************************************
Void RklStep(double x, double y[], RUNGE *f, double h,
double *xh, double yh[] )
BEGIN
Int
k,j;
Double
w[NDE],
a[5],
z[NDE];
a[0] = a[l] = a[4] = 0.5*h;
a[2] = a[3] = h;
*xh - x;
for( k=0; k<f - > nmb_eq; k++ ) {yh[k] - w[k] - y[k]; }
for( j=0; j; j++ ) {
/* Right part

*/

/* for(j)

*/

(f->fun)( *xh, w, z, f->prm );


*xh - x + a[j];
for( k=0; k < f- > nmb_eq; k++ ) {
yh[k] += a[j+l]*z[k]/3.;
w[k] - y[k] + a[j]*z[k];
}
}
END

186
#include < stdio.h >
#include < process.h >
#include runge.h
/ ****************************************************************************
* Name Runge
integratessyslem of differential equations.
*
*
*
* Synopsis
void Runge( x,y, f, prim, xfin, yfin, eps, eta );
*
*
double x
Start point of integration.
*
*
double y[]
The vector of initial values of sought
*
*
functions.
*
*
RUNGE *f
Pointer to the data structure with pointer
*
*
to the working function calculating values of
*
*
right parts of differential equations system
*
*
and pointer to the parameters structure used
*
*
in the working function.
*
*
int prim
Start parameter. It is equal to 1 on the first
*
*
call. Then it is equal to 0.
*
*
double xfin
Stop integration point.
*
*
double yfin[]
Vector of solutions at the point xfin.
*
*
double eps
Relative error of Runge-Kutta method.
*
*
double eta
Absolute error of Runge-Kutta method.
*
*
* Description
This function is included in the Runge-Kutta program complex.
*
*
Runge integrates system of differential equations using
*
*
automatic choice of integration step.
*
* Returns
Nothing.
*
*****************************************************************************
void
Runge ( double x, double y[], RUNGE *f, int prim,
double xfin, double yfin[], double eps, double eta )
BEGIN
int
i,k;
/* Counters.
*/
int
out;
/* Flag for return.
*/
static int
ss;
/* Parameter for acceleration of
*/
/* calculation.
*/
static double
hs;
/* Start value of integration step.
*/
double
h;
/* Current step of integration.
*/
double
yl [NDE],
/* Arrays for estimation of error of
*/
y2[NDE],
/* sought functions
*/
y3[NDE];
/*
*/
double
xl, x2, x3;
/* Intermediate points where yl[],y2[],
*/
/* y3[] calculated.
*/
if( prim ) {
/* The first call.
*/
if(f->nmb_eq NDE ) {
/* Check the number of equations.
*/
/* Print message.
*/
prinlf(RUNGE: The number of equtions more than %i!\n",NDE );
exit (1);
/* Terminate program.
*/
}
h=xfin- x; ss = 0;
/* Calculate initial step.
*/
}

187
else { h = hs; out = 0; }

/* Restore step from the previous call.


/* Main Ioop

*/
*/

for( ;; ) {
if( ( ( x + 2.01*h - xfin ) > 0 ) = ( h > 0 ) ) {
hs = h;
/* Remember step to calculate nexl point.
out = 1;
/* Set flag of return.
h = 0.5*( xfin - x );

*/
*/

}
/* Calculation wilh double step.

*/

RklStep( x, y, f, 2.*h, &xl, yl );


lab:
RkIStep( x, y, f, h, &x2,y2 );
RklStep( x2,y2, f, h, &x3,y3 );
for( k*=0; k < f - >nmb_eq; k++ ) {
/* Check condition of approach accuracy.
if( Compai(yl[k], y3[k], eta ) eps ) {
h *=0.5; out=0; xl = x2;
for(i=0; i<f - > mb_eq; i++ ) yl [i] = y2[i];
goto lab;

*/

}
}
x = x3;
if( out ) {
for (k=0; k<f - > nmb_eq; k++ ) yfin[k] - y3[k];
return;
}
for(k=0; k<f->nmb_eq; k++) y[k] = y3[k];
/* AcceIaration of calculation
/* (increasing of step) after 5
/* repetitions.

*/
*/
*/

/* for ( ;; )

*/

if (s=5) {ss=0; h*=2.;}


else ++ss;
}
END

188
ANNEX 4
EXAMPLE FOR TESTING PROGRAM
Program complex SHAPE is developed in order to calculate specular
reflectors with point sources. It is applicable for calculation of two types of
reflectors; rotationally symmetric, and cylindrical. Both reflector types are
defined by their profile-section. Program SHAPE uses program RUNGE
(Annex 3) to calculate a reflector profile.
To debug programs similar to SHAPE we present the data on reflector
which reproduce the following illuminance disstribution

E ( x ) 4(1 ( x /(2 f )) 2 ) 2 (1 ( x / H ) 2 ),
where f and H are the parameters.
Obviuosly, the desired profile is a parabola with a focal distance f, while an
isotropic light source is disposed at the height H above an area being lit.
Appying formulas from Ch. 1, one can varify that paraboloid produces in the
ccntre the illuminance four times greater than a flat mirror.
Assuming 2f=H=lm and having normalized the function E(x) to a
maximim, we obtain

E ( x) (1 / 5){4(1 x 2 ) 2 (1 x 2 ) 1}.
We also set the following:
(1) the acceptance angle of reflector is equal to /2;
(2) the radius of lit circle is equal to l m;
(3) the reflectance factor is equal to 1.
We calculate luminous fluxes:
(1) the luminous flux falling onto reflector
/2

Fs

1sin d 1.0;
0

(2) the direct flux falling on the target area


3 / 2

1
2

Fd x(1 x )

dx 1 1 / 2 0.2928932188 ;

(3) the flux within prescribed curve E(x)

189
2

Fe (1 / 5) 4 x(1 x ) dx x(1 x 2 ) 3 / 2 dx
0

(1 / 5)(1 f d ) 0.2585786439.
From the balance of fluxes we get

Fs Fd / Fe 5.0 .
SAMPLE OF REFLECTOR CALCULATION FOR PARABOLIC CASE
1. INPUT DATA.
Luminous intensity distribution of the light source
N
1
2
3
4
5
6
7
8
9
10

FI [deg]
0.000
20.000
40.000
60.000
80.000
100.000
120.000
140.000
160.000
180.000

I [r.u.]
1.0000
1.0000
1.0000
1.0000
1.0000
1.0000
1.0000
1.0000
1.0000
1.0000

Illuminance distribution on the working plane


N
1
2
3
4
5
6

X [m]
0.0000
0.0204
0.0408
0.0612
0.0816
0.1020

E[r.u.]
5.00000
4.99605
4.98421
4.96458
4.93730
4.90257

190
N
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

X [m]
0.1224
0.1429
0.1633
0.1837
0.2041
0.2245
0.2449
0.2653
0.2857
0.3061
0.3265
0.3469
0.3673
0.3878
0.4082
0.4286
0.4490
0.4694
0.4898
0.5102
0.5306
0.5510
0.5714
0.5918
0.6122
0.6327
0.6531
0.6735
0.6939
0.7143
0.7347
0.7551
0.7755
0.7959
0.8163
0.8367
0.8571

E[r.u.]
4.86062
4.81175
4.75630
4.69463
4.62715
4.55429
4.47649
4.39423
4.30797
4.21818
4.12535
4.02995
3.93242
3.83321
3.73276
3.63145
3.52969
3.42782
3.32618
3.22508
3.12479
3.02557
2.92766
2.83124
2.73651
2.64361
2.55267
2.46381
2.37712
2.29266
2.21049
2.13066
2.05319
1.97809
1.90535
1.83499
1.76696

191
N
44
45
46
47
48
49
50

X [m]
0.8776
0.8980
0.9184
0.9388
0.9592
0.9796
1.0000

Reflectance
Initial angle
Initial point
Initial radius
Final angle
Final point
Height of the light center

1.0
0.0
0.0
500.0
90.0
1.0
1.0

E[r.u.]
1.70126
1.63784
1.57667
1.51771
1.46090
1.40620
1.35355

[deg]
[m ]
[mm]
[deg]
[m]
[m]

2. POLAR COORDINATES OF REFLECTOR.


REFLECTED RAY WITH SYMMETRY AXIS.
N

FI [deg]

1
2
3
4
5
6
7
8
9
10
11
12
13

0.00
2.00
4.00
6.00
8.00
10.00
12.00
14.00
16.00
18.00
20.00
22.00
24.00

ALFA
[deg]
0.00000
0.00000
-0.00000
-0.00000
-0.00000
-0.00000
-0.00000
-0.00000
-0.00000
-0.00000
-0.00000
-0.00000
-0.00000

ANGLE

OF

R [mm ]

REL. ERROR [%]

500.00000
500.15234
500.60973
501.37329
502.44488
503.82713
505.52345
507.53802
509.87586
512.54282
515.54560
518.89185
522.59015

0.000000
0.000000
-0.000000
0.000000
0.000000
-0.000001
-0.000000
-0.000001
-0.000001
0.000001
-0.000000
-0.000000
0.000001

192

N
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

FI
[deg ]
26.00
28.00
30.00
32.00
34.00
36.00
38.00
40.00
42.00
44.00
46.00
48.00
50.00
52.00
54.00
56.00
58.00
60.00
62.00
64.00
66.00
68.00
70.00
72.00
74.00
76.00
78.00
80.00
82.00
84.00
86.00
88.00
90.00

ALFA
[deg]
-0.00000
-0.00000
-0.00000
-0.00000
-0.00000
-0.00000
-0.00000
-0.00000
-0.00000
-0.00000
-0.00000
-0.00000
-0.00000
-0.00000
-0.00000
-0.00000
-0.00000
-0.00000
-0.00001
-0.00001
-0.00001
-0.00001
-0.00001
-0.00001
-0.00001
-0.00002
-0.00000
-0.00000
-0.00000
-0.00000
-0.00001
-0.00001
-0.00001

R
[mm]
526.65006
531.08223
535.89838
541.11146
546.73562
552.78640
559.28075
566.23717
573.67580
581.61860
590.08943
599.11429
608.72142
618.94154
629.80809
641.35746
653.62926
666.66667
680.51674
695.23085
710.86511
727.48087
745.14530
763.93202
783.92185
805.20361
827.87509
852.04409
877.82972
905.36372
934.79219
966.27740
1000.00000

REL. ERROR
[%]
-0.000000
0.000001
-0.000001
0.000000
-0.000001
-0.000001
-0.000000
0.000001
0.000000
0.000001
-0.000001
-0.000000
0.000001
0.000000
-0.000000
0.000000
-0.000000
0.000000
-0.000000
-0.000001
-0.000000
0.000000
0.000000
-0.000000
-0.000000
-0.000001
-0.000000
-0.000001
0.000000
-0.000000
-0.000000
-0.000001
0.000000

Potrebbero piacerti anche